Adaptive variational quantum computing approaches for Green’s functions and nonlinear susceptibilities

Martin Mootz [email protected] Ames National Laboratory, U.S. Department of Energy, Ames, Iowa 50011, USA    Thomas Iadecola Ames National Laboratory, U.S. Department of Energy, Ames, Iowa 50011, USA    Yong-Xin Yao [email protected] Ames National Laboratory, U.S. Department of Energy, Ames, Iowa 50011, USA
Abstract

We present and benchmark quantum computing approaches for calculating real-time single-particle Green’s functions and nonlinear susceptibilities of Hamiltonian systems. The approaches leverage adaptive variational quantum algorithms for state preparation and propagation. Using automatically generated compact circuits, the dynamical evolution is performed over sufficiently long times to achieve adequate frequency resolution of the response functions. We showcase accurate Green’s function calculations using a statevector simulator on classical hardware for Fermi-Hubbard chains of 4 and 6 sites, with maximal ansatz circuit depths of 65 and 424 layers, respectively, and for the molecule LiH with a maximal ansatz circuit depth of 81 layers. Additionally, we consider an antiferromagnetic quantum spin-1 model that incorporates the Dzyaloshinskii-Moriya interaction to illustrate calculations of the third-order nonlinear susceptibilities, which can be measured in two-dimensional coherent spectroscopy experiments. These results demonstrate that real-time approaches using adaptive parameterized circuits to evaluate linear and nonlinear response functions can be feasible with near-term quantum processors.

\SectionNumbersOn\alsoaffiliation

Department of Physics and Astronomy, Iowa State University, Ames, Iowa 50011, USA \alsoaffiliationDepartment of Physics and Astronomy, Iowa State University, Ames, Iowa 50011, USA

1 Introduction

In the pursuit of understanding and predicting the properties of quantum systems, the development of efficient computational methodologies has become imperative. Among the plethora of tools available, the calculation of Green’s functions 1, 2 and related high-order nonlinear susceptibilities 3, 4, 5 provide crucial insights into dynamical responses, correlation effects, and excitations within complex quantum materials and molecular systems. Specifically, many-particle Green’s functions describe the correlations and interactions among particles in quantum systems via correlation functions or response functions and also facilitate the development of advanced theoretical frameworks for describing complex quantum phenomena such as the GW method 6 and dynamical mean-field theory 7, 8, 9. These methods rely heavily on Green’s functions to investigate the rich phenomena exhibited by complex materials, including superconductivity 10, magnetism 11, and topological phases 12. Central to the analysis of Green’s functions is the concept of spectral functions, which provide a direct link between theoretical calculations and measurable quantities in experiments, such as photoemission spectra 13, optical conductivity 14, or neutron scattering cross-sections 15. In contrast, nonlinear susceptibilities 3, 4, 5 offer a systematic approach to analyze the nonlinear optical response of quantum systems to multiple interacting laser pulses. Specifically, two-dimensional coherent spectroscopy (2DCS) 16, 17, 18, 19, 20, 21, 22, 23, 24, 25, 26, 27, 28, 29, 30, 31 directly probes nonlinear susceptibilities by measuring the time-dependent coherent response to two laser pulses.

Traditional methods for computing Green’s functions and nonlinear susceptibilities, such as exact diagonalization 32, density-matrix renormalization group 33, or quantum Monte Carlo simulation 34, often face challenges in simulating large-scale systems at low temperatures. Nevertheless, in recent years the advent of quantum computing 35 has opened up new avenues for tackling these challenges. These simulations hold promise in providing valuable insights into the properties of complex quantum materials that go beyond what is possible with classical computers 36. So far, these simulations have been conducted on noisy intermediate-scale quantum devices 37, 38, 39, 40, 41, which are restricted by the available number of qubits and by hardware noise. Several methods have been proposed to calculate Green’s functions and multi-time correlation functions 42 in frequency space using quantum-phase estimation 43, 44, 45, 46, and in the time domain using the Suzuki-Trotter decomposition of the time evolution operator 47, 48, 49, 50. However, these techniques generally demand deep quantum circuits and large numbers of controlled operations, making them impractical for noisy intermediate-scale quantum hardware. To address these issues, several quantum circuit compression algorithms have been introduced. These algorithms calculate Green’s functions in the frequency domain using the variational quantum eigensolver 51, 52, or in the time domain by simplifying the time evolution unitary operation using the coupled cluster Green’s function method 53, Cartan decomposition 54, or variational quantum dynamics simulation algorithms 55, 51, 56. The latter involves the preparation of a variational ansatz state to approximate the exact time-evolved state of the system. The equation of motion governing the time evolution of the variational parameters is derived based on the McLachlan variational principle 55, 57, 58, which aims to minimize the distance between the variational state and the exact time-evolved state.

Nonetheless, the efficacy of variational quantum dynamics simulations crucially depends on the flexibility of the variational ansatz to faithfully represent the dynamical states of the system. Using Hamiltonian variational ansatz (HVA) 51, 56, the accuracy of the real-time Green’s function can be improved by increasing the number of layers, i.e., the depth of the ansatz. However, a large number of layers can be required to precisely describe the quantum state dynamics over sufficiently long times to achieve adequate resolution of correlation function in frequency space, leading to large circuit depths. Adaptive variational algorithms, such as the adaptive variational quantum dynamics simulation (AVQDS) 59 approach, can automatically generate problem-specific ansätze with reduced complexity compared to general problem-agnostic fixed ansätze 59, 60, 61. In AVQDS, the McLachlan distance, a metric to measure the difference between the variational and exact state evolutions, is maintained below a predefined threshold throughout the time evolution by adaptively adding new parameterized unitaries chosen from a predetermined operator pool to the variational ansatz. This method has been applied in ref 62 to calculate the single-particle Green’s function using a Hadamard-test circuit to compute state overlaps. The result in ref 62 demonstrates that less quantum resources are required compared to variational quantum dynamics simulations with HVA, even as the accuracy over long simulation times is improved. However, the Hadamard-test circuit involves controlled multi-qubit rotation gates for state propagation, which can substantially increase the circuit complexity.

In this work, we employ the AVQDS algorithm to calculate single-particle Green’s functions and nonlinear susceptibilities along the real-time axis. In contrast to the controlled-unitaries-required (CUR) approach for overlap testing 62, we adopt the method presented in ref 51, which is controlled-unitaries-liberated (CUL) and applies to generic reference states besides the ground state. Instead, the method requires the parametrized circuit to directly simulate the dynamics of two quantum states, as opposed to one in the CUR method 56, 62. The CUL calculation is conveniently performed on circuits with an ancilla qubit plus a few controlled Pauli gates to mix the two quantum states, followed by state evolution. Although the simultaneous propagation of two states instead of one demands more flexible circuits, it allows one to leverage the variational degrees of freedom already existing in the initial parameterized state (e.g., the ground state) for variational state propagation using more compact circuits. We apply the CUL approach to calculate the single-particle Green’s functions of fermionic models including molecules using the AVQDS approach instead of variational quantum dynamics simulations with HVA as in ref 51. Additionally, we extend this method to calculate nonlinear susceptibilities of quantum spin models, which depend on two times and thus requires double application of the AVQDS approach to evolve the quantum states. The presented AVQDS simulations are performed using statevector simulators on classical hardware to validate our methods using well-established classical approaches. These simulations also allow us to estimate the quantum resources required for specific applications. Such estimates, typically characterized by metrics like the number of CNOT gates in a circuit for near-term quantum computing, provide critical insights before transitioning to actual quantum hardware

To illustrate the calculation of Green’s functions with the AVQDS approach, we study the single-particle Green’s functions of Fermi-Hubbard chains with N=4𝑁4N=4italic_N = 4 and N=6𝑁6N=6italic_N = 6 sites, corresponding respectively to 8 and 12 qubits. Additionally, we determine the Green’s function of the molecule LiH where the encoding of the Hamiltonian requires 10 qubits. Using statevector simulations, we calculate the Green’s function dynamics in momentum space and discuss the required quantum resources for near-term applications measured by number of CNOT gates and circuit layers. Additionally, we compute the spectral function and compare the results obtained with the circuit simulator to exact results derived from the Lehmann representation of the Green’s function. To demonstrate the validity of our method to calculate high-order susceptibilities measured in two-dimensional coherent spectroscopy experiments, we consider an antiferromagnetic quantum high-spin model that incorporates the Heisenberg exchange and Dzyaloshinskii-Moriya interactions 63. Specifically, we calculate the third-order nonlinear susceptibility in the two-dimensional (2D) time and frequency domain for a two-site spin-1 model and compare with exact numerical results.

The paper is organized as follows. The AVQDS algorithm is briefly discussed in section 2. In section 3 we present the algorithm for calculating single-particle Green’s functions and the method for simulating nonlinear susceptibilities. Section 4 presents various applications of these methods, demonstrating their utility in practical scenarios. Specifically, in section 4.1 we calculate the Green’s function and corresponding spectral functions for N=4𝑁4N=4italic_N = 4 and N=6𝑁6N=6italic_N = 6 Fermi-Hubbard chains and benchmark the performance of the AVQDS approach. In section 4.2, we present the Green’s function calculation for the molecule LiH. In section 4.3, we demonstrate the algorithm for simulating nonlinear susceptibilities by calculating the third-order susceptibility of a two-site spin-1 model. Finally, we conclude in section 5 with a summary of findings and an outlook.

2 AVQDS algorithm

In this work, we employ the AVQDS algorithm as a quantum computing approach to calculate Green’s functions and high-order nonlinear susceptibilities. Below, we briefly discuss the key points of AVQDS, while a more detailed discussion can be found in refs 59, 63. The algorithm can be naturally extended from real-time dynamics to the adaptive variational quantum imaginary-time evolution (AVQITE) approach 60, which is adopted for ground state preparation in the following calculations. We begin by considering a quantum system initially in the pure state |ΨketΨ|\Psi\rangle| roman_Ψ ⟩, whose quantum dynamics is governed by a generally time-independent Hamiltonian ^^\hat{\mathcal{H}}over^ start_ARG caligraphic_H end_ARG. The dynamics of the system’s density matrix ρ=|ΨΨ|𝜌ketΨbraΨ\rho=|\Psi\rangle\langle\Psi|italic_ρ = | roman_Ψ ⟩ ⟨ roman_Ψ | is then determined by the von Neumann equation:

dρdt=i[^,ρ].d𝜌d𝑡i^𝜌\displaystyle\frac{\mathrm{d}\rho}{\mathrm{d}t}=-\mathrm{i}\left[\hat{\mathcal% {H}},\rho\right]\,.divide start_ARG roman_d italic_ρ end_ARG start_ARG roman_d italic_t end_ARG = - roman_i [ over^ start_ARG caligraphic_H end_ARG , italic_ρ ] . (1)

In variational quantum dynamics simulations, the state |ΨketΨ|\Psi\rangle| roman_Ψ ⟩ is parameterized as |Ψ[𝜽]ketΨdelimited-[]𝜽|\Psi[\boldsymbol{\theta}]\rangle| roman_Ψ [ bold_italic_θ ] ⟩, with 𝜽(t)𝜽𝑡\boldsymbol{\theta}(t)bold_italic_θ ( italic_t ) being a real-valued time-dependent variational parameter vector of dimension Nθsubscript𝑁𝜃N_{\theta}italic_N start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT 55. The evolution of 𝜽𝜽\boldsymbol{\theta}bold_italic_θ is determined by equations of motion derived from the McLachlan variational principle 64, which minimizes the squared McLachlan distance 2superscript2\mathcal{L}^{2}caligraphic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT as the Frobenius norm (𝒪Tr[𝒪𝒪]norm𝒪tracesuperscript𝒪𝒪\|\mathcal{O}\|\equiv\Tr[\mathcal{O}^{\dagger}\mathcal{O}]∥ caligraphic_O ∥ ≡ roman_Tr [ caligraphic_O start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT caligraphic_O ]) between exact and variational evolving states:

2superscript2\displaystyle\mathcal{L}^{2}caligraphic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT μρ[𝜽]θμθ˙μ+i[^,ρ]2absentsuperscriptnormsubscript𝜇𝜌delimited-[]𝜽subscript𝜃𝜇subscript˙𝜃𝜇i^𝜌2\displaystyle\equiv\bigg{\|}\sum_{\mu}\frac{\partial\rho[\boldsymbol{\theta}]}% {\partial\theta_{\mu}}\dot{\theta}_{\mu}+\mathrm{i}\left[\hat{\mathcal{H}},% \rho\right]\bigg{\|}^{2}≡ ∥ ∑ start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT divide start_ARG ∂ italic_ρ [ bold_italic_θ ] end_ARG start_ARG ∂ italic_θ start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT end_ARG over˙ start_ARG italic_θ end_ARG start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT + roman_i [ over^ start_ARG caligraphic_H end_ARG , italic_ρ ] ∥ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT
=μνMμνθ˙μθ˙ν2μVμθ˙μ+2var𝜽[^],absentsubscript𝜇𝜈subscript𝑀𝜇𝜈subscript˙𝜃𝜇subscript˙𝜃𝜈2subscript𝜇subscript𝑉𝜇subscript˙𝜃𝜇2subscriptvar𝜽delimited-[]^\displaystyle=\sum_{\mu\nu}M_{\mu\nu}\dot{\theta}_{\mu}\dot{\theta}_{\nu}-2% \sum_{\mu}V_{\mu}\dot{\theta}_{\mu}+2\,\mathrm{var}_{\boldsymbol{\theta}}[\hat% {\mathcal{H}}]\,,= ∑ start_POSTSUBSCRIPT italic_μ italic_ν end_POSTSUBSCRIPT italic_M start_POSTSUBSCRIPT italic_μ italic_ν end_POSTSUBSCRIPT over˙ start_ARG italic_θ end_ARG start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT over˙ start_ARG italic_θ end_ARG start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT - 2 ∑ start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT italic_V start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT over˙ start_ARG italic_θ end_ARG start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT + 2 roman_var start_POSTSUBSCRIPT bold_italic_θ end_POSTSUBSCRIPT [ over^ start_ARG caligraphic_H end_ARG ] , (2)

where the Nθ×Nθsubscript𝑁𝜃subscript𝑁𝜃N_{\theta}\times N_{\theta}italic_N start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT × italic_N start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT matrix M𝑀Mitalic_M and vector V𝑉Vitalic_V of dimension Nθsubscript𝑁𝜃N_{\theta}italic_N start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT are defined as

Mμ,νsubscript𝑀𝜇𝜈absent\displaystyle M_{\mu,\nu}\equivitalic_M start_POSTSUBSCRIPT italic_μ , italic_ν end_POSTSUBSCRIPT ≡ Tr[ρ[𝜽]θμρ[𝜽]θν]=2Re[Ψ[𝜽]|θμ|Ψ[𝜽]θν\displaystyle\,\mathrm{Tr}\left[\frac{\partial\rho[\boldsymbol{\theta}]}{% \partial\theta_{\mu}}\frac{\partial\rho[\boldsymbol{\theta}]}{\partial\theta_{% \nu}}\right]=2\,\mathrm{Re}\left[\frac{\partial\langle\Psi[\boldsymbol{\theta}% ]|}{\partial\theta_{\mu}}\frac{\partial|\Psi[\boldsymbol{\theta}]\rangle}{% \partial\theta_{\nu}}\right.roman_Tr [ divide start_ARG ∂ italic_ρ [ bold_italic_θ ] end_ARG start_ARG ∂ italic_θ start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT end_ARG divide start_ARG ∂ italic_ρ [ bold_italic_θ ] end_ARG start_ARG ∂ italic_θ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG ] = 2 roman_Re [ divide start_ARG ∂ ⟨ roman_Ψ [ bold_italic_θ ] | end_ARG start_ARG ∂ italic_θ start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT end_ARG divide start_ARG ∂ | roman_Ψ [ bold_italic_θ ] ⟩ end_ARG start_ARG ∂ italic_θ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG
+Ψ[𝜽]|θμ|Ψ[𝜽]Ψ[𝜽]|θν|Ψ[𝜽]],\displaystyle\left.+\frac{\partial\langle\Psi[\boldsymbol{\theta}]|}{\partial% \theta_{\mu}}|\Psi[\boldsymbol{\theta}]\rangle\frac{\partial\langle\Psi[% \boldsymbol{\theta}]|}{\partial\theta_{\nu}}|\Psi[\boldsymbol{\theta}]\rangle% \right]\,,+ divide start_ARG ∂ ⟨ roman_Ψ [ bold_italic_θ ] | end_ARG start_ARG ∂ italic_θ start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT end_ARG | roman_Ψ [ bold_italic_θ ] ⟩ divide start_ARG ∂ ⟨ roman_Ψ [ bold_italic_θ ] | end_ARG start_ARG ∂ italic_θ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG | roman_Ψ [ bold_italic_θ ] ⟩ ] , (3)
Vμ=subscript𝑉𝜇absent\displaystyle V_{\mu}=italic_V start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT =  2Im[Ψ[𝜽]|θμ^|Ψ[𝜽]+Ψ[𝜽]||Ψ[𝜽]θμ^𝜽],2Imdelimited-[]braΨdelimited-[]𝜽subscript𝜃𝜇^ketΨdelimited-[]𝜽braΨdelimited-[]𝜽ketΨdelimited-[]𝜽subscript𝜃𝜇subscriptdelimited-⟨⟩^𝜽\displaystyle\,2\,\mathrm{Im}\left[\frac{\partial\langle\Psi[\boldsymbol{% \theta}]|}{\partial\theta_{\mu}}\hat{\mathcal{H}}|\Psi[\boldsymbol{\theta}]% \rangle+\langle\Psi[\boldsymbol{\theta}]|\frac{\partial|\Psi[\boldsymbol{% \theta}]\rangle}{\partial\theta_{\mu}}\langle\hat{\mathcal{H}}\rangle_{% \boldsymbol{\theta}}\right]\,,2 roman_Im [ divide start_ARG ∂ ⟨ roman_Ψ [ bold_italic_θ ] | end_ARG start_ARG ∂ italic_θ start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT end_ARG over^ start_ARG caligraphic_H end_ARG | roman_Ψ [ bold_italic_θ ] ⟩ + ⟨ roman_Ψ [ bold_italic_θ ] | divide start_ARG ∂ | roman_Ψ [ bold_italic_θ ] ⟩ end_ARG start_ARG ∂ italic_θ start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT end_ARG ⟨ over^ start_ARG caligraphic_H end_ARG ⟩ start_POSTSUBSCRIPT bold_italic_θ end_POSTSUBSCRIPT ] , (4)

with ^𝜽=Ψ[𝜽]|^|Ψ[𝜽]subscriptdelimited-⟨⟩^𝜽quantum-operator-productΨdelimited-[]𝜽^Ψdelimited-[]𝜽\langle\hat{\mathcal{H}}\rangle_{\boldsymbol{\theta}}=\langle\Psi[\boldsymbol{% \theta}]|\hat{\mathcal{H}}|\Psi[\boldsymbol{\theta}]\rangle⟨ over^ start_ARG caligraphic_H end_ARG ⟩ start_POSTSUBSCRIPT bold_italic_θ end_POSTSUBSCRIPT = ⟨ roman_Ψ [ bold_italic_θ ] | over^ start_ARG caligraphic_H end_ARG | roman_Ψ [ bold_italic_θ ] ⟩. The real symmetric matrix M𝑀Mitalic_M is directly related to the quantum Fisher information matrix 65, with the second term within the bracket accounting for the global phase contribution 55. In the last term of eq 2, var𝜽[^]=^2𝜽^𝜽2subscriptvar𝜽delimited-[]^subscriptdelimited-⟨⟩superscript^2𝜽subscriptsuperscriptdelimited-⟨⟩^2𝜽\mathrm{var}_{\boldsymbol{\theta}}[\hat{\mathcal{H}}]=\langle\hat{\mathcal{H}}% ^{2}\rangle_{\boldsymbol{\theta}}-\langle\hat{\mathcal{H}}\rangle^{2}_{% \boldsymbol{\theta}}roman_var start_POSTSUBSCRIPT bold_italic_θ end_POSTSUBSCRIPT [ over^ start_ARG caligraphic_H end_ARG ] = ⟨ over^ start_ARG caligraphic_H end_ARG start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT bold_italic_θ end_POSTSUBSCRIPT - ⟨ over^ start_ARG caligraphic_H end_ARG ⟩ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_italic_θ end_POSTSUBSCRIPT is the variance of ^^\hat{\mathcal{H}}over^ start_ARG caligraphic_H end_ARG in the variational state |Ψ[𝜽]ketΨdelimited-[]𝜽|\Psi[\boldsymbol{\theta}]\rangle| roman_Ψ [ bold_italic_θ ] ⟩. The minimization of 2superscript2\mathcal{L}^{2}caligraphic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT with respect to {θ˙μ}subscript˙𝜃𝜇\{\dot{\theta}_{\mu}\}{ over˙ start_ARG italic_θ end_ARG start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT } leads to the equation:

νMμνθ˙ν=Vμ,subscript𝜈subscript𝑀𝜇𝜈subscript˙𝜃𝜈subscript𝑉𝜇\displaystyle\sum_{\nu}M_{\mu\nu}\dot{\theta}_{\nu}=V_{\mu}\,,∑ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT italic_M start_POSTSUBSCRIPT italic_μ italic_ν end_POSTSUBSCRIPT over˙ start_ARG italic_θ end_ARG start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT = italic_V start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT , (5)

which governs the dynamics of the variational parameters. The optimized McLachlan distance for the variational ansatz |Ψ[𝜽]ketΨdelimited-[]𝜽|\Psi[\boldsymbol{\theta}]\rangle| roman_Ψ [ bold_italic_θ ] ⟩,

L2=2var𝜽[^]μVμθ˙μ,superscript𝐿22subscriptvar𝜽delimited-[]^subscript𝜇subscript𝑉𝜇subscript˙𝜃𝜇\displaystyle L^{2}=2\,\mathrm{var}_{\boldsymbol{\theta}}[\hat{\mathcal{H}}]-% \sum_{\mu}V_{\mu}\dot{\theta}_{\mu}\,,italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = 2 roman_var start_POSTSUBSCRIPT bold_italic_θ end_POSTSUBSCRIPT [ over^ start_ARG caligraphic_H end_ARG ] - ∑ start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT italic_V start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT over˙ start_ARG italic_θ end_ARG start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT , (6)

is a metric quantifying the accuracy of the variational dynamics.

In the AVQDS framework, the ansatz takes a pseudo-Trotter form:

|Ψ[𝜽]=μ=1Nθeiθμ𝒜^μ|φ0.ketΨdelimited-[]𝜽superscriptsubscriptproduct𝜇1subscript𝑁𝜃superscripteisubscript𝜃𝜇subscript^𝒜𝜇ketsubscript𝜑0\displaystyle|\Psi[\boldsymbol{\theta}]\rangle=\prod_{\mu=1}^{N_{\theta}}% \mathrm{e}^{-\mathrm{i}\theta_{\mu}\hat{\mathcal{A}}_{\mu}}|\varphi_{0}\rangle\,.| roman_Ψ [ bold_italic_θ ] ⟩ = ∏ start_POSTSUBSCRIPT italic_μ = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT end_POSTSUPERSCRIPT roman_e start_POSTSUPERSCRIPT - roman_i italic_θ start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT over^ start_ARG caligraphic_A end_ARG start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT end_POSTSUPERSCRIPT | italic_φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ⟩ . (7)

Here, 𝒜^μsubscript^𝒜𝜇\hat{\mathcal{A}}_{\mu}over^ start_ARG caligraphic_A end_ARG start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT (μ=1,,Nθ)\mu=1,\cdots,N_{\theta})italic_μ = 1 , ⋯ , italic_N start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT ) represent Hermitian generators, while |φ0ketsubscript𝜑0\ket{\varphi_{0}}| start_ARG italic_φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ⟩ denotes the reference state. To maintain the McLachlan distance L2superscript𝐿2L^{2}italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT below a given threshold Lcut2subscriptsuperscript𝐿2cutL^{2}_{\mathrm{cut}}italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_cut end_POSTSUBSCRIPT throughout the time evolution, operators are dynamically selected from a predefined pool of generators, which expands the set of unitaries in the ansatz eq 7. The selection criterion is to add to the ansatz the operator that maximally reduces L2superscript𝐿2L^{2}italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, so as to achieve accurate dynamics simulations with the minimal number of unitaries. In the initial proposal of the AVQDS approach 59, operators were chosen one by one; here we adopt a modified strategy. Specifically, multiple operators, rather than a single one, are simultaneously chosen according to the L2superscript𝐿2L^{2}italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT-criterion at each iteration, subject to the constraint that these operators act on disjoint sets of qubits. This approach takes the spatial compactness of the circuits into account, and significantly reduces the circuits depth with only a small change in the total number of CNOT gates, The ansatz expansion is done by first calculating the McLachlan distance Lν2subscriptsuperscript𝐿2𝜈L^{2}_{\nu}italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT for a new variational ansatz of the form eiθν𝒜^ν|Ψ[𝜽]superscript𝑒𝑖subscript𝜃𝜈subscript^𝒜𝜈ketΨdelimited-[]𝜽e^{-i\theta_{\nu}\hat{\mathcal{A}}_{\nu}}|\Psi[\boldsymbol{\theta}]\rangleitalic_e start_POSTSUPERSCRIPT - italic_i italic_θ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT over^ start_ARG caligraphic_A end_ARG start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_POSTSUPERSCRIPT | roman_Ψ [ bold_italic_θ ] ⟩ for all generators 𝒜^νsubscript^𝒜𝜈\hat{\mathcal{A}}_{\nu}over^ start_ARG caligraphic_A end_ARG start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT from a predefined operator pool of size Npsubscript𝑁pN_{\mathrm{p}}italic_N start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT. The resulting Lν2subscriptsuperscript𝐿2𝜈L^{2}_{\nu}italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT are ordered in ascending order. In the first step, the operator 𝒜^νsubscript^𝒜𝜈\hat{\mathcal{A}}_{\nu}over^ start_ARG caligraphic_A end_ARG start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT leading to the smallest Lν2subscriptsuperscript𝐿2𝜈L^{2}_{\nu}italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT is added to the ansatz eq 7, which increases NθNθ+1subscript𝑁𝜃subscript𝑁𝜃1N_{\theta}\rightarrow N_{\theta}+1italic_N start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT → italic_N start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT + 1 in eq 7. In the next step, the operator with the next smallest Lν2subscriptsuperscript𝐿2𝜈L^{2}_{\nu}italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT is appended, provided it acts on different qubits than the first operator. This process continues by adding operators with successively larger or equal Lν2subscriptsuperscript𝐿2𝜈L^{2}_{\nu}italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT that have disjoint supports from all previously appended operators during this iteration, until all qubits are covered or no more suitable operators are found. Note that this method does not introduce additional Lν2subscriptsuperscript𝐿2𝜈L^{2}_{\nu}italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT measurements, as L2superscript𝐿2L^{2}italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT for each pool operator is measured once per iteration. The variational parameters θνsubscript𝜃𝜈\theta_{\nu}italic_θ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT are set to zero for all operators appended to the ansatz eq 7 during the iteration step. This does not alter the ansatz state but can modify the McLachlan distance L2superscript𝐿2L^{2}italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT due to a non-zero derivative with respect to θνsubscript𝜃𝜈\theta_{\nu}italic_θ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT. Finally, L2superscript𝐿2L^{2}italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT is calculated for the new ansatz and compared with Lcut2subscriptsuperscript𝐿2cutL^{2}_{\mathrm{cut}}italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_cut end_POSTSUBSCRIPT. This adaptive procedure is repeated until the McLachlan distance L2superscript𝐿2L^{2}italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT of the new ansatz falls below Lcut2subscriptsuperscript𝐿2cutL^{2}_{\mathrm{cut}}italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_cut end_POSTSUBSCRIPT.

Following the adaptive ansatz expansion procedure, the variational parameters are evolved in time according to eq 5, which corresponds to a system of linear ordinary differential equations. In the simulations performed in this work, we solve eq 5 using a fourth-order Runge-Kutta method. This method provides a higher accuracy with a truncation error of order (δt)4superscript𝛿𝑡4(\delta t)^{4}( italic_δ italic_t ) start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT, compared to the Euler method with a truncation error of order (δt)2superscript𝛿𝑡2(\delta t)^{2}( italic_δ italic_t ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. While this method involves computational overhead due to each time step requiring four micro-time steps to update the variational parameters (compared to Euler’s method, which updates parameters in a single step) it still substantially reduces the total number of time steps and circuit complexity, as demonstrated in ref 63. Additionally, the time step δt𝛿𝑡\delta titalic_δ italic_t in the AVQDS approach is dynamically adjusted to ensure that max0μ<Nθ|δθμ|subscript0𝜇subscript𝑁𝜃𝛿subscript𝜃𝜇\max_{0\leq\mu<N_{\theta}}|\delta\theta_{\mu}|roman_max start_POSTSUBSCRIPT 0 ≤ italic_μ < italic_N start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT end_POSTSUBSCRIPT | italic_δ italic_θ start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT | remains below a predefined maximal step size δθmax𝛿subscript𝜃max\delta\theta_{\mathrm{max}}italic_δ italic_θ start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT, where δθμ=θμ(t+δt)θμ(t)𝛿subscript𝜃𝜇subscript𝜃𝜇𝑡𝛿𝑡subscript𝜃𝜇𝑡\delta\theta_{\mu}=\theta_{\mu}(t+\delta t)-\theta_{\mu}(t)italic_δ italic_θ start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT = italic_θ start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT ( italic_t + italic_δ italic_t ) - italic_θ start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT ( italic_t ). In the simulations we set δθmax=0.01𝛿subscript𝜃max0.01\delta\theta_{\mathrm{max}}=0.01italic_δ italic_θ start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = 0.01. Furthermore, to mitigate potential numerical issues associated with inverting the matrix M𝑀Mitalic_M in eq 5, we employ the Tikhonov regularization approach, where a small number of δ=106𝛿superscript106\delta=10^{-6}italic_δ = 10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT is added to the diagonal of M𝑀Mitalic_M. This stabilizes the matrix inversion process, especially when M𝑀Mitalic_M possesses a high condition number.

To estimate the cost of measuring M𝑀Mitalic_M, V𝑉Vitalic_V, ^θsubscriptdelimited-⟨⟩^𝜃\langle\hat{\mathcal{H}}\rangle_{\theta}⟨ over^ start_ARG caligraphic_H end_ARG ⟩ start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT, and ^2θsubscriptdelimited-⟨⟩superscript^2𝜃\langle\hat{\mathcal{H}}^{2}\rangle_{\theta}⟨ over^ start_ARG caligraphic_H end_ARG start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT on quantum hardware, we assume that the wavefunction ansatz has Nθsubscript𝑁𝜃N_{\theta}italic_N start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT variational parameters. In addition, we assume that the Hamiltonian consists of NHsubscript𝑁HN_{\mathrm{H}}italic_N start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT Pauli strings, which also corresponds to the number of generators in the operator pool when the Hamiltonian operator pool is chosen, as is the case in the Green’s function calculation for the Fermi-Hubbard chains and the LiH molecule in sections 4.1 and 4.2. As shown in ref 59, the upper bound on the number of distinct direct measurement circuits and generalized Hadamard test circuits is then given by (NH+2)Nθ+NH+NH2subscript𝑁H2subscript𝑁𝜃subscript𝑁Hsuperscriptsubscript𝑁H2(N_{\mathrm{H}}+2)N_{\theta}+N_{\mathrm{H}}+N_{\mathrm{H}}^{2}( italic_N start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT + 2 ) italic_N start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT + italic_N start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT + italic_N start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT and NH(Nθ1)+Nθ(Nθ1)/2subscript𝑁Hsubscript𝑁𝜃1subscript𝑁𝜃subscript𝑁𝜃12N_{\mathrm{H}}(N_{\theta}-1)+N_{\theta}(N_{\theta}-1)/2italic_N start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT ( italic_N start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT - 1 ) + italic_N start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT ( italic_N start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT - 1 ) / 2, respectively. The adaptive ansatz expansion procedure requires an additional NH(Nθ1)subscript𝑁Hsubscript𝑁𝜃1N_{\mathrm{H}}(N_{\theta}-1)italic_N start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT ( italic_N start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT - 1 ) generalized Hadamard test circuits. For example, for the N=4𝑁4N=4italic_N = 4-site (N=6𝑁6N=6italic_N = 6-site) Fermi-Hubbard chain, the Hamiltonian consists of NH=16subscript𝑁H16N_{\mathrm{H}}=16italic_N start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT = 16 (NH=26subscript𝑁H26N_{\mathrm{H}}=26italic_N start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT = 26) terms while the maximum Nθsubscript𝑁𝜃N_{\theta}italic_N start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT among all Ir,sp,qsubscriptsuperscript𝐼𝑝𝑞𝑟𝑠I^{p,q}_{r,s}italic_I start_POSTSUPERSCRIPT italic_p , italic_q end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_r , italic_s end_POSTSUBSCRIPT at time t=10𝑡10t=10italic_t = 10 is Nθ=221subscript𝑁𝜃221N_{\theta}=221italic_N start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT = 221 (Nθ=2021subscript𝑁𝜃2021N_{\theta}=2021italic_N start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT = 2021). In this situation, a single Runge-Kutta time step requires 1.3×1051.3superscript1051.3\times 10^{5}1.3 × 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT (8.8×1068.8superscript1068.8\times 10^{6}8.8 × 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT) measurements for N=4𝑁4N=4italic_N = 4 (N=6𝑁6N=6italic_N = 6) while the adaptive ansatz expansion procedure demands an additional 3.5×1033.5superscript1033.5\times 10^{3}3.5 × 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT (5.2×1045.2superscript1045.2\times 10^{4}5.2 × 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT) measurements for N=4𝑁4N=4italic_N = 4 (N=6𝑁6N=6italic_N = 6). Note that each time step using the Runge-Kutta method involves four measurements of the matrix M𝑀Mitalic_M, vector V𝑉Vitalic_V, ^θsubscriptdelimited-⟨⟩^𝜃\langle\hat{\mathcal{H}}\rangle_{\theta}⟨ over^ start_ARG caligraphic_H end_ARG ⟩ start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT, and ^2θsubscriptdelimited-⟨⟩superscript^2𝜃\langle\hat{\mathcal{H}}^{2}\rangle_{\theta}⟨ over^ start_ARG caligraphic_H end_ARG start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT.

Regarding the sensitivity to noise of the AVQDS approach, we have studied the impact of noise on the AVQITE approach for ground state preparation in ref 66. Specifically, we investigated how sampling noise—coherent errors resulting from imperfect gate operations, and stochastic errors caused by qubit decoherence, dephasing, and relaxation—affect the AVQITE simulations. We considered a fixed uniform single-qubit gate error rate of 104superscript10410^{-4}10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT, which corresponds to the value realized in current hardware, and a uniform two-qubit error rate in the range [104,102]superscript104superscript102[10^{-4},10^{-2}][ 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT , 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT ]. Additionally, we used 214superscript2142^{14}2 start_POSTSUPERSCRIPT 14 end_POSTSUPERSCRIPT shots for each measurement circuit. We found that, for a two-qubit noise level of 102superscript10210^{-2}10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT relevant for current hardware, the ground state energy error is about 12%percent1212\%12 %. This decreases to 3%percent33\%3 % for a two-qubit gate error rate of 103superscript10310^{-3}10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT. We expect comparable errors for the dynamics simulations using AVQDS.

To benchmark the performance of AVQDS, we compare the AVQDS results with results utilizing exact diagonalization for state propagation:

|Ψ[t+δt]=eiδt^|Ψ[t].ketΨdelimited-[]𝑡𝛿𝑡superscriptei𝛿𝑡^ketΨdelimited-[]𝑡\displaystyle|\Psi[t+\delta t]\rangle=\mathrm{e}^{-\mathrm{i}\delta t\hat{% \mathcal{H}}}|\Psi[t]\rangle\,.| roman_Ψ [ italic_t + italic_δ italic_t ] ⟩ = roman_e start_POSTSUPERSCRIPT - roman_i italic_δ italic_t over^ start_ARG caligraphic_H end_ARG end_POSTSUPERSCRIPT | roman_Ψ [ italic_t ] ⟩ . (8)

This simulation is performed on a uniformly discretized time grid with a step size of δt=0.002𝛿𝑡0.002\delta t=0.002italic_δ italic_t = 0.002. Since our study focuses on time-independent Hamiltonians, the error due to finite time mesh here is zero.

3 Methods

3.1 Computation of Green’s functions with AVQDS

We begin by considering a time-independent fermionic Hamiltonian ^^\hat{\mathcal{H}}over^ start_ARG caligraphic_H end_ARG, expressed in terms of fermionic operators c^psubscriptsuperscript^𝑐𝑝\hat{c}^{\dagger}_{p}over^ start_ARG italic_c end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT and c^psubscriptsuperscript^𝑐absent𝑝\hat{c}^{\phantom{\dagger}}_{p}over^ start_ARG italic_c end_ARG start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT that create and annihilate an electron in orbital site p=(j,σ)𝑝𝑗𝜎p=(j,\sigma)italic_p = ( italic_j , italic_σ ). Here j𝑗jitalic_j is the site index and σ𝜎\sigmaitalic_σ corresponds to the spin of the fermion. The dynamics of the system at zero temperature is described by the retarded single-particle Green’s function:

Gp,qR(t)=iΘ(t)subscriptsuperscript𝐺R𝑝𝑞𝑡iΘ𝑡\displaystyle G^{\mathrm{R}}_{p,q}(t)=-\mathrm{i}\Theta(t)italic_G start_POSTSUPERSCRIPT roman_R end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_p , italic_q end_POSTSUBSCRIPT ( italic_t ) = - roman_i roman_Θ ( italic_t ) [G|c^p(t)c^q(0)|G\displaystyle\left[\bra{\mathrm{G}}\hat{c}^{\phantom{\dagger}}_{p}(t)\hat{c}^{% \dagger}_{q}(0)\ket{\mathrm{G}}\right.[ ⟨ start_ARG roman_G end_ARG | over^ start_ARG italic_c end_ARG start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ( italic_t ) over^ start_ARG italic_c end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT ( 0 ) | start_ARG roman_G end_ARG ⟩
+G|c^q(0)c^p(t)|G].\displaystyle\left.+\bra{\mathrm{G}}\hat{c}^{\dagger}_{q}(0)\hat{c}^{\phantom{% \dagger}}_{p}(t)\ket{\mathrm{G}}\right]\,.+ ⟨ start_ARG roman_G end_ARG | over^ start_ARG italic_c end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT ( 0 ) over^ start_ARG italic_c end_ARG start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ( italic_t ) | start_ARG roman_G end_ARG ⟩ ] . (9)

Here, c^p(t)=ei^tc^pei^tsubscriptsuperscript^𝑐absent𝑝𝑡superscriptei^𝑡subscriptsuperscript^𝑐absent𝑝superscriptei^𝑡\hat{c}^{\phantom{\dagger}}_{p}(t)=\mathrm{e}^{\mathrm{i}\hat{\mathcal{H}}t}% \hat{c}^{\phantom{\dagger}}_{p}\mathrm{e}^{-\mathrm{i}\hat{\mathcal{H}}t}over^ start_ARG italic_c end_ARG start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ( italic_t ) = roman_e start_POSTSUPERSCRIPT roman_i over^ start_ARG caligraphic_H end_ARG italic_t end_POSTSUPERSCRIPT over^ start_ARG italic_c end_ARG start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT roman_e start_POSTSUPERSCRIPT - roman_i over^ start_ARG caligraphic_H end_ARG italic_t end_POSTSUPERSCRIPT corresponds to the Heisenberg representation of the fermionic operator c^psubscriptsuperscript^𝑐absent𝑝\hat{c}^{\phantom{\dagger}}_{p}over^ start_ARG italic_c end_ARG start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT, Θ(t)Θ𝑡\Theta(t)roman_Θ ( italic_t ) is the Heaviside step function, and |GketG\ket{\mathrm{G}}| start_ARG roman_G end_ARG ⟩ denotes the ground state of Hamiltonian ^^\hat{\mathcal{H}}over^ start_ARG caligraphic_H end_ARG.

(a)

\Qcircuit@C=1.3em @R=1.3em \lstick|0⟩ & \gateH \ctrl1 \qw \gateX \ctrl1 \gateH \meter\qw
\lstick|φ_0⟩ \gateU_G\gateP_β \gatee^-i^Ht \qw \gateP_α \qw \qw\gategroup12240.7em–

(b)

\Qcircuit@C=1.0em @R=1.3em \lstick|0⟩ & \gateH \ctrl1 \qw \gateX \ctrl1 \gateH \meter\qw
\lstick|φ_0⟩ \gateU_G[θ^1]\gateP_β \gateU_t[θ^2] \qw \gateP_α \qw \qw\gategroup12240.7em–

Figure 1: Hadamard test circuit to compute the Green’s function. (a) Circuit to measure the Green’s function component Iα,βp,qsuperscriptsubscript𝐼𝛼𝛽𝑝𝑞I_{\alpha,\beta}^{p,q}italic_I start_POSTSUBSCRIPT italic_α , italic_β end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_p , italic_q end_POSTSUPERSCRIPT, eq 12, using the exact time-evolution operator ei^tsuperscriptei^𝑡\mathrm{e}^{-\mathrm{i}\hat{\mathcal{H}}t}roman_e start_POSTSUPERSCRIPT - roman_i over^ start_ARG caligraphic_H end_ARG italic_t end_POSTSUPERSCRIPT. The ancillary qubit, initially in the state |0ket0\ket{0}| start_ARG 0 end_ARG ⟩, is represented by the upper horizontal line. X𝑋Xitalic_X and H𝐻Hitalic_H denote Pauli-X𝑋Xitalic_X and Hadamard gates on the ancilla, respectively. A register of qubits for the physical system of interest, initially in a reference product state |φ0ketsubscript𝜑0\ket{\varphi_{0}}| start_ARG italic_φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ⟩, is denoted by the lower horizontal line. The application of the general multi-qubit Pauli gates Pαsubscript𝑃𝛼P_{\alpha}italic_P start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT and Pβsubscript𝑃𝛽P_{\beta}italic_P start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT is controlled by the ancilla qubit, while the unitary operator UGsubscript𝑈GU_{\mathrm{G}}italic_U start_POSTSUBSCRIPT roman_G end_POSTSUBSCRIPT prepares the ground state, |G=UG|φ0ketGsubscript𝑈Gketsubscript𝜑0\ket{\mathrm{G}}=U_{\mathrm{G}}\ket{\varphi_{0}}| start_ARG roman_G end_ARG ⟩ = italic_U start_POSTSUBSCRIPT roman_G end_POSTSUBSCRIPT | start_ARG italic_φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ⟩. (b) Measurement circuit for Iα,βp,qsuperscriptsubscript𝐼𝛼𝛽𝑝𝑞I_{\alpha,\beta}^{p,q}italic_I start_POSTSUBSCRIPT italic_α , italic_β end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_p , italic_q end_POSTSUPERSCRIPT using a variational state evolution circuit. The state propagation circuit highlighted with the dashed rectangle in (a) is replaced by the (adaptive) variational circuit in (b), where the angles in the parameterized unitaries UG[𝜽1]subscript𝑈Gdelimited-[]superscript𝜽1U_{\mathrm{G}}[\boldsymbol{\theta}^{1}]italic_U start_POSTSUBSCRIPT roman_G end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ] and Ut[𝜽2]subscript𝑈𝑡delimited-[]superscript𝜽2U_{t}[\boldsymbol{\theta}^{2}]italic_U start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] evolve with time. The results are obtained from Z𝑍Zitalic_Z-basis measurements on the ancilla qubit.

To calculate eq 3.1 using a quantum computer, we adopt the Jordan-Wigner transformation 67 to map the fermionic operators to Pauli operators as:

c^p=αηα(p)Pα(p).subscriptsuperscript^𝑐absent𝑝subscript𝛼subscriptsuperscript𝜂𝑝𝛼subscriptsuperscript𝑃𝑝𝛼\displaystyle\hat{c}^{\phantom{\dagger}}_{p}=\sum_{\alpha}\eta^{(p)}_{\alpha}P% ^{(p)}_{\alpha}\,.over^ start_ARG italic_c end_ARG start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_η start_POSTSUPERSCRIPT ( italic_p ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_P start_POSTSUPERSCRIPT ( italic_p ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT . (10)

Here, ηα(p)subscriptsuperscript𝜂𝑝𝛼\eta^{(p)}_{\alpha}italic_η start_POSTSUPERSCRIPT ( italic_p ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT are complex-valued numbers and Pα(p)subscriptsuperscript𝑃𝑝𝛼P^{(p)}_{\alpha}italic_P start_POSTSUPERSCRIPT ( italic_p ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT are Pauli words up to a weight of Nqsubscript𝑁𝑞N_{q}italic_N start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT, which is the number of qubits required to encode the fermionic Hamiltonian ^^\hat{\mathcal{H}}over^ start_ARG caligraphic_H end_ARG. Using transformation eq 10, eq 3.1 can be rewritten as:

Gp,qR(t)subscriptsuperscript𝐺R𝑝𝑞𝑡\displaystyle G^{\mathrm{R}}_{p,q}(t)italic_G start_POSTSUPERSCRIPT roman_R end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_p , italic_q end_POSTSUBSCRIPT ( italic_t ) =iΘ(t)α,βηα(p)ηβ(q)[G|ei^tPα(p)ei^tPβ(q)|G\displaystyle=-i\,\Theta(t)\sum_{\alpha,\beta}\eta^{(p)}_{\alpha}\eta^{(q)}_{% \beta}\left[\bra{\mathrm{G}}\mathrm{e}^{\mathrm{i}\hat{\mathcal{H}}t}P^{(p)}_{% \alpha}\mathrm{e}^{-\mathrm{i}\hat{\mathcal{H}}t}P^{(q)}_{\beta}\ket{\mathrm{G% }}\right.= - italic_i roman_Θ ( italic_t ) ∑ start_POSTSUBSCRIPT italic_α , italic_β end_POSTSUBSCRIPT italic_η start_POSTSUPERSCRIPT ( italic_p ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_η start_POSTSUPERSCRIPT ( italic_q ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT [ ⟨ start_ARG roman_G end_ARG | roman_e start_POSTSUPERSCRIPT roman_i over^ start_ARG caligraphic_H end_ARG italic_t end_POSTSUPERSCRIPT italic_P start_POSTSUPERSCRIPT ( italic_p ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT roman_e start_POSTSUPERSCRIPT - roman_i over^ start_ARG caligraphic_H end_ARG italic_t end_POSTSUPERSCRIPT italic_P start_POSTSUPERSCRIPT ( italic_q ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT | start_ARG roman_G end_ARG ⟩
+G|Pβ(q)ei^tPα(p)ei^t|G]\displaystyle\qquad\qquad\qquad\qquad\;\;\left.+\bra{\mathrm{G}}P^{(q)}_{\beta% }\mathrm{e}^{\mathrm{i}\hat{\mathcal{H}}t}P^{(p)}_{\alpha}\mathrm{e}^{-\mathrm% {i}\hat{\mathcal{H}}t}\ket{\mathrm{G}}\right]+ ⟨ start_ARG roman_G end_ARG | italic_P start_POSTSUPERSCRIPT ( italic_q ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT roman_e start_POSTSUPERSCRIPT roman_i over^ start_ARG caligraphic_H end_ARG italic_t end_POSTSUPERSCRIPT italic_P start_POSTSUPERSCRIPT ( italic_p ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT roman_e start_POSTSUPERSCRIPT - roman_i over^ start_ARG caligraphic_H end_ARG italic_t end_POSTSUPERSCRIPT | start_ARG roman_G end_ARG ⟩ ]
2iΘ(t)α,βηα(p)ηβ(q)Iα,βp,q,absent2iΘ𝑡subscript𝛼𝛽subscriptsuperscript𝜂𝑝𝛼subscriptsuperscript𝜂𝑞𝛽subscriptsuperscript𝐼𝑝𝑞𝛼𝛽\displaystyle\equiv-2\mathrm{i}\,\Theta(t)\sum_{\alpha,\beta}\eta^{(p)}_{% \alpha}\eta^{(q)}_{\beta}I^{p,q}_{\alpha,\beta}\,,≡ - 2 roman_i roman_Θ ( italic_t ) ∑ start_POSTSUBSCRIPT italic_α , italic_β end_POSTSUBSCRIPT italic_η start_POSTSUPERSCRIPT ( italic_p ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_η start_POSTSUPERSCRIPT ( italic_q ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT italic_I start_POSTSUPERSCRIPT italic_p , italic_q end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α , italic_β end_POSTSUBSCRIPT , (11)

with

Iα,βp,qRe[G|ei^t𝒫α(p)ei^tPβ(q)|G].subscriptsuperscript𝐼𝑝𝑞𝛼𝛽Redelimited-[]braGsuperscriptei^𝑡subscriptsuperscript𝒫𝑝𝛼superscriptei^𝑡subscriptsuperscript𝑃𝑞𝛽ketG\displaystyle I^{p,q}_{\alpha,\beta}\equiv\mathrm{Re}\left[\bra{\mathrm{G}}% \mathrm{e}^{\mathrm{i}\hat{\mathcal{H}}t}\mathcal{P}^{(p)}_{\alpha}\mathrm{e}^% {-\mathrm{i}\hat{\mathcal{H}}t}P^{(q)}_{\beta}\ket{\mathrm{G}}\right]\,.italic_I start_POSTSUPERSCRIPT italic_p , italic_q end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α , italic_β end_POSTSUBSCRIPT ≡ roman_Re [ ⟨ start_ARG roman_G end_ARG | roman_e start_POSTSUPERSCRIPT roman_i over^ start_ARG caligraphic_H end_ARG italic_t end_POSTSUPERSCRIPT caligraphic_P start_POSTSUPERSCRIPT ( italic_p ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT roman_e start_POSTSUPERSCRIPT - roman_i over^ start_ARG caligraphic_H end_ARG italic_t end_POSTSUPERSCRIPT italic_P start_POSTSUPERSCRIPT ( italic_q ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT | start_ARG roman_G end_ARG ⟩ ] . (12)

Equation 12 can be measured using a Hadamard test circuit on a quantum computer, as shown in Figure 1 a. The time evolution operator ei^tsuperscriptei^𝑡\mathrm{e}^{-\mathrm{i}\hat{\mathcal{H}}t}roman_e start_POSTSUPERSCRIPT - roman_i over^ start_ARG caligraphic_H end_ARG italic_t end_POSTSUPERSCRIPT can be approximated using Trotter decomposition 68, 69, 70 or variational quantum dynamics simulation algorithms 55, including AVQDS 59.

In this work, we first use the AVQITE algorithm to prepare the ground state of a Hamiltonian system, |G[𝜽1]=UG[𝜽1]|φ0ket𝐺delimited-[]superscript𝜽1subscript𝑈Gdelimited-[]superscript𝜽1ketsubscript𝜑0\ket{G[\boldsymbol{\theta}^{1}]}=U_{\mathrm{G}}[\boldsymbol{\theta}^{1}]\ket{% \varphi_{0}}| start_ARG italic_G [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ] end_ARG ⟩ = italic_U start_POSTSUBSCRIPT roman_G end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ] | start_ARG italic_φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ⟩, with a series of parameterized unitaries UG[𝜽1]subscript𝑈Gdelimited-[]superscript𝜽1U_{\mathrm{G}}[\boldsymbol{\theta}^{1}]italic_U start_POSTSUBSCRIPT roman_G end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ] applied to a reference product state |φ0ketsubscript𝜑0\ket{\varphi_{0}}| start_ARG italic_φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ⟩. The time propagation of the ancilla and physical register, as outlined by the dashed box in Figure 1 a, is achieved by a parameterized circuit highlighted by the dashed box in Figure 1 b. The additional unitaries Ut[𝜽2]subscript𝑈𝑡delimited-[]superscript𝜽2U_{t}[\boldsymbol{\theta}^{2}]italic_U start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] are automatically generated following the AVQDS algorithm. The rotation angles 𝜽1superscript𝜽1\boldsymbol{\theta}^{1}bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT and 𝜽2superscript𝜽2\boldsymbol{\theta}^{2}bold_italic_θ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT both evolve according to the equation of motion eq 5. In other words, AVQDS is adopted for the time propagation of the state

|Ψ=12|0|G+12|1Pβ|GketΨtensor-product12ket0ket𝐺tensor-product12ket1subscript𝑃𝛽ket𝐺\ket{\Psi}=\frac{1}{\sqrt{2}}\ket{0}\otimes\ket{G}+\frac{1}{\sqrt{2}}\ket{1}% \otimes P_{\beta}\ket{G}| start_ARG roman_Ψ end_ARG ⟩ = divide start_ARG 1 end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG | start_ARG 0 end_ARG ⟩ ⊗ | start_ARG italic_G end_ARG ⟩ + divide start_ARG 1 end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG | start_ARG 1 end_ARG ⟩ ⊗ italic_P start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT | start_ARG italic_G end_ARG ⟩ (13)

under a Hamiltonian that acts only on the physical register. As shown in Appendix A, the circuit in Figure 1 b measures the Green’s function component as

Iα,βp,qRe[G[𝜽1]|Ut[𝜽2]𝒫α(p)Ut[𝜽2]Pβ(q)|G[𝜽1]].subscriptsuperscript𝐼𝑝𝑞𝛼𝛽Redelimited-[]bra𝐺delimited-[]superscript𝜽1superscriptsubscript𝑈𝑡delimited-[]superscript𝜽2subscriptsuperscript𝒫𝑝𝛼subscript𝑈𝑡delimited-[]superscript𝜽2subscriptsuperscript𝑃𝑞𝛽ket𝐺delimited-[]superscript𝜽1I^{p,q}_{\alpha,\beta}\approx\mathrm{Re}\left[\bra{G[\boldsymbol{\theta}^{1}]}% U_{t}^{\dagger}[\boldsymbol{\theta}^{2}]\mathcal{P}^{(p)}_{\alpha}U_{t}[% \boldsymbol{\theta}^{2}]P^{(q)}_{\beta}\ket{G[\boldsymbol{\theta}^{1}]}\right].italic_I start_POSTSUPERSCRIPT italic_p , italic_q end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α , italic_β end_POSTSUBSCRIPT ≈ roman_Re [ ⟨ start_ARG italic_G [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ] end_ARG | italic_U start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] caligraphic_P start_POSTSUPERSCRIPT ( italic_p ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_U start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] italic_P start_POSTSUPERSCRIPT ( italic_q ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT | start_ARG italic_G [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ] end_ARG ⟩ ] . (14)

3.2 Computation of nonlinear susceptibilities with AVQDS

Susceptibility expansion

While full time-dependent non-equilibrium calculations are a powerful approach for capturing nonlinear effects, particularly when considering the entire spectrum of dynamics beyond weak signals, nonlinear susceptibilities offer meaningful and interpretable insights in scenarios where linear approximations fail but where a fully time-dependent non-equilibrium approach may not be necessary or practical. Nonlinear susceptibilities, in particular, provide a complementary perspective that is often more tractable and insightful, enabling the analysis of higher-order effects in a controlled manner. This approach makes it easier to isolate and understand specific nonlinear contributions and is also valuable for understanding the onset of nonlinear behavior before a full breakdown of the linear response.

Before we discuss the formalism, we begin with a brief discussion of the susceptibility expansion for nonlinear responses measured in 2DCS experiments. We consider an N𝑁Nitalic_N-site quantum spin system described by spin-s𝑠sitalic_s operators S^ipsubscriptsuperscript^𝑆𝑝𝑖\hat{S}^{p}_{i}over^ start_ARG italic_S end_ARG start_POSTSUPERSCRIPT italic_p end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT (p=x,y,z𝑝𝑥𝑦𝑧p=x,y,zitalic_p = italic_x , italic_y , italic_z) at site i𝑖iitalic_i. The transmitted magnetic field measured in 2DCS experiments on magnetic systems is determined by the magnetization 𝐌(t)𝐒^tot=Ψ[t]|𝐒^tot|Ψ[t]𝐌𝑡delimited-⟨⟩superscript^𝐒totquantum-operator-productΨdelimited-[]𝑡superscript^𝐒totΨdelimited-[]𝑡\mathbf{M}(t)\equiv\langle\hat{\mathbf{S}}^{\mathrm{tot}}\rangle=\langle\Psi[t% ]|\hat{\mathbf{S}}^{\mathrm{tot}}|\Psi[t]\ranglebold_M ( italic_t ) ≡ ⟨ over^ start_ARG bold_S end_ARG start_POSTSUPERSCRIPT roman_tot end_POSTSUPERSCRIPT ⟩ = ⟨ roman_Ψ [ italic_t ] | over^ start_ARG bold_S end_ARG start_POSTSUPERSCRIPT roman_tot end_POSTSUPERSCRIPT | roman_Ψ [ italic_t ] ⟩. Here, 𝐒^tot=j=1N𝐒^jsuperscript^𝐒totsuperscriptsubscript𝑗1𝑁subscript^𝐒𝑗\hat{\mathbf{S}}^{\mathrm{tot}}=\sum_{j=1}^{N}\hat{\mathbf{S}}_{j}over^ start_ARG bold_S end_ARG start_POSTSUPERSCRIPT roman_tot end_POSTSUPERSCRIPT = ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT over^ start_ARG bold_S end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT is the total spin operator and |Ψ[t]ketΨdelimited-[]𝑡|\Psi[t]\rangle| roman_Ψ [ italic_t ] ⟩ corresponds to the quantum state of the system at time t𝑡titalic_t. In 2DCS the quantum spin system is excited by two magnetic field pulses polarized along the β𝛽\betaitalic_β- and γ𝛾\gammaitalic_γ-directions, separated in time by τ𝜏\tauitalic_τ. The total applied magnetic field can be written as 𝐁(t)=B1β(t)𝜷+B2γ(tτ)𝜸𝐁𝑡subscriptsuperscript𝐵𝛽1𝑡𝜷subscriptsuperscript𝐵𝛾2𝑡𝜏𝜸\mathbf{B}(t)=B^{\beta}_{1}(t)\boldsymbol{\beta}+B^{\gamma}_{2}(t-\tau)% \boldsymbol{\gamma}bold_B ( italic_t ) = italic_B start_POSTSUPERSCRIPT italic_β end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_t ) bold_italic_β + italic_B start_POSTSUPERSCRIPT italic_γ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_t - italic_τ ) bold_italic_γ, with the two pulses centered at time t=0𝑡0t=0italic_t = 0 and t=τ𝑡𝜏t=\tauitalic_t = italic_τ, while 𝜷𝜷\boldsymbol{\beta}bold_italic_β and 𝜸𝜸\boldsymbol{\gamma}bold_italic_γ denote the unit vectors along the β𝛽\betaitalic_β and γ𝛾\gammaitalic_γ directions. The differential transmitted magnetic field along the α𝛼\alphaitalic_α direction, BNLαsubscriptsuperscript𝐵𝛼NLB^{\alpha}_{\mathrm{NL}}italic_B start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_NL end_POSTSUBSCRIPT, as measured in 2DCS experiments, is proportional to the nonlinear differential magnetization:

MNLα(t,τ)=M12α(t,τ)M1α(t)M2α(t,τ),subscriptsuperscript𝑀𝛼NL𝑡𝜏subscriptsuperscript𝑀𝛼12𝑡𝜏subscriptsuperscript𝑀𝛼1𝑡subscriptsuperscript𝑀𝛼2𝑡𝜏\displaystyle M^{\alpha}_{\mathrm{NL}}(t,\tau)=M^{\alpha}_{\mathrm{12}}(t,\tau% )-M^{\alpha}_{\mathrm{1}}(t)-M^{\alpha}_{\mathrm{2}}(t,\tau)\,,italic_M start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_NL end_POSTSUBSCRIPT ( italic_t , italic_τ ) = italic_M start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT ( italic_t , italic_τ ) - italic_M start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_t ) - italic_M start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_t , italic_τ ) , (15)

which depends on time t𝑡titalic_t and the inter-pulse delay τ𝜏\tauitalic_τ. Here, M12α(t,τ)subscriptsuperscript𝑀𝛼12𝑡𝜏M^{\alpha}_{12}(t,\tau)italic_M start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT ( italic_t , italic_τ ) denotes the magnetization dynamics induced by both pulses, while M1α(t)subscriptsuperscript𝑀𝛼1𝑡{M}^{\alpha}_{1}(t)italic_M start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_t ) and M2α(t,τ)subscriptsuperscript𝑀𝛼2𝑡𝜏{M}^{\alpha}_{2}(t,\tau)italic_M start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_t , italic_τ ) denote the magnetization dynamics induced by pulses 1 and 2, respectively. As demonstrated in ref 63, the 2DCS spectra obtained by a 2D Fourier transform of eq 15 can be interpreted by applying a susceptibility expansion 3, 4, 5 of MNLα(t,τ)subscriptsuperscript𝑀𝛼NL𝑡𝜏M^{\alpha}_{\mathrm{NL}}(t,\tau)italic_M start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_NL end_POSTSUBSCRIPT ( italic_t , italic_τ ). By approximating the pulse shapes by δ𝛿\deltaitalic_δ-functions, the applied magnetic field can be written as

𝐁(t)=A1βδ(t)𝜷+A2γδ(tτ)𝜸,𝐁𝑡superscriptsubscript𝐴1𝛽𝛿𝑡𝜷superscriptsubscript𝐴2𝛾𝛿𝑡𝜏𝜸\displaystyle\mathbf{B}(t)=A_{1}^{\beta}\,\delta(t)\boldsymbol{\beta}+A_{2}^{% \gamma}\,\delta(t-\tau)\boldsymbol{\gamma}\,,bold_B ( italic_t ) = italic_A start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_β end_POSTSUPERSCRIPT italic_δ ( italic_t ) bold_italic_β + italic_A start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_γ end_POSTSUPERSCRIPT italic_δ ( italic_t - italic_τ ) bold_italic_γ , (16)

where Aisubscript𝐴𝑖A_{i}italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT corresponds to the i𝑖iitalic_ith pulse area. As a result, the nonlinear magnetization density along the α𝛼\alphaitalic_α-direction at time t+τ𝑡𝜏t+\tauitalic_t + italic_τ can be expressed in terms of nonlinear susceptibilities χ(n)superscript𝜒𝑛\chi^{(n)}italic_χ start_POSTSUPERSCRIPT ( italic_n ) end_POSTSUPERSCRIPT of order n𝑛nitalic_n 4, 71, 5:

MNLα(t+τ)/Nsuperscriptsubscript𝑀NL𝛼𝑡𝜏𝑁\displaystyle M_{\mathrm{NL}}^{\alpha}(t+\tau)/Nitalic_M start_POSTSUBSCRIPT roman_NL end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT ( italic_t + italic_τ ) / italic_N
(M12α(t+τ)M1α(t+τ)M2α(t+τ))/Nabsentsubscriptsuperscript𝑀𝛼12𝑡𝜏subscriptsuperscript𝑀𝛼1𝑡𝜏subscriptsuperscript𝑀𝛼2𝑡𝜏𝑁\displaystyle\equiv(M^{\alpha}_{\mathrm{12}}(t+\tau)-M^{\alpha}_{\mathrm{1}}(t% +\tau)-M^{\alpha}_{\mathrm{2}}(t+\tau))/N≡ ( italic_M start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT ( italic_t + italic_τ ) - italic_M start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_t + italic_τ ) - italic_M start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_t + italic_τ ) ) / italic_N
=χαβγ(2)(t,τ)A1βA2γabsentsubscriptsuperscript𝜒2𝛼𝛽𝛾𝑡𝜏subscriptsuperscript𝐴𝛽1subscriptsuperscript𝐴𝛾2\displaystyle=\chi^{(2)}_{\alpha\beta\gamma}(t,\tau)\,A^{\beta}_{1}A^{\gamma}_% {2}= italic_χ start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α italic_β italic_γ end_POSTSUBSCRIPT ( italic_t , italic_τ ) italic_A start_POSTSUPERSCRIPT italic_β end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_A start_POSTSUPERSCRIPT italic_γ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT
+χαβγδ(3)(t,τ,0)(A1β)2A2γ+χαβγδ(3)(t,0,τ)A1β(A2γ)2subscriptsuperscript𝜒3𝛼𝛽𝛾𝛿𝑡𝜏0superscriptsubscriptsuperscript𝐴𝛽12subscriptsuperscript𝐴𝛾2subscriptsuperscript𝜒3𝛼𝛽𝛾𝛿𝑡0𝜏subscriptsuperscript𝐴𝛽1superscriptsubscriptsuperscript𝐴𝛾22\displaystyle+\chi^{(3)}_{\alpha\beta\gamma\delta}(t,\tau,0)\,(A^{\beta}_{1})^% {2}A^{\gamma}_{2}+\chi^{(3)}_{\alpha\beta\gamma\delta}(t,0,\tau)\,A^{\beta}_{1% }(A^{\gamma}_{2})^{2}+ italic_χ start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α italic_β italic_γ italic_δ end_POSTSUBSCRIPT ( italic_t , italic_τ , 0 ) ( italic_A start_POSTSUPERSCRIPT italic_β end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_A start_POSTSUPERSCRIPT italic_γ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + italic_χ start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α italic_β italic_γ italic_δ end_POSTSUBSCRIPT ( italic_t , 0 , italic_τ ) italic_A start_POSTSUPERSCRIPT italic_β end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_A start_POSTSUPERSCRIPT italic_γ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT
+𝒪(B4).𝒪superscript𝐵4\displaystyle+\mathcal{O}(B^{4})\,.+ caligraphic_O ( italic_B start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT ) . (17)

The second-order susceptibility in the above equation is explicitly given by 4, 71

χαβγ(2)(t,τ)=subscriptsuperscript𝜒2𝛼𝛽𝛾𝑡𝜏absent\displaystyle\chi^{(2)}_{\alpha\beta\gamma}(t,\tau)=italic_χ start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α italic_β italic_γ end_POSTSUBSCRIPT ( italic_t , italic_τ ) =
1NΘ(t)Θ(τ)[[M^α(t+τ),M^β(τ)],M^γ(0)],1𝑁Θ𝑡Θ𝜏delimited-⟨⟩superscript^𝑀𝛼𝑡𝜏superscript^𝑀𝛽𝜏superscript^𝑀𝛾0\displaystyle-\frac{1}{N}\Theta(t)\Theta(\tau)\langle\left[\left[\hat{M}^{% \alpha}(t+\tau),\hat{M}^{\beta}(\tau)\right],\hat{M}^{\gamma}(0)\right]\rangle\,,- divide start_ARG 1 end_ARG start_ARG italic_N end_ARG roman_Θ ( italic_t ) roman_Θ ( italic_τ ) ⟨ [ [ over^ start_ARG italic_M end_ARG start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT ( italic_t + italic_τ ) , over^ start_ARG italic_M end_ARG start_POSTSUPERSCRIPT italic_β end_POSTSUPERSCRIPT ( italic_τ ) ] , over^ start_ARG italic_M end_ARG start_POSTSUPERSCRIPT italic_γ end_POSTSUPERSCRIPT ( 0 ) ] ⟩ , (18)

while the third-order susceptibilities are defined by 5

χαβγδ(3)(t,τ,0)=subscriptsuperscript𝜒3𝛼𝛽𝛾𝛿𝑡𝜏0absent\displaystyle\chi^{(3)}_{\alpha\beta\gamma\delta}(t,\tau,0)=italic_χ start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α italic_β italic_γ italic_δ end_POSTSUBSCRIPT ( italic_t , italic_τ , 0 ) =
iNΘ(t)Θ(τ)[[[M^α(t+τ),M^β(τ)],M^γ(0)],M^δ(0)],i𝑁Θ𝑡Θ𝜏delimited-⟨⟩superscript^𝑀𝛼𝑡𝜏superscript^𝑀𝛽𝜏superscript^𝑀𝛾0superscript^𝑀𝛿0\displaystyle-\frac{\mathrm{i}}{N}\Theta(t)\Theta(\tau)\langle\left[\left[% \left[\hat{M}^{\alpha}(t+\tau),\hat{M}^{\beta}(\tau)\right],\hat{M}^{\gamma}(0% )\right],\hat{M}^{\delta}(0)\right]\rangle\,,- divide start_ARG roman_i end_ARG start_ARG italic_N end_ARG roman_Θ ( italic_t ) roman_Θ ( italic_τ ) ⟨ [ [ [ over^ start_ARG italic_M end_ARG start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT ( italic_t + italic_τ ) , over^ start_ARG italic_M end_ARG start_POSTSUPERSCRIPT italic_β end_POSTSUPERSCRIPT ( italic_τ ) ] , over^ start_ARG italic_M end_ARG start_POSTSUPERSCRIPT italic_γ end_POSTSUPERSCRIPT ( 0 ) ] , over^ start_ARG italic_M end_ARG start_POSTSUPERSCRIPT italic_δ end_POSTSUPERSCRIPT ( 0 ) ] ⟩ , (19)
χαβγδ(3)(t,0,τ)=subscriptsuperscript𝜒3𝛼𝛽𝛾𝛿𝑡0𝜏absent\displaystyle\chi^{(3)}_{\alpha\beta\gamma\delta}(t,0,\tau)=italic_χ start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α italic_β italic_γ italic_δ end_POSTSUBSCRIPT ( italic_t , 0 , italic_τ ) =
iNΘ(t)Θ(τ)[[[M^α(t+τ),M^β(τ)],M^γ(τ)],M^δ(0)].i𝑁Θ𝑡Θ𝜏delimited-⟨⟩superscript^𝑀𝛼𝑡𝜏superscript^𝑀𝛽𝜏superscript^𝑀𝛾𝜏superscript^𝑀𝛿0\displaystyle-\frac{\mathrm{i}}{N}\Theta(t)\Theta(\tau)\langle\left[\left[% \left[\hat{M}^{\alpha}(t+\tau),\hat{M}^{\beta}(\tau)\right],\hat{M}^{\gamma}(% \tau)\right],\hat{M}^{\delta}(0)\right]\rangle\,.- divide start_ARG roman_i end_ARG start_ARG italic_N end_ARG roman_Θ ( italic_t ) roman_Θ ( italic_τ ) ⟨ [ [ [ over^ start_ARG italic_M end_ARG start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT ( italic_t + italic_τ ) , over^ start_ARG italic_M end_ARG start_POSTSUPERSCRIPT italic_β end_POSTSUPERSCRIPT ( italic_τ ) ] , over^ start_ARG italic_M end_ARG start_POSTSUPERSCRIPT italic_γ end_POSTSUPERSCRIPT ( italic_τ ) ] , over^ start_ARG italic_M end_ARG start_POSTSUPERSCRIPT italic_δ end_POSTSUPERSCRIPT ( 0 ) ] ⟩ . (20)

In this paper, we demonstrate the calculation of the third-order susceptibility χαβγδ(3)(t,τ,0)subscriptsuperscript𝜒3𝛼𝛽𝛾𝛿𝑡𝜏0\chi^{(3)}_{\alpha\beta\gamma\delta}(t,\tau,0)italic_χ start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α italic_β italic_γ italic_δ end_POSTSUBSCRIPT ( italic_t , italic_τ , 0 ) using a quantum computing approach. The second-order χαβγ(2)(t,τ)subscriptsuperscript𝜒2𝛼𝛽𝛾𝑡𝜏\chi^{(2)}_{\alpha\beta\gamma}(t,\tau)italic_χ start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α italic_β italic_γ end_POSTSUBSCRIPT ( italic_t , italic_τ ) and third-order χαβγδ(3)(t,0,τ)subscriptsuperscript𝜒3𝛼𝛽𝛾𝛿𝑡0𝜏\chi^{(3)}_{\alpha\beta\gamma\delta}(t,0,\tau)italic_χ start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α italic_β italic_γ italic_δ end_POSTSUBSCRIPT ( italic_t , 0 , italic_τ ) as well as higher-order susceptibilities can be evaluated similarly. This approach is complementary to the method of direct observable dynamics simulations we developed earlier by disentangling contributions from quantum processes of different orders 63.

Formalism

We next discuss how χαβγδ(3)(t,τ,0)subscriptsuperscript𝜒3𝛼𝛽𝛾𝛿𝑡𝜏0\chi^{(3)}_{\alpha\beta\gamma\delta}(t,\tau,0)italic_χ start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α italic_β italic_γ italic_δ end_POSTSUBSCRIPT ( italic_t , italic_τ , 0 ) can be measured on a quantum computer. By using M^α=j=1NS^jαsuperscript^𝑀𝛼superscriptsubscript𝑗1𝑁subscriptsuperscript^𝑆𝛼𝑗\hat{M}^{\alpha}=\sum_{j=1}^{N}\hat{S}^{\alpha}_{j}over^ start_ARG italic_M end_ARG start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT = ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT over^ start_ARG italic_S end_ARG start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT, the third-order susceptibility eq 3.2 can be written as:

χαβγδ(3)(t,τ,0)subscriptsuperscript𝜒3𝛼𝛽𝛾𝛿𝑡𝜏0\displaystyle\chi^{(3)}_{\alpha\beta\gamma\delta}(t,\tau,0)italic_χ start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α italic_β italic_γ italic_δ end_POSTSUBSCRIPT ( italic_t , italic_τ , 0 )
=2NΘ(t)Θ(τ)j,k,l,m=1NIm[S^jα(t+τ)S^kβ(τ)Slγ(0)Smδ(0)\displaystyle=\frac{2}{N}\Theta(t)\Theta(\tau)\sum_{j,k,l,m=1}^{N}\mathrm{Im}% \left[\langle\hat{S}^{\alpha}_{j}(t+\tau)\hat{S}^{\beta}_{k}(\tau)S^{\gamma}_{% l}(0){S}^{\delta}_{m}(0)\rangle\right.= divide start_ARG 2 end_ARG start_ARG italic_N end_ARG roman_Θ ( italic_t ) roman_Θ ( italic_τ ) ∑ start_POSTSUBSCRIPT italic_j , italic_k , italic_l , italic_m = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT roman_Im [ ⟨ over^ start_ARG italic_S end_ARG start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t + italic_τ ) over^ start_ARG italic_S end_ARG start_POSTSUPERSCRIPT italic_β end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_τ ) italic_S start_POSTSUPERSCRIPT italic_γ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ( 0 ) italic_S start_POSTSUPERSCRIPT italic_δ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT ( 0 ) ⟩
+Smδ(0)Slγ(0)S^jα(t+τ)S^kβ(τ)delimited-⟨⟩subscriptsuperscript𝑆𝛿𝑚0subscriptsuperscript𝑆𝛾𝑙0subscriptsuperscript^𝑆𝛼𝑗𝑡𝜏subscriptsuperscript^𝑆𝛽𝑘𝜏\displaystyle\qquad\qquad\qquad\qquad\quad\left.+\langle S^{\delta}_{m}(0){S}^% {\gamma}_{l}(0)\hat{S}^{\alpha}_{j}(t+\tau)\hat{S}^{\beta}_{k}(\tau)\rangle\right.+ ⟨ italic_S start_POSTSUPERSCRIPT italic_δ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT ( 0 ) italic_S start_POSTSUPERSCRIPT italic_γ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ( 0 ) over^ start_ARG italic_S end_ARG start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t + italic_τ ) over^ start_ARG italic_S end_ARG start_POSTSUPERSCRIPT italic_β end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_τ ) ⟩
S^lγ(0)S^jα(t+τ)S^kβ(τ)Smδ(0)delimited-⟨⟩subscriptsuperscript^𝑆𝛾𝑙0subscriptsuperscript^𝑆𝛼𝑗𝑡𝜏subscriptsuperscript^𝑆𝛽𝑘𝜏subscriptsuperscript𝑆𝛿𝑚0\displaystyle\qquad\qquad\qquad\qquad\quad\left.-\langle\hat{S}^{\gamma}_{l}(0% )\hat{S}^{\alpha}_{j}(t+\tau)\hat{S}^{\beta}_{k}(\tau)S^{\delta}_{m}(0)\rangle\right.- ⟨ over^ start_ARG italic_S end_ARG start_POSTSUPERSCRIPT italic_γ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ( 0 ) over^ start_ARG italic_S end_ARG start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t + italic_τ ) over^ start_ARG italic_S end_ARG start_POSTSUPERSCRIPT italic_β end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_τ ) italic_S start_POSTSUPERSCRIPT italic_δ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT ( 0 ) ⟩
S^lδ(0)S^jα(t+τ)S^kβ(τ)Smγ(0)].\displaystyle\qquad\qquad\qquad\qquad\quad\left.-\langle\hat{S}^{\delta}_{l}(0% )\hat{S}^{\alpha}_{j}(t+\tau)\hat{S}^{\beta}_{k}(\tau)S^{\gamma}_{m}(0)\rangle% \right]\,.- ⟨ over^ start_ARG italic_S end_ARG start_POSTSUPERSCRIPT italic_δ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ( 0 ) over^ start_ARG italic_S end_ARG start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t + italic_τ ) over^ start_ARG italic_S end_ARG start_POSTSUPERSCRIPT italic_β end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_τ ) italic_S start_POSTSUPERSCRIPT italic_γ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT ( 0 ) ⟩ ] . (21)

To evaluate eq 3.2 for a generic spin-s𝑠sitalic_s model using a quantum computer, one can either use a qudit-based quantum device 72, 73 or bosonic quantum devices 74, 75 where the number of qudit levels/qumodes corresponds to the number of spin states 2s+12𝑠12s+12 italic_s + 1, or map the spin-s𝑠sitalic_s levels to qubits using transformations such as the Gray code or binary encoding 63, 76. In this paper, we choose the latter approach as qubit-based platforms are currently the most widely available. The transformation of spin-s𝑠sitalic_s operators to multi-qubit operators can be written as

S^jα=p=1nαηj,pαPj,pα.subscriptsuperscript^𝑆𝛼𝑗superscriptsubscript𝑝1subscript𝑛𝛼subscriptsuperscript𝜂𝛼𝑗𝑝subscriptsuperscript𝑃𝛼𝑗𝑝\displaystyle\hat{S}^{\alpha}_{j}=\sum_{p=1}^{n_{\alpha}}\eta^{\alpha}_{j,p}P^% {\alpha}_{j,p}\,.over^ start_ARG italic_S end_ARG start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_p = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_η start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j , italic_p end_POSTSUBSCRIPT italic_P start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j , italic_p end_POSTSUBSCRIPT . (22)

Here, the transformation contains nαsubscript𝑛𝛼n_{\alpha}italic_n start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT terms and the index j=1,,N𝑗1𝑁j=1,\dots,Nitalic_j = 1 , … , italic_N labels the physical site; ηj,pαsubscriptsuperscript𝜂𝛼𝑗𝑝\eta^{\alpha}_{j,p}italic_η start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j , italic_p end_POSTSUBSCRIPT are real-valued coefficients and Pj,pαsubscriptsuperscript𝑃𝛼𝑗𝑝P^{\alpha}_{j,p}italic_P start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j , italic_p end_POSTSUBSCRIPT are Pauli words. The encoding of a spin-s𝑠sitalic_s site requires nqsubscript𝑛𝑞n_{q}italic_n start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT qubits such that the system contains Nq=nqNsubscript𝑁𝑞subscript𝑛𝑞𝑁N_{q}=n_{q}Nitalic_N start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT = italic_n start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT italic_N qubits in total. Using S^jα(t)=ei^tS^jαei^tsubscriptsuperscript^𝑆𝛼𝑗𝑡superscriptei^𝑡subscriptsuperscript^𝑆𝛼𝑗superscriptei^𝑡\hat{S}^{\alpha}_{j}(t)=\mathrm{e}^{\mathrm{i}\hat{\mathcal{H}}t}\hat{S}^{% \alpha}_{j}\mathrm{e}^{-\mathrm{i}\hat{\mathcal{H}}t}over^ start_ARG italic_S end_ARG start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t ) = roman_e start_POSTSUPERSCRIPT roman_i over^ start_ARG caligraphic_H end_ARG italic_t end_POSTSUPERSCRIPT over^ start_ARG italic_S end_ARG start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT roman_e start_POSTSUPERSCRIPT - roman_i over^ start_ARG caligraphic_H end_ARG italic_t end_POSTSUPERSCRIPT and eq 22, eq 3.2 becomes

χαβγδ(3)(t,τ,0)subscriptsuperscript𝜒3𝛼𝛽𝛾𝛿𝑡𝜏0\displaystyle\chi^{(3)}_{\alpha\beta\gamma\delta}(t,\tau,0)italic_χ start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α italic_β italic_γ italic_δ end_POSTSUBSCRIPT ( italic_t , italic_τ , 0 )
=2NΘ(t)Θ(τ)j,k,l,m=1Np=1nαq=1nβr=1nγs=1nδηj,pαηk,qβηl,rγηm,sδabsent2𝑁Θ𝑡Θ𝜏superscriptsubscript𝑗𝑘𝑙𝑚1𝑁superscriptsubscript𝑝1subscript𝑛𝛼superscriptsubscript𝑞1subscript𝑛𝛽superscriptsubscript𝑟1subscript𝑛𝛾superscriptsubscript𝑠1subscript𝑛𝛿subscriptsuperscript𝜂𝛼𝑗𝑝subscriptsuperscript𝜂𝛽𝑘𝑞subscriptsuperscript𝜂𝛾𝑙𝑟subscriptsuperscript𝜂𝛿𝑚𝑠\displaystyle=\frac{2}{N}\Theta(t)\Theta(\tau)\sum_{j,k,l,m=1}^{N}\sum_{p=1}^{% n_{\alpha}}\sum_{q=1}^{n_{\beta}}\sum_{r=1}^{n_{\gamma}}\sum_{s=1}^{n_{\delta}% }\eta^{\alpha}_{j,p}\eta^{\beta}_{k,q}\eta^{\gamma}_{l,r}\eta^{\delta}_{m,s}= divide start_ARG 2 end_ARG start_ARG italic_N end_ARG roman_Θ ( italic_t ) roman_Θ ( italic_τ ) ∑ start_POSTSUBSCRIPT italic_j , italic_k , italic_l , italic_m = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_p = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_q = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_r = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_s = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_η start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j , italic_p end_POSTSUBSCRIPT italic_η start_POSTSUPERSCRIPT italic_β end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k , italic_q end_POSTSUBSCRIPT italic_η start_POSTSUPERSCRIPT italic_γ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l , italic_r end_POSTSUBSCRIPT italic_η start_POSTSUPERSCRIPT italic_δ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_m , italic_s end_POSTSUBSCRIPT
×Im[ei^(t+τ)Pj,pαei^tPk,qβei^τPl,rγPm,sδ\displaystyle\times\mathrm{Im}\left[\langle\mathrm{e}^{\mathrm{i}\hat{\mathcal% {H}}(t+\tau)}P^{\alpha}_{j,p}\mathrm{e}^{-\mathrm{i}\hat{\mathcal{H}}t}P^{% \beta}_{k,q}e^{-\mathrm{i}\hat{\mathcal{H}}\tau}P^{\gamma}_{l,r}P^{\delta}_{m,% s}\rangle\right.× roman_Im [ ⟨ roman_e start_POSTSUPERSCRIPT roman_i over^ start_ARG caligraphic_H end_ARG ( italic_t + italic_τ ) end_POSTSUPERSCRIPT italic_P start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j , italic_p end_POSTSUBSCRIPT roman_e start_POSTSUPERSCRIPT - roman_i over^ start_ARG caligraphic_H end_ARG italic_t end_POSTSUPERSCRIPT italic_P start_POSTSUPERSCRIPT italic_β end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k , italic_q end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - roman_i over^ start_ARG caligraphic_H end_ARG italic_τ end_POSTSUPERSCRIPT italic_P start_POSTSUPERSCRIPT italic_γ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l , italic_r end_POSTSUBSCRIPT italic_P start_POSTSUPERSCRIPT italic_δ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_m , italic_s end_POSTSUBSCRIPT ⟩
+Pm,sδPl,rγei^(t+τ)Pj,pαei^tPk,qβei^τdelimited-⟨⟩subscriptsuperscript𝑃𝛿𝑚𝑠subscriptsuperscript𝑃𝛾𝑙𝑟superscriptei^𝑡𝜏subscriptsuperscript𝑃𝛼𝑗𝑝superscriptei^𝑡subscriptsuperscript𝑃𝛽𝑘𝑞superscriptei^𝜏\displaystyle\qquad\left.+\langle P^{\delta}_{m,s}P^{\gamma}_{l,r}\mathrm{e}^{% \mathrm{i}\hat{\mathcal{H}}(t+\tau)}P^{\alpha}_{j,p}\mathrm{e}^{-\mathrm{i}% \hat{\mathcal{H}}t}P^{\beta}_{k,q}\mathrm{e}^{-\mathrm{i}\hat{\mathcal{H}}\tau% }\rangle\right.+ ⟨ italic_P start_POSTSUPERSCRIPT italic_δ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_m , italic_s end_POSTSUBSCRIPT italic_P start_POSTSUPERSCRIPT italic_γ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l , italic_r end_POSTSUBSCRIPT roman_e start_POSTSUPERSCRIPT roman_i over^ start_ARG caligraphic_H end_ARG ( italic_t + italic_τ ) end_POSTSUPERSCRIPT italic_P start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j , italic_p end_POSTSUBSCRIPT roman_e start_POSTSUPERSCRIPT - roman_i over^ start_ARG caligraphic_H end_ARG italic_t end_POSTSUPERSCRIPT italic_P start_POSTSUPERSCRIPT italic_β end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k , italic_q end_POSTSUBSCRIPT roman_e start_POSTSUPERSCRIPT - roman_i over^ start_ARG caligraphic_H end_ARG italic_τ end_POSTSUPERSCRIPT ⟩
Pl,rγei^(t+τ)Pj,pαei^tPk,qβei^τPm,sδdelimited-⟨⟩subscriptsuperscript𝑃𝛾𝑙𝑟superscriptei^𝑡𝜏subscriptsuperscript𝑃𝛼𝑗𝑝superscriptei^𝑡subscriptsuperscript𝑃𝛽𝑘𝑞superscriptei^𝜏subscriptsuperscript𝑃𝛿𝑚𝑠\displaystyle\qquad\left.-\langle P^{\gamma}_{l,r}\mathrm{e}^{\mathrm{i}\hat{% \mathcal{H}}(t+\tau)}P^{\alpha}_{j,p}\mathrm{e}^{-\mathrm{i}\hat{\mathcal{H}}t% }P^{\beta}_{k,q}\mathrm{e}^{-\mathrm{i}\hat{\mathcal{H}}\tau}P^{\delta}_{m,s}% \rangle\right.- ⟨ italic_P start_POSTSUPERSCRIPT italic_γ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l , italic_r end_POSTSUBSCRIPT roman_e start_POSTSUPERSCRIPT roman_i over^ start_ARG caligraphic_H end_ARG ( italic_t + italic_τ ) end_POSTSUPERSCRIPT italic_P start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j , italic_p end_POSTSUBSCRIPT roman_e start_POSTSUPERSCRIPT - roman_i over^ start_ARG caligraphic_H end_ARG italic_t end_POSTSUPERSCRIPT italic_P start_POSTSUPERSCRIPT italic_β end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k , italic_q end_POSTSUBSCRIPT roman_e start_POSTSUPERSCRIPT - roman_i over^ start_ARG caligraphic_H end_ARG italic_τ end_POSTSUPERSCRIPT italic_P start_POSTSUPERSCRIPT italic_δ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_m , italic_s end_POSTSUBSCRIPT ⟩
Pl,rδei^(t+τ)Pj,pαei^tPk,qβei^τPm,sγ].\displaystyle\qquad\left.-\langle P^{\delta}_{l,r}\mathrm{e}^{\mathrm{i}\hat{% \mathcal{H}}(t+\tau)}P^{\alpha}_{j,p}\mathrm{e}^{-\mathrm{i}\hat{\mathcal{H}}t% }P^{\beta}_{k,q}\mathrm{e}^{-\mathrm{i}\hat{\mathcal{H}}\tau}P^{\gamma}_{m,s}% \rangle\right]\,.- ⟨ italic_P start_POSTSUPERSCRIPT italic_δ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l , italic_r end_POSTSUBSCRIPT roman_e start_POSTSUPERSCRIPT roman_i over^ start_ARG caligraphic_H end_ARG ( italic_t + italic_τ ) end_POSTSUPERSCRIPT italic_P start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j , italic_p end_POSTSUBSCRIPT roman_e start_POSTSUPERSCRIPT - roman_i over^ start_ARG caligraphic_H end_ARG italic_t end_POSTSUPERSCRIPT italic_P start_POSTSUPERSCRIPT italic_β end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k , italic_q end_POSTSUBSCRIPT roman_e start_POSTSUPERSCRIPT - roman_i over^ start_ARG caligraphic_H end_ARG italic_τ end_POSTSUPERSCRIPT italic_P start_POSTSUPERSCRIPT italic_γ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_m , italic_s end_POSTSUBSCRIPT ⟩ ] . (23)

All the terms in eq 3.2 can be measured using the circuit in Figure 2 a. Nevertheless, to reduce the circuit depth for near-term applications, we adopt the generalized CUL circuit shown in Figure 2 b for nonlinear correlation function simulations. In the CUL circuit, the state propagation for the system plus ancilla due to ei^τsuperscriptei^𝜏\mathrm{e}^{-\mathrm{i}\hat{\mathcal{H}}\tau}roman_e start_POSTSUPERSCRIPT - roman_i over^ start_ARG caligraphic_H end_ARG italic_τ end_POSTSUPERSCRIPT maps to the evolution of parameters in UG[𝜽1]subscript𝑈Gdelimited-[]superscript𝜽1U_{\textrm{G}}[\boldsymbol{\theta}^{1}]italic_U start_POSTSUBSCRIPT G end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ] and Uτ[𝜽3]subscript𝑈𝜏delimited-[]superscript𝜽3U_{\tau}[\boldsymbol{\theta}^{3}]italic_U start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ]; while the state propagation due to ei^tsuperscriptei^𝑡\mathrm{e}^{-\mathrm{i}\hat{\mathcal{H}}t}roman_e start_POSTSUPERSCRIPT - roman_i over^ start_ARG caligraphic_H end_ARG italic_t end_POSTSUPERSCRIPT maps to parameter updating in UG[𝜽1]subscript𝑈Gdelimited-[]superscript𝜽1U_{\textrm{G}}[\boldsymbol{\theta}^{1}]italic_U start_POSTSUBSCRIPT G end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ], Uτ[𝜽3]subscript𝑈𝜏delimited-[]superscript𝜽3U_{\tau}[\boldsymbol{\theta}^{3}]italic_U start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ], and Ut[𝜽2]subscript𝑈𝑡delimited-[]superscript𝜽2U_{t}[\boldsymbol{\theta}^{2}]italic_U start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ]. We adopt AVQITE to generate UG[𝜽1]subscript𝑈Gdelimited-[]superscript𝜽1U_{\textrm{G}}[\boldsymbol{\theta}^{1}]italic_U start_POSTSUBSCRIPT G end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ] for ground state preparation, and AVQDS to generate Uτ[𝜽3]subscript𝑈𝜏delimited-[]superscript𝜽3U_{\tau}[\boldsymbol{\theta}^{3}]italic_U start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ] and Ut[𝜽2]subscript𝑈𝑡delimited-[]superscript𝜽2U_{t}[\boldsymbol{\theta}^{2}]italic_U start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] for state evolution. More detailed discussions are given in Appendix C on measuring the different terms within the square brackets of eq 3.2 using the CUL circuit in Figure 2 b. Specifically, the CUL circuit measures the imaginary part of G[𝜽1]|P0P1Uτ[𝜽3]Ut[𝜽2]P2Ut[𝜽2]P3Uτ[𝜽3]P4P5|G[𝜽1]bra𝐺delimited-[]superscript𝜽1subscript𝑃0subscript𝑃1superscriptsubscript𝑈𝜏delimited-[]superscript𝜽3superscriptsubscript𝑈𝑡delimited-[]superscript𝜽2subscript𝑃2subscript𝑈𝑡delimited-[]superscript𝜽2subscript𝑃3subscript𝑈𝜏delimited-[]superscript𝜽3subscript𝑃4subscript𝑃5ket𝐺delimited-[]superscript𝜽1\bra{G[\boldsymbol{\theta}^{1}]}P_{0}P_{1}U_{\tau}^{\dagger}[\boldsymbol{% \theta}^{3}]U_{t}^{\dagger}[\boldsymbol{\theta}^{2}]P_{2}\,U_{t}[\boldsymbol{% \theta}^{2}]P_{3}\,U_{\tau}[\boldsymbol{\theta}^{3}]P_{4}P_{5}\ket{G[% \boldsymbol{\theta}^{1}]}⟨ start_ARG italic_G [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ] end_ARG | italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_U start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ] italic_U start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] italic_P start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_U start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] italic_P start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT italic_U start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ] italic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT | start_ARG italic_G [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ] end_ARG ⟩. By setting P0=P1=INqsubscript𝑃0subscript𝑃1superscript𝐼tensor-productabsentsubscript𝑁𝑞P_{0}=P_{1}=I^{\otimes N_{q}}italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = italic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_I start_POSTSUPERSCRIPT ⊗ italic_N start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT end_POSTSUPERSCRIPT, P2=Pj,pαsubscript𝑃2subscriptsuperscript𝑃𝛼𝑗𝑝P_{2}=P^{\alpha}_{j,p}italic_P start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = italic_P start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j , italic_p end_POSTSUBSCRIPT, P3=Pk,qβsubscript𝑃3subscriptsuperscript𝑃𝛽𝑘𝑞P_{3}=P^{\beta}_{k,q}italic_P start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT = italic_P start_POSTSUPERSCRIPT italic_β end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k , italic_q end_POSTSUBSCRIPT, P4=Pl,rγsubscript𝑃4subscriptsuperscript𝑃𝛾𝑙𝑟P_{4}=P^{\gamma}_{l,r}italic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT = italic_P start_POSTSUPERSCRIPT italic_γ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l , italic_r end_POSTSUBSCRIPT, and P5=Pm,sδsubscript𝑃5subscriptsuperscript𝑃𝛿𝑚𝑠P_{5}=P^{\delta}_{m,s}italic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT = italic_P start_POSTSUPERSCRIPT italic_δ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_m , italic_s end_POSTSUBSCRIPT (P0=Pm,sδsubscript𝑃0subscriptsuperscript𝑃𝛿𝑚𝑠P_{0}=P^{\delta}_{m,s}italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = italic_P start_POSTSUPERSCRIPT italic_δ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_m , italic_s end_POSTSUBSCRIPT, P1=Pl,rγsubscript𝑃1subscriptsuperscript𝑃𝛾𝑙𝑟P_{1}=P^{\gamma}_{l,r}italic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_P start_POSTSUPERSCRIPT italic_γ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l , italic_r end_POSTSUBSCRIPT, P2=Pj,pαsubscript𝑃2subscriptsuperscript𝑃𝛼𝑗𝑝P_{2}=P^{\alpha}_{j,p}italic_P start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = italic_P start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j , italic_p end_POSTSUBSCRIPT, P3=Pk,qβsubscript𝑃3subscriptsuperscript𝑃𝛽𝑘𝑞P_{3}=P^{\beta}_{k,q}italic_P start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT = italic_P start_POSTSUPERSCRIPT italic_β end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k , italic_q end_POSTSUBSCRIPT, and P4=P5=INqsubscript𝑃4subscript𝑃5superscript𝐼tensor-productabsentsubscript𝑁𝑞P_{4}=P_{5}=I^{\otimes N_{q}}italic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT = italic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT = italic_I start_POSTSUPERSCRIPT ⊗ italic_N start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT end_POSTSUPERSCRIPT), the first (second) term within the square brackets of eq 3.2 is calculated, while the third (fourth) contributions follow by setting P0=P5=INqsubscript𝑃0subscript𝑃5superscript𝐼tensor-productabsentsubscript𝑁𝑞P_{0}=P_{5}=I^{\otimes N_{q}}italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = italic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT = italic_I start_POSTSUPERSCRIPT ⊗ italic_N start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT end_POSTSUPERSCRIPT, P1=Pl,rγ(δ)subscript𝑃1subscriptsuperscript𝑃𝛾𝛿𝑙𝑟P_{1}=P^{\gamma(\delta)}_{l,r}italic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_P start_POSTSUPERSCRIPT italic_γ ( italic_δ ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l , italic_r end_POSTSUBSCRIPT, P2=Pj,pαsubscript𝑃2subscriptsuperscript𝑃𝛼𝑗𝑝P_{2}=P^{\alpha}_{j,p}italic_P start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = italic_P start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j , italic_p end_POSTSUBSCRIPT, P3=Pk,qβsubscript𝑃3subscriptsuperscript𝑃𝛽𝑘𝑞P_{3}=P^{\beta}_{k,q}italic_P start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT = italic_P start_POSTSUPERSCRIPT italic_β end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k , italic_q end_POSTSUBSCRIPT, and P4=Pm,sδ(γ)subscript𝑃4subscriptsuperscript𝑃𝛿𝛾𝑚𝑠P_{4}=P^{\delta(\gamma)}_{m,s}italic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT = italic_P start_POSTSUPERSCRIPT italic_δ ( italic_γ ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_m , italic_s end_POSTSUBSCRIPT. We note that the CUL circuit in Figure 2 b can be simplified for specific settings of {Pi}i=05superscriptsubscriptsubscript𝑃𝑖𝑖05\{P_{i}\}_{i=0}^{5}{ italic_P start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT } start_POSTSUBSCRIPT italic_i = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT gates. For instance, if Pauli gate Pi=INqsubscript𝑃𝑖superscript𝐼tensor-productabsentsubscript𝑁𝑞P_{i}=I^{\otimes N_{q}}italic_P start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_I start_POSTSUPERSCRIPT ⊗ italic_N start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT end_POSTSUPERSCRIPT, it can be removed trivially. If P4=P0subscript𝑃4subscript𝑃0P_{4}=P_{0}italic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT = italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, the two controlled P4subscript𝑃4P_{4}italic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT and P0subscript𝑃0P_{0}italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT gates can be merged to a single P0subscript𝑃0P_{0}italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT gate.

(a)

\Qcircuit@C=0.75em @R=1.5em \lstick|0⟩ & \gateH \gateS \ctrl1 \ctrl1 \gateX \ctrl1 \ctrl1 \gateX \ctrl1 \gateX \ctrl1 \gateH \meter\qw
\lstick|φ_0⟩ \gateU_G \qw \gateP_5 \gateP_4 \qw\gateP_0 \gateP_1 \gatee^-i^Hτ \gateP_3 \gatee^-i^Ht \gateP_2 \qw \qw

(b)

\Qcircuit@C=0.75em @R=1.5em \lstick|0⟩ & \gateH \gateS \ctrl1 \ctrl1 \gateX \ctrl1 \ctrl1 \gateX \ctrl1 \gateX \ctrl1 \gateH \meter\qw
\lstick|φ_0⟩ \gateU_G[θ^1] \qw \gateP_5 \gateP_4 \qw\gateP_0 \gateP_1 \gateU_τ(θ^3) \gateP_3 \gateU_t(θ^2) \gateP_2 \qw \qw

Figure 2: Quantum circuit for measuring the third-order susceptibility eq 3.2. (a) Circuit utilizing exact time-evolution gates. The upper horizontal line represents the ancillary qubit, initially in the state |0ket0\ket{0}| start_ARG 0 end_ARG ⟩. X𝑋Xitalic_X, H𝐻Hitalic_H, and S𝑆Sitalic_S correspond to the Pauli-X𝑋Xitalic_X, Hadamard rotation, and S𝑆Sitalic_S-gate operations on the ancilla. The lower horizontal line denotes the qubit register representing the quantum spin system, initially in a reference product state |φ0ketsubscript𝜑0\ket{\varphi_{0}}| start_ARG italic_φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ⟩, which is rotated to the ground state |G=UG|φ0ket𝐺subscript𝑈Gketsubscript𝜑0\ket{G}=U_{\mathrm{G}}\ket{\varphi_{0}}| start_ARG italic_G end_ARG ⟩ = italic_U start_POSTSUBSCRIPT roman_G end_POSTSUBSCRIPT | start_ARG italic_φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ⟩ by a unitary circuit UGsubscript𝑈GU_{\mathrm{G}}italic_U start_POSTSUBSCRIPT roman_G end_POSTSUBSCRIPT. Multiple controlled Pauli gates are used to measure different terms in eq 3.2. The expectation value Im[G|P0P1ei^(t+τ)P2ei^tP3ei^τP4P5|G]Imdelimited-[]bra𝐺subscript𝑃0subscript𝑃1superscriptei^𝑡𝜏subscript𝑃2superscriptei^𝑡subscript𝑃3superscriptei^𝜏subscript𝑃4subscript𝑃5ket𝐺\mathrm{Im}[\bra{G}P_{0}P_{1}\mathrm{e}^{\mathrm{i}\hat{\mathcal{H}}(t+\tau)}P% _{2}\,\mathrm{e}^{-\mathrm{i}\hat{\mathcal{H}}t}P_{3}\,\mathrm{e}^{-\mathrm{i}% \hat{\mathcal{H}}\tau}P_{4}P_{5}\ket{G}]roman_Im [ ⟨ start_ARG italic_G end_ARG | italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT roman_e start_POSTSUPERSCRIPT roman_i over^ start_ARG caligraphic_H end_ARG ( italic_t + italic_τ ) end_POSTSUPERSCRIPT italic_P start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT roman_e start_POSTSUPERSCRIPT - roman_i over^ start_ARG caligraphic_H end_ARG italic_t end_POSTSUPERSCRIPT italic_P start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT roman_e start_POSTSUPERSCRIPT - roman_i over^ start_ARG caligraphic_H end_ARG italic_τ end_POSTSUPERSCRIPT italic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT | start_ARG italic_G end_ARG ⟩ ] is obtained by Pauli-Z𝑍Zitalic_Z measurement on the ancilla qubit. (b) The CUL circuit to measure the third-order susceptibility, which is an adaptive variational circuit equivalent to (a). The ground state is prepared using a parameterized circuit UG[𝜽1]subscript𝑈Gdelimited-[]superscript𝜽1U_{\mathrm{G}}[\boldsymbol{\theta}^{1}]italic_U start_POSTSUBSCRIPT roman_G end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ], the state propagation by ei^τsuperscriptei^𝜏\mathrm{e}^{-\mathrm{i}\hat{\mathcal{H}}\tau}roman_e start_POSTSUPERSCRIPT - roman_i over^ start_ARG caligraphic_H end_ARG italic_τ end_POSTSUPERSCRIPT is achieved by evolving the angles in Uτ[𝜽3]subscript𝑈𝜏delimited-[]superscript𝜽3U_{\tau}[\boldsymbol{\theta}^{3}]italic_U start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ] and UG[𝜽1]subscript𝑈Gdelimited-[]superscript𝜽1U_{\mathrm{G}}[\boldsymbol{\theta}^{1}]italic_U start_POSTSUBSCRIPT roman_G end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ], and state propagation by ei^tsuperscriptei^𝑡\mathrm{e}^{-\mathrm{i}\hat{\mathcal{H}}t}roman_e start_POSTSUPERSCRIPT - roman_i over^ start_ARG caligraphic_H end_ARG italic_t end_POSTSUPERSCRIPT is achieved by evolving angles in Ut[𝜽2]subscript𝑈𝑡delimited-[]superscript𝜽2U_{t}[\boldsymbol{\theta}^{2}]italic_U start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ], Uτ[𝜽3]subscript𝑈𝜏delimited-[]superscript𝜽3U_{\tau}[\boldsymbol{\theta}^{3}]italic_U start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ], and UG[𝜽1]subscript𝑈Gdelimited-[]superscript𝜽1U_{\mathrm{G}}[\boldsymbol{\theta}^{1}]italic_U start_POSTSUBSCRIPT roman_G end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ]. The parameterized unitaries UG[𝜽1]subscript𝑈Gdelimited-[]superscript𝜽1U_{\mathrm{G}}[\boldsymbol{\theta}^{1}]italic_U start_POSTSUBSCRIPT roman_G end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ], Ut[𝜽2]subscript𝑈𝑡delimited-[]superscript𝜽2U_{t}[\boldsymbol{\theta}^{2}]italic_U start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ], and Uτ[𝜽3]subscript𝑈𝜏delimited-[]superscript𝜽3U_{\tau}[\boldsymbol{\theta}^{3}]italic_U start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ] are automatically generated using adaptive variational algorithms.

The calculation of the different terms of eq 3.2 involves four main steps: (i) Preparation of the ground state of the quantum spin Hamiltonian, |G[𝜽1]=UG[𝜽1]|φ0ket𝐺delimited-[]superscript𝜽1subscript𝑈Gdelimited-[]superscript𝜽1ketsubscript𝜑0\ket{G[\boldsymbol{\theta}^{1}]}=U_{\mathrm{G}}[\boldsymbol{\theta}^{1}]\ket{% \varphi_{0}}| start_ARG italic_G [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ] end_ARG ⟩ = italic_U start_POSTSUBSCRIPT roman_G end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ] | start_ARG italic_φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ⟩ where |φ0ketsubscript𝜑0\ket{\varphi_{0}}| start_ARG italic_φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ⟩ is a reference product state, using adaptive variational algorithms like qubit ADAPT-VQE 77 or AVQITE 60. This step is analogous to step (i) for the calculation of the Green’s function using the quantum circuit in Figure 1 b. (ii) Generating and evolving the parameterized unitary circuit Uτ(𝜽3(τ))subscript𝑈𝜏superscript𝜽3𝜏U_{\tau}(\boldsymbol{\theta}^{3}(\tau))italic_U start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT ( bold_italic_θ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ( italic_τ ) ) by applying the AVQDS approach discussed in section 2 to propagate the quantum state

|Ψ=12ketΨ12\displaystyle\ket{\Psi}=\frac{1}{\sqrt{2}}| start_ARG roman_Ψ end_ARG ⟩ = divide start_ARG 1 end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG [|1P1P0|G[𝜽1(τ)]\displaystyle\left[\ket{1}\otimes P_{1}P_{0}\ket{G[\boldsymbol{\theta}^{1}(% \tau)]}\right.[ | start_ARG 1 end_ARG ⟩ ⊗ italic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT | start_ARG italic_G [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ( italic_τ ) ] end_ARG ⟩
+i|0P4P5|G[𝜽1(τ)]],\displaystyle\left.+\mathrm{i}\ket{0}\otimes P_{4}P_{5}\ket{G[\boldsymbol{% \theta}^{1}(\tau)]}\right]\,,+ roman_i | start_ARG 0 end_ARG ⟩ ⊗ italic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT | start_ARG italic_G [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ( italic_τ ) ] end_ARG ⟩ ] , (24)

where the parameters 𝜽1(τ)superscript𝜽1𝜏\boldsymbol{\theta}^{1}(\tau)bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ( italic_τ ) also evolve from the ground state solution to leverage their degrees of freedom for dynamics simulations. Here, eq 24 corresponds to the state in the circuit in Figure 2 b after applying the controlled P1subscript𝑃1P_{1}italic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT operation. (iii) Generating and evolving the parameterized unitary circuit Ut(𝜽2(t))subscript𝑈𝑡superscript𝜽2𝑡U_{t}(\boldsymbol{\theta}^{2}(t))italic_U start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ( bold_italic_θ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_t ) ) by applying the AVQDS approach to propagate the state

|Ψ=12ketΨ12\displaystyle\ket{\Psi}=\frac{1}{\sqrt{2}}| start_ARG roman_Ψ end_ARG ⟩ = divide start_ARG 1 end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG [|0Uτ(𝜽3(t))P1P0|G[𝜽1(t)]\displaystyle\left[\ket{0}\otimes U_{\tau}(\boldsymbol{\theta}^{3}(t))P_{1}P_{% 0}\ket{G[\boldsymbol{\theta}^{1}(t)]}\right.[ | start_ARG 0 end_ARG ⟩ ⊗ italic_U start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT ( bold_italic_θ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ( italic_t ) ) italic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT | start_ARG italic_G [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ( italic_t ) ] end_ARG ⟩
+i|1P3Uτ(𝜽3(t))P4P5|G[𝜽1(t)]],\displaystyle\left.+\mathrm{i}\ket{1}\otimes P_{3}U_{\tau}(\boldsymbol{\theta}% ^{3}(t))P_{4}P_{5}\ket{G[\boldsymbol{\theta}^{1}(t)]}\right]\,,+ roman_i | start_ARG 1 end_ARG ⟩ ⊗ italic_P start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT italic_U start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT ( bold_italic_θ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ( italic_t ) ) italic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT | start_ARG italic_G [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ( italic_t ) ] end_ARG ⟩ ] , (25)

where 𝜽1(t)superscript𝜽1𝑡\boldsymbol{\theta}^{1}(t)bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ( italic_t ) and 𝜽3(t)superscript𝜽3𝑡\boldsymbol{\theta}^{3}(t)bold_italic_θ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ( italic_t ) also vary to facilitate the state propagation using compact circuits. (iv) Repeat the quantum circuit evaluations in Figure 2 b for the different contributions in eq 3.2, which in total involves 4N4nαnβnγnδ4superscript𝑁4subscript𝑛𝛼subscript𝑛𝛽subscript𝑛𝛾subscript𝑛𝛿4N^{4}n_{\alpha}n_{\beta}n_{\gamma}n_{\delta}4 italic_N start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT italic_n start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT terms.

4 Applications

4.1 Single-particle Green’s function of Fermi-Hubbard chains

Model

To benchmark the performance of the AVQDS approach in calculating the Green’s function as outlined in section 3.1 and to compare the results with alternative approaches in refs 51, 62, 56, we study the one-dimensional Fermi-Hubbard model of N𝑁Nitalic_N sites in its particle-hole-symmetric form:

^^\displaystyle\hat{\mathcal{H}}over^ start_ARG caligraphic_H end_ARG =ti,j,σ(c^i,σc^j,σ+h.c.)\displaystyle=-t\sum_{\langle i,j\rangle,\sigma}\left(\hat{c}^{\dagger}_{i,% \sigma}\hat{c}^{\phantom{\dagger}}_{j,\sigma}+\mathrm{h.c.}\right)= - italic_t ∑ start_POSTSUBSCRIPT ⟨ italic_i , italic_j ⟩ , italic_σ end_POSTSUBSCRIPT ( over^ start_ARG italic_c end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i , italic_σ end_POSTSUBSCRIPT over^ start_ARG italic_c end_ARG start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j , italic_σ end_POSTSUBSCRIPT + roman_h . roman_c . )
+Ujnj,nj,U2j,σnj,σ.𝑈subscript𝑗subscript𝑛𝑗subscript𝑛𝑗𝑈2subscript𝑗𝜎subscript𝑛𝑗𝜎\displaystyle+U\sum_{j}n_{j,\uparrow}n_{j,\downarrow}-\frac{U}{2}\sum_{j,% \sigma}n_{j,\sigma}\,.+ italic_U ∑ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_j , ↑ end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_j , ↓ end_POSTSUBSCRIPT - divide start_ARG italic_U end_ARG start_ARG 2 end_ARG ∑ start_POSTSUBSCRIPT italic_j , italic_σ end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_j , italic_σ end_POSTSUBSCRIPT . (26)

Here, c^j,σsubscriptsuperscript^𝑐absent𝑗𝜎\hat{c}^{\phantom{\dagger}}_{j,\sigma}over^ start_ARG italic_c end_ARG start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j , italic_σ end_POSTSUBSCRIPT denotes the annihilation operator for a fermion with spin σ𝜎\sigmaitalic_σ at site j𝑗jitalic_j, nj,σsubscript𝑛𝑗𝜎n_{j,\sigma}italic_n start_POSTSUBSCRIPT italic_j , italic_σ end_POSTSUBSCRIPT represents the number operator, t𝑡titalic_t denotes the hopping amplitude between neighboring sites, and U𝑈Uitalic_U corresponds to the on-site interaction strength. We focus on N=4𝑁4N=4italic_N = 4 and N=6𝑁6N=6italic_N = 6-site chains with open boundary conditions. Specifically, we set the hopping parameter t=1𝑡1t=1italic_t = 1 and the on-site interaction is fixed at U=4.0𝑈4.0U=4.0italic_U = 4.0. To map Hamiltonian eq 4.1 to multi-qubit operators, we utilize the Jordan-Wigner transformation 67. The encoding of Hamiltonian eq 4.1 requires Nq=8subscript𝑁𝑞8N_{q}=8italic_N start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT = 8 and Nq=12subscript𝑁𝑞12N_{q}=12italic_N start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT = 12 qubits for N=4𝑁4N=4italic_N = 4 and N=6𝑁6N=6italic_N = 6-site systems, respectively.

Refer to caption
Figure 3: Numerical simulation of the AVQDS approach for computing the single-particle Green’s function of Fermi-Hubbard model. Examples of Iα,βp,q(t)subscriptsuperscript𝐼𝑝𝑞𝛼𝛽𝑡I^{p,q}_{\alpha,\beta}(t)italic_I start_POSTSUPERSCRIPT italic_p , italic_q end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α , italic_β end_POSTSUBSCRIPT ( italic_t ) dynamics for six different combinations of p𝑝pitalic_p, q𝑞qitalic_q, α𝛼\alphaitalic_α, and β𝛽\betaitalic_β, obtained by evaluating the quantum circuit in Figure 1 b for Fermi-Hubbard chains with (a) N=4𝑁4N=4italic_N = 4- and (b) N=6𝑁6N=6italic_N = 6-sites. The results obtained with the AVQDS approach (solid lines) are compared with those of the exact simulations (black dashed lines) obtained via exact diagonalization eq 8. The corresponding infidelities 1f1𝑓1-f1 - italic_f in (c) and (d) demonstrate the high accuracy of AVQDS in calculating the Green’s function components Iα,βp,q(t)subscriptsuperscript𝐼𝑝𝑞𝛼𝛽𝑡I^{p,q}_{\alpha,\beta}(t)italic_I start_POSTSUPERSCRIPT italic_p , italic_q end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α , italic_β end_POSTSUBSCRIPT ( italic_t ), achieving a fidelity of at least 99.93 %percent\%% for N=4𝑁4N=4italic_N = 4 and 99.64 %percent\%% for N=6𝑁6N=6italic_N = 6. The corresponding number of CNOT gates in (e) and (f) increases from an initial count of 270 (2402) to a maximum of 610 (6148) at the final simulation time of t=10𝑡10t=10italic_t = 10 for N=4𝑁4N=4italic_N = 4 (N=6𝑁6N=6italic_N = 6). The circuit depth in (g) and (h) grows from 26 (147) to a maximum circuit depth of 65 (424) at t=10𝑡10t=10italic_t = 10 for N=4𝑁4N=4italic_N = 4 (N=6𝑁6N=6italic_N = 6). Note that the pair I1,1(0,),(0,)subscriptsuperscript𝐼0011I^{(0,\uparrow),(0,\uparrow)}_{1,1}italic_I start_POSTSUPERSCRIPT ( 0 , ↑ ) , ( 0 , ↑ ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 1 , 1 end_POSTSUBSCRIPT (green line) and I1,1(1,),(0,)subscriptsuperscript𝐼1011I^{(1,\uparrow),(0,\uparrow)}_{1,1}italic_I start_POSTSUPERSCRIPT ( 1 , ↑ ) , ( 0 , ↑ ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 1 , 1 end_POSTSUBSCRIPT (yellow line) as well as the pair I2,2(0,),(0,)subscriptsuperscript𝐼0022I^{(0,\uparrow),(0,\uparrow)}_{2,2}italic_I start_POSTSUPERSCRIPT ( 0 , ↑ ) , ( 0 , ↑ ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 2 , 2 end_POSTSUBSCRIPT (dark blue line) and I1,2(0,),(0,)subscriptsuperscript𝐼0012I^{(0,\downarrow),(0,\uparrow)}_{1,2}italic_I start_POSTSUPERSCRIPT ( 0 , ↓ ) , ( 0 , ↑ ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 1 , 2 end_POSTSUBSCRIPT (orange line) have the same Pβsubscript𝑃𝛽P_{\beta}italic_P start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT but different Pαsubscript𝑃𝛼P_{\alpha}italic_P start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT in the circuit of Figure 1 b. As a result, the circuits for measuring these pairs involve exactly the same parameterized unitaries UG[𝜽1]subscript𝑈Gdelimited-[]superscript𝜽1U_{\mathrm{G}}[\boldsymbol{\theta}^{1}]italic_U start_POSTSUBSCRIPT roman_G end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ] and Ut[𝜽2]subscript𝑈𝑡delimited-[]superscript𝜽2U_{t}[\boldsymbol{\theta}^{2}]italic_U start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] in Figure 1 b, and therefore the same evolution of the number of CNOTs, circuit depth, and infidelity. However, the dynamics of these pairs in (a) and (b) are distinct due to different Pαsubscript𝑃𝛼P_{\alpha}italic_P start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT.

Ground state preparation

The ground state is prepared using the AVQITE approach 60, where the distance between the actual imaginary time-evolved state and the variational ansatz state is minimized in accordance with McLachlan’s variational principle 64. Analogous to AVQDS, the McLachlan distance is kept below a predefined threshold during the imaginary time evolution by adaptively appending new parameterized unitaries to the variational ansatz. The generators of the unitaries are chosen from a predetermined operator pool. The obtained ground state ansatz also takes the pseudo-Trotter form eq 7, allowing straightforward combination with the AVQDS approach.

In our simulations, we adopt an operator pool derived from the unitary coupled-cluster singles and doubles excitation operators 77, 78, or the related qubit excitation operators 79. Specifically, the pool can be written as:

𝒫={σ^ipσ^jq}{σ^ipσ^jqσ^krσ^ls},𝒫superscriptsubscript^𝜎𝑖𝑝superscriptsubscript^𝜎𝑗𝑞superscriptsubscript^𝜎𝑖𝑝superscriptsubscript^𝜎𝑗𝑞superscriptsubscript^𝜎𝑘𝑟superscriptsubscript^𝜎𝑙𝑠\displaystyle\mathcal{P}=\{\hat{\sigma}_{i}^{p}\hat{\sigma}_{j}^{q}\}\cup\{% \hat{\sigma}_{i}^{p}\hat{\sigma}_{j}^{q}\hat{\sigma}_{k}^{r}\hat{\sigma}_{l}^{% s}\}\,,caligraphic_P = { over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_p end_POSTSUPERSCRIPT over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_q end_POSTSUPERSCRIPT } ∪ { over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_p end_POSTSUPERSCRIPT over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_q end_POSTSUPERSCRIPT over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT } , (27)

subject to the constraint that only Pauli strings with odd number of σ^ysuperscript^𝜎𝑦\hat{\sigma}^{y}over^ start_ARG italic_σ end_ARG start_POSTSUPERSCRIPT italic_y end_POSTSUPERSCRIPT are included. Here p,q,r,s{x,y}𝑝𝑞𝑟𝑠𝑥𝑦p,q,r,s\in\{x,y\}italic_p , italic_q , italic_r , italic_s ∈ { italic_x , italic_y }, i,j,k,l𝑖𝑗𝑘𝑙i,j,k,litalic_i , italic_j , italic_k , italic_l run over all qubits, and σ^xsuperscript^𝜎𝑥\hat{\sigma}^{x}over^ start_ARG italic_σ end_ARG start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT and σ^ysuperscript^𝜎𝑦\hat{\sigma}^{y}over^ start_ARG italic_σ end_ARG start_POSTSUPERSCRIPT italic_y end_POSTSUPERSCRIPT represent Pauli operators. The pool comprises Np=616subscript𝑁p616N_{\mathrm{p}}=616italic_N start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT = 616 generators for the model with N=4𝑁4N=4italic_N = 4 and Np=4092subscript𝑁p4092N_{\mathrm{p}}=4092italic_N start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT = 4092 generators for N=6𝑁6N=6italic_N = 6. Regarding the reference state |φ0ketsubscript𝜑0\ket{\varphi_{0}}| start_ARG italic_φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ⟩, we opt for a product state with spin-up electrons occupying fermionic orbital sites 1,,N/21𝑁21,\cdots,N/21 , ⋯ , italic_N / 2 and spin-down electrons occupying sites N/2+1,,N𝑁21𝑁N/2+1,\cdots,Nitalic_N / 2 + 1 , ⋯ , italic_N, which is subsequently converted to a product state in the qubit representation. Although there are many other forms of reference states 80, such as a product state from Hartree-Fock or qubit mean-field calculations 81, this choice is consistent with ref 62, which allows a fair comparison on quantum resources between the CUR 62 and our CUL approaches.

Simulation results

We first analyze the performance and quantum resource requirements for the algorithm outlined in section 3.1. For the operator pool in the AVQDS calculations, we utilize the Hamiltonian operator pool, which encompasses all individual Pauli strings contained in the qubit representation of the Hamiltonian. As a result, the pool comprises Np=16subscript𝑁p16N_{\mathrm{p}}=16italic_N start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT = 16 (Np=26subscript𝑁p26N_{\mathrm{p}}=26italic_N start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT = 26) operators for N=4𝑁4N=4italic_N = 4 (N=6𝑁6N=6italic_N = 6) model. To compute the Green’s function Gp,qR(t)subscriptsuperscript𝐺R𝑝𝑞𝑡G^{\mathrm{R}}_{p,q}(t)italic_G start_POSTSUPERSCRIPT roman_R end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_p , italic_q end_POSTSUBSCRIPT ( italic_t ) using eq 11, the quantum circuit in Figure 1 b needs to be calculated for all α,β𝛼𝛽\alpha,\betaitalic_α , italic_β-components in eq 11. Since the expression of the creation or annihilation operators according to eq 10 consists of n=2𝑛2n=2italic_n = 2 terms after the Jordan-Wigner transformation, computing the Green’s function via eq 11 involves n2=4superscript𝑛24n^{2}=4italic_n start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = 4 simulations of the quantum circuit in Figure 1 b for a given p,q𝑝𝑞p,qitalic_p , italic_q-component.

Figure 3 a and 3 b show examples of the Iα,βp,qsubscriptsuperscript𝐼𝑝𝑞𝛼𝛽I^{p,q}_{\alpha,\beta}italic_I start_POSTSUPERSCRIPT italic_p , italic_q end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α , italic_β end_POSTSUBSCRIPT dynamics for different combinations of p𝑝pitalic_p, q𝑞qitalic_q, α𝛼\alphaitalic_α, and β𝛽\betaitalic_β, obtained by evaluating the quantum circuit presented in Figure 1 b for the N=4𝑁4N=4italic_N = 4- and N=6𝑁6N=6italic_N = 6-site Fermi-Hubbard chains. The results of the AVQDS approach (solid lines) are plotted together with the corresponding outcomes of the exact simulation (black dashed lines) obtained via exact diagonalization eq 8. The AVQDS results closely align with those of the exact simulations for all presented Iα,βp,qsubscriptsuperscript𝐼𝑝𝑞𝛼𝛽I^{p,q}_{\alpha,\beta}italic_I start_POSTSUPERSCRIPT italic_p , italic_q end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α , italic_β end_POSTSUBSCRIPT. To quantify the precision of the AVQDS calculations, we plot the corresponding infidelity in Figure 3 c and 3 d, defined as 1f=1|Ψ[𝜽(t)]|Ψexact(t)|21𝑓1superscriptinner-productΨdelimited-[]𝜽𝑡subscriptΨexact𝑡21-f=1-|\langle\Psi[\boldsymbol{\theta}(t)]|\Psi_{\mathrm{exact}}(t)\rangle|^{2}1 - italic_f = 1 - | ⟨ roman_Ψ [ bold_italic_θ ( italic_t ) ] | roman_Ψ start_POSTSUBSCRIPT roman_exact end_POSTSUBSCRIPT ( italic_t ) ⟩ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. Here, |Ψexact[t]=Iei^t|ΨketsubscriptΨexactdelimited-[]𝑡tensor-product𝐼superscriptei^𝑡ketΨ|\Psi_{\mathrm{exact}}[t]\rangle=I\otimes\mathrm{e}^{-\mathrm{i}\hat{\mathcal{% H}}t}\ket{\Psi}| roman_Ψ start_POSTSUBSCRIPT roman_exact end_POSTSUBSCRIPT [ italic_t ] ⟩ = italic_I ⊗ roman_e start_POSTSUPERSCRIPT - roman_i over^ start_ARG caligraphic_H end_ARG italic_t end_POSTSUPERSCRIPT | start_ARG roman_Ψ end_ARG ⟩ with |ΨketΨ\ket{\Psi}| start_ARG roman_Ψ end_ARG ⟩ defined in eq 13 is calculated using exact diagonalization eq 8, while |Ψ[𝜽(t)]ketΨdelimited-[]𝜽𝑡|\Psi[\boldsymbol{\theta}(t)]\rangle| roman_Ψ [ bold_italic_θ ( italic_t ) ] ⟩ represents the corresponding variational wavefunction obtained through AVQDS. The infidelity remains below 7.1×1047.1superscript1047.1\times 10^{-4}7.1 × 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT (3.6×103)3.6\times 10^{-3})3.6 × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT ) within the studied time window for the N=4𝑁4N=4italic_N = 4-site (N=6𝑁6N=6italic_N = 6-site) simulations. This demonstrates that on a statevector simulator AVQDS can accurately calculate the real-time Green’s function components Iα,βp,qsubscriptsuperscript𝐼𝑝𝑞𝛼𝛽I^{p,q}_{\alpha,\beta}italic_I start_POSTSUPERSCRIPT italic_p , italic_q end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α , italic_β end_POSTSUBSCRIPT using the quantum circuit presented in Figure 1 b.

The above results are obtained from noiseless statevector simulations and therefore neglect the effects of gate errors as well as finite sampling of the quantum state when measuring the matrix M𝑀Mitalic_M and vector V𝑉Vitalic_V in eqs 2 and 4. Thus, it is important to estimate the quantum resources necessary for AVQDS calculations on noisy intermediate-scale quantum devices. Our focus here is on the quantum circuit complexity quantified by the number of two-qubit CNOT gates. To simplify the analysis, we assume full qubit connectivity, as found in trapped-ion architecture, where implementing the multi-qubit rotation gate eiθ𝒜^superscriptei𝜃^𝒜\mathrm{e}^{-\mathrm{i}\theta\hat{\mathcal{A}}}roman_e start_POSTSUPERSCRIPT - roman_i italic_θ over^ start_ARG caligraphic_A end_ARG end_POSTSUPERSCRIPT demands 2(p1)2𝑝12(p-1)2 ( italic_p - 1 ) CNOT gates for a Pauli string 𝒜^^𝒜\hat{\mathcal{A}}over^ start_ARG caligraphic_A end_ARG containing p𝑝pitalic_p Pauli operators. Figure 3 e and 3 f show the growth of the CNOT number with time for the Iα,βp,qsubscriptsuperscript𝐼𝑝𝑞𝛼𝛽I^{p,q}_{\alpha,\beta}italic_I start_POSTSUPERSCRIPT italic_p , italic_q end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α , italic_β end_POSTSUBSCRIPT simulations presented in Figure 3 a and 3 b. The AVQITE ground-state preparation achieves an infidelity of approximately 106superscript10610^{-6}10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT and 3.6×1053.6superscript1053.6\times 10^{-5}3.6 × 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT for N=4𝑁4N=4italic_N = 4 and N=6𝑁6N=6italic_N = 6, respectively. The number of variational parameters Nθsubscript𝑁𝜃N_{\theta}italic_N start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT required for the ground state preparation is 65 for N=4𝑁4N=4italic_N = 4 and and 551 for N=6𝑁6N=6italic_N = 6. Specifically, the associated variational ansatz eq 7 incorporates 30 two- and 35 four-qubit rotation gates for N=4𝑁4N=4italic_N = 4, resulting in 270 CNOTs, and 226 two- and 325 four-qubit rotation gates for N=6𝑁6N=6italic_N = 6, yielding 2402 CNOTs. The number of CNOTs increases rapidly up to t3𝑡3t\approx 3italic_t ≈ 3 for the N=4𝑁4N=4italic_N = 4 simulations, before saturating at prolonged duration, while it exhibits a sublinear increase during the entire simulation up to t=10𝑡10t=10italic_t = 10 for the N=6𝑁6N=6italic_N = 6 simulations. The number of CNOTs rises from the initial 270 (2402) for ground state preparation to a maximum number of 610 (6148) at the final simulation time of t=10𝑡10t=10italic_t = 10 for N=4𝑁4N=4italic_N = 4 (N=6𝑁6N=6italic_N = 6). The resulting number of CNOTs at the final simulation time is comparable for the different components Iα,βp,qsubscriptsuperscript𝐼𝑝𝑞𝛼𝛽I^{p,q}_{\alpha,\beta}italic_I start_POSTSUPERSCRIPT italic_p , italic_q end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α , italic_β end_POSTSUBSCRIPT. In addition to the CNOT count and number of qubits, the circuit depth determined by the number of layers represents a critical metric for noisy intermediate-scale quantum devices. Figure 3 g and 3 h present the evolution of the circuit depth with time for the Iα,βp,qsubscriptsuperscript𝐼𝑝𝑞𝛼𝛽I^{p,q}_{\alpha,\beta}italic_I start_POSTSUPERSCRIPT italic_p , italic_q end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α , italic_β end_POSTSUBSCRIPT dynamics presented in Figure 3 a and 3 b. The depths of the state-preparation circuits are 26 and 147 for N=4𝑁4N=4italic_N = 4 and N=6𝑁6N=6italic_N = 6, respectively. During the time evolution, the circuit depth grows to a maximum of 65 (424) at the final simulation time of t=10𝑡10t=10italic_t = 10 for N=4𝑁4N=4italic_N = 4 (N=6𝑁6N=6italic_N = 6). The observed saturation of the circuit depth at long simulation times for N=4𝑁4N=4italic_N = 4 and sublinear increase for N=6𝑁6N=6italic_N = 6 point to a significant advantage compared to Trotterization, where the circuit depth is determined by the number and weight of the Hamiltonian terms and grows linearly with the number of time steps 59, 63. We also note that on today’s quantum processors, circuits of up to 60 layers of two-qubit gates can be successfully executed, as demonstrated in ref82, which is close to the circuit depth required for simulating the Green’s function of the four-site Hubbard model in Figure 3.

For comparison, if variational quantum dynamics simulation with HVA is used in the CUL Green’s function calculations, the number of two-qubit rotation gates required for a single layer HVA is 8N3/2+N4N8superscript𝑁32𝑁4𝑁8N^{3/2}+N-4\sqrt{N}8 italic_N start_POSTSUPERSCRIPT 3 / 2 end_POSTSUPERSCRIPT + italic_N - 4 square-root start_ARG italic_N end_ARG for the Fermi-Hubbard model with N𝑁Nitalic_N sites and open boundary conditions 51. For instance, with N=4𝑁4N=4italic_N = 4 and 16 layers of HVA as used in ref 51, the estimated number of two-qubit rotation gates is 960. Consequently, variational quantum dynamics simulation with a 16-layer HVA exceeds the maximum number of required CNOT gates of our AVQDS approach by approximately 50 %percent\%%, despite yielding less accurate results at long simulation times compared to our results based on the AVQDS approach. This demonstrates that the calculations of the Green’s function with AVQDS lead to much shallower quantum circuits relative to HVA. To compare with the CUR approach 62, we subtract the CNOTs in the ground state preparation circuit from the CNOT counts in Figure 3 e, as only the number of CNOTs accumulated during the time evolution is presented in ref 62. After this adjustment, the maximum increase in the CNOT count during the time evolution is 340 within the studied time window of [0,10]010[0,10][ 0 , 10 ], which is about the same as that in ref 62. We note that for the overlap test in CUR method, an additional 2Np2subscript𝑁p2N_{\textrm{p}}2 italic_N start_POSTSUBSCRIPT p end_POSTSUBSCRIPT CNOTs are needed, which is about 2×70=1402701402\times 70=1402 × 70 = 140. Here Npsubscript𝑁pN_{\textrm{p}}italic_N start_POSTSUBSCRIPT p end_POSTSUBSCRIPT is the number of parameters introduced in the single-state AVQDS calculations in the CUR approach. Comparable CNOT counts but deeper quantum circuits are reported for CUR calculations using HVA for variational state propagation 56.

Refer to caption
Figure 4: Single-particle Green’s function in momentum space and spectral function. Dynamics of the real and imaginary parts of GkR(t)subscriptsuperscript𝐺R𝑘𝑡G^{\mathrm{R}}_{k}(t)italic_G start_POSTSUPERSCRIPT roman_R end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_t ) at momentum k=0𝑘0k=0italic_k = 0 for Fermi-Hubbard model with (a) N=4𝑁4N=4italic_N = 4- and (b) N=6𝑁6N=6italic_N = 6-sites. The real part (red circles) and imaginary part (cyan diamonds) of Gk=0R(t)subscriptsuperscript𝐺R𝑘0𝑡G^{\mathrm{R}}_{k=0}(t)italic_G start_POSTSUPERSCRIPT roman_R end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT ( italic_t ) obtained with the AVQDS approach agree well with the corresponding results of the exact simulations (solid black lines). (c), (d) Spectral function Ak=0(ω)subscript𝐴𝑘0𝜔A_{k=0}(\omega)italic_A start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT ( italic_ω ) obtained by Fourier transforming the dynamics presented in (a) and (b) using the Padé approximation and calculating eq 29. The result is shown for three different tmaxsubscript𝑡maxt_{\mathrm{max}}italic_t start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT used within the Padé approximation. As a comparison, the exact result for the spectral function based on eq 4.1 is plotted as a shaded area. tmax=7subscript𝑡max7t_{\mathrm{max}}=7italic_t start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = 7 is sufficient to accurately reproduce the main features in Ak=0(ω)subscript𝐴𝑘0𝜔A_{k=0}(\omega)italic_A start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT ( italic_ω ). (e), (f) |Sν,0|2superscriptsubscript𝑆𝜈02|S_{\nu,0}|^{2}| italic_S start_POSTSUBSCRIPT italic_ν , 0 end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (black dots) and |S0,ν|2superscriptsubscript𝑆0𝜈2|S_{0,\nu}|^{2}| italic_S start_POSTSUBSCRIPT 0 , italic_ν end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (red dots) as a function of the energy differences E0Eνsubscript𝐸0subscript𝐸𝜈E_{0}-E_{\nu}italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT and EνE0subscript𝐸𝜈subscript𝐸0E_{\nu}-E_{0}italic_E start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, respectively, for (e) N=4𝑁4N=4italic_N = 4 and (f) N=6𝑁6N=6italic_N = 6. Only the dominant transition amplitudes with |Sν,μ|2>0.02superscriptsubscript𝑆𝜈𝜇20.02|S_{\nu,\mu}|^{2}>0.02| italic_S start_POSTSUBSCRIPT italic_ν , italic_μ end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT > 0.02 are shown. The peaks in the spectral functions originate from transitions between the ground state with energy E0subscript𝐸0E_{0}italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT to the excited states with energies Eνsubscript𝐸𝜈E_{\nu}italic_E start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT, and vice versa.

It is important to note that the quantum resource estimates for Green’s function calculations using the proposed quantum computing approach include the cost of ground state preparation as well as that of state propagation for the combined ancilla and physical qubit system. With a focus on near-term quantum algorithms, we utilize adaptive variational quantum algorithms for the calculations leveraging the problem-specific compact ansätze they automatically generate. Although the number of parameters needed for the variational wavefunction for large-size systems can be large, we emphasize that the adaptive variational algorithms we adopt only evolve the parameters along the real or imaginary time axis, without requiring explicit high-dimensional parameter optimizations. The adaptive algorithms construct and adjust compact ansätze along the dynamical path. This ensures that the exact state evolution can be followed closely, which often allows substantially more time steps to be taken than for a fixed ansatz with a restricted number of variational parameters.

Due to the nature of the adaptive algorithms, it is hard to derive a generic analytical upper bound for the required quantum resources and the practical performance may be problem-dependent. Nevertheless, substantial efforts have been devoted to numerically benchmarking the adaptive variational algorithms. For instance, highly compact ground state ansätze with much lower circuit depth than the problem-agnostic unitary coupled cluster ansatz have been demonstrated with AVQITE applications to molecular systems 60. AVQDS has been shown capable of simulating quantum dynamics with circuits of depth over two orders of magnitude less than standard Trotter circuits 59, 63. More importantly, the system-size scaling of the quantum circuit complexity characterized by the number of CNOT gates has also been numerically studied with a set of spin models, including high-spin systems. Although a definitive asymptotic scaling cannot be extracted due to limited system sizes, the numerical benchmarks suggest a polynomial growth of quantum circuit depth for larger-size systems, in contrast to the exponential scaling of exact diagonalization methods. The adaptive variational quantum algorithms can naturally leverage the computing power of quantum processing units, namely the efficient representation of the quantum many-body state and unitary state evolution. This renders the proposed quantum computing approach advantageous over classical computing for simulating quantum many-body systems.

Momentum-space Green’s function and spectral function

To further demonstrate the accuracy of CUL Green’s function calculations using AVQDS, we study the retarded Green’s function in momentum space, which has the following form:

Gk,σRsubscriptsuperscript𝐺R𝑘𝜎\displaystyle G^{\mathrm{R}}_{k,\sigma}italic_G start_POSTSUPERSCRIPT roman_R end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k , italic_σ end_POSTSUBSCRIPT =1Ni,j=1NG(i,σ),(j,σ)Reik(ij).absent1𝑁superscriptsubscript𝑖𝑗1𝑁subscriptsuperscript𝐺R𝑖𝜎𝑗𝜎superscriptei𝑘𝑖𝑗\displaystyle=\frac{1}{N}\sum_{i,j=1}^{N}G^{\mathrm{R}}_{(i,\sigma),(j,\sigma)% }\mathrm{e}^{-\mathrm{i}k(i-j)}\,.= divide start_ARG 1 end_ARG start_ARG italic_N end_ARG ∑ start_POSTSUBSCRIPT italic_i , italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT italic_G start_POSTSUPERSCRIPT roman_R end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ( italic_i , italic_σ ) , ( italic_j , italic_σ ) end_POSTSUBSCRIPT roman_e start_POSTSUPERSCRIPT - roman_i italic_k ( italic_i - italic_j ) end_POSTSUPERSCRIPT . (28)

To calculate eq 28 using eq 11, Iα,β(i,σ),(j,σ)(t)subscriptsuperscript𝐼𝑖𝜎𝑗𝜎𝛼𝛽𝑡I^{(i,\sigma),(j,\sigma)}_{\alpha,\beta}(t)italic_I start_POSTSUPERSCRIPT ( italic_i , italic_σ ) , ( italic_j , italic_σ ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α , italic_β end_POSTSUBSCRIPT ( italic_t ) is evaluated for all i𝑖iitalic_i, j𝑗jitalic_j, σ𝜎\sigmaitalic_σ, α𝛼\alphaitalic_α, and β𝛽\betaitalic_β. In general, the t𝑡titalic_t-stepping in the dynamics of Iα,β(i,σ),(j,σ)(t)subscriptsuperscript𝐼𝑖𝜎𝑗𝜎𝛼𝛽𝑡I^{(i,\sigma),(j,\sigma)}_{\alpha,\beta}(t)italic_I start_POSTSUPERSCRIPT ( italic_i , italic_σ ) , ( italic_j , italic_σ ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α , italic_β end_POSTSUBSCRIPT ( italic_t ) is not equidistant due to the dynamically adjusted δt𝛿𝑡\delta titalic_δ italic_t in the AVQDS approach and is also component-dependent. In practice, we obtain Iα,β(i,σ),(j,σ)(t)subscriptsuperscript𝐼𝑖𝜎𝑗𝜎𝛼𝛽𝑡I^{(i,\sigma),(j,\sigma)}_{\alpha,\beta}(t)italic_I start_POSTSUPERSCRIPT ( italic_i , italic_σ ) , ( italic_j , italic_σ ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α , italic_β end_POSTSUBSCRIPT ( italic_t ) on a uniform t𝑡titalic_t-mesh using linear interpolation before calculating Gk,σRsubscriptsuperscript𝐺R𝑘𝜎G^{\mathrm{R}}_{k,\sigma}italic_G start_POSTSUPERSCRIPT roman_R end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k , italic_σ end_POSTSUBSCRIPT via eq 28. For simplicity, we set GkRGk,σRsubscriptsuperscript𝐺R𝑘subscriptsuperscript𝐺R𝑘𝜎G^{\mathrm{R}}_{k}\equiv G^{\mathrm{R}}_{k,\sigma}italic_G start_POSTSUPERSCRIPT roman_R end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ≡ italic_G start_POSTSUPERSCRIPT roman_R end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k , italic_σ end_POSTSUBSCRIPT due to spin rotation symmetry. Figure 4 a and 4 b show the real and imaginary parts of GkR(t)subscriptsuperscript𝐺R𝑘𝑡G^{\mathrm{R}}_{k}(t)italic_G start_POSTSUPERSCRIPT roman_R end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_t ) at momentum k=0𝑘0k=0italic_k = 0 for Fermi-Hubbard chains with N=4𝑁4N=4italic_N = 4- and N=6𝑁6N=6italic_N = 6-sites, respectively. The real and imaginary parts of Gk=0Rsubscriptsuperscript𝐺R𝑘0G^{\mathrm{R}}_{k=0}italic_G start_POSTSUPERSCRIPT roman_R end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT obtained with the AVQDS approach are shown as red circles and cyan diamonds, respectively, which agree well with the exact simulation results represented by solid black lines.

We next study the spectral function, which is defined by

Ak(ω)=1πIm[GkR(ω)],subscript𝐴𝑘𝜔1𝜋Imdelimited-[]subscriptsuperscript𝐺R𝑘𝜔\displaystyle A_{k}(\omega)=-\frac{1}{\pi}\mathrm{Im}\left[G^{\mathrm{R}}_{k}(% \omega)\right]\,,italic_A start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_ω ) = - divide start_ARG 1 end_ARG start_ARG italic_π end_ARG roman_Im [ italic_G start_POSTSUPERSCRIPT roman_R end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_ω ) ] , (29)

where GkR(ω)subscriptsuperscript𝐺R𝑘𝜔G^{\mathrm{R}}_{k}(\omega)italic_G start_POSTSUPERSCRIPT roman_R end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_ω ) is the Fourier transform of GkR(t)subscriptsuperscript𝐺R𝑘𝑡G^{\mathrm{R}}_{k}(t)italic_G start_POSTSUPERSCRIPT roman_R end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_t ),

GkR(ω)=dtei(ω+iε)tGkR(t),subscriptsuperscript𝐺R𝑘𝜔superscriptsubscriptdifferential-d𝑡superscriptei𝜔i𝜀𝑡subscriptsuperscript𝐺R𝑘𝑡\displaystyle G^{\mathrm{R}}_{k}(\omega)=\int_{-\infty}^{\infty}\mathrm{d}t\,% \mathrm{e}^{\mathrm{i}(\omega+\mathrm{i}\varepsilon)t}G^{\mathrm{R}}_{k}(t)\,,italic_G start_POSTSUPERSCRIPT roman_R end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_ω ) = ∫ start_POSTSUBSCRIPT - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT roman_d italic_t roman_e start_POSTSUPERSCRIPT roman_i ( italic_ω + roman_i italic_ε ) italic_t end_POSTSUPERSCRIPT italic_G start_POSTSUPERSCRIPT roman_R end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_t ) , (30)

with infinitesimal ε𝜀\varepsilonitalic_ε to guarantee convergence of the integral. The calculation of eq 30 is challenging, as the accurate computation of the spectral function requires long simulation times. Reference 83 presents a quantum computing approach for accurately calculating spectral functions by reducing noise and extending time-domain results through denoising the imaginary time response functions using Hankel projections. Here, however, we adopt the Padé approximation for spectral analysis 84, 62, which has been demonstrated to accelerate the convergence of Fourier transforms with simulation time. To proceed, one first casts the discrete form of the Fourier transform eq 30 as a power series expansion:

GkR(ω)=n=0NTGkR(tn)zn,subscriptsuperscript𝐺R𝑘𝜔superscriptsubscript𝑛0subscript𝑁Tsubscriptsuperscript𝐺R𝑘subscript𝑡𝑛superscript𝑧𝑛\displaystyle G^{\mathrm{R}}_{k}(\omega)=\sum_{n=0}^{N_{\textrm{T}}}G^{\mathrm% {R}}_{k}(t_{n})z^{n}\,,italic_G start_POSTSUPERSCRIPT roman_R end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_ω ) = ∑ start_POSTSUBSCRIPT italic_n = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT T end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_G start_POSTSUPERSCRIPT roman_R end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) italic_z start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT , (31)

where tn=nδtsubscript𝑡𝑛𝑛𝛿𝑡t_{n}=n\delta titalic_t start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = italic_n italic_δ italic_t is the n𝑛nitalic_nth point in a uniform time mesh [0,T]0𝑇[0,T][ 0 , italic_T ] with time step size δt=T/NT𝛿𝑡𝑇subscript𝑁T\delta t=T/N_{\textrm{T}}italic_δ italic_t = italic_T / italic_N start_POSTSUBSCRIPT T end_POSTSUBSCRIPT with NTsubscript𝑁TN_{\textrm{T}}italic_N start_POSTSUBSCRIPT T end_POSTSUBSCRIPT even, and zn=(ei(ω+iε)δt)nsuperscript𝑧𝑛superscriptsuperscriptei𝜔i𝜀𝛿𝑡𝑛z^{n}=\left(\mathrm{e}^{\mathrm{i}(\omega+\mathrm{i}\varepsilon)\delta t}% \right)^{n}italic_z start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT = ( roman_e start_POSTSUPERSCRIPT roman_i ( italic_ω + roman_i italic_ε ) italic_δ italic_t end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT. We choose a diagonal Padé approximation 85, where GkR(ω)subscriptsuperscript𝐺R𝑘𝜔G^{\mathrm{R}}_{k}(\omega)italic_G start_POSTSUPERSCRIPT roman_R end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_ω ) is expressed as a ratio of two polynomials of equal order:

GkR(ω)=n=0NT/2anzn1+n=1NT/2bnzn.subscriptsuperscript𝐺R𝑘𝜔superscriptsubscript𝑛0subscript𝑁T2subscript𝑎𝑛superscript𝑧𝑛1superscriptsubscript𝑛1subscript𝑁T2subscript𝑏𝑛superscript𝑧𝑛\displaystyle G^{\mathrm{R}}_{k}(\omega)=\frac{\sum_{n=0}^{N_{\textrm{T}}/2}a_% {n}z^{n}}{1+\sum_{n=1}^{N_{\textrm{T}}/2}b_{n}z^{n}}\,.italic_G start_POSTSUPERSCRIPT roman_R end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_ω ) = divide start_ARG ∑ start_POSTSUBSCRIPT italic_n = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT T end_POSTSUBSCRIPT / 2 end_POSTSUPERSCRIPT italic_a start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_z start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_ARG start_ARG 1 + ∑ start_POSTSUBSCRIPT italic_n = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT T end_POSTSUBSCRIPT / 2 end_POSTSUPERSCRIPT italic_b start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_z start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_ARG . (32)

The coefficients of these polynomials, ansubscript𝑎𝑛a_{n}italic_a start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT and bnsubscript𝑏𝑛b_{n}italic_b start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT, are obtained by solving the system of linear equations obtained from matching orders of znsuperscript𝑧𝑛z^{n}italic_z start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT in eqs 31 and 32, as implemented in SciPy 86. The resulting rational function eq 32 can then be used as an approximation to the original function. Since the coefficients are independent of frequency, GkR(ω)subscriptsuperscript𝐺R𝑘𝜔G^{\mathrm{R}}_{k}(\omega)italic_G start_POSTSUPERSCRIPT roman_R end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_ω ) can be calculated for any frequency based on eq 32, in contrast to the Fast Fourier transform (FFT) where the spectral resolution is determined by the maximum simulation time tmaxsubscript𝑡maxt_{\mathrm{max}}italic_t start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT. However, the accuracy of the Padé approximant depends on tmaxsubscript𝑡maxt_{\mathrm{max}}italic_t start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT, as illustrated below.

Figure 4 c and 4 d show the spectral function Ak=0(ω)subscript𝐴𝑘0𝜔A_{k=0}(\omega)italic_A start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT ( italic_ω ) obtained by calculating the Fourier transform of the real-time components presented in Figure 4 a and 4 b, using the Padé approximation with a damping factor of ε=0.3𝜀0.3\varepsilon=0.3italic_ε = 0.3 (tied to tmaxsubscript𝑡maxt_{\mathrm{max}}italic_t start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT) and evaluating eq 29. The results are shown for three different values of tmaxsubscript𝑡maxt_{\mathrm{max}}italic_t start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT used in the Padé approximation. For comparison, we also present the exact result for the spectral function (shaded area), obtained using the Lehmann representation of the Green’s function, as derived in Appendix B. The simulated spectral function based on the AVQDS approach agrees well with the exact Ak=0(ω)subscript𝐴𝑘0𝜔A_{k=0}(\omega)italic_A start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT ( italic_ω ). In particular, a simulation time of tmax=7subscript𝑡max7t_{\mathrm{max}}=7italic_t start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = 7 is already sufficient to accurately reproduce the main features in Ak=0(ω)subscript𝐴𝑘0𝜔A_{k=0}(\omega)italic_A start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT ( italic_ω ) for the damping factor ε=0.3𝜀0.3\varepsilon=0.3italic_ε = 0.3 considered in the Fourier transformation eq 30. Compared to the discrete Fourier transform eq 31, the Padé approximation requires a smaller tmaxsubscript𝑡maxt_{\mathrm{max}}italic_t start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT to accurately calculate the spectral function, as demonstrated in Appendix E.

To identify the origin of the peaks in the spectral function Ak=0(ω)subscript𝐴𝑘0𝜔A_{k=0}(\omega)italic_A start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT ( italic_ω ), we study the Lehmann representation of the Green’s function in more detail. For k=0𝑘0k=0italic_k = 0, the Lehmann representation of the Green’s function eq B derived in Appendix B simplifies to

Gk=0R(ω)=1Nν[|S0,ν|2E0Eν+ω+iε\displaystyle G^{\mathrm{R}}_{k=0}(\omega)=\frac{1}{N}\sum_{\nu}\left[\frac{|S% _{0,\nu}|^{2}}{E_{0}-E_{\nu}+\omega+\mathrm{i}\,\varepsilon}\right.italic_G start_POSTSUPERSCRIPT roman_R end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT ( italic_ω ) = divide start_ARG 1 end_ARG start_ARG italic_N end_ARG ∑ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT [ divide start_ARG | italic_S start_POSTSUBSCRIPT 0 , italic_ν end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT + italic_ω + roman_i italic_ε end_ARG
+|Sν,0|2EνE0+ω+iε],\displaystyle\qquad\qquad\qquad\qquad\left.+\frac{|S_{\nu,0}|^{2}}{E_{\nu}-E_{% 0}+\omega+\mathrm{i}\,\varepsilon}\right]\,,+ divide start_ARG | italic_S start_POSTSUBSCRIPT italic_ν , 0 end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_E start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + italic_ω + roman_i italic_ε end_ARG ] ,
Sν,μpTν,μp,subscript𝑆𝜈𝜇subscript𝑝subscriptsuperscript𝑇𝑝𝜈𝜇\displaystyle S_{\nu,\mu}\equiv\sum_{p}T^{p}_{\nu,\mu}\,,italic_S start_POSTSUBSCRIPT italic_ν , italic_μ end_POSTSUBSCRIPT ≡ ∑ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT italic_T start_POSTSUPERSCRIPT italic_p end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_ν , italic_μ end_POSTSUBSCRIPT , (33)

where Tμ,νpΨμ|c^p|Ψνsubscriptsuperscript𝑇𝑝𝜇𝜈brasubscriptΨ𝜇subscriptsuperscript^𝑐absent𝑝ketsubscriptΨ𝜈T^{p}_{\mu,\nu}\equiv\bra{\Psi_{\mu}}\hat{c}^{\phantom{\dagger}}_{p}\ket{\Psi_% {\nu}}italic_T start_POSTSUPERSCRIPT italic_p end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_μ , italic_ν end_POSTSUBSCRIPT ≡ ⟨ start_ARG roman_Ψ start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT end_ARG | over^ start_ARG italic_c end_ARG start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT | start_ARG roman_Ψ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG ⟩ are the transition matrix elements between eigenstates |ΨνketsubscriptΨ𝜈\ket{\Psi_{\nu}}| start_ARG roman_Ψ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG ⟩ and |ΨμketsubscriptΨ𝜇\ket{\Psi_{\mu}}| start_ARG roman_Ψ start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT end_ARG ⟩ of the Hamiltonian eq 4.1. From this expression, it is evident that the peaks in the spectral function arise from transitions between the ground state with energy E0subscript𝐸0E_{0}italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT to the excited states with energies Eνsubscript𝐸𝜈E_{\nu}italic_E start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT by adding an electron [first term in Gk=0Rsubscriptsuperscript𝐺R𝑘0G^{\mathrm{R}}_{k=0}italic_G start_POSTSUPERSCRIPT roman_R end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT, eq 4.1] or removing an electron [second term in Gk=0Rsubscriptsuperscript𝐺R𝑘0G^{\mathrm{R}}_{k=0}italic_G start_POSTSUPERSCRIPT roman_R end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT, eq 4.1], and vice versa. The spectral weight of these peaks in Ak=0(ω)subscript𝐴𝑘0𝜔A_{k=0}(\omega)italic_A start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT ( italic_ω ) is determined by Sμ,νsubscript𝑆𝜇𝜈S_{\mu,\nu}italic_S start_POSTSUBSCRIPT italic_μ , italic_ν end_POSTSUBSCRIPT. Figure 4 e and 4 f present |Sν,0|2superscriptsubscript𝑆𝜈02|S_{\nu,0}|^{2}| italic_S start_POSTSUBSCRIPT italic_ν , 0 end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (black dots) and |S0,ν|2superscriptsubscript𝑆0𝜈2|S_{0,\nu}|^{2}| italic_S start_POSTSUBSCRIPT 0 , italic_ν end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (red dots) as a function of the energy differences E0Eνsubscript𝐸0subscript𝐸𝜈E_{0}-E_{\nu}italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT and EνE0subscript𝐸𝜈subscript𝐸0E_{\nu}-E_{0}italic_E start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, respectively, for N=4𝑁4N=4italic_N = 4 (Figure 4 e) and N=6𝑁6N=6italic_N = 6 (Figure 4 f). In the case of N=4𝑁4N=4italic_N = 4, the energetically close peaks around ω2.2𝜔2.2\omega\approx-2.2italic_ω ≈ - 2.2 in the spectral function in Figure 4 c result from transitions between degenerate excited states ν=20,,23𝜈2023\nu=20,...,23italic_ν = 20 , … , 23 to the ground state as well as from the transition of the degenerate states ν=32,,35𝜈3235\nu=32,...,35italic_ν = 32 , … , 35 to the ground state. The states ν=20,,23𝜈2023\nu=20,...,23italic_ν = 20 , … , 23 have energy E0Eν=1.955subscript𝐸0subscript𝐸𝜈1.955E_{0}-E_{\nu}=-1.955italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT = - 1.955 while E0Eν=32,,35=2.55subscript𝐸0subscript𝐸𝜈32352.55E_{0}-E_{\nu=32,...,35}=-2.55italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT italic_ν = 32 , … , 35 end_POSTSUBSCRIPT = - 2.55. The dominant signal at positive frequencies, ω3.7𝜔3.7\omega\approx 3.7italic_ω ≈ 3.7, in Figure 4 c results from transitions between the ground state and degenerate excited states ν=51,,54𝜈5154\nu=51,...,54italic_ν = 51 , … , 54 with energy EνE0=3.71subscript𝐸𝜈subscript𝐸03.71E_{\nu}-E_{0}=3.71italic_E start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 3.71. For N=6𝑁6N=6italic_N = 6, the peaks around ω2.5𝜔2.5\omega\approx-2.5italic_ω ≈ - 2.5 in the spectral function in Figure 4 d stem from transitions between excited states ν=49,,52𝜈4952\nu=49,...,52italic_ν = 49 , … , 52, ν=72,,75𝜈7275\nu=72,...,75italic_ν = 72 , … , 75, and ν=115,,118𝜈115118\nu=115,...,118italic_ν = 115 , … , 118 to the ground state. The states ν=49,,52𝜈4952\nu=49,...,52italic_ν = 49 , … , 52 (ν=72,,75𝜈7275\nu=72,...,75italic_ν = 72 , … , 75) have energy E0Eν=1.98subscript𝐸0subscript𝐸𝜈1.98E_{0}-E_{\nu}=-1.98italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT = - 1.98 (E0Eν=2.38subscript𝐸0subscript𝐸𝜈2.38E_{0}-E_{\nu}=-2.38italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT = - 2.38) while E0Eν=115,,118=2.87subscript𝐸0subscript𝐸𝜈1151182.87E_{0}-E_{\nu=115,...,118}=-2.87italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT italic_ν = 115 , … , 118 end_POSTSUBSCRIPT = - 2.87. The dominant signal at positive frequencies, ω3.6𝜔3.6\omega\approx 3.6italic_ω ≈ 3.6, in Figure 4 d mainly results from transitions between the ground state and degenerate excited states ν=223,,226𝜈223226\nu=223,...,226italic_ν = 223 , … , 226 with energy EνE0=3.65subscript𝐸𝜈subscript𝐸03.65E_{\nu}-E_{0}=3.65italic_E start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 3.65. As a result, the spectral function provides direct insights into the energy levels and excitations of fermions in the studied Fermi-Hubbard model.

To characterize the excited states contributing to the dominant peaks in the spectral functions in more detail, we calculate the expectation values of the total spin operator S^2=(S^x)2+(S^y)2+(S^z)2superscript^𝑆2superscriptsuperscript^𝑆𝑥2superscriptsuperscript^𝑆𝑦2superscriptsuperscript^𝑆𝑧2\hat{S}^{2}=(\hat{S}^{x})^{2}+(\hat{S}^{y})^{2}+(\hat{S}^{z})^{2}over^ start_ARG italic_S end_ARG start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = ( over^ start_ARG italic_S end_ARG start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + ( over^ start_ARG italic_S end_ARG start_POSTSUPERSCRIPT italic_y end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + ( over^ start_ARG italic_S end_ARG start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT for these states. Here, the spin operators expressed in terms of fermionic creation (c^superscript^𝑐\hat{c}^{\dagger}over^ start_ARG italic_c end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT) and annihilation (c^superscript^𝑐absent\hat{c}^{\phantom{\dagger}}over^ start_ARG italic_c end_ARG start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT) operators are given by

S^α=i=1Ns,s=,c^i,s(σs,sα)c^i,s,superscript^𝑆𝛼superscriptsubscript𝑖1𝑁subscriptformulae-sequence𝑠superscript𝑠subscriptsuperscript^𝑐𝑖𝑠subscriptsuperscript𝜎𝛼𝑠superscript𝑠subscriptsuperscript^𝑐absent𝑖superscript𝑠\displaystyle\hat{S}^{\alpha}=\sum_{i=1}^{N}\sum_{s,s^{\prime}=\uparrow,% \downarrow}\hat{c}^{\dagger}_{i,s}\left(\sigma^{\alpha}_{s,s^{\prime}}\right)% \hat{c}^{\phantom{\dagger}}_{i,s^{\prime}},over^ start_ARG italic_S end_ARG start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT = ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_s , italic_s start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = ↑ , ↓ end_POSTSUBSCRIPT over^ start_ARG italic_c end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i , italic_s end_POSTSUBSCRIPT ( italic_σ start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_s , italic_s start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ) over^ start_ARG italic_c end_ARG start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i , italic_s start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT , (34)

where α=x,y,z𝛼𝑥𝑦𝑧\alpha=x,y,zitalic_α = italic_x , italic_y , italic_z denotes three spatial components, σαsuperscript𝜎𝛼\sigma^{\alpha}italic_σ start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT are the Pauli matrices, and i𝑖iitalic_i is a site index. The calculation of S^2=Ψν|S^2|Ψν=S(S+1)delimited-⟨⟩superscript^𝑆2brasubscriptΨ𝜈superscript^𝑆2ketsubscriptΨ𝜈𝑆𝑆1\langle\hat{S}^{2}\rangle=\bra{\Psi_{\nu}}\hat{S}^{2}\ket{\Psi_{\nu}}=S(S+1)⟨ over^ start_ARG italic_S end_ARG start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ = ⟨ start_ARG roman_Ψ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG | over^ start_ARG italic_S end_ARG start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT | start_ARG roman_Ψ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG ⟩ = italic_S ( italic_S + 1 ) then yields the total spin quantum number S𝑆Sitalic_S of the eigenstate |ΨνketsubscriptΨ𝜈\ket{\Psi_{\nu}}| start_ARG roman_Ψ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG ⟩. To determine S𝑆Sitalic_S, we obtain the eigenstates |ΨνketsubscriptΨ𝜈\ket{\Psi_{\nu}}| start_ARG roman_Ψ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG ⟩ via exact diagonalization of the Hamiltonian and then calculate Ψν|S^2|ΨνbrasubscriptΨ𝜈superscript^𝑆2ketsubscriptΨ𝜈\bra{\Psi_{\nu}}\hat{S}^{2}\ket{\Psi_{\nu}}⟨ start_ARG roman_Ψ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG | over^ start_ARG italic_S end_ARG start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT | start_ARG roman_Ψ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG ⟩ after transforming the fermionic operators in eq (34) to Pauli operators using the Jordan-Wigner transformation. To further characterize the excited states, we also calculate the particle number of the state |ΨνketsubscriptΨ𝜈\ket{\Psi_{\nu}}| start_ARG roman_Ψ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG ⟩, which follows from

Ne=i=1Nσ=,Ψν|c^i,σc^i,σ|Ψν.subscript𝑁𝑒superscriptsubscript𝑖1𝑁subscript𝜎brasubscriptΨ𝜈subscriptsuperscript^𝑐𝑖𝜎subscriptsuperscript^𝑐absent𝑖𝜎ketsubscriptΨ𝜈\displaystyle N_{e}=\sum_{i=1}^{N}\sum_{\sigma=\uparrow,\downarrow}\bra{\Psi_{% \nu}}\hat{c}^{\dagger}_{i,\sigma}\hat{c}^{\phantom{\dagger}}_{i,\sigma}\ket{% \Psi_{\nu}}\,.italic_N start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_σ = ↑ , ↓ end_POSTSUBSCRIPT ⟨ start_ARG roman_Ψ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG | over^ start_ARG italic_c end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i , italic_σ end_POSTSUBSCRIPT over^ start_ARG italic_c end_ARG start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i , italic_σ end_POSTSUBSCRIPT | start_ARG roman_Ψ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG ⟩ . (35)

For the N=4𝑁4N=4italic_N = 4 site Fermi-Hubbard chain, we find that four degenerate excited states contribute to each of the three dominant peaks in the spectral function (two dominant peaks at negative frequencies, ω1.96𝜔1.96\omega\approx-1.96italic_ω ≈ - 1.96 and ω2.55𝜔2.55\omega\approx-2.55italic_ω ≈ - 2.55, and one dominant peak at positive frequency, ω3.71𝜔3.71\omega\approx 3.71italic_ω ≈ 3.71). Two of these four excited states have particle number of Ne=3subscript𝑁𝑒3N_{e}=3italic_N start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT = 3, while the remaining two have Ne=5subscript𝑁𝑒5N_{e}=5italic_N start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT = 5. It is important to note that the peaks in the spectral function result from transitions between the ground state and excited states with one electron removed or added, as the ground state has Ne=4subscript𝑁𝑒4N_{e}=4italic_N start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT = 4 due to the considered half-filling. For these states, the possible values for S𝑆Sitalic_S are S=12𝑆12S=\frac{1}{2}italic_S = divide start_ARG 1 end_ARG start_ARG 2 end_ARG and S=32𝑆32S=\frac{3}{2}italic_S = divide start_ARG 3 end_ARG start_ARG 2 end_ARG for Ne=3subscript𝑁𝑒3N_{e}=3italic_N start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT = 3 and S=12𝑆12S=\frac{1}{2}italic_S = divide start_ARG 1 end_ARG start_ARG 2 end_ARG, S=32𝑆32S=\frac{3}{2}italic_S = divide start_ARG 3 end_ARG start_ARG 2 end_ARG, and S=52𝑆52S=\frac{5}{2}italic_S = divide start_ARG 5 end_ARG start_ARG 2 end_ARG for Ne=5subscript𝑁𝑒5N_{e}=5italic_N start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT = 5. The calculated S^2delimited-⟨⟩superscript^𝑆2\langle\hat{S}^{2}\rangle⟨ over^ start_ARG italic_S end_ARG start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ of all excited states contributing to the three dominant signals in the spectral function is S(S+1)=34𝑆𝑆134S(S+1)=\frac{3}{4}italic_S ( italic_S + 1 ) = divide start_ARG 3 end_ARG start_ARG 4 end_ARG. This corresponds to a spin quantum number S=12𝑆12S=\frac{1}{2}italic_S = divide start_ARG 1 end_ARG start_ARG 2 end_ARG and a degeneracy of 2S+1=22𝑆122S+1=22 italic_S + 1 = 2 for both the Ne=3subscript𝑁𝑒3N_{e}=3italic_N start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT = 3 and Ne=5subscript𝑁𝑒5N_{e}=5italic_N start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT = 5 states, giving a total degeneracy of four.

Similarly, for N=6𝑁6N=6italic_N = 6 sites, four states contribute to each of the four dominant peaks in the spectral function (three dominant peaks at negative frequencies (ω2.87,2.38,1.98𝜔2.872.381.98\omega\approx-2.87,-2.38,-1.98italic_ω ≈ - 2.87 , - 2.38 , - 1.98) and one dominant signal at positive frequency (ω3.65𝜔3.65\omega\approx 3.65italic_ω ≈ 3.65)). Two of these four excited states have a particle number of Ne=5subscript𝑁𝑒5N_{e}=5italic_N start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT = 5 while the remaining two have Ne=7subscript𝑁𝑒7N_{e}=7italic_N start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT = 7. These states also have a spin quantum number S=12𝑆12S=\frac{1}{2}italic_S = divide start_ARG 1 end_ARG start_ARG 2 end_ARG and a degeneracy of 2S+1=22𝑆122S+1=22 italic_S + 1 = 2 for both the Ne=5subscript𝑁𝑒5N_{e}=5italic_N start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT = 5 and Ne=7subscript𝑁𝑒7N_{e}=7italic_N start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT = 7 states, resulting in a total degeneracy of four.

Calculation of the spectral function with Prony approximation and compressive sensing

In the above Green’s function analysis using the Padé approximation 32, finite broadening with ε=0.3𝜀0.3\varepsilon=0.3italic_ε = 0.3 is applied to mimic the infinite lattice with continuous energy bands. However, rigorously speaking, our calculations being carried out for finite systems at zero temperature implies zero broadening for the energy levels. Therefore, it is interesting to investigate and compare approaches in addition to the Padé approximation for modelling delta-function peaks. Specifically, we include the Prony approximation 87, 88 and compressive sensing 89, 90. In the Prony approximation, a signal is decomposed into a sum of damped exponentials. When applied to the Green’s function G(t)𝐺𝑡G(t)italic_G ( italic_t ), this method allows the identification of the complex frequencies (poles) and corresponding residues that characterize the system’s spectral properties. Meanwhile, compressive sensing is a technique used to recover sparse signals from incomplete or noisy data. Detailed discussions of how these methods are employed to obtain the spectral function from the Green’s function in the time domain are given in Appendices F and G for the Prony approximation and compressive sensing, respectively.

Refer to caption
Figure 5: Comparison of different signal processing techniques for spectral function calculation. (a), (b) Spectral function Ak=0(ω)subscript𝐴𝑘0𝜔A_{k=0}(\omega)italic_A start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT ( italic_ω ) obtained by Fourier transforming the dynamics presented in Figure 4 a and 4 b, using a damping factor of ε=0.3𝜀0.3\varepsilon=0.3italic_ε = 0.3 in the Fourier transformation. The result of the Padé approximation (shaded area) is shown together with the results of Prony approximation (blue line) and compressive sensing (orange line). (c), (d) The corresponding results without broadening (ε=0𝜀0\varepsilon=0italic_ε = 0) for (c) N=4𝑁4N=4italic_N = 4 and (d) N=6𝑁6N=6italic_N = 6.

We compare the performances of the Padé approximation, the Prony approximation, and compressive sensing in calculating the spectral function Ak=0(ω)subscript𝐴𝑘0𝜔A_{k=0}(\omega)italic_A start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT ( italic_ω ) for the N=4𝑁4N=4italic_N = 4 and N=6𝑁6N=6italic_N = 6-site Fermi-Hubbard chains. For the Padé and Prony approximations we use uniform time and frequency meshes with Nt=501subscript𝑁𝑡501N_{t}=501italic_N start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT = 501 time steps and Nω=1001subscript𝑁𝜔1001N_{\omega}=1001italic_N start_POSTSUBSCRIPT italic_ω end_POSTSUBSCRIPT = 1001 frequency points. For the compressive sensing method we consider sparser uniform time and frequency meshes of sizes Nt=101subscript𝑁𝑡101N_{t}=101italic_N start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT = 101 and Nω=201subscript𝑁𝜔201N_{\omega}=201italic_N start_POSTSUBSCRIPT italic_ω end_POSTSUBSCRIPT = 201. Figure 5 a and 5 b show the spectral functions Ak=0(ω)subscript𝐴𝑘0𝜔A_{k=0}(\omega)italic_A start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT ( italic_ω ) obtained by calculating the Fourier transform of the real-time Green’s function components presented in Figure 4 a and 4 b, using a damping factor of ε=0.3𝜀0.3\varepsilon=0.3italic_ε = 0.3 in the Fourier transformation. The result of the Padé approximation (shaded area) is compared with the results of the Prony approximation (blue line) and compressive sensing (orange line). The corresponding results without broadening (ε=0𝜀0\varepsilon=0italic_ε = 0) are presented in Figure 5 c and 5 d. For finite broadening, all three methods produce similar results, with the spectral functions showing all dominant signals consistent with Figure 4 e and 4 f. Nevertheless, the results from compressive sensing exhibit multiple small artificial spikes. In contrast, for zero broadening (Figure 5 c and 5 d), compressive sensing produces the most accurate results for the delta peaks when compared to Figure 4 e and 4 f, highlighting its effectiveness in resolving sharp spectral features. In particular, in contrast to the Padé and Prony approximations, the compressive sensing method generates Ak=0(ω)subscript𝐴𝑘0𝜔A_{k=0}(\omega)italic_A start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT ( italic_ω ) with spectral weights more consistent with the exact analysis results shown in Figure 4 e and 4 f. For instance, considering the three main spectral peaks at ω=2.55,1.96,3.71𝜔2.551.963.71\omega=-2.55,-1.96,3.71italic_ω = - 2.55 , - 1.96 , 3.71 of N=4-site model, the ratio of spectral weights is 1:0.30:0.03 for the Padé approximation and 1:0.28:0.16 for the Prony approximation; while the ratio becomes 1:1.18:0.17 with compressive sensing, which is much closer to the ratio 1:1.21:0.25 obtained from the exact analysis. This demonstrates that the compressive sensing approach is most effective for resolving delta peaks.

4.2 Single-particle Green’s function of molecule LiH

The method introduced in section 3.1 can naturally be used to calculate the single-particle Green’s functions of molecules, providing valuable information about molecular ionization potential, electron affinities and excitation energies. To demonstrate this, we compute the Green’s function of the molecule LiH. We use the PySCF quantum chemistry package to generate the molecular Hamiltonian 91. We adopt the minimal STO-3G basis set, treating the 1s1𝑠1s1 italic_s-orbital of Li as a core orbital, with a bond length of 1.547 Å close to equilibrium. The resulting molecular Hamiltonian expressed in the Hartree-Fock molecular orbital basis consisting of 5 spatial orbitals with 2 electron filling, is transformed to the qubit representation using the Jordan-Wigner encoding requiring in total 10 qubits. Similar to calculations of the Fermi-Hubbard model, we use AVQITE to prepare the ground state for LiH. The AVQITE calculation adopts the unitary coupled-cluster singles and doubles excitation operator pool 27, which consists of Np=144subscript𝑁p144N_{\mathrm{p}}=144italic_N start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT = 144 generators. As reference state we choose the Hartree-Fock product state. For the operator pool in the AVQDS calculations, we consider the Hamiltonian operator pool, which comprises Np=276subscript𝑁p276N_{\mathrm{p}}=276italic_N start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT = 276 operators.

Figure 6 a presents examples of the simulated Iα,βp,qsubscriptsuperscript𝐼𝑝𝑞𝛼𝛽I^{p,q}_{\alpha,\beta}italic_I start_POSTSUPERSCRIPT italic_p , italic_q end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α , italic_β end_POSTSUBSCRIPT dynamics for five different components, where calculations using the adaptive variational quantum circuit (solid lines) presented in Figure 1 b agree very well with the reference results (black dashed lines) from exact diagonalization. The corresponding infidelity in Figure 6 b stays below 9.3×1059.3superscript1059.3\times 10^{-5}9.3 × 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT within the studied time window which confirms that AVQDS can be used to accurately calculate the real-time Green’s function components.

The required near-term quantum resources in terms of CNOT gates are plotted in Figure 6 c for the Iα,βp,qsubscriptsuperscript𝐼𝑝𝑞𝛼𝛽I^{p,q}_{\alpha,\beta}italic_I start_POSTSUPERSCRIPT italic_p , italic_q end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α , italic_β end_POSTSUBSCRIPT simulations presented in Figure 6 a. AVQITE prepares the ground state with an infidelity of 1f=1.8×1061𝑓1.8superscript1061-f=1.8\times 10^{-6}1 - italic_f = 1.8 × 10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT. The corresponding variational ansatz includes 6 two- and 11 four-qubit rotation gates, yielding 78 CNOT gates assuming full qubit connectivity. The number of CNOTs initially increases rapidly during the time evolution, and tends to saturate at later times. The maximal number of CNOT gates at the final simulation time is in the range of [640, 874] across all the components. The corresponding dynamics of the circuit depth is plotted in Figure 6 d. The depth of the ground state circuit is 14 and grows to 81 at the final simulation time of t=20𝑡20t=20italic_t = 20.

Refer to caption
Figure 6: Numerical simulation of the AVQDS approach for computing the single-particle Green’s function of the molecule LiH. (a) Examples of Iα,βp,q(t)subscriptsuperscript𝐼𝑝𝑞𝛼𝛽𝑡I^{p,q}_{\alpha,\beta}(t)italic_I start_POSTSUPERSCRIPT italic_p , italic_q end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α , italic_β end_POSTSUBSCRIPT ( italic_t ) dynamics for five different combinations of p𝑝pitalic_p, q𝑞qitalic_q, α𝛼\alphaitalic_α, and β𝛽\betaitalic_β (solid lines), obtained by evaluating the AVQDS quantum circuit in Figure 1 b. The exact simulations (black dashed lines) via exact diagonalization (eq 8) are also shown for comparison. (b) The corresponding infidelities 1f1𝑓1-f1 - italic_f illustrate the high accuracy of AVQDS in calculating the Green’s function components Iα,βp,q(t)subscriptsuperscript𝐼𝑝𝑞𝛼𝛽𝑡I^{p,q}_{\alpha,\beta}(t)italic_I start_POSTSUPERSCRIPT italic_p , italic_q end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α , italic_β end_POSTSUBSCRIPT ( italic_t ), with an infidelity below 9.3×1059.3superscript1059.3\times 10^{-5}9.3 × 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT during the studied time window. (c) The corresponding number of CNOT gates, which increases from an initial count of 78 to a maximum of 874 at the final simulation time of t=20𝑡20t=20italic_t = 20. (d) The corresponding circuit depth, which increases from 14 to a maximum circuit depth of 81 at t=20𝑡20t=20italic_t = 20.

We next present the trace of the spectral function of LiH, Tr[A(ω)]=1πTr[Im[GR(ω)]]Trdelimited-[]𝐴𝜔1𝜋Trdelimited-[]Imdelimited-[]superscript𝐺R𝜔\mathrm{Tr}\left[A(\omega)\right]=-\frac{1}{\pi}\mathrm{Tr}\left[\mathrm{Im}[G% ^{\mathrm{R}}(\omega)]\right]roman_Tr [ italic_A ( italic_ω ) ] = - divide start_ARG 1 end_ARG start_ARG italic_π end_ARG roman_Tr [ roman_Im [ italic_G start_POSTSUPERSCRIPT roman_R end_POSTSUPERSCRIPT ( italic_ω ) ] ], which represents the total density of states. Here, GR(ω)=dtei(ω+iε)tGR(t)superscript𝐺R𝜔superscriptsubscriptdifferential-d𝑡superscriptei𝜔i𝜀𝑡superscript𝐺R𝑡G^{\mathrm{R}}(\omega)=\int_{-\infty}^{\infty}\mathrm{d}t\,\mathrm{e}^{\mathrm% {i}(\omega+\mathrm{i}\,\varepsilon)t}G^{\mathrm{R}}(t)italic_G start_POSTSUPERSCRIPT roman_R end_POSTSUPERSCRIPT ( italic_ω ) = ∫ start_POSTSUBSCRIPT - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT roman_d italic_t roman_e start_POSTSUPERSCRIPT roman_i ( italic_ω + roman_i italic_ε ) italic_t end_POSTSUPERSCRIPT italic_G start_POSTSUPERSCRIPT roman_R end_POSTSUPERSCRIPT ( italic_t ) is a 10×10101010\times 1010 × 10 (including spin degeneracy) matrix at each ω𝜔\omegaitalic_ω point. Figure 7 a shows the result obtained with compressive sensing without broadening in the Fourier transformation. For comparison, in Figure 7 b we also show p|Mν,0p|2subscript𝑝superscriptsubscriptsuperscript𝑀𝑝𝜈02\sum_{p}|M^{p}_{\nu,0}|^{2}∑ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT | italic_M start_POSTSUPERSCRIPT italic_p end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_ν , 0 end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (black dots) and p|M0,νp|2subscript𝑝superscriptsubscriptsuperscript𝑀𝑝0𝜈2\sum_{p}|M^{p}_{0,\nu}|^{2}∑ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT | italic_M start_POSTSUPERSCRIPT italic_p end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 , italic_ν end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (red dots) as a function of the energy differences E0Eνsubscript𝐸0subscript𝐸𝜈E_{0}-E_{\nu}italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT and EνE0subscript𝐸𝜈subscript𝐸0E_{\nu}-E_{0}italic_E start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, respectively. The spectral function shows all dominant transitions present in Figure 7 b. Specifically, Tr[A(ω)]Trdelimited-[]𝐴𝜔\mathrm{Tr}\left[A(\omega)\right]roman_Tr [ italic_A ( italic_ω ) ] exhibits three dominant peaks. The peak around ω0.27𝜔0.27\omega\approx-0.27italic_ω ≈ - 0.27 Ha in the spectral function in Figure 7 a results from transitions between the ground state and doubly degenerate excited states with spin S=1/2𝑆12S=1/2italic_S = 1 / 2 and electron number of Ne=1subscript𝑁𝑒1N_{e}=1italic_N start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT = 1. For comparison, the ground state has S=0𝑆0S=0italic_S = 0 while Ne=2subscript𝑁𝑒2N_{e}=2italic_N start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT = 2 due to the treatment of the Li 1s1𝑠1s1 italic_s orbital as a frozen core orbital. The dominant signals at positive frequencies, ω0.08𝜔0.08\omega\approx 0.08italic_ω ≈ 0.08 Ha and ω0.16𝜔0.16\omega\approx 0.16italic_ω ≈ 0.16 Ha, in Figure 7 a arise from transitions between the ground state and doubly degenerate excited states with spin S=1/2𝑆12S=1/2italic_S = 1 / 2 but Ne=3subscript𝑁𝑒3N_{e}=3italic_N start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT = 3. As a result, the spectral function provides direct insights into the excitation levels of the molecular system similar to the study of the Fermi-Hubbard models in section 4.1.

Refer to caption
Figure 7: Spectral function of LiH. (a) Trace of spectral function Tr[A(ω)]Trdelimited-[]𝐴𝜔\mathrm{Tr}\left[A(\omega)\right]roman_Tr [ italic_A ( italic_ω ) ] calculated with compressive sensing, which shows several discrete delta peaks. (b) p|Mν,0p|2subscript𝑝superscriptsubscriptsuperscript𝑀𝑝𝜈02\sum_{p}|M^{p}_{\nu,0}|^{2}∑ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT | italic_M start_POSTSUPERSCRIPT italic_p end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_ν , 0 end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (black dots) and p|M0,νp|2subscript𝑝superscriptsubscriptsuperscript𝑀𝑝0𝜈2\sum_{p}|M^{p}_{0,\nu}|^{2}∑ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT | italic_M start_POSTSUPERSCRIPT italic_p end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 , italic_ν end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (red dots) as a function of the energy differences E0Eνsubscript𝐸0subscript𝐸𝜈E_{0}-E_{\nu}italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT and EνE0subscript𝐸𝜈subscript𝐸0E_{\nu}-E_{0}italic_E start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, respectively. Only the dominant transition amplitudes with p|Mν,μp|2>0.02subscript𝑝superscriptsubscriptsuperscript𝑀𝑝𝜈𝜇20.02\sum_{p}|M^{p}_{\nu,\mu}|^{2}>0.02∑ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT | italic_M start_POSTSUPERSCRIPT italic_p end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_ν , italic_μ end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT > 0.02 are shown. The peaks in the spectral functions originate from transitions between the ground state with energy E0subscript𝐸0E_{0}italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT to the excited states with energies Eνsubscript𝐸𝜈E_{\nu}italic_E start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT, and vice versa.

4.3 Third-order susceptibility of two-site quantum spin-1 model

Model

To demonstrate the calculation of the third-order susceptibility using the CUL circuit presented in section 3.2, we investigate a higher-spin model that has been used to describe 2DCS experiments on rare-earth orthoferrites 63. The Hamiltonian is given by:

^=^absent\displaystyle\hat{\mathcal{H}}=over^ start_ARG caligraphic_H end_ARG = Ji=1N1𝐒^i𝐒^i+1𝐃i=1N1𝐒^i×𝐒^i+1.𝐽superscriptsubscript𝑖1𝑁1subscript^𝐒𝑖subscript^𝐒𝑖1𝐃superscriptsubscript𝑖1𝑁1subscript^𝐒𝑖subscript^𝐒𝑖1\displaystyle\,J\sum_{i=1}^{N-1}\hat{\mathbf{S}}_{i}\cdot\hat{\mathbf{S}}_{i+1% }-\mathbf{D}\cdot\sum_{i=1}^{N-1}\hat{\mathbf{S}}_{i}\times\hat{\mathbf{S}}_{i% +1}\,.italic_J ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N - 1 end_POSTSUPERSCRIPT over^ start_ARG bold_S end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ⋅ over^ start_ARG bold_S end_ARG start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT - bold_D ⋅ ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N - 1 end_POSTSUPERSCRIPT over^ start_ARG bold_S end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT × over^ start_ARG bold_S end_ARG start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT . (36)

The first term of the Hamiltonian characterizes the antiferromagnetic coupling between nearest neighbors with an exchange constant J>0𝐽0J>0italic_J > 0. The second term in eq 36 describes the Dzyaloshinskii-Moriya (DM) interaction with an antisymmetric exchange vector 𝐃𝐃\mathbf{D}bold_D, which we assume to be aligned along the y𝑦yitalic_y-direction, i.e., 𝐃=D𝐲𝐃𝐷𝐲\mathbf{D}=D\mathbf{y}bold_D = italic_D bold_y.

In the simulations, we consider a two-site spin-1 model for demonstration. Note that the quantum resource scaling with respect to system size and spin magnitude s𝑠sitalic_s for the ground state and dynamics simulations of this model have been numerically studied in refs 63, 76. We define the energy unit by setting the coupling constant J𝐽Jitalic_J to one, while using a Dzyaloshinskii-Moriya interaction strength of D=0.2𝐷0.2D=0.2italic_D = 0.2. The inset of Figure 8(e) shows the energy levels Ensubscript𝐸𝑛E_{n}italic_E start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT of the analyzed two-site spin-1 model obtained via exact diagonalization. The energy difference between the ground state (n=0𝑛0n=0italic_n = 0) and the first excited state (n=1𝑛1n=1italic_n = 1) determines the magnon frequency ωAF=E1E00.16subscript𝜔AFsubscript𝐸1subscript𝐸00.16\omega_{\mathrm{AF}}=E_{1}-E_{0}\approx 0.16italic_ω start_POSTSUBSCRIPT roman_AF end_POSTSUBSCRIPT = italic_E start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≈ 0.16, represented by the double arrow. Notably, the eigenenergies of states n=1,2,3𝑛123n=1,2,3italic_n = 1 , 2 , 3 as well as the states n=4,,8𝑛48n=4,\cdots,8italic_n = 4 , ⋯ , 8 are nearly degenerate and the energy difference between En=1,2,3subscript𝐸𝑛123E_{n=1,2,3}italic_E start_POSTSUBSCRIPT italic_n = 1 , 2 , 3 end_POSTSUBSCRIPT and En=4,,8subscript𝐸𝑛48E_{n=4,\cdots,8}italic_E start_POSTSUBSCRIPT italic_n = 4 , ⋯ , 8 end_POSTSUBSCRIPT is approximately 2ωAF2subscript𝜔AF2\omega_{\mathrm{AF}}2 italic_ω start_POSTSUBSCRIPT roman_AF end_POSTSUBSCRIPT (see inset of Figure 8), indicating a quasi-harmonic energy spectrum of the two-site spin-1 model, as discussed in more detail in ref 63.

To map the spin-1 operators to multi-qubit operators, we use the Gray code 92 which provides shallower quantum circuits compared to the binary encoding in calculating two-dimensional coherent spectra, as demonstrated in ref 63. The encoding of the spin-1 operators requires nq=2subscript𝑛𝑞2n_{q}=2italic_n start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT = 2 qubits for a single spin such that the Hamiltonian eq 36 for N=2𝑁2N=2italic_N = 2 sites is represented by Nq=nqN=4subscript𝑁𝑞subscript𝑛𝑞𝑁4N_{q}=n_{q}N=4italic_N start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT = italic_n start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT italic_N = 4 qubits.

Refer to caption
Figure 8: Numerical simulation of the AVQDS algorithm for calculating the third-order nonlinear susceptibility of the two-site spin-1 model. (a)–(c) Four examples of Ipqrsjklm(t,τ)subscriptsuperscript𝐼𝑗𝑘𝑙𝑚𝑝𝑞𝑟𝑠𝑡𝜏I^{jklm}_{pqrs}(t,\tau)italic_I start_POSTSUPERSCRIPT italic_j italic_k italic_l italic_m end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_p italic_q italic_r italic_s end_POSTSUBSCRIPT ( italic_t , italic_τ ) (a) as a function of τ𝜏\tauitalic_τ at fixed tJ=0𝑡𝐽0tJ=0italic_t italic_J = 0, and as a function of time t𝑡titalic_t at fixed (b) τJ=0𝜏𝐽0\tau J=0italic_τ italic_J = 0 and (c) τJ=10𝜏𝐽10\tau J=10italic_τ italic_J = 10. The dynamics obtained by the AVQDS approach (solid lines) accurately reproduce the exact dynamics (dashed black lines) for all presented Ipqrsjklmsubscriptsuperscript𝐼𝑗𝑘𝑙𝑚𝑝𝑞𝑟𝑠I^{jklm}_{pqrs}italic_I start_POSTSUPERSCRIPT italic_j italic_k italic_l italic_m end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_p italic_q italic_r italic_s end_POSTSUBSCRIPT. (d)–(f) The corresponding dynamics of the required number of CNOT gates. The number of CNOT gates significantly increases only at earlier times and saturates after a time of about τJ=1.5𝜏𝐽1.5\tau J=1.5italic_τ italic_J = 1.5 in (d) and tJ=1.3𝑡𝐽1.3tJ=1.3italic_t italic_J = 1.3 in (e) and (f). The saturated number of CNOTs at the final simulation time of tJ=20𝑡𝐽20tJ=20italic_t italic_J = 20 falls within the range of 128 to 216. Inset of (e): Eigenenergies Ensubscript𝐸𝑛E_{n}italic_E start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT of the quantum spin Hamiltonian in eq 36. The Hamiltonian exhibits (2s+1)N=9superscript2𝑠1𝑁9(2s+1)^{N}=9( 2 italic_s + 1 ) start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT = 9 eigenstates for N=2𝑁2N=2italic_N = 2 sites and spin-s=1𝑠1s=1italic_s = 1. The magnon frequency ωAF=E1E00.16subscript𝜔AFsubscript𝐸1subscript𝐸00.16\omega_{\mathrm{AF}}=E_{1}-E_{0}\approx 0.16italic_ω start_POSTSUBSCRIPT roman_AF end_POSTSUBSCRIPT = italic_E start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≈ 0.16 is given by the energy difference between the ground state (n=0𝑛0n=0italic_n = 0) and the first excited state (n=1𝑛1n=1italic_n = 1). The states n=1,2,3𝑛123n=1,2,3italic_n = 1 , 2 , 3 as well as n=4,,8𝑛48n=4,\cdots,8italic_n = 4 , ⋯ , 8 are nearly degenerate.

Ground state preparation

To prepare the ground state |GketG\ket{\mathrm{G}}| start_ARG roman_G end_ARG ⟩ of the quantum-spin Hamiltonian eq 36, we employ the AVQITE method 60. We utilize |φ0=|0Nqketsubscript𝜑0superscriptket0tensor-productabsent𝑁q\ket{\varphi_{0}}=\ket{0}^{\otimes N\mathrm{q}}| start_ARG italic_φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ⟩ = | start_ARG 0 end_ARG ⟩ start_POSTSUPERSCRIPT ⊗ italic_N roman_q end_POSTSUPERSCRIPT as our reference state. For the ground state preparation, we adopt the following operator pool:

𝒫={σiy}i=1Nq{σiyσi+1z}i=1Nq1{σizσi+1y}i=1Nq1.𝒫superscriptsubscriptsubscriptsuperscript𝜎𝑦𝑖𝑖1subscript𝑁𝑞superscriptsubscriptsubscriptsuperscript𝜎𝑦𝑖subscriptsuperscript𝜎𝑧𝑖1𝑖1subscript𝑁𝑞1superscriptsubscriptsubscriptsuperscript𝜎𝑧𝑖subscriptsuperscript𝜎𝑦𝑖1𝑖1subscript𝑁𝑞1\displaystyle\mathcal{P}=\{\sigma^{y}_{i}\}_{i=1}^{N_{q}}\cup\{\sigma^{y}_{i}% \sigma^{z}_{i+1}\}_{i=1}^{N_{q}-1}\cup\{\sigma^{z}_{i}\sigma^{y}_{i+1}\}_{i=1}% ^{N_{q}-1}\,.caligraphic_P = { italic_σ start_POSTSUPERSCRIPT italic_y end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT } start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT end_POSTSUPERSCRIPT ∪ { italic_σ start_POSTSUPERSCRIPT italic_y end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_σ start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT } start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT - 1 end_POSTSUPERSCRIPT ∪ { italic_σ start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_σ start_POSTSUPERSCRIPT italic_y end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT } start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT - 1 end_POSTSUPERSCRIPT . (37)

For the third-order susceptibility calculation with AVQDS, we utilize the following pool:

𝒫=𝒫absent\displaystyle\mathcal{P}=caligraphic_P = {Ai:A{σx,σy,σz}, 1iNq}conditional-setsubscript𝐴𝑖formulae-sequence𝐴superscript𝜎𝑥superscript𝜎𝑦superscript𝜎𝑧1𝑖subscript𝑁q\displaystyle\{A_{i}:A\in\{\sigma^{x},\sigma^{y},\sigma^{z}\},\,1\leq i\leq N_% {\mathrm{q}}\}{ italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT : italic_A ∈ { italic_σ start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT , italic_σ start_POSTSUPERSCRIPT italic_y end_POSTSUPERSCRIPT , italic_σ start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT } , 1 ≤ italic_i ≤ italic_N start_POSTSUBSCRIPT roman_q end_POSTSUBSCRIPT }
{AiBj:A,B{σx,σy,σz}, 1i<jNq}.conditional-setsubscript𝐴𝑖subscript𝐵𝑗formulae-sequence𝐴𝐵superscript𝜎𝑥superscript𝜎𝑦superscript𝜎𝑧1𝑖𝑗subscript𝑁q\displaystyle\cup\{A_{i}B_{j}:A,B\in\{\sigma^{x},\sigma^{y},\sigma^{z}\},\,1% \leq i<j\leq N_{\mathrm{q}}\}\,.∪ { italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_B start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT : italic_A , italic_B ∈ { italic_σ start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT , italic_σ start_POSTSUPERSCRIPT italic_y end_POSTSUPERSCRIPT , italic_σ start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT } , 1 ≤ italic_i < italic_j ≤ italic_N start_POSTSUBSCRIPT roman_q end_POSTSUBSCRIPT } . (38)

This pool contains all possible one- and two-qubit Pauli words.

Simulation results

To demonstrate the calculation of the third-order susceptibility using the CUL circuits with AVQDS presented in section 3.2, we focus on the third-order susceptibility χzzzz(3)(t,τ,0)superscriptsubscript𝜒𝑧𝑧𝑧𝑧3𝑡𝜏0\chi_{zzzz}^{(3)}(t,\tau,0)italic_χ start_POSTSUBSCRIPT italic_z italic_z italic_z italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT ( italic_t , italic_τ , 0 ). This susceptibility is particularly relevant for analyzing 2DCS experiments employing a collinear two-pulse geometry, where the applied magnetic field consists of two copropagating pulses polarized along the z𝑧zitalic_z-direction. In such a scenario, the third and fourth terms within the square brackets in eq 3.2 become identical. Since the transformation eq 22 involves nz=2subscript𝑛𝑧2n_{z}=2italic_n start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = 2 terms for S^jzsubscriptsuperscript^𝑆𝑧𝑗\hat{S}^{z}_{j}over^ start_ARG italic_S end_ARG start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT and spin s=1𝑠1s=1italic_s = 1 63, the summation in eq 3.2 encompasses 3N4nz4=7683superscript𝑁4superscriptsubscript𝑛𝑧47683N^{4}n_{z}^{4}=7683 italic_N start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT italic_n start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT = 768 terms, each of which needs to be calculated using the CUL circuit depicted in Figure 2 b.

To illustrate the calculation of χzzzz(3)(t,τ,0)superscriptsubscript𝜒𝑧𝑧𝑧𝑧3𝑡𝜏0\chi_{zzzz}^{(3)}(t,\tau,0)italic_χ start_POSTSUBSCRIPT italic_z italic_z italic_z italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT ( italic_t , italic_τ , 0 ) with the CUL circuit in Figure 2 b, we use the first term within the square brackets of eq 3.2 as an example:

Ipqrsjklm(t,τ)subscriptsuperscript𝐼𝑗𝑘𝑙𝑚𝑝𝑞𝑟𝑠𝑡𝜏\displaystyle I^{jklm}_{pqrs}(t,\tau)italic_I start_POSTSUPERSCRIPT italic_j italic_k italic_l italic_m end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_p italic_q italic_r italic_s end_POSTSUBSCRIPT ( italic_t , italic_τ )
Im[ei^(t+τ)Pj,pzei^tPk,qzei^τPl,rzPm,sz]absentImdelimited-[]delimited-⟨⟩superscriptei^𝑡𝜏subscriptsuperscript𝑃𝑧𝑗𝑝superscriptei^𝑡subscriptsuperscript𝑃𝑧𝑘𝑞superscriptei^𝜏subscriptsuperscript𝑃𝑧𝑙𝑟subscriptsuperscript𝑃𝑧𝑚𝑠\displaystyle\equiv\mathrm{Im}[\langle\mathrm{e}^{\mathrm{i}\hat{\mathcal{H}}(% t+\tau)}P^{z}_{j,p}\mathrm{e}^{-\mathrm{i}\hat{\mathcal{H}}t}P^{z}_{k,q}% \mathrm{e}^{-\mathrm{i}\hat{\mathcal{H}}\tau}P^{z}_{l,r}P^{z}_{m,s}\rangle]≡ roman_Im [ ⟨ roman_e start_POSTSUPERSCRIPT roman_i over^ start_ARG caligraphic_H end_ARG ( italic_t + italic_τ ) end_POSTSUPERSCRIPT italic_P start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j , italic_p end_POSTSUBSCRIPT roman_e start_POSTSUPERSCRIPT - roman_i over^ start_ARG caligraphic_H end_ARG italic_t end_POSTSUPERSCRIPT italic_P start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k , italic_q end_POSTSUBSCRIPT roman_e start_POSTSUPERSCRIPT - roman_i over^ start_ARG caligraphic_H end_ARG italic_τ end_POSTSUPERSCRIPT italic_P start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l , italic_r end_POSTSUBSCRIPT italic_P start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_m , italic_s end_POSTSUBSCRIPT ⟩ ]
Im[G[𝜽1]|Uτ(𝜽2)Ut(𝜽3)Pj,pzUt(𝜽3)Pk,qz\displaystyle\approx\mathrm{Im}[\bra{G[\boldsymbol{\theta}^{1}]}U_{\tau}^{% \dagger}(\boldsymbol{\theta}^{2})U_{t}^{\dagger}(\boldsymbol{\theta}^{3})P^{z}% _{j,p}U_{t}(\boldsymbol{\theta}^{3})P^{z}_{k,q}≈ roman_Im [ ⟨ start_ARG italic_G [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ] end_ARG | italic_U start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ( bold_italic_θ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) italic_U start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ( bold_italic_θ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ) italic_P start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j , italic_p end_POSTSUBSCRIPT italic_U start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ( bold_italic_θ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ) italic_P start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k , italic_q end_POSTSUBSCRIPT
×Uτ(𝜽2)Pl,rzPm,sz|G[𝜽1]].\displaystyle\qquad\qquad\quad\times U_{\tau}(\boldsymbol{\theta}^{2})P^{z}_{l% ,r}P^{z}_{m,s}\ket{G[\boldsymbol{\theta}^{1}]}]\,.× italic_U start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT ( bold_italic_θ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) italic_P start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l , italic_r end_POSTSUBSCRIPT italic_P start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_m , italic_s end_POSTSUBSCRIPT | start_ARG italic_G [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ] end_ARG ⟩ ] . (39)

We first evolve the state in eq 24 with P0=P1=INqsubscript𝑃0subscript𝑃1superscript𝐼tensor-productabsentsubscript𝑁𝑞P_{0}=P_{1}=I^{\otimes N_{q}}italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = italic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_I start_POSTSUPERSCRIPT ⊗ italic_N start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT end_POSTSUPERSCRIPT, P4=Pl,rzsubscript𝑃4subscriptsuperscript𝑃𝑧𝑙𝑟P_{4}=P^{z}_{l,r}italic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT = italic_P start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l , italic_r end_POSTSUBSCRIPT, and P5=Pm,szsubscript𝑃5subscriptsuperscript𝑃𝑧𝑚𝑠P_{5}=P^{z}_{m,s}italic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT = italic_P start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_m , italic_s end_POSTSUBSCRIPT up to time τ𝜏\tauitalic_τ using AVQDS. After applying the X𝑋Xitalic_X-gate and controlled P3=Pk,qzsubscript𝑃3subscriptsuperscript𝑃𝑧𝑘𝑞P_{3}=P^{z}_{k,q}italic_P start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT = italic_P start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k , italic_q end_POSTSUBSCRIPT gate, we further propagate the state in eq 25 to t[0,tmax]𝑡0subscript𝑡maxt\in[0,t_{\mathrm{max}}]italic_t ∈ [ 0 , italic_t start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT ], for which we adopt a uniform time mesh. Similar circuit simulations are repeated for τ[0,τmax]𝜏0subscript𝜏max\tau\in[0,\tau_{\mathrm{max}}]italic_τ ∈ [ 0 , italic_τ start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT ] with a uniform mesh.

Figure 8 a shows four examples of Ipqrsjklm(t,τ)subscriptsuperscript𝐼𝑗𝑘𝑙𝑚𝑝𝑞𝑟𝑠𝑡𝜏I^{jklm}_{pqrs}(t,\tau)italic_I start_POSTSUPERSCRIPT italic_j italic_k italic_l italic_m end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_p italic_q italic_r italic_s end_POSTSUBSCRIPT ( italic_t , italic_τ ) as a function of τ𝜏\tauitalic_τ at fixed t=0𝑡0t=0italic_t = 0, while Figure 8 b and 8 c present the corresponding Ipqrsjklm(t,τ)subscriptsuperscript𝐼𝑗𝑘𝑙𝑚𝑝𝑞𝑟𝑠𝑡𝜏I^{jklm}_{pqrs}(t,\tau)italic_I start_POSTSUPERSCRIPT italic_j italic_k italic_l italic_m end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_p italic_q italic_r italic_s end_POSTSUBSCRIPT ( italic_t , italic_τ ) as a function of t𝑡titalic_t at fixed τ=0𝜏0\tau=0italic_τ = 0 and τJ=10𝜏𝐽10\tau J=10italic_τ italic_J = 10, respectively. The results obtained from the AVQDS approach (solid lines) are compared with the exact simulation results obtained by evolving the states in eqs 24 and 25 using exact diagonalization eq 8 (dashed black lines). The results from the CUL circuit simulations agree with the exact dynamics for all presented Ipqrsjklmsubscriptsuperscript𝐼𝑗𝑘𝑙𝑚𝑝𝑞𝑟𝑠I^{jklm}_{pqrs}italic_I start_POSTSUPERSCRIPT italic_j italic_k italic_l italic_m end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_p italic_q italic_r italic_s end_POSTSUBSCRIPT, with fidelities exceeding 99.99%percent99.9999.99\%99.99 %.

The corresponding number of CNOT gates as a function of time are shown in Figure 8 d8 f. The initial 22222222 CNOTs at t=τ=0𝑡𝜏0t=\tau=0italic_t = italic_τ = 0 are determined by the ground state ansatz, which is obtained using AVQITE with an infidelity of about 1010superscript101010^{-10}10 start_POSTSUPERSCRIPT - 10 end_POSTSUPERSCRIPT. Throughout the time evolution, the number of CNOT gates only increases rapidly at very early times, followed by saturation after a time of about τJ=1.5𝜏𝐽1.5\tau J=1.5italic_τ italic_J = 1.5 (Figure 8 d) and tJ=1.3𝑡𝐽1.3tJ=1.3italic_t italic_J = 1.3 (Figure 8 e and 8 f). The saturated number of CNOTs at the final simulation time of tJ=20𝑡𝐽20tJ=20italic_t italic_J = 20 ranges from 128 to 216 for the various Ipqrsjklm(t,τ)subscriptsuperscript𝐼𝑗𝑘𝑙𝑚𝑝𝑞𝑟𝑠𝑡𝜏I^{jklm}_{pqrs}(t,\tau)italic_I start_POSTSUPERSCRIPT italic_j italic_k italic_l italic_m end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_p italic_q italic_r italic_s end_POSTSUBSCRIPT ( italic_t , italic_τ ).

Refer to caption
Figure 9: Third-order susceptibility of the two-site spin-1 system in the two-dimensional time and frequency domains. (a) χzzzz(3)(t,τ,0)subscriptsuperscript𝜒3𝑧𝑧𝑧𝑧𝑡𝜏0\chi^{(3)}_{zzzz}(t,\tau,0)italic_χ start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_z italic_z italic_z italic_z end_POSTSUBSCRIPT ( italic_t , italic_τ , 0 ) as a function of times t𝑡titalic_t and τ𝜏\tauitalic_τ calculated using the CUL circuit, which agrees with the exact result in (b) obtained by evaluating eq D. Slices shown in (c) and (d) are indicated by horizontal and vertical dashed lines. (c), (d) Slices of χzzzz(3)(t,τ,0)subscriptsuperscript𝜒3𝑧𝑧𝑧𝑧𝑡𝜏0\chi^{(3)}_{zzzz}(t,\tau,0)italic_χ start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_z italic_z italic_z italic_z end_POSTSUBSCRIPT ( italic_t , italic_τ , 0 ) from (a) and (b) at two fixed times τ𝜏\tauitalic_τ (c) and t𝑡titalic_t (d). The AVQDS results (circles) accurately reproduce the exact dynamics (solid black lines). (e), (f) Two-dimensional (2D) Fourier transform of χzzzz(3)(t,τ,0)subscriptsuperscript𝜒3𝑧𝑧𝑧𝑧𝑡𝜏0\chi^{(3)}_{zzzz}(t,\tau,0)italic_χ start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_z italic_z italic_z italic_z end_POSTSUBSCRIPT ( italic_t , italic_τ , 0 ) from (a) and (b), respectively. Both 2D spectra are in excellent agreement.

To obtain the third-order susceptibility χzzzz(3)(t,τ,0)subscriptsuperscript𝜒3𝑧𝑧𝑧𝑧𝑡𝜏0\chi^{(3)}_{zzzz}(t,\tau,0)italic_χ start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_z italic_z italic_z italic_z end_POSTSUBSCRIPT ( italic_t , italic_τ , 0 ), we simulated the different contributions in eq 3.2 using the CUL circuit in Figure 2 b. Simulations were performed up to τmaxJ=40subscript𝜏max𝐽40\tau_{\textrm{max}}J=40italic_τ start_POSTSUBSCRIPT max end_POSTSUBSCRIPT italic_J = 40 with a step size of δτ=0.5𝛿𝜏0.5\delta\tau=0.5italic_δ italic_τ = 0.5, which provides sufficient resolution of the signals in the 2D spectrum of the third-order susceptibility. To calculate χzzzz(3)(t,τ,0)subscriptsuperscript𝜒3𝑧𝑧𝑧𝑧𝑡𝜏0\chi^{(3)}_{zzzz}(t,\tau,0)italic_χ start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_z italic_z italic_z italic_z end_POSTSUBSCRIPT ( italic_t , italic_τ , 0 ), we interpolated the different contributions in eq 3.2 to a 401×401401401401\times 401401 × 401 uniform mesh of t𝑡titalic_t and τ𝜏\tauitalic_τ with tmaxJ=τmaxJ=40subscript𝑡max𝐽subscript𝜏max𝐽40t_{\textrm{max}}J=\tau_{\textrm{max}}J=40italic_t start_POSTSUBSCRIPT max end_POSTSUBSCRIPT italic_J = italic_τ start_POSTSUBSCRIPT max end_POSTSUBSCRIPT italic_J = 40 before performing the summation over the components in eq 3.2. Figure 9 a shows χzzzz(3)(t,τ,0)subscriptsuperscript𝜒3𝑧𝑧𝑧𝑧𝑡𝜏0\chi^{(3)}_{zzzz}(t,\tau,0)italic_χ start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_z italic_z italic_z italic_z end_POSTSUBSCRIPT ( italic_t , italic_τ , 0 ) as a function of times t𝑡titalic_t and τ𝜏\tauitalic_τ from statevector simulations of the CUL circuits. The corresponding result for the exact dynamics, derived in Appendix D, is presented in Figure 9 b. The third-order susceptibility obtained by the CUL circuits agrees well with the exact simulation result. This is further demonstrated by the slices at fixed τ𝜏\tauitalic_τ and fixed t𝑡titalic_t plotted in Figure 9 c and 9 d, where the CUL circuit results (circles) match the exact result (solid black line) over the full range of simulation times.

Figure 9 e and 9 f show the resulting 2D spectra of the third-order susceptibility, χzzzz(3)(ωt,ωτ,0)subscriptsuperscript𝜒3𝑧𝑧𝑧𝑧subscript𝜔𝑡subscript𝜔𝜏0\chi^{(3)}_{zzzz}(\omega_{t},\omega_{\tau},0)italic_χ start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_z italic_z italic_z italic_z end_POSTSUBSCRIPT ( italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT , italic_ω start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT , 0 ), obtained after performing the 2D discrete Fourier transformation of χzzzz(3)(t,τ,0)subscriptsuperscript𝜒3𝑧𝑧𝑧𝑧𝑡𝜏0\chi^{(3)}_{zzzz}(t,\tau,0)italic_χ start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_z italic_z italic_z italic_z end_POSTSUBSCRIPT ( italic_t , italic_τ , 0 ) from Figure9 a and 9 b using the Padé approximation in both the t𝑡titalic_t and τ𝜏\tauitalic_τ-directions. The 2D spectrum of the third-order susceptibility transformed from the real-time CUL circuit results agrees well with that from exact diagonalization. The 2D spectrum exhibits four dominant peaks at (ωt,ωτ)=(0,±ωAF)subscript𝜔𝑡subscript𝜔𝜏0plus-or-minussubscript𝜔AF(\omega_{t},\omega_{\tau})=(0,\pm\omega_{\mathrm{AF}})( italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT , italic_ω start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT ) = ( 0 , ± italic_ω start_POSTSUBSCRIPT roman_AF end_POSTSUBSCRIPT ), (ωt,ωτ)=(0,0)subscript𝜔𝑡subscript𝜔𝜏00(\omega_{t},\omega_{\tau})=(0,0)( italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT , italic_ω start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT ) = ( 0 , 0 ), and (ωt,ωτ)=(ωAF,ωAF)subscript𝜔𝑡subscript𝜔𝜏subscript𝜔AFsubscript𝜔AF(\omega_{t},\omega_{\tau})=(\omega_{\mathrm{AF}},\omega_{\mathrm{AF}})( italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT , italic_ω start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT ) = ( italic_ω start_POSTSUBSCRIPT roman_AF end_POSTSUBSCRIPT , italic_ω start_POSTSUBSCRIPT roman_AF end_POSTSUBSCRIPT ). As demonstrated in ref 63 and discussed in Appendix D, these signals originate from transitions between the ground state |Ψ0ketsubscriptΨ0|\Psi_{0}\rangle| roman_Ψ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ⟩ and the first excited state |Ψ1ketsubscriptΨ1|\Psi_{1}\rangle| roman_Ψ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ⟩, as well as between |Ψ1ketsubscriptΨ1|\Psi_{1}\rangle| roman_Ψ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ⟩ and the third excited state |Ψ3ketsubscriptΨ3|\Psi_{3}\rangle| roman_Ψ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ⟩. Such signals directly manifest in the 2DCS spectra, underscoring the significance of nonlinear susceptibilities in interpreting 2DCS experiments.

5 Conclusion and outlook

In this work, we presented and benchmarked quantum computing methods to calculate single-particle Green’s functions and nonlinear susceptibilities, adopting controlled-unitaries-liberated circuits and adaptive variational algorithms for ground state preparation and real-time propagation. For the computation of single-particle Green’s functions, we followed the CUL approach outlined in ref 51, utilizing AVQITE to generate ground state circuits and AVQDS to approximate the time-evolution operator instead of CUL with HVA as in ref 51. To illustrate the CUL approach, we computed the single-particle Green’s function of Fermi-Hubbard chains with N=4𝑁4N=4italic_N = 4 and N=6𝑁6N=6italic_N = 6 sites. Specifically, we evaluated the real-time Green’s function in momentum space as well as its corresponding spectral function, and compared the results with exact diagonalization calculations using the Lehmann representation of the Green’s function. Our findings demonstrate that the CUL approach with AVQDS can accurately simulate the dynamics of Green’s functions over sufficiently long times and obtain reliable spectral functions. The CUL quantum circuits required for the simulation of the Green’s functions are shallower compared to those obtained with HVA in ref 51. Furthermore, in comparison to the calculation of the single-particle Green’s function with the CUR method using AVQDS in ref 62, our results demonstrate that utilizing the CUL approach with AVQDS can reduce the circuit complexity by bypassing the state overlap measurement. Comparable CNOT counts but deeper quantum circuits are required for CUR with HVA 56. As another example to demonstrate the general applicability of the CUL approach with AVQDS, we applied it to accurately evaluate the Green’s function of the molecule LiH, which amounts to an estimated circuit depth up to 81 layers.

We also extended the CUL method from calculating single-particle Green’s functions to evaluating nonlinear susceptibilities, which is crucial to explain the two-dimensional coherent spectroscopy experiments on semiconductors 16, 17, 18, superconductors 23, 24, topological systems 93, quantum spin systems 94, 63, and molecular systems 95, 96. Here, the nonlinear susceptibilities depend on two times such that their computation via the quantum circuit presented in Figure 2 b requires the application of the AVQDS algorithm for two times. To demonstrate the validity of our method, we studied an antiferromagnetic quantum spin model including a Dzyaloshinskii-Moriya interaction. We calculated the third-order nonlinear susceptibility in the 2D time and frequency domains for a two-site spin-1 model and compared the results with numerical exact data. The third-order susceptibility calculated using the quantum computing approach agrees well with the exact result which confirms the accuracy of the CUL method in evaluating higher-order correlation functions using AVQDS for state evolution.

Several modifications of the presented approach are feasible, where the quantum circuit complexity can potentially be reduced by trading for more measurements. For example, the Hadamard test circuits could be replaced with direct measurement circuits as shown in ref 97. An interesting direction for future work is motivated by the quantum algorithm proposed in ref 50, which was developed to study long-time dynamics and correlation functions of driven-dissipative quantum systems, where the steady state of the time evolution is stable to perturbations. Compared with the standard Hadamard test circuit where an ancilla qubit is required to maintain entanglement and coherence with the system qubits during the entire simulation time, the key idea is to set the ancilla to the |+ket\ket{+}| start_ARG + end_ARG ⟩ state before applying the ancilla-controlled Pauli gates and measure the ancilla immediately after the ancilla-system entangling gates at each of the specified times to decouple the ancilla from the system qubits. The n𝑛nitalic_n-point correlation function can then be constructed based on the measurement outcomes. The ancilla qubit is only entangled with the system qubits for a minimal amount of time when the ancilla-system entangling gates are applied. This promises a great advantage when the bottleneck for successful execution of the measurement circuits is the coherence requirement of the ancilla qubit, like the steady state of a driven-dissipative quantum system. For simulations of isolated quantum systems as in this work, no clear advantage is expected due to the stringent coherence requirement for both ancilla and system qubits 50. However, a potential advantage may still be achieved thanks to the adaptive variational quantum algorithms, which generate highly compressed quantum circuits for dynamics simulations and substantially reduce the coherence time requirement. By adopting mid-circuit measurements for the ancilla qubit which disentangles the ancilla from the system qubits prior to time evolution, the state propagation only needs to be applied to the system qubits, which can potentially simplify the AVQDS circuits.

The presented algorithms for computing single-particle Green’s functions and nonlinear susceptibilities can readily be extended to calculate higher-order multi-time correlation functions that depend on more than two times. Here, it would be interesting to compare the quantum resources required by the CUL method used in this paper against those required by the CUR approach considered in refs 56, 62. In addition, the approach can be naturally extended to finite temperature calculations by adopting techniques such as a quantum version of the minimally entangled typical thermal states algorithm 66, 98, 99.

Regarding the practical implementation of the AVQDS approach for computing high-order correlation functions on quantum hardware, it is essential to investigate the impact of hardware noise and shot noise resulting from a finite number of measurements. Here, the incorporation of error mitigation methods plays a crucial role 100, 101.

Appendix A Quantum circuit for calculating single-particle Green’s functions

In this appendix, we demonstrate that the quantum circuit in Figure 1 b evaluates eq 14. The system qubits are initialized in a reference product state |φ0ketsubscript𝜑0\ket{\varphi_{0}}| start_ARG italic_φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ⟩, which can be straightforwardly prepared on a quantum computer. Then, the ground state is prepared by applying the unitary operator UG[𝜽1]subscript𝑈Gdelimited-[]superscript𝜽1U_{\mathrm{G}}[\boldsymbol{\theta}^{1}]italic_U start_POSTSUBSCRIPT roman_G end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ] following AVQITE, yielding |G[𝜽1]=UG[𝜽1]|φ0ket𝐺delimited-[]superscript𝜽1subscript𝑈Gdelimited-[]superscript𝜽1ketsubscript𝜑0\ket{G[\boldsymbol{\theta}^{1}]}=U_{\mathrm{G}}[\boldsymbol{\theta}^{1}]\ket{% \varphi_{0}}| start_ARG italic_G [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ] end_ARG ⟩ = italic_U start_POSTSUBSCRIPT roman_G end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ] | start_ARG italic_φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ⟩. The initial state of the ancilla qubit is |0ket0\ket{0}| start_ARG 0 end_ARG ⟩, which becomes (|0+|1)/2ket0ket12(\ket{0}+\ket{1})/\sqrt{2}( | start_ARG 0 end_ARG ⟩ + | start_ARG 1 end_ARG ⟩ ) / square-root start_ARG 2 end_ARG after applying the Hadamard gate. The quantum state after applying UGsubscript𝑈GU_{\mathrm{G}}italic_U start_POSTSUBSCRIPT roman_G end_POSTSUBSCRIPT on the system qubits and H𝐻Hitalic_H on the ancilla qubit reads:

|Ψ=12[|0|G[𝜽1]+|1|G[𝜽1]].ketΨ12delimited-[]tensor-productket0ket𝐺delimited-[]superscript𝜽1tensor-productket1ket𝐺delimited-[]superscript𝜽1\displaystyle\ket{\Psi}=\frac{1}{\sqrt{2}}\left[\ket{0}\otimes\ket{G[% \boldsymbol{\theta}^{1}]}+\ket{1}\otimes\ket{G[\boldsymbol{\theta}^{1}]}\right% ]\,.| start_ARG roman_Ψ end_ARG ⟩ = divide start_ARG 1 end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG [ | start_ARG 0 end_ARG ⟩ ⊗ | start_ARG italic_G [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ] end_ARG ⟩ + | start_ARG 1 end_ARG ⟩ ⊗ | start_ARG italic_G [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ] end_ARG ⟩ ] . (40)

After applying the controlled Pβsubscript𝑃𝛽P_{\beta}italic_P start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT operation controlled by the ancilla qubit, the state is given by 12|0|G+12|1Pβ|Gtensor-product12ket0ket𝐺tensor-product12ket1subscript𝑃𝛽ket𝐺\frac{1}{\sqrt{2}}\ket{0}\otimes\ket{G}+\frac{1}{\sqrt{2}}\ket{1}\otimes P_{% \beta}\ket{G}divide start_ARG 1 end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG | start_ARG 0 end_ARG ⟩ ⊗ | start_ARG italic_G end_ARG ⟩ + divide start_ARG 1 end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG | start_ARG 1 end_ARG ⟩ ⊗ italic_P start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT | start_ARG italic_G end_ARG ⟩. The time-evolving state under AVQDS with time-dependent parameters 𝜽1(t)superscript𝜽1𝑡\boldsymbol{\theta}^{1}(t)bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ( italic_t ) and 𝜽2(t)superscript𝜽2𝑡\boldsymbol{\theta}^{2}(t)bold_italic_θ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_t ) becomes:

|Ψ=ketΨabsent\displaystyle\ket{\Psi}=| start_ARG roman_Ψ end_ARG ⟩ = 12|0Ut[𝜽2]|G[𝜽1]tensor-product12ket0subscript𝑈𝑡delimited-[]superscript𝜽2ket𝐺delimited-[]superscript𝜽1\displaystyle\frac{1}{\sqrt{2}}\ket{0}\otimes U_{t}[\boldsymbol{\theta}^{2}]% \ket{G[\boldsymbol{\theta}^{1}]}divide start_ARG 1 end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG | start_ARG 0 end_ARG ⟩ ⊗ italic_U start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] | start_ARG italic_G [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ] end_ARG ⟩ (41)
+12|1Ut[𝜽2]Pβ|G[𝜽1].tensor-product12ket1subscript𝑈𝑡delimited-[]superscript𝜽2subscript𝑃𝛽ket𝐺delimited-[]superscript𝜽1\displaystyle+\frac{1}{\sqrt{2}}\ket{1}\otimes U_{t}[\boldsymbol{\theta}^{2}]P% _{\beta}\ket{G[\boldsymbol{\theta}^{1}]}\,.+ divide start_ARG 1 end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG | start_ARG 1 end_ARG ⟩ ⊗ italic_U start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] italic_P start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT | start_ARG italic_G [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ] end_ARG ⟩ .

Applying the X𝑋Xitalic_X-gate on the ancilla qubit followed by the controlled Pαsubscript𝑃𝛼P_{\alpha}italic_P start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT operation yields

|Ψ=ketΨabsent\displaystyle\ket{\Psi}=| start_ARG roman_Ψ end_ARG ⟩ = 12|1PαUt[𝜽2]|G[𝜽1]tensor-product12ket1subscript𝑃𝛼subscript𝑈𝑡delimited-[]superscript𝜽2ket𝐺delimited-[]superscript𝜽1\displaystyle\frac{1}{\sqrt{2}}\ket{1}\otimes P_{\alpha}U_{t}[\boldsymbol{% \theta}^{2}]\ket{G[\boldsymbol{\theta}^{1}]}divide start_ARG 1 end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG | start_ARG 1 end_ARG ⟩ ⊗ italic_P start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_U start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] | start_ARG italic_G [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ] end_ARG ⟩ (42)
+12|0Ut[𝜽2]Pβ|G[𝜽1].tensor-product12ket0subscript𝑈𝑡delimited-[]superscript𝜽2subscript𝑃𝛽ket𝐺delimited-[]superscript𝜽1\displaystyle+\frac{1}{\sqrt{2}}\ket{0}\otimes U_{t}[\boldsymbol{\theta}^{2}]P% _{\beta}\ket{G[\boldsymbol{\theta}^{1}]}\,.+ divide start_ARG 1 end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG | start_ARG 0 end_ARG ⟩ ⊗ italic_U start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] italic_P start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT | start_ARG italic_G [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ] end_ARG ⟩ .

After the application of the Hadamard gate on the ancilla, the quantum state is given by

|ΨketΨ\displaystyle\ket{\Psi}| start_ARG roman_Ψ end_ARG ⟩ =12|0(PαUt[𝜽2]+Ut[𝜽2]Pβ)|G[𝜽1]absenttensor-product12ket0subscript𝑃𝛼subscript𝑈𝑡delimited-[]superscript𝜽2subscript𝑈𝑡delimited-[]superscript𝜽2subscript𝑃𝛽ket𝐺delimited-[]superscript𝜽1\displaystyle=\frac{1}{2}\ket{0}\otimes\left(P_{\alpha}U_{t}[\boldsymbol{% \theta}^{2}]+U_{t}[\boldsymbol{\theta}^{2}]P_{\beta}\right)\ket{G[\boldsymbol{% \theta}^{1}]}= divide start_ARG 1 end_ARG start_ARG 2 end_ARG | start_ARG 0 end_ARG ⟩ ⊗ ( italic_P start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_U start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] + italic_U start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] italic_P start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT ) | start_ARG italic_G [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ] end_ARG ⟩
12|1(PαUt[𝜽2]Ut[𝜽2]Pβ)|G[𝜽1].tensor-product12ket1subscript𝑃𝛼subscript𝑈𝑡delimited-[]superscript𝜽2subscript𝑈𝑡delimited-[]superscript𝜽2subscript𝑃𝛽ket𝐺delimited-[]superscript𝜽1\displaystyle-\frac{1}{2}\ket{1}\otimes\left(P_{\alpha}U_{t}[\boldsymbol{% \theta}^{2}]-U_{t}[\boldsymbol{\theta}^{2}]P_{\beta}\right)\ket{G[\boldsymbol{% \theta}^{1}]}\,.- divide start_ARG 1 end_ARG start_ARG 2 end_ARG | start_ARG 1 end_ARG ⟩ ⊗ ( italic_P start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_U start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] - italic_U start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] italic_P start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT ) | start_ARG italic_G [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ] end_ARG ⟩ . (43)

Finally, performing a Z𝑍Zitalic_Z measurement on the ancilla qubit, Ψ|σ^zINq|Ψtensor-productbraΨsuperscript^𝜎𝑧superscript𝐼tensor-productabsentsubscript𝑁𝑞ketΨ\bra{\Psi}\hat{\sigma}^{z}\otimes I^{\otimes N_{q}}\ket{\Psi}⟨ start_ARG roman_Ψ end_ARG | over^ start_ARG italic_σ end_ARG start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT ⊗ italic_I start_POSTSUPERSCRIPT ⊗ italic_N start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT end_POSTSUPERSCRIPT | start_ARG roman_Ψ end_ARG ⟩, yields Iα,βp,q=2p|01subscriptsuperscript𝐼𝑝𝑞𝛼𝛽2subscript𝑝ket01I^{p,q}_{\alpha,\beta}=2p_{\ket{0}}-1italic_I start_POSTSUPERSCRIPT italic_p , italic_q end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α , italic_β end_POSTSUBSCRIPT = 2 italic_p start_POSTSUBSCRIPT | start_ARG 0 end_ARG ⟩ end_POSTSUBSCRIPT - 1 as defined in eq 14, where p|0subscript𝑝ket0p_{\ket{0}}italic_p start_POSTSUBSCRIPT | start_ARG 0 end_ARG ⟩ end_POSTSUBSCRIPT is the probability that the ancilla is measured to be in the state |0ket0\ket{0}| start_ARG 0 end_ARG ⟩.

Appendix B Lehmann representation of spectral function

In this appendix, we derive the Lehmann representation of the spectral function. Using the completeness relation I^=ν|ΨνΨν|^𝐼subscript𝜈ketsubscriptΨ𝜈brasubscriptΨ𝜈\hat{I}=\sum_{\nu}\ket{\Psi_{\nu}}\bra{\Psi_{\nu}}over^ start_ARG italic_I end_ARG = ∑ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT | start_ARG roman_Ψ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG ⟩ ⟨ start_ARG roman_Ψ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG |, where |ΨνketsubscriptΨ𝜈\ket{\Psi_{\nu}}| start_ARG roman_Ψ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG ⟩ are the eigenstates of the fermionic Hamiltonian ^^\hat{\mathcal{H}}over^ start_ARG caligraphic_H end_ARG with energies Eνsubscript𝐸𝜈E_{\nu}italic_E start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT, eq 3.1 can be written as

Gp,qR(t)=iΘ(t)subscriptsuperscript𝐺R𝑝𝑞𝑡iΘ𝑡\displaystyle G^{\mathrm{R}}_{p,q}(t)=-\mathrm{i}\,\Theta(t)italic_G start_POSTSUPERSCRIPT roman_R end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_p , italic_q end_POSTSUBSCRIPT ( italic_t ) = - roman_i roman_Θ ( italic_t ) [ei(E0Eν)tT0,νp(T0,νq)\displaystyle\left[\mathrm{e}^{\mathrm{i}(E_{0}-E_{\nu})t}T^{p}_{0,\nu}\left(T% ^{q}_{0,\nu}\right)^{\star}\right.[ roman_e start_POSTSUPERSCRIPT roman_i ( italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ) italic_t end_POSTSUPERSCRIPT italic_T start_POSTSUPERSCRIPT italic_p end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 , italic_ν end_POSTSUBSCRIPT ( italic_T start_POSTSUPERSCRIPT italic_q end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 , italic_ν end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT ⋆ end_POSTSUPERSCRIPT
ei(E0Eν)tTν,0p(Tν,0q)],\displaystyle\left.\mathrm{e}^{\mathrm{i}(E_{0}-E_{\nu})t}T^{p}_{\nu,0}\left(T% ^{q}_{\nu,0}\right)^{\star}\right]\,,roman_e start_POSTSUPERSCRIPT roman_i ( italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ) italic_t end_POSTSUPERSCRIPT italic_T start_POSTSUPERSCRIPT italic_p end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_ν , 0 end_POSTSUBSCRIPT ( italic_T start_POSTSUPERSCRIPT italic_q end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_ν , 0 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT ⋆ end_POSTSUPERSCRIPT ] , (44)

with transition matrix elements Tμ,νpΨμ|c^p|Ψνsubscriptsuperscript𝑇𝑝𝜇𝜈brasubscriptΨ𝜇subscriptsuperscript^𝑐absent𝑝ketsubscriptΨ𝜈T^{p}_{\mu,\nu}\equiv\bra{\Psi_{\mu}}\hat{c}^{\phantom{\dagger}}_{p}\ket{\Psi_% {\nu}}italic_T start_POSTSUPERSCRIPT italic_p end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_μ , italic_ν end_POSTSUBSCRIPT ≡ ⟨ start_ARG roman_Ψ start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT end_ARG | over^ start_ARG italic_c end_ARG start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT | start_ARG roman_Ψ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG ⟩. By applying the Fourier transform

Gp,qR(ω)=dtei(ω+iε)tGp,qR(t)subscriptsuperscript𝐺R𝑝𝑞𝜔superscriptsubscriptdifferential-d𝑡superscriptei𝜔i𝜀𝑡subscriptsuperscript𝐺R𝑝𝑞𝑡\displaystyle G^{\mathrm{R}}_{p,q}(\omega)=\int_{-\infty}^{\infty}\mathrm{d}t% \,\mathrm{e}^{\mathrm{i}(\omega+\mathrm{i}\,\varepsilon)t}G^{\mathrm{R}}_{p,q}% (t)italic_G start_POSTSUPERSCRIPT roman_R end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_p , italic_q end_POSTSUBSCRIPT ( italic_ω ) = ∫ start_POSTSUBSCRIPT - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT roman_d italic_t roman_e start_POSTSUPERSCRIPT roman_i ( italic_ω + roman_i italic_ε ) italic_t end_POSTSUPERSCRIPT italic_G start_POSTSUPERSCRIPT roman_R end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_p , italic_q end_POSTSUBSCRIPT ( italic_t ) (45)

with an infinitesimal ε𝜀\varepsilonitalic_ε to guarantee convergence of the integral, we find

Gp,qR(ω)=ν[T0,νp(T0,νq)E0Eν+ω+iε+Tν,0p(Tν,0q)EνE0+ω+iε].subscriptsuperscript𝐺R𝑝𝑞𝜔subscript𝜈delimited-[]subscriptsuperscript𝑇𝑝0𝜈superscriptsubscriptsuperscript𝑇𝑞0𝜈subscript𝐸0subscript𝐸𝜈𝜔i𝜀subscriptsuperscript𝑇𝑝𝜈0superscriptsubscriptsuperscript𝑇𝑞𝜈0subscript𝐸𝜈subscript𝐸0𝜔i𝜀\displaystyle G^{\mathrm{R}}_{p,q}(\omega)=\sum_{\nu}\left[\frac{T^{p}_{0,\nu}% \left(T^{q}_{0,\nu}\right)^{\star}}{E_{0}-E_{\nu}+\omega+\mathrm{i}\,% \varepsilon}+\frac{T^{p}_{\nu,0}\left(T^{q}_{\nu,0}\right)^{\star}}{E_{\nu}-E_% {0}+\omega+\mathrm{i}\,\varepsilon}\right]\,.italic_G start_POSTSUPERSCRIPT roman_R end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_p , italic_q end_POSTSUBSCRIPT ( italic_ω ) = ∑ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT [ divide start_ARG italic_T start_POSTSUPERSCRIPT italic_p end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 , italic_ν end_POSTSUBSCRIPT ( italic_T start_POSTSUPERSCRIPT italic_q end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 , italic_ν end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT ⋆ end_POSTSUPERSCRIPT end_ARG start_ARG italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT + italic_ω + roman_i italic_ε end_ARG + divide start_ARG italic_T start_POSTSUPERSCRIPT italic_p end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_ν , 0 end_POSTSUBSCRIPT ( italic_T start_POSTSUPERSCRIPT italic_q end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_ν , 0 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT ⋆ end_POSTSUPERSCRIPT end_ARG start_ARG italic_E start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + italic_ω + roman_i italic_ε end_ARG ] . (46)

Finally, the retarded Green’s function in momentum space is given by

GkRsubscriptsuperscript𝐺R𝑘\displaystyle G^{\mathrm{R}}_{k}italic_G start_POSTSUPERSCRIPT roman_R end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT =1Np,qGp,qReik(pq)absent1𝑁subscript𝑝𝑞subscriptsuperscript𝐺R𝑝𝑞superscriptei𝑘𝑝𝑞\displaystyle=\frac{1}{N}\sum_{p,q}G^{\mathrm{R}}_{p,q}\mathrm{e}^{-\mathrm{i}% \,k(p-q)}= divide start_ARG 1 end_ARG start_ARG italic_N end_ARG ∑ start_POSTSUBSCRIPT italic_p , italic_q end_POSTSUBSCRIPT italic_G start_POSTSUPERSCRIPT roman_R end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_p , italic_q end_POSTSUBSCRIPT roman_e start_POSTSUPERSCRIPT - roman_i italic_k ( italic_p - italic_q ) end_POSTSUPERSCRIPT
=1Np,qν[T0,νp(T0,νq)E0Eν+ω+iε\displaystyle=\frac{1}{N}\sum_{p,q}\sum_{\nu}\left[\frac{T^{p}_{0,\nu}\left(T^% {q}_{0,\nu}\right)^{\star}}{E_{0}-E_{\nu}+\omega+\mathrm{i}\,\varepsilon}\right.= divide start_ARG 1 end_ARG start_ARG italic_N end_ARG ∑ start_POSTSUBSCRIPT italic_p , italic_q end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT [ divide start_ARG italic_T start_POSTSUPERSCRIPT italic_p end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 , italic_ν end_POSTSUBSCRIPT ( italic_T start_POSTSUPERSCRIPT italic_q end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 , italic_ν end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT ⋆ end_POSTSUPERSCRIPT end_ARG start_ARG italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT + italic_ω + roman_i italic_ε end_ARG
+Tν,0p(Tν,0q)EνE0+ω+iε]eik(pq),\displaystyle\left.+\frac{T^{p}_{\nu,0}\left(T^{q}_{\nu,0}\right)^{\star}}{E_{% \nu}-E_{0}+\omega+i\varepsilon}\right]\mathrm{e}^{-\mathrm{i}\,k(p-q)}\,,+ divide start_ARG italic_T start_POSTSUPERSCRIPT italic_p end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_ν , 0 end_POSTSUBSCRIPT ( italic_T start_POSTSUPERSCRIPT italic_q end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_ν , 0 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT ⋆ end_POSTSUPERSCRIPT end_ARG start_ARG italic_E start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + italic_ω + italic_i italic_ε end_ARG ] roman_e start_POSTSUPERSCRIPT - roman_i italic_k ( italic_p - italic_q ) end_POSTSUPERSCRIPT , (47)

which simplifies to eq 4.1 for k=0𝑘0k=0italic_k = 0.

Appendix C Quantum circuit for calculating the third-order susceptibility

In this appendix, we demonstrate that the quantum circuit presented in Figure 2 b yields the imaginary part of the expectation value G[𝜽1]|P0P1Uτ[𝜽3]Ut[𝜽2]P2Ut[𝜽2]P3Uτ[𝜽3]P4P5|G[𝜽1]bra𝐺delimited-[]superscript𝜽1subscript𝑃0subscript𝑃1superscriptsubscript𝑈𝜏delimited-[]superscript𝜽3superscriptsubscript𝑈𝑡delimited-[]superscript𝜽2subscript𝑃2subscript𝑈𝑡delimited-[]superscript𝜽2subscript𝑃3subscript𝑈𝜏delimited-[]superscript𝜽3subscript𝑃4subscript𝑃5ket𝐺delimited-[]superscript𝜽1\bra{G[\boldsymbol{\theta}^{1}]}P_{0}P_{1}U_{\tau}^{\dagger}[\boldsymbol{% \theta}^{3}]U_{t}^{\dagger}[\boldsymbol{\theta}^{2}]P_{2}\,U_{t}[\boldsymbol{% \theta}^{2}]P_{3}\,U_{\tau}[\boldsymbol{\theta}^{3}]P_{4}P_{5}\ket{G[% \boldsymbol{\theta}^{1}]}⟨ start_ARG italic_G [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ] end_ARG | italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_U start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ] italic_U start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] italic_P start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_U start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] italic_P start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT italic_U start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ] italic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT | start_ARG italic_G [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ] end_ARG ⟩. The system is initialized in a reference product state |φ0ketsubscript𝜑0\ket{\varphi_{0}}| start_ARG italic_φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ⟩. The ground state is prepared by applying the unitary operator UG[𝜽1]subscript𝑈Gdelimited-[]superscript𝜽1U_{\mathrm{G}}[\boldsymbol{\theta}^{1}]italic_U start_POSTSUBSCRIPT roman_G end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ] using AVQITE, leading to |G[𝜽1]=UG[𝜽1]|φ0ket𝐺delimited-[]superscript𝜽1subscript𝑈Gdelimited-[]superscript𝜽1ketsubscript𝜑0\ket{G[\boldsymbol{\theta}^{1}]}=U_{\mathrm{G}}[\boldsymbol{\theta}^{1}]\ket{% \varphi_{0}}| start_ARG italic_G [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ] end_ARG ⟩ = italic_U start_POSTSUBSCRIPT roman_G end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ] | start_ARG italic_φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ⟩. The ancilla qubit, initially in the state |0ket0\ket{0}| start_ARG 0 end_ARG ⟩, is given by 21/2(|0+i|1)superscript212ket0iket12^{-1/2}(\ket{0}+\mathrm{i}\ket{1})2 start_POSTSUPERSCRIPT - 1 / 2 end_POSTSUPERSCRIPT ( | start_ARG 0 end_ARG ⟩ + roman_i | start_ARG 1 end_ARG ⟩ ) after applying the Hadamard gate H𝐻Hitalic_H and S𝑆Sitalic_S-gate. As a result, the quantum state before applying the controlled P5subscript𝑃5P_{5}italic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT gate reads

|Ψ=12[|0|G[𝜽1]+i|1|G[𝜽1]].ketΨ12delimited-[]tensor-productket0ket𝐺delimited-[]superscript𝜽1tensor-productiket1ket𝐺delimited-[]superscript𝜽1\displaystyle\ket{\Psi}=\frac{1}{\sqrt{2}}\left[\ket{0}\otimes\ket{G[% \boldsymbol{\theta}^{1}]}+\mathrm{i}\ket{1}\otimes\ket{G[\boldsymbol{\theta}^{% 1}]}\right]\,.| start_ARG roman_Ψ end_ARG ⟩ = divide start_ARG 1 end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG [ | start_ARG 0 end_ARG ⟩ ⊗ | start_ARG italic_G [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ] end_ARG ⟩ + roman_i | start_ARG 1 end_ARG ⟩ ⊗ | start_ARG italic_G [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ] end_ARG ⟩ ] . (48)

Applying the controlled P5subscript𝑃5P_{5}italic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT and P4subscript𝑃4P_{4}italic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT operations controlled by the ancilla qubit leads to the quantum state

|Ψ=12[|0|G[𝜽1]+i|1P4P5|G[𝜽1]].ketΨ12delimited-[]tensor-productket0ket𝐺delimited-[]superscript𝜽1tensor-productiket1subscript𝑃4subscript𝑃5ket𝐺delimited-[]superscript𝜽1\displaystyle\ket{\Psi}=\frac{1}{\sqrt{2}}\left[\ket{0}\otimes\ket{G[% \boldsymbol{\theta}^{1}]}+\mathrm{i}\ket{1}\otimes P_{4}P_{5}\ket{G[% \boldsymbol{\theta}^{1}]}\right]\,.| start_ARG roman_Ψ end_ARG ⟩ = divide start_ARG 1 end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG [ | start_ARG 0 end_ARG ⟩ ⊗ | start_ARG italic_G [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ] end_ARG ⟩ + roman_i | start_ARG 1 end_ARG ⟩ ⊗ italic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT | start_ARG italic_G [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ] end_ARG ⟩ ] . (49)

Applying the X𝑋Xitalic_X gate on the ancilla qubit followed by the controlled P0subscript𝑃0P_{0}italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and P1subscript𝑃1P_{1}italic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT gates yields

|Ψ=12[|1P1P0|G[𝜽1]+i|0P4P5|G[𝜽1]].ketΨ12delimited-[]tensor-productket1subscript𝑃1subscript𝑃0ket𝐺delimited-[]superscript𝜽1tensor-productiket0subscript𝑃4subscript𝑃5ket𝐺delimited-[]superscript𝜽1\displaystyle\ket{\Psi}=\frac{1}{\sqrt{2}}\left[\ket{1}\otimes P_{1}P_{0}\ket{% G[\boldsymbol{\theta}^{1}]}+\mathrm{i}\ket{0}\otimes P_{4}P_{5}\ket{G[% \boldsymbol{\theta}^{1}]}\right]\,.| start_ARG roman_Ψ end_ARG ⟩ = divide start_ARG 1 end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG [ | start_ARG 1 end_ARG ⟩ ⊗ italic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT | start_ARG italic_G [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ] end_ARG ⟩ + roman_i | start_ARG 0 end_ARG ⟩ ⊗ italic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT | start_ARG italic_G [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ] end_ARG ⟩ ] . (50)

The time-evolving state under AVQDS with time-dependent variational parameters 𝜽1(τ)superscript𝜽1𝜏\boldsymbol{\theta}^{1}(\tau)bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ( italic_τ ) and 𝜽3(τ)superscript𝜽3𝜏\boldsymbol{\theta}^{3}(\tau)bold_italic_θ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ( italic_τ ) becomes

|Ψ=12ketΨ12\displaystyle\ket{\Psi}=\frac{1}{\sqrt{2}}| start_ARG roman_Ψ end_ARG ⟩ = divide start_ARG 1 end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG [|1Uτ[𝜽3]P1P0|G[𝜽1]\displaystyle\left[\ket{1}\otimes U_{\tau}[\boldsymbol{\theta}^{3}]P_{1}P_{0}% \ket{G[\boldsymbol{\theta}^{1}]}\right.[ | start_ARG 1 end_ARG ⟩ ⊗ italic_U start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ] italic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT | start_ARG italic_G [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ] end_ARG ⟩
+i|0Uτ[𝜽3]P4P5|G[𝜽1]].\displaystyle\left.+\mathrm{i}\ket{0}\otimes U_{\tau}[\boldsymbol{\theta}^{3}]% P_{4}P_{5}\ket{G[\boldsymbol{\theta}^{1}]}\right]\,.+ roman_i | start_ARG 0 end_ARG ⟩ ⊗ italic_U start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ] italic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT | start_ARG italic_G [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ] end_ARG ⟩ ] . (51)

Application of the X𝑋Xitalic_X gate followed by the controlled P3subscript𝑃3P_{3}italic_P start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT gate leads to

|Ψ=12ketΨ12\displaystyle\ket{\Psi}=\frac{1}{\sqrt{2}}| start_ARG roman_Ψ end_ARG ⟩ = divide start_ARG 1 end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG [|0Uτ[𝜽3]P1P0|G[𝜽1]\displaystyle\left[\ket{0}\otimes U_{\tau}[\boldsymbol{\theta}^{3}]P_{1}P_{0}% \ket{G[\boldsymbol{\theta}^{1}]}\right.[ | start_ARG 0 end_ARG ⟩ ⊗ italic_U start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ] italic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT | start_ARG italic_G [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ] end_ARG ⟩
+i|1P3Uτ[𝜽3]P4P5|G[𝜽1]].\displaystyle\left.+\mathrm{i}\ket{1}\otimes P_{3}U_{\tau}[\boldsymbol{\theta}% ^{3}]P_{4}P_{5}\ket{G[\boldsymbol{\theta}^{1}]}\right]\,.+ roman_i | start_ARG 1 end_ARG ⟩ ⊗ italic_P start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT italic_U start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ] italic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT | start_ARG italic_G [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ] end_ARG ⟩ ] . (52)

Propagating this state in time using AVQDS with time-dependent variational parameters 𝜽1(t)superscript𝜽1𝑡\boldsymbol{\theta}^{1}(t)bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ( italic_t ), 𝜽3(t)superscript𝜽3𝑡\boldsymbol{\theta}^{3}(t)bold_italic_θ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ( italic_t ), and 𝜽2(t)superscript𝜽2𝑡\boldsymbol{\theta}^{2}(t)bold_italic_θ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_t ) produces

|Ψ=12ketΨ12\displaystyle\ket{\Psi}=\frac{1}{\sqrt{2}}| start_ARG roman_Ψ end_ARG ⟩ = divide start_ARG 1 end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG [|0Ut[𝜽2]Uτ[𝜽3]P1P0|G[𝜽1]\displaystyle\left[\ket{0}\otimes U_{t}[\boldsymbol{\theta}^{2}]U_{\tau}[% \boldsymbol{\theta}^{3}]P_{1}P_{0}\ket{G[\boldsymbol{\theta}^{1}]}\right.[ | start_ARG 0 end_ARG ⟩ ⊗ italic_U start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] italic_U start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ] italic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT | start_ARG italic_G [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ] end_ARG ⟩
+i|1Ut[𝜽2]P3Uτ[𝜽3]P4P5|G[𝜽1]].\displaystyle\left.+\mathrm{i}\ket{1}\otimes U_{t}[\boldsymbol{\theta}^{2}]P_{% 3}U_{\tau}[\boldsymbol{\theta}^{3}]P_{4}P_{5}\ket{G[\boldsymbol{\theta}^{1}]}% \right]\,.+ roman_i | start_ARG 1 end_ARG ⟩ ⊗ italic_U start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] italic_P start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT italic_U start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ] italic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT | start_ARG italic_G [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ] end_ARG ⟩ ] . (53)

The application of the X𝑋Xitalic_X gate followed by the controlled P2subscript𝑃2P_{2}italic_P start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT gate results in the state

|Ψ=12ketΨ12\displaystyle\ket{\Psi}=\frac{1}{\sqrt{2}}| start_ARG roman_Ψ end_ARG ⟩ = divide start_ARG 1 end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG [|1P2Ut[𝜽2]Uτ[𝜽3]P1P0|G[𝜽1]\displaystyle\left[\ket{1}\otimes P_{2}U_{t}[\boldsymbol{\theta}^{2}]U_{\tau}[% \boldsymbol{\theta}^{3}]P_{1}P_{0}\ket{G[\boldsymbol{\theta}^{1}]}\right.[ | start_ARG 1 end_ARG ⟩ ⊗ italic_P start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_U start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] italic_U start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ] italic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT | start_ARG italic_G [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ] end_ARG ⟩
+i|0Ut[𝜽2]P3Uτ[𝜽3]P4P5|G[𝜽1]].\displaystyle\left.+\mathrm{i}\ket{0}\otimes U_{t}[\boldsymbol{\theta}^{2}]P_{% 3}U_{\tau}[\boldsymbol{\theta}^{3}]P_{4}P_{5}\ket{G[\boldsymbol{\theta}^{1}]}% \right]\,.+ roman_i | start_ARG 0 end_ARG ⟩ ⊗ italic_U start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] italic_P start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT italic_U start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ] italic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT | start_ARG italic_G [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ] end_ARG ⟩ ] . (54)

After executing the Hadamard gate on the ancilla qubit, the quantum state is given by

|ΨketΨ\displaystyle\ket{\Psi}| start_ARG roman_Ψ end_ARG ⟩ =12|0[P2Ut[𝜽2]Uτ[𝜽3]P1P0\displaystyle=\frac{1}{2}\ket{0}\otimes\left[P_{2}U_{t}[\boldsymbol{\theta}^{2% }]U_{\tau}[\boldsymbol{\theta}^{3}]P_{1}P_{0}\right.= divide start_ARG 1 end_ARG start_ARG 2 end_ARG | start_ARG 0 end_ARG ⟩ ⊗ [ italic_P start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_U start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] italic_U start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ] italic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT
+iUt[𝜽2]P3Uτ[𝜽3]P4P5]|G[𝜽1]\displaystyle\qquad\qquad\left.+iU_{t}[\boldsymbol{\theta}^{2}]P_{3}U_{\tau}[% \boldsymbol{\theta}^{3}]P_{4}P_{5}\right]\ket{G[\boldsymbol{\theta}^{1}]}+ italic_i italic_U start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] italic_P start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT italic_U start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ] italic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT ] | start_ARG italic_G [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ] end_ARG ⟩
12|1[P2Ut[𝜽2]Uτ[𝜽3]P1P0\displaystyle-\frac{1}{2}\ket{1}\otimes\left[P_{2}U_{t}[\boldsymbol{\theta}^{2% }]U_{\tau}[\boldsymbol{\theta}^{3}]P_{1}P_{0}\right.- divide start_ARG 1 end_ARG start_ARG 2 end_ARG | start_ARG 1 end_ARG ⟩ ⊗ [ italic_P start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_U start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] italic_U start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ] italic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT
iUt[𝜽2]P3Uτ[𝜽3]P4P5]|G[𝜽1].\displaystyle\qquad\qquad\left.-iU_{t}[\boldsymbol{\theta}^{2}]P_{3}U_{\tau}[% \boldsymbol{\theta}^{3}]P_{4}P_{5}\right]\ket{G[\boldsymbol{\theta}^{1}]}\,.- italic_i italic_U start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] italic_P start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT italic_U start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ] italic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT ] | start_ARG italic_G [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ] end_ARG ⟩ . (55)

Performing a Z𝑍Zitalic_Z measurement on the ancilla qubit yields

Im[G[𝜽1]|P0P1Uτ[𝜽3]Ut[𝜽2]P2Ut[𝜽2]P3Uτ[𝜽3]P4P5|G[𝜽1]]Imdelimited-[]bra𝐺delimited-[]superscript𝜽1subscript𝑃0subscript𝑃1superscriptsubscript𝑈𝜏delimited-[]superscript𝜽3superscriptsubscript𝑈𝑡delimited-[]superscript𝜽2subscript𝑃2subscript𝑈𝑡delimited-[]superscript𝜽2subscript𝑃3subscript𝑈𝜏delimited-[]superscript𝜽3subscript𝑃4subscript𝑃5ket𝐺delimited-[]superscript𝜽1\displaystyle\mathrm{Im}[\bra{G[\boldsymbol{\theta}^{1}]}P_{0}P_{1}U_{\tau}^{% \dagger}[\boldsymbol{\theta}^{3}]U_{t}^{\dagger}[\boldsymbol{\theta}^{2}]P_{2}% \,U_{t}[\boldsymbol{\theta}^{2}]P_{3}\,U_{\tau}[\boldsymbol{\theta}^{3}]P_{4}P% _{5}\ket{G[\boldsymbol{\theta}^{1}]}]roman_Im [ ⟨ start_ARG italic_G [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ] end_ARG | italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_U start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ] italic_U start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] italic_P start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_U start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] italic_P start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT italic_U start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT [ bold_italic_θ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ] italic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT | start_ARG italic_G [ bold_italic_θ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ] end_ARG ⟩ ]
=2p|11,absent2subscript𝑝ket11\displaystyle=2p_{\ket{1}}-1\,,= 2 italic_p start_POSTSUBSCRIPT | start_ARG 1 end_ARG ⟩ end_POSTSUBSCRIPT - 1 , (56)

where p|1subscript𝑝ket1p_{\ket{1}}italic_p start_POSTSUBSCRIPT | start_ARG 1 end_ARG ⟩ end_POSTSUBSCRIPT is the probability for the ancilla to be measured in state |1ket1\ket{1}| start_ARG 1 end_ARG ⟩.

Appendix D Exact dynamics and spectrum of third-order susceptibility

In this appendix, we derive analytical expressions for the third-order susceptibility in time and frequency domains. A more detailed discussion about the linear, second, and third-order susceptibilities of the quantum spin model eq 36 can be found in ref 63.

By applying the completeness relation I^=ν|ΨνΨν|^𝐼subscript𝜈ketsubscriptΨ𝜈brasubscriptΨ𝜈\hat{I}=\sum_{\nu}\ket{\Psi_{\nu}}\bra{\Psi_{\nu}}over^ start_ARG italic_I end_ARG = ∑ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT | start_ARG roman_Ψ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG ⟩ ⟨ start_ARG roman_Ψ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG | with eigenstates |ΨνketsubscriptΨ𝜈\ket{\Psi_{\nu}}| start_ARG roman_Ψ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG ⟩ of the quantum spin Hamiltonian ^^\hat{\mathcal{H}}over^ start_ARG caligraphic_H end_ARG, eq 3.2 for α=β=γ=δ=z𝛼𝛽𝛾𝛿𝑧\alpha=\beta=\gamma=\delta=zitalic_α = italic_β = italic_γ = italic_δ = italic_z can be written as

χzzzz(3)(t,τ,0)=2NΘ(t)Θ(τ)μνλF0,μzFμ,νzFν,λzFλ,0zsubscriptsuperscript𝜒3𝑧𝑧𝑧𝑧𝑡𝜏02𝑁Θ𝑡Θ𝜏subscript𝜇𝜈𝜆subscriptsuperscript𝐹𝑧0𝜇subscriptsuperscript𝐹𝑧𝜇𝜈subscriptsuperscript𝐹𝑧𝜈𝜆subscriptsuperscript𝐹𝑧𝜆0\displaystyle\chi^{(3)}_{zzzz}(t,\tau,0)=\frac{2}{N}\Theta(t)\Theta(\tau)\sum_% {\mu\nu\lambda}F^{z}_{0,\mu}F^{z}_{\mu,\nu}F^{z}_{\nu,\lambda}F^{z}_{\lambda,0}italic_χ start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_z italic_z italic_z italic_z end_POSTSUBSCRIPT ( italic_t , italic_τ , 0 ) = divide start_ARG 2 end_ARG start_ARG italic_N end_ARG roman_Θ ( italic_t ) roman_Θ ( italic_τ ) ∑ start_POSTSUBSCRIPT italic_μ italic_ν italic_λ end_POSTSUBSCRIPT italic_F start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 , italic_μ end_POSTSUBSCRIPT italic_F start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_μ , italic_ν end_POSTSUBSCRIPT italic_F start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_ν , italic_λ end_POSTSUBSCRIPT italic_F start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_λ , 0 end_POSTSUBSCRIPT
×Im[eiE0(t+τ)iEμtiEντ+eiEν(t+τ)iEλtiE0τ\displaystyle\times\mathrm{Im}\left[\mathrm{e}^{\mathrm{i}\,E_{0}(t+\tau)-% \mathrm{i}\,E_{\mu}t-\mathrm{i}\,E_{\nu}\tau}+\mathrm{e}^{\mathrm{i}\,E_{\nu}(% t+\tau)-\mathrm{i}\,E_{\lambda}t-\mathrm{i}\,E_{0}\tau}\right.× roman_Im [ roman_e start_POSTSUPERSCRIPT roman_i italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_t + italic_τ ) - roman_i italic_E start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT italic_t - roman_i italic_E start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT italic_τ end_POSTSUPERSCRIPT + roman_e start_POSTSUPERSCRIPT roman_i italic_E start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ( italic_t + italic_τ ) - roman_i italic_E start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT italic_t - roman_i italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_τ end_POSTSUPERSCRIPT
2eiEμ(t+τ)iEνtiEλτ].\displaystyle\qquad\quad\left.-2\,\mathrm{e}^{\mathrm{i}\,E_{\mu}(t+\tau)-% \mathrm{i}\,E_{\nu}t-\mathrm{i}\,E_{\lambda}\tau}\right]\,.- 2 roman_e start_POSTSUPERSCRIPT roman_i italic_E start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT ( italic_t + italic_τ ) - roman_i italic_E start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT italic_t - roman_i italic_E start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT italic_τ end_POSTSUPERSCRIPT ] . (57)

Here, Eνsubscript𝐸𝜈E_{\nu}italic_E start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT denotes the eigenenergies of ^^\hat{\mathcal{H}}over^ start_ARG caligraphic_H end_ARG and Fj,kzΨj|S^z|Ψksubscriptsuperscript𝐹𝑧𝑗𝑘quantum-operator-productsubscriptΨ𝑗superscript^𝑆𝑧subscriptΨ𝑘F^{z}_{j,k}\equiv\langle\Psi_{j}|\hat{S}^{z}|\Psi_{k}\rangleitalic_F start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j , italic_k end_POSTSUBSCRIPT ≡ ⟨ roman_Ψ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT | over^ start_ARG italic_S end_ARG start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT | roman_Ψ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ⟩ corresponds to the magnetic dipole matrix elements along the z𝑧zitalic_z-direction. Figure 9 b shows eq D calculated for the two-site spin-1 model. The eigenenergies and eigenstates Eμsubscript𝐸𝜇E_{\mu}italic_E start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT and ΨνsubscriptΨ𝜈\Psi_{\nu}roman_Ψ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT used to calculate χzzzz(3)(t,τ,0)subscriptsuperscript𝜒3𝑧𝑧𝑧𝑧𝑡𝜏0\chi^{(3)}_{zzzz}(t,\tau,0)italic_χ start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_z italic_z italic_z italic_z end_POSTSUBSCRIPT ( italic_t , italic_τ , 0 ) were obtained using exact diagonalization.

To transform χzzzz(3)(t,τ,0)subscriptsuperscript𝜒3𝑧𝑧𝑧𝑧𝑡𝜏0\chi^{(3)}_{zzzz}(t,\tau,0)italic_χ start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_z italic_z italic_z italic_z end_POSTSUBSCRIPT ( italic_t , italic_τ , 0 ) to the 2D frequency space, the 2D Fourier transform

χzzzz(3)(ωt,ωτ,0)subscriptsuperscript𝜒3𝑧𝑧𝑧𝑧subscript𝜔𝑡subscript𝜔𝜏0\displaystyle\chi^{(3)}_{zzzz}(\omega_{t},\omega_{\tau},0)italic_χ start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_z italic_z italic_z italic_z end_POSTSUBSCRIPT ( italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT , italic_ω start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT , 0 )
=0dt0dτχzzzz(3)(t,τ,0)ei(ωt+i 0+)tei(ωτ+i 0+)τabsentsuperscriptsubscript0differential-d𝑡superscriptsubscript0differential-d𝜏subscriptsuperscript𝜒3𝑧𝑧𝑧𝑧𝑡𝜏0superscripteisubscript𝜔𝑡isuperscript 0𝑡superscripteisubscript𝜔𝜏isuperscript 0𝜏\displaystyle=\int_{0}^{\infty}\mathrm{d}t\int_{0}^{\infty}\mathrm{d}\tau\,% \chi^{(3)}_{zzzz}(t,\tau,0)\mathrm{e}^{\mathrm{i}\,(\omega_{t}+\mathrm{i}\,0^{% +})t}\mathrm{e}^{\mathrm{i}\,(\omega_{\tau}+\mathrm{i}\,0^{+})\tau}= ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT roman_d italic_t ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT roman_d italic_τ italic_χ start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_z italic_z italic_z italic_z end_POSTSUBSCRIPT ( italic_t , italic_τ , 0 ) roman_e start_POSTSUPERSCRIPT roman_i ( italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT + roman_i 0 start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT ) italic_t end_POSTSUPERSCRIPT roman_e start_POSTSUPERSCRIPT roman_i ( italic_ω start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT + roman_i 0 start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT ) italic_τ end_POSTSUPERSCRIPT (58)

is applied, yielding 63

χzzzz(3)(ωt,ωτ,0)subscriptsuperscript𝜒3𝑧𝑧𝑧𝑧subscript𝜔𝑡subscript𝜔𝜏0\displaystyle\chi^{(3)}_{zzzz}(\omega_{t},\omega_{\tau},0)italic_χ start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_z italic_z italic_z italic_z end_POSTSUBSCRIPT ( italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT , italic_ω start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT , 0 )
=iNμ,ν,λSzzzzμ,ν,λ[L0,μ(ωt)L0,ν(ωτ)Lμ,ν(ωt)L0,ν(ωτ)\displaystyle=\frac{\mathrm{i}}{N}\sum_{\mu,\nu,\lambda}S^{\mu,\nu,\lambda}_{% zzzz}\left[L_{0,\mu}(\omega_{t})L_{0,\nu}(\omega_{\tau})-L_{\mu,\nu}(\omega_{t% })L_{0,\nu}(\omega_{\tau})\right.= divide start_ARG roman_i end_ARG start_ARG italic_N end_ARG ∑ start_POSTSUBSCRIPT italic_μ , italic_ν , italic_λ end_POSTSUBSCRIPT italic_S start_POSTSUPERSCRIPT italic_μ , italic_ν , italic_λ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_z italic_z italic_z italic_z end_POSTSUBSCRIPT [ italic_L start_POSTSUBSCRIPT 0 , italic_μ end_POSTSUBSCRIPT ( italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ) italic_L start_POSTSUBSCRIPT 0 , italic_ν end_POSTSUBSCRIPT ( italic_ω start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT ) - italic_L start_POSTSUBSCRIPT italic_μ , italic_ν end_POSTSUBSCRIPT ( italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ) italic_L start_POSTSUBSCRIPT 0 , italic_ν end_POSTSUBSCRIPT ( italic_ω start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT )
2Lμ,ν(ωt)Lμ,λ(ωτ)+2Lν,λ(ωt)Lμ,λ(ωτ)2subscript𝐿𝜇𝜈subscript𝜔𝑡subscript𝐿𝜇𝜆subscript𝜔𝜏2subscript𝐿𝜈𝜆subscript𝜔𝑡subscript𝐿𝜇𝜆subscript𝜔𝜏\displaystyle\qquad\qquad\qquad\left.-2L_{\mu,\nu}(\omega_{t})L_{\mu,\lambda}(% \omega_{\tau})+2L_{\nu,\lambda}(\omega_{t})L_{\mu,\lambda}(\omega_{\tau})\right.- 2 italic_L start_POSTSUBSCRIPT italic_μ , italic_ν end_POSTSUBSCRIPT ( italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ) italic_L start_POSTSUBSCRIPT italic_μ , italic_λ end_POSTSUBSCRIPT ( italic_ω start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT ) + 2 italic_L start_POSTSUBSCRIPT italic_ν , italic_λ end_POSTSUBSCRIPT ( italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ) italic_L start_POSTSUBSCRIPT italic_μ , italic_λ end_POSTSUBSCRIPT ( italic_ω start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT )
+Lν,λ(ωt)Lν,0(ωτ)Lλ,0(ωt)Lν,0(ωτ)],\displaystyle\qquad\qquad\qquad\left.+L_{\nu,\lambda}(\omega_{t})L_{\nu,0}(% \omega_{\tau})-L_{\lambda,0}(\omega_{t})L_{\nu,0}(\omega_{\tau})\right]\,,+ italic_L start_POSTSUBSCRIPT italic_ν , italic_λ end_POSTSUBSCRIPT ( italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ) italic_L start_POSTSUBSCRIPT italic_ν , 0 end_POSTSUBSCRIPT ( italic_ω start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT ) - italic_L start_POSTSUBSCRIPT italic_λ , 0 end_POSTSUBSCRIPT ( italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ) italic_L start_POSTSUBSCRIPT italic_ν , 0 end_POSTSUBSCRIPT ( italic_ω start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT ) ] , (59)

where

Szzzzμ,ν,λF0,μzFμ,νzFν,λzFλ,0z,subscriptsuperscript𝑆𝜇𝜈𝜆𝑧𝑧𝑧𝑧subscriptsuperscript𝐹𝑧0𝜇subscriptsuperscript𝐹𝑧𝜇𝜈subscriptsuperscript𝐹𝑧𝜈𝜆subscriptsuperscript𝐹𝑧𝜆0\displaystyle S^{\mu,\nu,\lambda}_{zzzz}\equiv F^{z}_{0,\mu}F^{z}_{\mu,\nu}F^{% z}_{\nu,\lambda}F^{z}_{\lambda,0}\,,italic_S start_POSTSUPERSCRIPT italic_μ , italic_ν , italic_λ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_z italic_z italic_z italic_z end_POSTSUBSCRIPT ≡ italic_F start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 , italic_μ end_POSTSUBSCRIPT italic_F start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_μ , italic_ν end_POSTSUBSCRIPT italic_F start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_ν , italic_λ end_POSTSUBSCRIPT italic_F start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_λ , 0 end_POSTSUBSCRIPT , (60)

and

Lμ,ν(ω)=1ω+i 0++EμEνsubscript𝐿𝜇𝜈𝜔1𝜔isuperscript 0subscript𝐸𝜇subscript𝐸𝜈\displaystyle L_{\mu,\nu}(\omega)=\frac{1}{\omega+\mathrm{i}\,0^{+}+E_{\mu}-E_% {\nu}}italic_L start_POSTSUBSCRIPT italic_μ , italic_ν end_POSTSUBSCRIPT ( italic_ω ) = divide start_ARG 1 end_ARG start_ARG italic_ω + roman_i 0 start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + italic_E start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG (61)

with 0+superscript00^{+}0 start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT denoting an infinitesimal positive quantity. According to eq 59, the peaks in the 2D spectra of the third-order susceptibility in Figure 9 e and 9 f emerge at energies determined by differences between eigenenergies EμEνsubscript𝐸𝜇subscript𝐸𝜈E_{\mu}-E_{\nu}italic_E start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT along both the ωtsubscript𝜔𝑡\omega_{t}italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT- and ωτsubscript𝜔𝜏\omega_{\tau}italic_ω start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT-axes. Consequently, these energy positions along ωtsubscript𝜔𝑡\omega_{t}italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT and ωτsubscript𝜔𝜏\omega_{\tau}italic_ω start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT in χzzzz(3)(ωt,ωτ,0)subscriptsuperscript𝜒3𝑧𝑧𝑧𝑧subscript𝜔𝑡subscript𝜔𝜏0\chi^{(3)}_{zzzz}(\omega_{t},\omega_{\tau},0)italic_χ start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_z italic_z italic_z italic_z end_POSTSUBSCRIPT ( italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT , italic_ω start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT , 0 ) describe two transitions between distinct eigenstates. The magnitude of the peaks in the 2D spectra is governed by Szzzzμ,ν,λsubscriptsuperscript𝑆𝜇𝜈𝜆𝑧𝑧𝑧𝑧S^{\mu,\nu,\lambda}_{zzzz}italic_S start_POSTSUPERSCRIPT italic_μ , italic_ν , italic_λ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_z italic_z italic_z italic_z end_POSTSUBSCRIPT, along with the count of contributing transitions to the signals.

Refer to caption
Figure 10: Spectral functions calculated with Padé approximation vs. discrete Fourier transformation. (a), (b) Spectral function Ak=0(ω)subscript𝐴𝑘0𝜔A_{k=0}(\omega)italic_A start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT ( italic_ω ) obtained by Fourier transforming the Green’s function dynamics presented in Figure 4 a for N=4𝑁4N=4italic_N = 4 and 4 b for N=6𝑁6N=6italic_N = 6 and calculating eq 29. The spectral function obtained by using the discrete Fourier transformation eq 31 (shaded area) is show alongside the one calculated using the Padé approximation (blue line). These results are obtained for a maximum simulation time of tmax=7subscript𝑡max7t_{\mathrm{max}}=7italic_t start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = 7 and a damping factor of ε=0.3𝜀0.3\varepsilon=0.3italic_ε = 0.3 in the Fourier transformation. (c), (d) The corresponding results for tmax=10subscript𝑡max10t_{\mathrm{max}}=10italic_t start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = 10. Compared to the spectral function obtained with Padé approximation, Ak=0(ω)subscript𝐴𝑘0𝜔A_{k=0}(\omega)italic_A start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT ( italic_ω ) calculated with discrete Fourier transformation requires a larger maximum simulation time to achieve comparable accuracy. (e), (f) |Sν,0|2superscriptsubscript𝑆𝜈02|S_{\nu,0}|^{2}| italic_S start_POSTSUBSCRIPT italic_ν , 0 end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (black dots) and |S0,ν|2superscriptsubscript𝑆0𝜈2|S_{0,\nu}|^{2}| italic_S start_POSTSUBSCRIPT 0 , italic_ν end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (red dots) as a function of the energy differences E0Eνsubscript𝐸0subscript𝐸𝜈E_{0}-E_{\nu}italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT and EνE0subscript𝐸𝜈subscript𝐸0E_{\nu}-E_{0}italic_E start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, respectively, for (e) N=4𝑁4N=4italic_N = 4 and (f) N=6𝑁6N=6italic_N = 6. Only the dominant transition amplitudes with |Sν,μ|2>0.02superscriptsubscript𝑆𝜈𝜇20.02|S_{\nu,\mu}|^{2}>0.02| italic_S start_POSTSUBSCRIPT italic_ν , italic_μ end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT > 0.02 are shown. The peaks in the spectral functions result from transitions between the ground state with energy E0subscript𝐸0E_{0}italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT to the excited states with energies Eνsubscript𝐸𝜈E_{\nu}italic_E start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT, and vice versa.

Appendix E Spectral function calculated with Padé approximation vs. discrete Fourier transformation

In this appendix, we demonstrate that the Padé approximation accelerates the convergence of the Fourier transform with respect to the required maximum simulation time compared to discrete Fourier transformation. Figure 10 a and 10 b present the spectral function Ak=0(ω)subscript𝐴𝑘0𝜔A_{k=0}(\omega)italic_A start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT ( italic_ω ) obtained by calculating the Fourier transform of the real-time Green’s function shown in Figure 4 a for N=4𝑁4N=4italic_N = 4 and Figure 4 b for N=6𝑁6N=6italic_N = 6, and evaluating eq 29. The spectral function obtained using the discrete Fourier transformation eq 31 (shaded area) is compared with the one calculated using the Padé approximation (blue line). These results are obtained for a maximum simulation time of tmax=7subscript𝑡max7t_{\mathrm{max}}=7italic_t start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = 7 and a damping factor of ε=0.3𝜀0.3\varepsilon=0.3italic_ε = 0.3 in the Fourier transformation. The corresponding results for tmax=10subscript𝑡max10t_{\mathrm{max}}=10italic_t start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = 10 are shown in Figure 10 c and 10 d. The spectral function obtained with the Padé approximation using tmax=7subscript𝑡max7t_{\mathrm{max}}=7italic_t start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = 7 resolves the dominant transitions between the ground state with energy E0subscript𝐸0E_{0}italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT to the excited states with energies Eνsubscript𝐸𝜈E_{\nu}italic_E start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT, and vice versa. This is seen by comparing the spectral functions in Figure 10 a and 10 b with the corresponding transition amplitudes |Sν,0|2superscriptsubscript𝑆𝜈02|S_{\nu,0}|^{2}| italic_S start_POSTSUBSCRIPT italic_ν , 0 end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (black dots) and |S0,ν|2superscriptsubscript𝑆0𝜈2|S_{0,\nu}|^{2}| italic_S start_POSTSUBSCRIPT 0 , italic_ν end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (red dots) plotted in Figure 10 e and 10 f. In contrast, the spectral function obtained with discrete Fourier transformation using tmax=7subscript𝑡max7t_{\mathrm{max}}=7italic_t start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = 7 does not resolve all dominant transitions in Figure 10 e and 10 f. For example, the splitting of the main peak in Figure 10 a is not resolved in the spectral function calculated with the discrete Fourier transformation (shaded area). This demonstrates that the accurate calculation of the spectral function with the discrete Fourier transformation requires a larger tmaxsubscript𝑡maxt_{\mathrm{max}}italic_t start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT. Specifically, for a maximum simulation time of tmax=10subscript𝑡max10t_{\mathrm{max}}=10italic_t start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = 10 (Figure 10 c and 10 d), the spectral functions obtained with the discrete Fourier transformation show comparable accuracy to those calculated with the Padé approximation.

Appendix F Prony approximation

The Prony approximation is a powerful technique for decomposing a signal into a sum of damped exponentials. When applied to the Green’s function G(t)𝐺𝑡G(t)italic_G ( italic_t ), this method allows the identification of the complex frequencies (poles) and corresponding residues that characterize the system’s spectral properties. Specifically, the Green’s functions, sampled at Ntsubscript𝑁𝑡N_{t}italic_N start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT discrete time points tnsubscript𝑡𝑛t_{n}italic_t start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT, is expressed as a sum of L𝐿Litalic_L complex exponentials:

G(tn)=i=1Lcieγitn,𝐺subscript𝑡𝑛superscriptsubscript𝑖1𝐿subscript𝑐𝑖superscript𝑒subscript𝛾𝑖subscript𝑡𝑛\displaystyle G(t_{n})=\sum_{i=1}^{L}c_{i}e^{-\gamma_{i}t_{n}}\,,italic_G ( italic_t start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) = ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_L end_POSTSUPERSCRIPT italic_c start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - italic_γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT end_POSTSUPERSCRIPT , (62)

where cisubscript𝑐𝑖c_{i}italic_c start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT are complex coefficients and γi=αi+iωisubscript𝛾𝑖subscript𝛼𝑖𝑖subscript𝜔𝑖\gamma_{i}=\alpha_{i}+i\omega_{i}italic_γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT + italic_i italic_ω start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT are the complex frequencies with damping factor αisubscript𝛼𝑖\alpha_{i}italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT (related to broadening), and oscillation frequency ωisubscript𝜔𝑖\omega_{i}italic_ω start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT (related to the energy levels). A linear prediction model is then constructed where each data point is predicted using a linear combination of the previous L𝐿Litalic_L data points:

G(tn)=i=1LaiG(tni)for nL.formulae-sequence𝐺subscript𝑡𝑛superscriptsubscript𝑖1𝐿subscript𝑎𝑖𝐺subscript𝑡𝑛𝑖for 𝑛𝐿\displaystyle G(t_{n})=-\sum_{i=1}^{L}a_{i}G(t_{n-i})\quad\text{for }n\geq L\,.italic_G ( italic_t start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) = - ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_L end_POSTSUPERSCRIPT italic_a start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_G ( italic_t start_POSTSUBSCRIPT italic_n - italic_i end_POSTSUBSCRIPT ) for italic_n ≥ italic_L . (63)

The coefficients aisubscript𝑎𝑖a_{i}italic_a start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT are determined by solving a linear system obtained from this model. The roots of the characteristic polynomial

p(z)=zL+a1zL1++aL𝑝𝑧superscript𝑧𝐿subscript𝑎1superscript𝑧𝐿1subscript𝑎𝐿\displaystyle p(z)=z^{L}+a_{1}z^{L-1}+\dots+a_{L}italic_p ( italic_z ) = italic_z start_POSTSUPERSCRIPT italic_L end_POSTSUPERSCRIPT + italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_z start_POSTSUPERSCRIPT italic_L - 1 end_POSTSUPERSCRIPT + ⋯ + italic_a start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT (64)

yields zi=eγiΔtsubscript𝑧𝑖superscript𝑒subscript𝛾𝑖Δ𝑡z_{i}=e^{\gamma_{i}\Delta t}italic_z start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_e start_POSTSUPERSCRIPT italic_γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT (i=1,,L𝑖1𝐿i=1,\dots,Litalic_i = 1 , … , italic_L) where Δt=t1t0Δ𝑡subscript𝑡1subscript𝑡0\Delta t=t_{1}-t_{0}roman_Δ italic_t = italic_t start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT represents the uniform time step of the signal. Once the zisubscript𝑧𝑖z_{i}italic_z start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT (or equivalently γisubscript𝛾𝑖\gamma_{i}italic_γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT) are determined, the residues cisubscript𝑐𝑖c_{i}italic_c start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT are calculated by solving a set of linear equations derived from the original signal data using the previously found γisubscript𝛾𝑖\gamma_{i}italic_γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. These residues represent the weights of each exponential component in the Green’s function. The spectral function A(ω)𝐴𝜔A(\omega)italic_A ( italic_ω ) is then constructed from the poles and residues. For each pole γisubscript𝛾𝑖\gamma_{i}italic_γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, a corresponding delta-like contribution is added to the spectral function:

A(ω)=1πIm[j=1Lcji(ωωj)+αj].𝐴𝜔1𝜋Imdelimited-[]superscriptsubscript𝑗1𝐿subscript𝑐𝑗𝑖𝜔subscript𝜔𝑗subscript𝛼𝑗\displaystyle A(\omega)=-\frac{1}{\pi}\mathrm{Im}\left[\sum_{j=1}^{L}\frac{c_{% j}}{i(\omega-\omega_{j})+\alpha_{j}}\right]\,.italic_A ( italic_ω ) = - divide start_ARG 1 end_ARG start_ARG italic_π end_ARG roman_Im [ ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_L end_POSTSUPERSCRIPT divide start_ARG italic_c start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG start_ARG italic_i ( italic_ω - italic_ω start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) + italic_α start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG ] . (65)

This representation effectively captures the spectral features, with broadening determined by the real part αjsubscript𝛼𝑗\alpha_{j}italic_α start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT of the poles and the location of the peaks given by ωjsubscript𝜔𝑗\omega_{j}italic_ω start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT. In the simulations, we include L=100𝐿100L=100italic_L = 100 terms in the expansion of the Prony approximation (eq (62)) which is sufficient to capture all dominant signals in the spectral function.

Appendix G Compressive sensing

Compressive sensing is a technique used to recover sparse signals from incomplete or noisy data. In the context of our analysis, the spectral function A(ω)𝐴𝜔A(\omega)italic_A ( italic_ω ) is assumed to be sparse, meaning it can be represented as a sum of a few significant delta-like peaks. To apply compressive sensing, we start with the discretized time-domain Green’s function G(tn)𝐺subscript𝑡𝑛G(t_{n})italic_G ( italic_t start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ), representing it as a linear combination of possible spectral components. This leads to a system of linear equations, where the goal is to find the spectral function A(ω)𝐴𝜔A(\omega)italic_A ( italic_ω ) that best fits the observed data:

G(tn)=A(ω)eiωtn𝑑ω.𝐺subscript𝑡𝑛superscriptsubscript𝐴𝜔superscript𝑒𝑖𝜔subscript𝑡𝑛differential-d𝜔\displaystyle G(t_{n})=\int_{-\infty}^{\infty}A(\omega)e^{-i\omega t_{n}}d\omega.italic_G ( italic_t start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) = ∫ start_POSTSUBSCRIPT - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_A ( italic_ω ) italic_e start_POSTSUPERSCRIPT - italic_i italic_ω italic_t start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_d italic_ω . (66)

This equation can be discretized over a grid of Nωsubscript𝑁𝜔N_{\omega}italic_N start_POSTSUBSCRIPT italic_ω end_POSTSUBSCRIPT frequencies ωmsubscript𝜔𝑚\omega_{m}italic_ω start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT to yield the following linear system:

G(tn)m=1NωA(ωm)eiωmtn,𝐺subscript𝑡𝑛superscriptsubscript𝑚1subscript𝑁𝜔𝐴subscript𝜔𝑚superscript𝑒𝑖subscript𝜔𝑚subscript𝑡𝑛\displaystyle G(t_{n})\approx\sum_{m=1}^{N_{\omega}}A(\omega_{m})e^{-i\omega_{% m}t_{n}},italic_G ( italic_t start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) ≈ ∑ start_POSTSUBSCRIPT italic_m = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT italic_ω end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_A ( italic_ω start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT ) italic_e start_POSTSUPERSCRIPT - italic_i italic_ω start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT end_POSTSUPERSCRIPT , (67)

or in matrix form 𝐆=𝚽𝐀𝐆𝚽𝐀\mathbf{G}=\mathbf{\Phi}\mathbf{A}bold_G = bold_Φ bold_A, where 𝐆𝐆\mathbf{G}bold_G is the vector of sampled values G(tn)𝐺subscript𝑡𝑛G(t_{n})italic_G ( italic_t start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ), 𝐀𝐀\mathbf{A}bold_A is the vector of spectral coefficients A(ωm)𝐴subscript𝜔𝑚A(\omega_{m})italic_A ( italic_ω start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT ), and 𝚽𝚽\mathbf{\Phi}bold_Φ is the measurement matrix with elements Φn,m=eiωmtnsubscriptΦ𝑛𝑚superscript𝑒𝑖subscript𝜔𝑚subscript𝑡𝑛\Phi_{n,m}=e^{-i\omega_{m}t_{n}}roman_Φ start_POSTSUBSCRIPT italic_n , italic_m end_POSTSUBSCRIPT = italic_e start_POSTSUPERSCRIPT - italic_i italic_ω start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT end_POSTSUPERSCRIPT. The matrix 𝚽𝚽\mathbf{\Phi}bold_Φ has dimensions Nt×Nωsubscript𝑁𝑡subscript𝑁𝜔N_{t}\times N_{\omega}italic_N start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT × italic_N start_POSTSUBSCRIPT italic_ω end_POSTSUBSCRIPT. Using compressive sensing, we solve the underdetermined system 𝐆=𝚽𝐀𝐆𝚽𝐀\mathbf{G}=\mathbf{\Phi}\mathbf{A}bold_G = bold_Φ bold_A by enforcing sparsity on the solution vector 𝐀𝐀\mathbf{A}bold_A. This is done by minimizing the 1subscript1\ell_{1}roman_ℓ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT-norm of 𝐀𝐀\mathbf{A}bold_A, subject to the constraint that the reconstructed Green’s function matches the observed data within a certain tolerance. This problem can be formulated as:

𝐀min=argmin𝐀(𝐆𝚽𝐀22+λ𝐀1),subscript𝐀minsubscriptargmin𝐀superscriptsubscriptnorm𝐆𝚽𝐀22𝜆subscriptnorm𝐀1\displaystyle\mathbf{A}_{\mathrm{min}}=\text{argmin}_{\mathbf{A}}\left(\|% \mathbf{G}-\mathbf{\Phi}\mathbf{A}\|_{2}^{2}+\lambda\|\mathbf{A}\|_{1}\right),bold_A start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT = argmin start_POSTSUBSCRIPT bold_A end_POSTSUBSCRIPT ( ∥ bold_G - bold_Φ bold_A ∥ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_λ ∥ bold_A ∥ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) , (68)

where λ𝜆\lambdaitalic_λ is a regularization parameter controlling the trade-off between the fidelity to the data (first term) and the sparsity of the solution (second term). The solution to this optimization problem is obtained using Python’s LASSO (Least Absolute Shrinkage and Selection Operator) function, which efficiently enforces sparsity by shrinking small coefficients to zero while preserving the significant ones. The resulting spectral function 𝐀min(ω)subscript𝐀min𝜔\mathbf{A}_{\mathrm{min}}(\omega)bold_A start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT ( italic_ω ) is highly accurate in reconstructing sharp features, such as delta peaks, even when the data contains noise. This method is particularly effective for cases with zero broadening, where it outperforms both the Padé and Prony approximations as demonstrated in section 4.1. In the calculation of the spectral functions, we use a regularization parameter of λ=2×103𝜆2superscript103\lambda=2\times 10^{-3}italic_λ = 2 × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT, which results in an mean squared error of less than 0.3%percent0.30.3~{}\%0.3 % between the actual Green’s function in the time domain and the one predicted by the spectral function using LASSO. As a general trend, the number of observable peaks in the A(ω)𝐴𝜔A(\omega)italic_A ( italic_ω ) increases with decreasing λ𝜆\lambdaitalic_λ.

{acknowledgement}

We acknowledge useful discussions with P. P. Orth and N. Gomes. We are grateful to the reviewers for many insightful suggestions, including additional signal processing techniques and an alternative way to measure response functions by leveraging mid-circuit measurements of the ancilla qubit. This work was supported by the U.S. Department of Energy (DOE), Office of Science, Basic Energy Sciences, Materials Science and Engineering Division, including the grant of computer time at the National Energy Research Scientific Computing Center (NERSC) in Berkeley, California. The research was performed at the Ames National Laboratory, which is operated for the U.S. DOE by Iowa State University under Contract No. DE-AC02-07CH11358.

References

  • Bruus and Flensberg 2004 Bruus, H.; Flensberg, K. Many-Body Quantum Theory in Condensed Matter Physics: An Introduction; Oxford Graduate Texts; OUP Oxford, 2004.
  • Stefanucci and van Leeuwen 2013 Stefanucci, G.; van Leeuwen, R. Nonequilibrium Many-Body Theory of Quantum Systems: A Modern Introduction; Cambridge University Press, 2013.
  • Mukamel 1995 Mukamel, S. Principles of Nonlinear Optical Spectroscopy; Oxford series in optical and imaging sciences; Oxford University Press, 1995.
  • Wan and Armitage 2019 Wan, Y.; Armitage, N. P. Resolving Continua of Fractional Excitations by Spinon Echo in THz 2D Coherent Spectroscopy. Phys. Rev. Lett. 2019, 122, 257401.
  • Nandkishore et al. 2021 Nandkishore, R. M.; Choi, W.; Kim, Y. B. Spectroscopic fingerprints of gapped quantum spin liquids, both conventional and fractonic. Phys. Rev. Res. 2021, 3, 013254.
  • Aryasetiawan and Gunnarsson 1998 Aryasetiawan, F.; Gunnarsson, O. The GW method. Rep. Prog. Phys. 1998, 61, 237.
  • Georges et al. 1996 Georges, A.; Kotliar, G.; Krauth, W.; Rozenberg, M. J. Dynamical mean-field theory of strongly correlated fermion systems and the limit of infinite dimensions. Rev. Mod. Phys. 1996, 68, 13.
  • Kotliar et al. 2006 Kotliar, G.; Savrasov, S. Y.; Haule, K.; Oudovenko, V. S.; Parcollet, O.; Marianetti, C. Electronic structure calculations with dynamical mean-field theory. Rev. Mod. Phys. 2006, 78, 865.
  • Aoki et al. 2014 Aoki, H.; Tsuji, N.; Eckstein, M.; Kollar, M.; Oka, T.; Werner, P. Nonequilibrium dynamical mean-field theory and its applications. Rev. Mod. Phys. 2014, 86, 779–837.
  • Kopnin 2001 Kopnin, N. Theory of Nonequilibrium Superconductivity; International Series of Monographs on Physics; Clarendon Press, 2001.
  • Schollwöck et al. 2008 Schollwöck, U.; Richter, J.; Farnell, D.; Bishop, R. Quantum Magnetism; Lecture Notes in Physics; Springer Berlin Heidelberg, 2008.
  • Gurarie 2011 Gurarie, V. Single-particle Green’s functions and interacting topological insulators. Phys. Rev. B 2011, 83, 085426.
  • Damascelli 2004 Damascelli, A. Probing the Electronic Structure of Complex Systems by ARPES. Phys. Scr. 2004, 2004, 61.
  • Mahan 2012 Mahan, G. Many-Particle Physics; Physics of Solids and Liquids; Springer US, 2012.
  • Lovesey 1986 Lovesey, S. Theory of Neutron Scattering from Condensed Matter; International series of monographs on physics; Clarendon Press, 1986.
  • Kuehn et al. 2009 Kuehn, W.; Reimann, K.; Woerner, M.; Elsaesser, T. Phase-resolved two-dimensional spectroscopy based on collinear n-wave mixing in the ultrafast time domain. J. Chem. Phys. 2009, 130, 164503.
  • Kuehn et al. 2011 Kuehn, W.; Reimann, K.; Woerner, M.; Elsaesser, T.; Hey, R. Two-Dimensional Terahertz Correlation Spectra of Electronic Excitations in Semiconductor Quantum Wells. J. Phys. Chem. B 2011, 115, 5448–5455.
  • Junginger et al. 2012 Junginger, F.; Mayer, B.; Schmidt, C.; Schubert, O.; Mährlein, S.; Leitenstorfer, A.; Huber, R.; Pashkin, A. Nonperturbative Interband Response of a Bulk InSb Semiconductor Driven Off Resonantly by Terahertz Electromagnetic Few-Cycle Pulses. Phys. Rev. Lett. 2012, 109, 147403.
  • Woerner et al. 2013 Woerner, M.; Kuehn, W.; Bowlan, P.; Reimann, K.; Elsaesser, T. Ultrafast Two-Dimensional Terahertz Spectroscopy of Elementary Excitations in Solids. New J. Phys. 2013, 15, 025039.
  • Maag et al. 2016 Maag, T.; Bayer, A.; Baierl, S.; Hohenleutner, M.; Korn, T.; Schüller, C.; Schuh, D.; Bougeard, D.; Lange, C.; Huber, R.; Mootz, M.; Sipe, J. E.; Koch, S. W.; Kira, M. Coherent cyclotron motion beyond Kohn’s theorem. Nat. Phys. 2016, 12, 119–123.
  • Yang et al. 2018 Yang, X.; Vaswani, C.; Sundahl, C.; Mootz, M.; Gagel, P.; Luo, L.; Kang, J. H.; Orth, P. P.; Perakis, I. E.; Eom, C. B.; Wang, J. Terahertz-light quantum tuning of a metastable emergent phase hidden by superconductivity. Nat. Mater. 2018, 17, 586–591.
  • Johnson et al. 2019 Johnson, C. L.; Knighton, B. E.; Johnson, J. A. Distinguishing Nonlinear Terahertz Excitation Pathways with Two-Dimensional Spectroscopy. Phys. Rev. Lett. 2019, 122, 073901.
  • Luo et al. 2023 Luo, L.; Mootz, M.; Kang, J. H.; Huang, C.; Eom, K.; Lee, J. W.; Vaswani, C.; Collantes, Y. G.; Hellstrom, E. E.; Perakis, I. E.; Eom, C. B.; Wang, J. Quantum coherence tomography of light-controlled superconductivity. Nat. Phys. 2023, 19, 201.
  • Luo et al. 2019 Luo, L.; Yang, X.; Liu, X.; Liu, Z.; Vaswani, C.; Cheng, D.; Mootz, M.; Zhao, X.; Yao, Y.; Wang, C.-Z.; Ho, K.-M.; Perakis, I. E.; Dobrowolska, M.; Furdyna, J. K.; Wang, J. Ultrafast manipulation of topologically enhanced surface transport driven by mid-infrared and terahertz pulses in Bi2Se3. Nat. Commun. 2019, 10, 607.
  • Yang et al. 2019 Yang, X.; Vaswani, C.; Sundahl, C.; Mootz, M.; Luo, L.; Kang, J. H.; Perakis, I. E.; Eom, C. B.; Wang, J. Lightwave-driven gapless superconductivity and forbidden quantum beats by terahertz symmetry breaking. Nat. Photon. 2019, 13, 707–713.
  • Vaswani et al. 2020 Vaswani, C.; Mootz, M.; Sundahl, C.; Mudiyanselage, D. H.; Kang, J. H.; Yang, X.; Cheng, D.; Huang, C.; Kim, R. H. J.; Liu, Z.; Luo, L.; Perakis, I. E.; Eom, C. B.; Wang, J. Terahertz Second-Harmonic Generation from Lightwave Acceleration of Symmetry-Breaking Nonlinear Supercurrents. Phys. Rev. Lett. 2020, 124, 207003.
  • Vaswani et al. 2020 Vaswani, C. et al. Light-Driven Raman Coherence as a Nonthermal Route to Ultrafast Topology Switching in a Dirac Semimetal. Phys. Rev. X 2020, 10, 021013.
  • Yang et al. 2020 Yang, X.; Luo, L.; Vaswani, C.; Zhao, X.; Yao, Y.; Cheng, D.; Liu, Z.; Kim, R. H. J.; Liu, X.; Dobrowolska-Furdyna, M.; Furdyna, J. K.; Perakis, I. E.; Wang, C.; Ho, K.; Wang, J. Light control of surface–bulk coupling by terahertz vibrational coherence in a topological insulator. npj Quantum Mater. 2020, 5, 13.
  • Mahmood et al. 2021 Mahmood, F.; Chaudhuri, D.; Gopalakrishnan, S.; Nandkishore, R.; Armitage, N. P. Observation of a Marginal Fermi Glass. Nat. Phys. 2021, 17, 627–631.
  • Vaswani et al. 2021 Vaswani, C.; Kang, J. H.; Mootz, M.; Luo, L.; Yang, X.; Sundahl, C.; Cheng, D.; Huang, C.; Kim, R. H. J.; Liu, Z.; Collantes, Y. G.; Hellstrom, E. E.; Perakis, I. E.; Eom, C. B.; Wang, J. Light quantum control of persisting Higgs modes in iron-based superconductors. Nat. Commun. 2021, 12, 258.
  • Song et al. 2023 Song, B.; Yang, X.; Sundahl, C.; Kang, J.-H.; Mootz, M.; Yao, Y.; Perakis, I.; Luo, L.; Eom, C.; Wang, J. Ultrafast Martensitic Phase Transition Driven by Intense Terahertz Pulses. Ultrafast Sci. 2023, 3, 0007.
  • Iskakov and Danilov 2018 Iskakov, S.; Danilov, M. Exact diagonalization library for quantum electron models. Comput. Phys. Commun. 2018, 225, 128–139.
  • Schollwoeck 2011 Schollwoeck, U. The density-matrix renormalization group in the age of matrix product states. Ann. Phys. 2011, 326, 96–192.
  • Gubernatis et al. 2016 Gubernatis, J.; Kawashima, N.; Werner, P. Quantum Monte Carlo Methods: Algorithms for Lattice Models; Cambridge University Press, 2016.
  • Feynman 1982 Feynman, R. P. Simulating physics with computers. Int. J. Theor. Phys. 1982, 21, 467–488.
  • Daley et al. 2022 Daley, A. J.; Bloch, I.; Kokail, C.; Flannigan, S.; Pearson, N.; Troyer, M.; Zoller, P. Practical quantum advantage in quantum simulation. Nature 2022, 607, 667–676.
  • Preskill 2018 Preskill, J. Quantum Computing in the NISQ era and beyond. Quantum 2018, 2, 79.
  • Hempel et al. 2018 Hempel, C.; Maier, C.; Romero, J.; McClean, J.; Monz, T.; Shen, H.; Jurcevic, P.; Lanyon, B. P.; Love, P.; Babbush, R.; Aspuru-Guzik, A.; Blatt, R.; Roos, C. F. Quantum Chemistry Calculations on a Trapped-Ion Quantum Simulator. Phys. Rev. X 2018, 8, 031022.
  • Kandala et al. 2017 Kandala, A.; Mezzacapo, A.; Temme, K.; Takita, M.; Brink, M.; Chow, J. M.; Gambetta, J. M. Hardware-efficient variational quantum eigensolver for small molecules and quantum magnets. Nature 2017, 549, 242–246.
  • Francis et al. 2020 Francis, A.; Freericks, J. K.; Kemper, A. F. Quantum computation of magnon spectra. Phys. Rev. B 2020, 101, 014411.
  • Chen et al. 2022 Chen, I.-C.; Burdick, B.; Yao, Y.-X.; Orth, P. P.; Iadecola, T. Error-mitigated simulation of quantum many-body scars on quantum computers with pulse-level control. Phys. Rev. Res. 2022, 4, 043027.
  • Pedernales et al. 2014 Pedernales, J. S.; Di Candia, R.; Egusquiza, I. L.; Casanova, J.; Solano, E. Efficient Quantum Algorithm for Computing n𝑛nitalic_n-time Correlation Functions. Phys. Rev. Lett. 2014, 113, 020505.
  • Baker 2021 Baker, T. E. Lanczos recursion on a quantum computer for the Green’s function and ground state. Phys. Rev. A 2021, 103, 032404.
  • Wecker et al. 2015 Wecker, D.; Hastings, M. B.; Wiebe, N.; Clark, B. K.; Nayak, C.; Troyer, M. Solving strongly correlated electron models on a quantum computer. Phys. Rev. A 2015, 92, 062318.
  • Kosugi and Matsushita 2020 Kosugi, T.; Matsushita, Y.-i. Construction of Green’s functions on a quantum computer: Quasiparticle spectra of molecules. Phys. Rev. A 2020, 101, 012330.
  • Roggero and Carlson 2019 Roggero, A.; Carlson, J. Dynamic linear response quantum algorithm. Phys. Rev. C 2019, 100, 034610.
  • Bauer et al. 2016 Bauer, B.; Wecker, D.; Millis, A. J.; Hastings, M. B.; Troyer, M. Hybrid quantum-classical approach to correlated materials. Phys. Rev. X 2016, 6, 031045.
  • Kreula et al. 2016 Kreula, J. M.; Clark, S. R.; Jaksch, D. Non-linear quantum-classical scheme to simulate non-equilibrium strongly correlated fermionic many-body dynamics. Sci. Rep. 2016, 6, 32940.
  • Chiesa et al. 2019 Chiesa, A.; Tacchino, F.; Grossi, M.; Santini, P.; Tavernelli, I.; Gerace, D.; Carretta, S. Quantum hardware simulating four-dimensional inelastic neutron scattering. Nat. Phys. 2019, 15, 455–459.
  • Del Re et al. 2024 Del Re, L.; Rost, B.; Foss-Feig, M.; Kemper, A. F.; Freericks, J. K. Robust Measurements of n𝑛nitalic_n-Point Correlation Functions of Driven-Dissipative Quantum Systems on a Digital Quantum Computer. Phys. Rev. Lett. 2024, 132, 100601.
  • Endo et al. 2020 Endo, S.; Kurata, I.; Nakagawa, Y. O. Calculation of the Green’s function on near-term quantum computers. Phys. Rev. Res. 2020, 2, 033281.
  • Chen et al. 2021 Chen, H.; Nusspickel, M.; Tilly, J.; Booth, G. H. Variational quantum eigensolver for dynamic correlation functions. Phys. Rev. A 2021, 104, 032405.
  • Keen et al. 2022 Keen, T.; Peng, B.; Kowalski, K.; Lougovski, P.; Johnston, S. Hybrid quantum-classical approach for coupled-cluster Green’s function theory. Quantum 2022, 6, 675.
  • Steckmann et al. 2023 Steckmann, T.; Keen, T.; Kökcü, E.; Kemper, A. F.; Dumitrescu, E. F.; Wang, Y. Mapping the metal-insulator phase diagram by algebraically fast-forwarding dynamics on a cloud quantum computer. Phys. Rev. Res. 2023, 5, 023198.
  • Yuan et al. 2019 Yuan, X.; Endo, S.; Zhao, Q.; Li, Y.; Benjamin, S. C. Theory of variational quantum simulation. Quantum 2019, 3, 191.
  • Libbi et al. 2022 Libbi, F.; Rizzo, J.; Tacchino, F.; Marzari, N.; Tavernelli, I. Effective calculation of the Green’s function in the time domain on near-term quantum processors. Phys. Rev. Res. 2022, 4, 043038.
  • Endo et al. 2020 Endo, S.; Sun, J.; Li, Y.; Benjamin, S. C.; Yuan, X. Variational quantum simulation of general processes. Phys. Rev. Lett. 2020, 125, 010501.
  • Nagano et al. 2023 Nagano, L.; Bapat, A.; Bauer, C. W. Quench dynamics of the Schwinger model via variational quantum algorithms. Phys. Rev. D 2023, 108, 034501.
  • Yao et al. 2021 Yao, Y.-X.; Gomes, N.; Zhang, F.; Wang, C.-Z.; Ho, K.-M.; Iadecola, T.; Orth, P. P. Adaptive Variational Quantum Dynamics Simulations. PRX Quantum 2021, 2, 030307.
  • Gomes et al. 2021 Gomes, N.; Mukherjee, A.; Zhang, F.; Iadecola, T.; Wang, C.-Z.; Ho, K.-M.; Orth, P. P.; Yao, Y.-X. Adaptive Variational Quantum Imaginary Time Evolution Approach for Ground State Preparation. Adv. Quantum Technol. 2021, 4, 2100114.
  • Grimsley et al. 2019 Grimsley, H. R.; Economou, S. E.; Barnes, E.; Mayhall, N. J. An adaptive variational algorithm for exact molecular simulations on a quantum computer. Nat. Commun. 2019, 10, 3007.
  • Gomes et al. 2023 Gomes, N.; Williams-Young, D. B.; de Jong, W. A. Computing the Many-Body Green’s Function with Adaptive Variational Quantum Dynamics. J. Chem. Theory Comput. 2023, 19, 3313–3323.
  • Mootz et al. 2024 Mootz, M.; Orth, P. P.; Huang, C.; Luo, L.; Wang, J.; Yao, Y.-X. Two-dimensional coherent spectrum of high-spin models via a quantum computing approach. Quantum Sci. Technol. 2024, 9, 035054.
  • McLachlan 1964 McLachlan, A. A variational solution of the time-dependent Schrodinger equation. Mol. Phys. 1964, 8, 39–44.
  • Meyer 2021 Meyer, J. J. Fisher Information in Noisy Intermediate-Scale Quantum Applications. Quantum 2021, 5, 539.
  • Getelina et al. 2023 Getelina, J. C.; Gomes, N.; Iadecola, T.; Orth, P. P.; Yao, Y.-X. Adaptive variational quantum minimally entangled typical thermal states for finite temperature simulations. SciPost Phys. 2023, 15, 102.
  • Jordan and Wigner 1993 Jordan, P.; Wigner, E. P. The Collected Works of Eugene Paul Wigner; Springer-Verlag Berlin Heidelberg New York, 1993; pp 109–129.
  • Lloyd 1996 Lloyd, S. Universal Quantum Simulators. Science 1996, 273, 1073–1078.
  • Smith et al. 2019 Smith, A.; Kim, M.; Pollmann, F.; Knolle, J. Simulating quantum many-body dynamics on a current digital quantum computer. npj Quantum Inf. 2019, 5, 1–13.
  • Childs et al. 2018 Childs, A. M.; Maslov, D.; Nam, Y.; Ross, N. J.; Su, Y. Toward the first quantum simulation with quantum speedup. PNAS 2018, 115, 9456–9461.
  • Choi et al. 2020 Choi, W.; Lee, K. H.; Kim, Y. B. Theory of Two-Dimensional Nonlinear Spectroscopy for the Kitaev Spin Liquid. Phys. Rev. Lett. 2020, 124, 117205.
  • Wang et al. 2020 Wang, Y.; Hu, Z.; Sanders, B. C.; Kais, S. Qudits and High-Dimensional Quantum Computing. Front. Phys. 2020, 8, 1.
  • Ogunkoya et al. 2024 Ogunkoya, O.; Kim, J.; Peng, B.; Barış Özgüler, A.; Alexeev, Y. Qutrit circuits and algebraic relations: A pathway to efficient spin-1 Hamiltonian simulation. Phys. Rev. A 2024, 109, 012426.
  • Dutta et al. 2024 Dutta, R. et al. Simulating Chemistry on Bosonic Quantum Devices. J. Chem. Theory Comput. 2024, 20, 6426–6441.
  • Stavenger et al. 2022 Stavenger, T. J.; Crane, E.; Smith, K. C.; Kang, C. T.; Girvin, S. M.; Wiebe, N. C2QA-bosonic qiskit. 2022 IEEE High Performance Extreme Computing Conference (HPEC). 2022; pp 1–8.
  • Getelina et al. 2024 Getelina, J. a. C.; Wang, C.-Z.; Iadecola, T.; Yao, Y.-X.; Orth, P. P. Adaptive variational ground state preparation for spin-1 models on qubit-based architectures. Phys. Rev. B 2024, 109, 085128.
  • Tang et al. 2021 Tang, H. L.; Shkolnikov, V.; Barron, G. S.; Grimsley, H. R.; Mayhall, N. J.; Barnes, E.; Economou, S. E. Qubit-ADAPT-VQE: An Adaptive Algorithm for Constructing Hardware-Efficient Ansätze on a Quantum Processor. PRX Quantum 2021, 2, 020310.
  • Gomes et al. 2020 Gomes, N.; Zhang, F.; Berthusen, N. F.; Wang, C.-Z.; Ho, K.-M.; Orth, P. P.; Yao, Y.-X. Efficient step-merged quantum imaginary time evolution algorithm for quantum chemistry. J. Chem. Theory Comput. 2020, 16, 6256–6266.
  • Yordanov et al. 2021 Yordanov, Y. S.; Armaos, V.; Barnes, C. H.; Arvidsson-Shukur, D. R. Qubit-excitation-based adaptive variational quantum eigensolver. Commun. Phys. 2021, 4, 1–11.
  • Mukherjee et al. 2023 Mukherjee, A.; Berthusen, N. F.; Getelina, J. C.; Orth, P. P.; Yao, Y.-X. Comparative study of adaptive variational quantum eigensolvers for multi-orbital impurity models. Commun. Phys. 2023, 6, 4.
  • Ryabinkin et al. 2018 Ryabinkin, I. G.; Yen, T.-C.; Genin, S. N.; Izmaylov, A. F. Qubit coupled cluster method: a systematic approach to quantum chemistry on a quantum computer. J. Chem. Theory Comput. 2018, 14, 6317–6326.
  • Kim et al. 2023 Kim, Y.; Eddins, A.; Anand, S.; Wei, K. X.; van den Berg, E.; Rosenblatt, S.; Nayfeh, H.; Wu, Y.; Zaletel, M.; Temme, K.; Kandala, A. Evidence for the Utility of Quantum Computing before Fault Tolerance. Nature 2023, 618, 500–505.
  • Yu et al. 2024 Yu, Y.; Kemper, A. F.; Yang, C.; Gull, E. Denoising of imaginary time response functions with Hankel projections. Phys. Rev. Res. 2024, 6, L032042.
  • Bruner et al. 2016 Bruner, A.; LaMaster, D.; Lopata, K. Accelerated broadband spectra using transition dipole decomposition and Padé approximants. J. Chem. Theory Comput. 2016, 12, 3741–3750.
  • Bruner et al. 2016 Bruner, A.; LaMaster, D.; Lopata, K. Accelerated Broadband Spectra Using Transition Dipole Decomposition and Padé Approximants. J. Chem. Theory Comput. 2016, 12, 3741–3750.
  • Virtanen et al. 2020 Virtanen, P. et al. SciPy 1.0: fundamental algorithms for scientific computing in Python. Nat. Methods 2020, 17, 261–272.
  • Gazizova et al. 2024 Gazizova, D.; Zhang, L.; Gull, E.; LeBlanc, J. P. F. Feynman diagrammatics based on discrete pole representations: A path to renormalized perturbation theories. Phys. Rev. B 2024, 110, 075158.
  • Zhang and Gull 2024 Zhang, L.; Gull, E. Minimal pole representation and controlled analytic continuation of Matsubara response functions. Phys. Rev. B 2024, 110, 035154.
  • Massa et al. 2015 Massa, A.; Rocca, P.; Oliveri, G. Compressive Sensing in Electromagnetics - A Review. IEEE Open J. Antennas Propag. 2015, 57, 224–238.
  • Qin et al. 2018 Qin, Z.; Fan, J.; Liu, Y.; Gao, Y.; Li, G. Y. Sparse Representation for Wireless Communications: A Compressive Sensing Approach. IEEE Signal Process. Mag. 2018, 35, 40–58.
  • Sun et al. 2018 Sun, Q.; Berkelbach, T. C.; Blunt, N. S.; Booth, G. H.; Guo, S.; Li, Z.; Liu, J.; McClain, J. D.; Sayfutyarova, E. R.; Sharma, S.; Wouters, S.; Chan, G. K.-L. PySCF: the Python-based simulations of chemistry framework. Wiley Interdiscip. Rev.: Comput. Mol. Sci. 2018, 8, e1340.
  • Di Matteo et al. 2021 Di Matteo, O.; McCoy, A.; Gysbers, P.; Miyagi, T.; Woloshyn, R. M.; Navrátil, P. Improving Hamiltonian encodings with the Gray code. Phys. Rev. A 2021, 103, 042405.
  • Bhandia et al. 2024 Bhandia, R.; Barbalas, D.; Xiao, R.; Chamorro, J. R.; Berry, T.; McQueen, T. M.; Samarth, N.; Armitage, N. P. Anomalous electronic energy relaxation and soft phonons in the Dirac semimetal Cd3As2subscriptCd3subscriptAs2{\mathrm{Cd}}_{3}{\mathrm{As}}_{2}roman_Cd start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT roman_As start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT. Phys. Rev. B 2024, 110, 075131.
  • Huang et al. 2024 Huang, C.; Luo, L.; Mootz, M.; Shang, J.; Man, P.; Su, L.; Perakis, I. E.; Yao, Y. X.; Wu, A.; Wang, J. Extreme terahertz magnon multiplication induced by resonant magnetic pulse pairs. Nat. Commun. 2024, 15, 3214.
  • Ghalgaoui et al. 2020 Ghalgaoui, A.; Koll, L.-M.; Schütte, B.; Fingerhut, B. P.; Reimann, K.; Woerner, M.; Elsaesser, T. Field-Induced Tunneling Ionization and Terahertz-Driven Electron Dynamics in Liquid Water. J. Phys. Chem. Lett. 2020, 11, 7717–7722.
  • Zhang et al. 2021 Zhang, Y.; Shi, J.; Li, X.; Coy, S. L.; Field, R. W.; Nelson, K. A. Nonlinear rotational spectroscopy reveals many-body interactions in water molecules. Proc. Natl. Acad. Sci. 2021, 118, e2020941118.
  • Mitarai and Fujii 2019 Mitarai, K.; Fujii, K. Methodology for replacing indirect measurements with direct measurements. Phys. Rev. Res. 2019, 1, 013006.
  • Motta et al. 2020 Motta, M.; Sun, C.; Tan, A. T.; O’Rourke, M. J.; Ye, E.; Minnich, A. J.; Brandão, F. G.; Chan, G. K.-L. Determining eigenstates and thermal states on a quantum computer using quantum imaginary time evolution. Nat. Phys. 2020, 16, 205–210.
  • White 2009 White, S. R. Minimally entangled typical quantum States at finite temperature. Phys. Rev. Lett. 2009, 102 19, 190601.
  • Cai et al. 2023 Cai, Z.; Babbush, R.; Benjamin, S. C.; Endo, S.; Huggins, W. J.; Li, Y.; McClean, J. R.; O’Brien, T. E. Quantum error mitigation. Rev. Mod. Phys. 2023, 95, 045005.
  • Getelina et al. 2024 Getelina, J. C.; Sharma, P.; Iadecola, T.; Orth, P. P.; Yao, Y.-X. Quantum subspace expansion in the presence of hardware noise. arXiv e-prints 2024, arXiv:2404.09132.