Simulating many-body open quantum systems (OQSs) is challenging due to the intricate interplay between the system and its environment, resulting in strong quantum correlations in both space and time. This Perspective presents an overview of recently developed theoretical methods using artificial intelligence (AI) and quantum computing (QC) to simulate the dynamics of these systems. We briefly introduce the dissipaton-embedded quantum master equation in second quantization, which provides a single master equation suitable for representation by neural quantum states or quantum circuits. The promising performance of AI- and QC-based approaches is demonstrated through preliminary research on simulating the quantum dissipative dynamics of many-body OQSs. We also discuss the limitations and future developments of these methods, which hold promise for overcoming the computational challenges associated with many-body OQS dynamics.

Many-body open quantum systems (OQSs) are ubiquitous in physics, chemistry, biology, and materials science. In an OQS, quantum correlations arise from the many-body quantum states formed by couplings between the system and its environment, as well as from the non-Markovian memory originating from the environment’s nontrivial energetic structure. Understanding and leveraging these quantum correlations in both space and time could facilitate advancements in thermoelectric transport in nano-devices,1–12 molecular spectroscopies13–20 and chemical reaction dynamics21–28 in condensed phase, quantum computing and quantum sensing utilizing surface-adsorbed molecular spins,29–40 etc. With ongoing advancements in these fields, quantum correlations are expected to play a crucial role in driving innovative discoveries and applications across science and technology.

The rapid progress of cutting-edge experiments demands greater accuracy, precision, robustness, and versatility from theoretical methods, especially in comprehensively describing realistic complex systems and relevant experimental conditions. To elucidate the underlying mechanisms behind experimental observations, various numerically exact methods for many-body OQSs have been developed. These methods include the hierarchical equations of motion (HEOM) approach for bosonic and fermionic environments,41–48 pseudomode theory,49–54 Davydov’s ansatz and its multistate extensions,55–60 stochastic equation of motion,61–70 quantum state diffusion,71 the hierarchy of stochastic pure states,72 and quantum Monte Carlo.73 Despite these advancements in simulating OQS dynamics, the “exponential wall” problem, as put forth by Kohn,74 significantly hinders the efficient simulation of increasingly complex many-body OQSs. This challenge arises because the number of degrees of freedom required for practical calculations grows exponentially with system size, many-body correlation strength, and the degree of environmental non-Markovianity.

To address this problem, efficient dimension reduction methods have been developed, including matrix product state (MPS),75 the density matrix renormalization group76 and its time-dependent extension,77 multi-configuration time-dependent Hartree and its multi-layer extension,78,79 and time-evolving matrix product operator.80,81 The application of MPS methods has been extended from isolated systems82–84 to OQSs.85–99 The combination of MPS and HEOM has enabled the simulation of dissipative dynamics in a large OQS consisting of 21 qubits.92 

In recent years, the growing demand for efficient processing and advanced problem-solving capabilities has led to the emergence of innovative methods based on artificial intelligence (AI)100–109 and quantum computing (QC).110–118 These technologies are reshaping the landscape of scientific and technological advancement, providing unprecedented opportunities in fields such as materials science, protein structure prediction, and quantum chemistry simulations.

In 2017, Carleo and Troyer100 pioneered the use of artificial neural networks to represent high-dimensional many-body wavefunctions of isolated quantum systems. This neural quantum state (NQS) approach enabled accurate calculations of ground-state energies and unitary time evolution, with computational costs increasing polynomially with system size. The NQS method was later extended to represent the reduced density matrices of OQSs, facilitating efficient simulations of quantum dissipative dynamics subject to the Markovian condition.101–105,119,120 Chen et al.121 have established the necessary and sufficient conditions for transforming tensor network states into restricted Boltzmann machines (RBMs) of specific architectures, showing that RBMs possess greater expressive power than tensor network states in representing quantum many-body states. Meanwhile, quantum computers,122–133 which are built of quantum mechanical elements that adhere to quantum laws, utilize quantum resources to manage complex quantum correlations more effectively. This capability holds great promise for solving problems that are impractical or impossible for classical computers.111,113–115,117,134,135

By harnessing the powerful representation capabilities of AI and the quantum parallelism inherent in QC,120 there is a promising prospect of breaking the exponential wall that has long constrained the simulation of many-body OQSs. However, current studies101,103–105,136–142 are limited and primarily focus on OQSs with weak quantum correlations in space or time. It is difficult for many existing numerically exact methods to integrate directly with AI and QC technologies. For instance, the hierarchical structure of the HEOM significantly complicates its mapping onto neural network models and quantum circuits. Therefore, there is an urgent need to develop novel theoretical frameworks, such as the recently developed dissipaton-embedded quantum master equation (DQME) theory, to better accommodate emerging AI and QC methods.

This Perspective focuses on AI- and QC-based theories and their applications in simulating many-body OQSs. The remainder of this paper is organized as follows: In Sec. II, a novel theoretical framework, the DQME in second quantization, is introduced for simulating the non-Markovian quantum dissipative dynamics of many-body OQSs. In Sec. III, we briefly discuss the computer software developed based on conventional algorithms for the simulation of OQSs. Sections IV and V discuss the representation of correlated quantum states and dissipative quantum dynamical equations using neural networks and quantum circuits, respectively. This is followed by the design of quantum machine learning algorithms for OQS dynamics in Sec. VI. We conclude by exploring the future prospects of AI- and QC-based approaches to many-body OQSs in Sec. VII.

We begin by briefly introducing the fermionic HEOM theory, which describes the quantum dissipative dynamics of OQSs coupled to non-Markovian environments, as illustrated in Fig. 1(a). For example, consider a single-orbital system coupled to a noninteracting electron reservoir. The extension to more complex OQSs is straightforward. The total Hamiltonian of the system plus environment is Htot = Hsys + Henv + Hcoup, where Hsys and Henv represent the system and the electron reservoir, respectively. The system–environment coupling Hamiltonian is Hcoup=ltld̂lâ+H.c.=F̂â+H.c. Here, F̂ltld̂l, with d̂l creating an electron in the lth band of the reservoir, â annihilating an electron from the system’s orbital, and tl representing the hopping integral between the reservoir’s lth band and the system’s orbital.

FIG. 1.

Arrows (1)–(3) illustrate the mappings from (a) the original non-Markovian environment to (b) the dissipatons in first quantization, to (c) those in second quantization, and to (d) the quantum circuit, respectively. [Reprinted with permission from Li et al., Phys. Rev. A 110, 032620 (2024). Copyright 2024 American Physical Society.]

FIG. 1.

Arrows (1)–(3) illustrate the mappings from (a) the original non-Markovian environment to (b) the dissipatons in first quantization, to (c) those in second quantization, and to (d) the quantum circuit, respectively. [Reprinted with permission from Li et al., Phys. Rev. A 110, 032620 (2024). Copyright 2024 American Physical Society.]

Close modal
The influence of non-Markovian memory from the electron reservoir is characterized by the hybridization correlation function Cσ(t). To construct the HEOM, Cσ(t) is decomposed into a sum of memory basis functions in the form of exponential functions (hereafter, we set e = = 1),
(1)
(2)
where σ = ±, σ̄σ, F̂+F̂, F̂F̂, and F̂σ(t)=eiĤenvtF̂σeiĤenvt. The coefficients {ηkσ} and exponents {γkσ} are complex numbers, and the value of K depends significantly on the reservoir’s temperature. Numerous memory decomposition schemes143–150 have been proposed to achieve an accurate decomposition of Eq. (2) with a minimal K, especially in low-temperature scenarios where non-Markovian effects are prominent.
The basic variables of the HEOM are the reduced density operator (RDO) and auxiliary density operators (ADOs). In the path-integral formulation, an ADO is defined by acting generating fields {Bjσ} on the RDO that represents the decoupled initial state of the system and environment, ρ0, as follows:151 
(3)
Here, Sf and Sb denote the forward and backward action functionals, respectively; Φ is the Feynman–Vernon influence functional;152 and Bj (Bj+) represents the transfer of a fermionic particle from the system (environment) to the environment (system) via the jth dissipation mode. Together, the RDO and ADOs form a high-dimensional reduced density tensor (RDT), the size of which increases exponentially with the number of memory basis functions, the terminal tier at which the hierarchy is truncated, and the dimension of the system.

The HEOM theory was first proposed by Tanimura and Kubo,41 and the detailed formalism for bosonic environments has been elaborated in Refs. 42–44 and 153–158. Both the fermionic and bosonic HEOM approaches have been used to benchmark results obtained from other methods for OQSs.44,52,53,68,159–161 As coupled differential equations, the HEOM typically involves a large number of ADOs that are distinct in their indices and parities. This inherent complexity makes it challenging to express the HEOM as a compact dynamical equation.

In 2014, Yan introduced the concept of the “dissipaton,”162 a statistical quasiparticle that captures the environment’s non-Markovian memory. This concept has led to the dissipaton equation of motion (DEOM) for describing OQS dynamics,157,158,162–164 where the dissipaton density operators serve as dynamical variables. The DEOM can be viewed as a generalization of the HEOM, as it allows for the evaluation of both the system’s observables and the hybrid system–environment response properties. However, like the HEOM, the complex hierarchical structure of the DEOM complicates its mapping onto neural networks and quantum circuits.

An alternative approach to utilizing the concept of dissipaton treats dissipatons as Brownian quasiparticles [red plates in Fig. 1(b)] interacting with the original OQS within a Markovian bath (light blue dots). This leads to the bosonic DQME in first quantization (DQME-FQ), expressed as follows:165 
(4)
where x=xk and Γk=γkxkxk+xk. The coefficients are given by ζk=(ηk+ηk̄*)/2 and ξk=(ηkηk̄*)/(2iζk). While Eq. (4) is fundamentally equivalent to the HEOM/DEOM, it offers the advantage of incorporating all system-plus-dissipaton degrees of freedom into a single dynamical equation. However, since the RDT of the dissipaton-embedded system depends on continuous coordinates, encoding the DQME-FQ for AI- and QC-based simulations remains challenging.

To address the limitations of the DQME-FQ, a second quantization formulation was developed that characterizes the exchange of particles and energy between the system and environment as the creation and annihilation of dissipatons.166  Figure 1(c) illustrates this concept, where the state of the environment is represented by a collection of dissipaton configurations, analogous to electron configurations representing the electronic structure of a chemical system. An “elementary” dissipation process is defined as the excitation from one dissipaton configuration to another, with each ADO in the HEOM/DEOM corresponding to such a process. This approach leads to the DQME in second quantization (DQME-SQ), which governs transitions between different dissipaton configurations. The theoretical construction of DQME-SQ is briefly outlined below.

Consider an OQS coupled to a bosonic environment, with the coupling Hamiltonian given by Hcoup=Q̂F̂, where Q̂ is a system operator. The dissipaton decomposition for the environment’s hybridization operators reads
(5)
with f̂k(t)f̂k(0)D=δkkηkeγkt. Here, the subscript D indicates that the expectation value is taken over the vacuum state of the dissipaton space. This decomposition ensures that the hybridization correlation function can be accurately recovered. The bosonic DQME-SQ, which describes the quantum dynamics of the dissipaton-embedded system, is expressed as166 
(6)
where ρ̃ is the RDT of the dissipaton-embedded system and N̂kb̂k+b̂k is the dissipaton number operator. The creation and annihilation operators, b̂k+ and b̂k, satisfy [b̂k,b̂k+]=δkk. On the right-hand side of Eq. (6), the second and last two terms account for the decay of dissipatons and their interaction with the system, respectively.
The fermionic DQME-SQ has been derived similarly,166 which is expressed as
(7)
where the coefficients are given by ζkσ=(ηkσηkσ̄*)14 and ξkσ=ηkσ/ζkσ.

The RDO of the system is obtained by projecting ρ̃ onto the dissipaton vacuum state.166 It is straightforward to generalize Eqs. (6) and (7) to cases where the system is coupled to the environment through multiple channels.166 The DQME-SQ theory is formally exact if the environment follows Gaussian statistics and the exponential decomposition of hybridization correlation function holds.

We highlight the advantages of DQME-SQ as follows: First, the quasiparticle description for environments allows for the treatment of both system and environment degrees of freedom in a unified manner. Second, compared to the continuous-variable representation of the RDT in DQME-FQ [cf. Eq. (4)], the dissipaton degrees of freedom are discrete in DQME-SQ, enabling easier encoding into neural quantum states or qubit states. Third, the DQME-SQ is formulated in a time-local structure and takes a much simpler form of a single differential equation rather than a set of coupled equations, as compared to HEOM. Finally, by utilizing the inherent symmetry and sparsity of RDT, its mapping to neural networks and quantum circuits and the associated sampling procedure can be greatly simplified. These advantages make DQME-SQ particularly suitable for representation in neural quantum states or quantum circuits [Fig. 1(d)] for AI- or QC-based simulations of many-body OQSs.

The pseudomode theory was first proposed by Garraway49 and has since evolved into a formally exact approach to non-Markovian OQS dynamics, based on the exponential decomposition of the hybridization correlation function.50–54 In the framework of pseudomode theory, the exponential decomposition leads to a finite set of discrete, unphysical pseudomodes. Formulated as a Lindblad quantum master equation (QME), the pseudomode approach has been used to simulate the non-Markovian dynamics of both bosonic50 and fermionic OQSs.167 The relationship and differences between the DQME-SQ and pseudomode methods are discussed in Ref. 166.

Besides methodological advancements, computer software tools47,48,94,168–173 capable of accurately and efficiently calculating the stationary and dynamic properties of OQSs are crucial. Such software tools provide a golden standard for assessing results obtained from AI- or QC-based approaches. For instance, Johansson et al. and Shammah et al. have constructed and developed the Quantum Toolbox in Python (QuTiP), an object-oriented open-source framework, for simulating the dynamics of OQSs.174–177 Recently, the QuTiP-BoFiN library has been integrated with the QuTiP169 to implement HEOM calculations for both bosonic and fermionic environments. Guan et al. have developed a Python package mpsqd,171 which utilizes the MPS and the matrix product operator to represent the ADO/RDOs and the generalized Liouvillian of the HEOM, respectively, offering a robust and flexible framework to efficiently and accurately simulate the OQSs dynamics. Furthermore, we have developed a general-purpose program, HEOM for QUantum Impurity with a Correlated Kernel (HEOM-QUICK),47,48,147,178 and widely used to investigate thermoelectric transport,7–12 Kondo effect,46,179–187 and spin excitations188,189 in strongly correlated OQSs coupled to fermionic environments.

These software tools rely on conventional algorithms, which limit their simulations due to the exponential wall problem. For example, in the practical implementations of the HEOM-QUICK program, about 4NSNEL elements in the ADOs are processed in the computer memory, where NS denotes the number of single-particle states, NE denotes the dissipaton states, and L denotes the maximum allowed occupancy of dissipatons. Although various numerical algorithms have been designed to improve the efficiency of HEOM-QUICK,47,48 simulating OQSs of high degrees of freedom in the ultra-low-temperature regime is still a challenging task. Therefore, there is an urgent need to develop and refine AI- and QC-based algorithms for many-body OQSs, along with their corresponding software implementations.

Data-driven machine learning approaches, which develop and optimize models using large training datasets from conventional methods, have proven effective for simulating the quantum dynamics of OQSs.190–203 Here, we focus on a different class of machine learning approaches that do not rely on large external datasets but, instead, incorporate physical laws or constraints into their architecture and training process.204 In the context of OQSs, quantum correlations in both space and time are encoded in the models, while the quantum dynamical equation serves as a precise physical constraint guiding the training.205,206

Correlated quantum states in both isolated and open systems have been effectively represented using the expressive power of machine learning models. For instance, neural networks such as RBMs have been employed to represent the many-body wavefunction of pure states in isolated systems, resulting in the NQS representation Ψθ(n);100,109 see Table I. Here, θ denotes the set of variational parameters in the neural network and the vector n=(n1,n2,) indicates a specific occupancy pattern for the Fock states or a configuration of particles within the system. The NQS approach has since been extended to represent the RDOs of mixed states in Markovian OQSs, denoted as ρθ(n,n). Nagy and Savona et al. have introduced an RBM ansatz to express an RDO as103 
(8)
where θ ≡ (ai, bm, cl, Wli, Xmi); see Ref. 103 for details.
TABLE I.

An overview of NQS representation for basic variables in the quantum dynamical equations for isolated and open quantum systems.

NQS variablesDynamical equationType of system
Wavefunction: Ψθ(n) Schrödinger equation Isolated system 
RDO: ρθ(n,n) Lindblad QME Markovian OQS 
RDT: ρ̃θ(n,n,m) DQME-SQ Non-Markovian OQS 
NQS variablesDynamical equationType of system
Wavefunction: Ψθ(n) Schrödinger equation Isolated system 
RDO: ρθ(n,n) Lindblad QME Markovian OQS 
RDT: ρ̃θ(n,n,m) DQME-SQ Non-Markovian OQS 

Time evolution and stationary states are obtained by accurately solving the Lindblad QME using variational or Markov-chain Monte Carlo algorithms.103–105,136 The effectiveness and accuracy of these algorithms have been demonstrated by simulating the dissipative dynamics of one- and two-dimensional spin lattice models,103–105 and the simulation results agree well with those obtained with conventional methods. Recently, autoregressive models have also been employed to represent many-body quantum states,207–209 enabling an accurate and efficient description of local correlations. However, most existing studies focus on the Markovian dynamics of OQSs in bosonic environments, while the simulation of non-Markovian dynamics is considered the “ultimate toughness”120 for AI-based approaches to OQS dynamics and remains largely unexplored.

The DQME-SQ theory, represented by Eqs. (6) and (7), provides an effective framework for simulating the non-Markovian dissipative dynamics of many-body OQSs.166,210 As illustrated in Fig. 2, a neural network can be constructed to represent the RDT, ρ̃θ(n,n,m), where m indicates the environment’s dissipaton configuration, while n and n denote particle configurations within the system. In contrast to the NQS approach for Markovian dynamics,103–105,136 the neural network model depicted in Fig. 2 explicitly incorporates non-Markovian memory through the dissipaton configurations. Moreover, accurately recovering system–environment correlations requires careful design of the interconnections between nodes; see Ref. 210 for more details.

FIG. 2.

Schematic illustration of a neural network representation of ρ̃, the RDT in the DQME-SQ approach, where θ denotes the parameters of the neural network model. The blue lines within the rectangle represent the dissipaton states of the environment, with linewidths indicating the characteristic decay rates of these states. The red lines in the circle represent the system’s single-particle states.

FIG. 2.

Schematic illustration of a neural network representation of ρ̃, the RDT in the DQME-SQ approach, where θ denotes the parameters of the neural network model. The blue lines within the rectangle represent the dissipaton states of the environment, with linewidths indicating the characteristic decay rates of these states. The red lines in the circle represent the system’s single-particle states.

Close modal
In the NQS approach, the problem of time evolution of ρ̃ is transformed into the evolution of the parameters θ = {θi} of the neural network model as follows:
(9)
where ρ̃θi are evaluated using the chain rule of analytic differentiation and θ̇i are determined by using the time-dependent variational principle (TDVP),104 in which the following loss function is minimized:
(10)
Here, L denotes the generalized Liouvillian representing the right-hand side of a quantum dynamical equation, such as Eq. (6) or Eq. (7), and ‖…‖2 denotes the two-norm. The time-dependent variational parameters are optimized by solving the large linear equations with conventional numerical algorithms, presenting an outstanding challenge for the long-time dynamics of both Markovian105 and non-Markovian OQSs.

In contrast to the HEOM/DEOM approach, where the number of ADOs grows exponentially with increasing system size or non-Markovianity of the environment,48 the NQS approach is expected to scale polynomially with the complexity of the system or environment. In addition, the efficiency of the NQS approach can be further enhanced by exploiting the sparsity and symmetry of the RDT in the DQME-SQ framework. Consequently, AI-based methods are anticipated to enable the investigation of highly complex many-body NQSs that are currently inaccessible with existing simulation techniques.

The quantum dynamics of OQSs are intrinsically nonunitary, while gate-based operations in quantum computers are unitary. Therefore, novel algorithms are needed to map nonunitary dissipative dynamics into the unitary framework of quantum algorithms. Following the proposal of the first quantum algorithm for simulating Markovian OQS dynamics116 by Kliesch et al., and its subsequent improvement by Childs and Li,211 substantial progress has been made in recent years.139–142,212–218 Various efficient algorithms have been proposed, including those based on imaginary-time evolution,219–221 the time-dependent variational principle,222 and duality quantum algorithms.223,224 General theoretical frameworks for the quantum simulation of OQSs are discussed in Ref. 118.

Some of these algorithms have been implemented on noisy intermediate-scale quantum (NISQ) computers.141,218,225–228 However, QC algorithms capable of simulating the non-Markovian dynamics of OQSs remain scarce. Below, we demonstrate that the compact form of the DQME-SQ enables straightforward encoding of correlated quantum states of many-body OQSs into quantum circuits, allowing for digital quantum simulations of non-Markovian OQS dynamics.166 

In the context of DQME-SQ, the RDT of a system coupled to a fermionic environment is mapped onto a pure qubit state as follows:
(11)
where μ,νHsys are the system states. The states |m⟩ and |n⟩ are defined as
(12)
(13)
with nk = 1 if (k, +) ∈ (j1, …, jn) and nk = 0 otherwise, and mk = 1 if (k, −) ∈ (j1, …, jn) and mk = 0 otherwise.
The time evolution of ρ̃ can be formally expressed as
(14)
where Λ is the non-Hermitian dynamical generator associated with Eq. (7), expressed as follows:
(15)
Here, Isys and ID denote the identity operators in the system and dissipaton spaces, respectively, and the superscript T denotes the transpose of an operator. Equations (11) and (15) constitute the quantum encoding protocol for the fermionic DQME-SQ. Once mapped onto quantum circuits, various quantum algorithms can be implemented to realize the propagation of ρ̃(t).

Figure 3 illustrates the design of a quantum circuit for solving the DQME-SQ, which employs the linear combination of unitaries (LCU) method and Trotter decomposition to split the unitary operators U0(δt) and U±(δt) into simpler quantum logic units for time propagation. The target bits store the RDT state, which is propagated by the control bit. It is noteworthy that both the number of qubits and the circuit depth increase linearly with the degrees of freedom of the dissipatons. This contrasts with the HEOM/DEOM approach implemented with classical algorithms, where computational costs grow exponentially with the system size or the complexity of the environment. Thus, the favorable scalability of the QC-based DQME-SQ method makes it a viable and efficient strategy for simulating non-Markovian OQS dynamics.

FIG. 3.

Quantum circuit for the time propagation of DQME-SQ. The circuit input is ρ̃(t)0. The quantum propagator is expressed as U0(δt)=eiΛ0δt, where Λ0 ≡ (Λ + Λ)/2. The propagators U±(δt)=±ieiϵ(IiΛ1δt) use a sufficiently small real parameter ϵ, with Λ1 ≡ (Λ − Λ)/2. Here, H represents the Hadamard gate and Z denotes the Z-measurement. More details can be found in Ref. 166. [Reprinted with permission from Li et al., Phys. Rev. A 110, 032620 (2024). Copyright 2024 American Physical Society.]

FIG. 3.

Quantum circuit for the time propagation of DQME-SQ. The circuit input is ρ̃(t)0. The quantum propagator is expressed as U0(δt)=eiΛ0δt, where Λ0 ≡ (Λ + Λ)/2. The propagators U±(δt)=±ieiϵ(IiΛ1δt) use a sufficiently small real parameter ϵ, with Λ1 ≡ (Λ − Λ)/2. Here, H represents the Hadamard gate and Z denotes the Z-measurement. More details can be found in Ref. 166. [Reprinted with permission from Li et al., Phys. Rev. A 110, 032620 (2024). Copyright 2024 American Physical Society.]

Close modal

The feasibility of DQME-SQ is demonstrated by implementing the LCU-Trotter algorithm on the Aer simulator of Qiskit.229 As depicted in Fig. 4, in both the low and high temperature regimes, the population dynamics of a spin-boson model with digital quantum simulations agree remarkably with those obtained with HEOM method using classical algorithms. The exactness and universality of DQME-SQ are further corroborated by simulations of more complex OQSs; see Ref. 166 for more details.

FIG. 4.

Digital quantum simulation of the time evolution of population difference P1P0 between two system states of a spin-boson model at low and high temperatures (T). Results obtained by the HEOM are also given for comparison. Ω is the energy difference between two system states. [Reprinted with permission from Li et al., Phys. Rev. A 110, 032620 (2024). Copyright 2024 American Physical Society.]

FIG. 4.

Digital quantum simulation of the time evolution of population difference P1P0 between two system states of a spin-boson model at low and high temperatures (T). Results obtained by the HEOM are also given for comparison. Ω is the energy difference between two system states. [Reprinted with permission from Li et al., Phys. Rev. A 110, 032620 (2024). Copyright 2024 American Physical Society.]

Close modal

Recently, Hu et al.,225,226 Wang et al.,218 Head-Marsden et al.,227 and Sun et al.141 have successfully implemented quantum algorithms on NISQ devices to simulate OQSs, ranging from simple218,225,227 to more realistic model systems.141,226 The dissipative dynamics simulated are comparable to those obtained using classical numerical methods. In addition, it has been suggested that the intrinsic quantum noise from quantum gates can be harnessed to efficiently model the dissipative environment in a controllable manner.141 Most of these studies focus on the Markovian dynamics of OQSs in bosonic environments, while the simulation of non-Markovian dynamics in fermionic environments remains to be thoroughly explored.

The promising performance of AI- and QC-based approaches naturally suggests the development of quantum machine learning methods,120,139 which harness the power of both AI and QC algorithms. For instance, Long et al. have studied the Markovian dissipative dynamics of an OQS described by the Lindblad quantum master equation,230 
(16)
where ρ denotes the RDO of the OQS and γj and Lj represent the dissipation rate and the jump operator associated with the jth dissipative channel, respectively.
To enable a quantum neural network representation of correlated quantum states in OQSs, Ref. 230 proposes a hybrid quantum–classical algorithm to express the RDO elements using an RBM, as illustrated in Fig. 5,
(17)
where h and h′ represent the binary nodes in the left and right hidden layers of the bipartite RBM, respectively, and q corresponds to the nodes in the ancilla layer. The amplitudes ψσ(q, h) and ψη(q, h′) are directly obtained from the RBM. They are further mapped onto pure quantum states in an extended Hilbert space using a purified neural network ansatz,105 
(18a)
(18b)
Here, C and C′ are normalization constants and |σ, R⟩ = |σ⟩ ⊗|R⟩ serve as the basis vector of the extended Hilbert space, with R = {q, h, h′}. Finally, the above quantum NQS approach results in a tensor state,
(19)
with which the RDO of the OQS is evaluated as ρ=RR|χ|R. The quantum circuit for preparing |ψL⟩ is shown in Fig. 6(a), while the counterpart for |ψR⟩ is similar, with the gate operations on the h and h′ qubits swapped.
FIG. 5.

Schematic illustration of the bipartite RBM used in Ref. 230 to represent the RDO elements. The binary units σ, h, q ∈ {1, − 1} are associated with the biases {a, b, c}. The weights between units in different layers, {W, U}, are represented by the connecting lines. [Reprinted with permission from Long et al., J. Chem. Phys. 161, 084105 (2024). Copyright 2024 AIP Publishing LLC.]

FIG. 5.

Schematic illustration of the bipartite RBM used in Ref. 230 to represent the RDO elements. The binary units σ, h, q ∈ {1, − 1} are associated with the biases {a, b, c}. The weights between units in different layers, {W, U}, are represented by the connecting lines. [Reprinted with permission from Long et al., J. Chem. Phys. 161, 084105 (2024). Copyright 2024 AIP Publishing LLC.]

Close modal
FIG. 6.

(a) Quantum circuit for preparing |ψL⟩. Block B1 (B2) inputs biases {ai} ({bj, ck}) for the spin (hidden and ancilla) qubits. Block E inputs the weights connecting different layers. Blocks B1 and B2 are composed of single-qubit Ry- and Rz-gates, and block E is composed of Rzz-gates. H represents the Hadamard gates. (b) Workflow for calculating the expectation value of system observables with the quantum NQS representation. [Reprinted with permission from Long et al., J. Chem. Phys. 161, 084105 (2024). Copyright 2024 AIP Publishing LLC.]

FIG. 6.

(a) Quantum circuit for preparing |ψL⟩. Block B1 (B2) inputs biases {ai} ({bj, ck}) for the spin (hidden and ancilla) qubits. Block E inputs the weights connecting different layers. Blocks B1 and B2 are composed of single-qubit Ry- and Rz-gates, and block E is composed of Rzz-gates. H represents the Hadamard gates. (b) Workflow for calculating the expectation value of system observables with the quantum NQS representation. [Reprinted with permission from Long et al., J. Chem. Phys. 161, 084105 (2024). Copyright 2024 AIP Publishing LLC.]

Close modal
The expectation value of any observable Ô can be evaluated on the quantum circuit through
(20)
where IR is the identity operator in the R space. The workflow is illustrated in Fig. 6(b). A conditional optimization approach was developed to reduce quantum measurement errors,230 particularly in cases where ⟨ψR|ψL⟩ is small. The feasibility of the proposed quantum NQS approach for OQS dynamics has been demonstrated through simulations of the spin-boson and transverse field Ising models. The results agree consistently with those obtained using classical methods; see Ref. 230 for more details.

An alternative hybrid quantum–classical algorithm based on the NQS representation to solve Eq. (16) has been proposed by Lee et al.139 The OQS is represented by an ensemble of pure state trajectories, whose time evolution is described by the stochastic Schrödinger equation simulated by the time-dependent variational Monte Carlo algorithm. To obtain the accurate expectation value of observables, a large number of trajectories are required to perform the ensemble averaging, resulting in significant computational overhead.

In this Perspective, we have discussed recent advances in simulating many-body OQS dynamics using AI- and QC-based methods. By mapping the non-Markovian environment to statistically independent dissipatons, the DQME-SQ framework is established, represented as a single master equation. This formulation allows for efficient representation through neural quantum states or quantum circuits. These capabilities are exemplified by recent simulations of non-Markovian OQS dynamics based on RBMs and QC algorithms, which offer a compact encoding of the RDT and achieve accuracy comparable to the conventional HEOM/DEOM methods. Furthermore, the RBM representation has been implemented using quantum algorithms to design a quantum NQS representation for simulating the Markovian dynamics of many-body OQSs. This quantum machine learning method circumvents the computational costs associated with classical algorithms and can be extended to simulate non-Markovian dynamics.

Despite the encouraging and promising progress of currently developed AI- and QC-based algorithms in simulating many-body OQSs, the implementations are primarily limited to simple model systems with relatively low degrees of freedom. This limitation arises from the need to sample a large number of dissipaton configurations, which strongly bottlenecks the accuracy and efficiency of the TDVP algorithm. In addition, the depth of quantum circuits remains too large for full quantum simulations. Consequently, these algorithmic limitations severely hinder their versatility in simulating realistic, complex many-body OQSs.

Practical limitations stemming from hardware-related factors also hinder the deployment of QC algorithms for OQS simulations on current quantum devices.131,132,231–233 Quantum error correction is not yet feasible on near-term devices due to the high resource overhead required, such as the number of qubits and error rates needed to implement fault-tolerant gates.231 This limits the depth and fidelity of quantum circuits that can be executed reliably. Current quantum devices suffer from high noise levels and short coherence times,131 which introduce errors and reduce the accuracy of quantum computations. In addition, there are limitations in qubit connectivity and gate fidelity,232 which can restrict the types of circuits that can be implemented efficiently. Moreover, classical control hardware and software need to be optimized to handle the high overhead of quantum error mitigation techniques for noisy quantum circuits.132,233 Since the simulations of non-Markovian OQS dynamics have rarely been implemented on real quantum computers, a systemic estimation of the accuracy and scale required for current or near-term devices still remains to be further explored.

Future developments of AI-based simulation methods for many-body OQSs should focus on enhancing the efficiency of neural network parameter determination by employing a more sophisticated variational ansatz and cost functions, exploring alternative representations beyond RBMs, and representing the generalized Liouvillian in Eq. (10) by the quantum neural propagator.205,206 For QC-based approaches to OQS dynamics, it will be highly desirable to implement the proposed algorithms on real quantum computers in the NISQ regime and beyond. To this end, algorithms should be further simplified to reduce the circuit depth, and systemic errors and measurement costs due to unitary decomposition approximations should be more carefully accommodated. These developments may allow for precise simulations of OQSs with larger system sizes and more complex environments and/or external fields.

The support from the National Natural Science Foundation of China (Grant Nos. 22393912, 22425301, 22203083, 22103073, and 22373091), Innovation Program for Quantum Science and Technology (Grant No. 2021ZD0303306), and Strategic Priority Research Program of the Chinese Academy of Sciences (Grant No. XDB0450101) is gratefully acknowledged.

The authors have no conflicts to disclose.

Lyuzhou Ye: Funding acquisition (equal); Writing – original draft (equal); Writing – review & editing (equal). Yao Wang: Funding acquisition (equal); Writing – original draft (equal); Writing – review & editing (equal). Xiao Zheng: Conceptualization (equal); Funding acquisition (equal); Supervision (equal); Writing – original draft (equal); Writing – review & editing (equal).

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

1.
Y.
Dubi
and
M.
Di Ventra
,
Rev. Mod. Phys.
83
,
131
(
2011
).
2.
K.
Uchida
,
S.
Daimon
,
R.
Iguchi
, and
E.
Saitoh
,
Nature
558
,
95
(
2018
).
3.
Z.
Huang
,
F.
Chen
,
R.
D’agosta
,
P. A.
Bennett
,
M.
Di Ventra
, and
N.
Tao
,
Nat. Nanotechnol.
2
,
698
(
2007
).
4.
K. L.
Grosse
,
M.-H.
Bae
,
F.
Lian
,
E.
Pop
, and
W. P.
King
,
Nat. Nanotechnol.
6
,
287
(
2011
).
5.
L.
Cui
,
R.
Miao
,
K.
Wang
,
D.
Thompson
,
L. A.
Zotti
,
J. C.
Cuevas
,
E.
Meyhofer
, and
P.
Reddy
,
Nat. Nanotechnol.
13
,
122
(
2018
).
6.
K.
Kang
and
T.
Lee
,
Nat. Nanotechnol.
13
,
97
(
2018
).
7.
L.
Ye
,
D.
Hou
,
R.
Wang
,
D.
Cao
,
X.
Zheng
, and
Y.
Yan
,
Phys. Rev. B
90
,
165116
(
2014
).
8.
L.
Ye
,
D.
Hou
,
X.
Zheng
,
Y.
Yan
, and
M.
Di Ventra
,
Phys. Rev. B
91
,
205106
(
2015
).
9.
L.
Ye
,
X.
Zheng
,
Y.
Yan
, and
M.
Di Ventra
,
Phys. Rev. B
94
,
245105
(
2016
).
10.
X.
Zeng
,
L.
Ye
,
D.
Zhang
,
R.-X.
Xu
,
X.
Zheng
, and
M.
Di Ventra
,
Phys. Rev. B
103
,
085411
(
2021
).
11.
X.
Zeng
,
L.
Ye
,
R.-X.
Xu
, and
X.
Zheng
,
Chin. J. Chem. Phys.
36
,
404
(
2023
).
12.
X.
Zeng
,
D.
Zhang
,
L.
Ye
,
L.
Cao
,
R.-X.
Xu
,
X.
Zheng
, and
M.
Di Ventra
,
Phys. Rev. B
109
,
115423
(
2024
).
13.
T.
Elsaesser
,
Acc. Chem. Res.
42
,
1220
(
2009
).
14.
T.
Hasegawa
,
Quantitative Infrared Spectroscopy for Understanding of a Condensed Matter
(
Springer
,
2017
).
15.
J. C.
Dean
and
G. D.
Scholes
,
Acc. Chem. Res.
50
,
2746
(
2017
).
16.
Y.
Ozaki
,
K. B.
Beć
,
Y.
Morisawa
,
S.
Yamamoto
,
I.
Tanabe
,
C. W.
Huck
, and
T. S.
Hofer
,
Chem. Soc. Rev.
50
,
10917
(
2021
).
17.
A. M.
Kelley
,
Condensed-phase Molecular Spectroscopy and Photophysics
(
John Wiley & Sons
,
2022
).
18.
J.
Xu
,
X.
Zhu
,
S.
Tan
,
Y.
Zhang
,
B.
Li
,
Y.
Tian
,
H.
Shan
,
X.
Cui
,
A.
Zhao
,
Z.
Dong
,
J.
Yang
,
Y.
Luo
,
B.
Wang
, and
J. G.
Hou
,
Science
371
,
818
(
2021
).
19.
S.
Liu
,
A.
Hammud
,
I.
Hamada
,
M.
Wolf
,
M.
Müller
, and
T.
Kumagai
,
Sci. Adv.
8
,
eabq5682
(
2022
).
20.
Z.
Wang
,
L.
Balembois
,
M.
Rančić
,
E.
Billaud
,
M.
Le Dantec
,
A.
Ferrier
,
P.
Goldner
,
S.
Bertaina
,
T.
Chanelière
,
D.
Esteve
,
D.
Vion
,
P.
Bertet
, and
E.
Flurin
,
Nature
619
,
276
(
2023
).
21.
W.
Domcke
,
H.
Koppel
, and
D. R.
Yarkony
,
Conical Intersections: Electronic Structure, Dynamics & Spectroscopy
(
World Scientific
,
2004
), Vol.
15
.
22.
K. S.
Peters
,
Acc. Chem. Res.
42
,
89
(
2009
).
23.
L.
Chen
and
Q.
Shi
,
J. Chem. Phys.
130
,
134505
(
2009
).
24.
R. P.
Bell
,
The Tunnel Effect in Chemistry
(
Springer
,
2013
).
25.
S.
Hu
,
S. C.
Sharma
,
A. D.
Scouras
,
A. V.
Soudackov
,
C. A. M.
Carr
,
S.
Hammes-Schiffer
,
T.
Alber
, and
J. P.
Klinman
,
J. Am. Chem. Soc.
136
,
8157
(
2014
).
26.
W.
Xie
,
Y.
Xu
,
L.
Zhu
, and
Q.
Shi
,
J. Chem. Phys.
140
,
174105
(
2014
).
27.
Y.
Liu
,
Y.
Yan
,
T.
Xing
, and
Q.
Shi
,
J. Phys. Chem. B
125
,
5959
(
2021
).
28.
V.
May
and
O.
Kühn
,
Charge and Energy Transfer Dynamics in Molecular Systems
(
John Wiley & Sons
,
2023
).
29.
S.
Loth
,
M.
Etzkorn
,
C. P.
Lutz
,
D. M.
Eigler
, and
A. J.
Heinrich
,
Science
329
,
1628
(
2010
).
30.
S.
Baumann
,
W.
Paul
,
T.
Choi
,
C. P.
Lutz
,
A.
Ardavan
, and
A. J.
Heinrich
,
Science
350
,
417
(
2015
).
31.
T.
Choi
,
W.
Paul
,
S.
Rolf-Pissarczyk
,
A. J.
Macdonald
,
F. D.
Natterer
,
K.
Yang
,
P.
Willke
,
C. P.
Lutz
, and
A. J.
Heinrich
,
Nat. Nanotechnol.
12
,
420
(
2017
).
32.
F. D.
Natterer
,
K.
Yang
,
W.
Paul
,
P.
Willke
,
T.
Choi
,
T.
Greber
,
A. J.
Heinrich
, and
C. P.
Lutz
,
Nature
543
,
226
(
2017
).
33.
W.
Paul
,
K.
Yang
,
S.
Baumann
,
N.
Romming
,
T.
Choi
,
C. P.
Lutz
, and
A. J.
Heinrich
,
Nat. Phys.
13
,
403
(
2017
).
34.
K.
Yang
,
Y.
Bae
,
W.
Paul
,
F. D.
Natterer
,
P.
Willke
,
J. L.
Lado
,
A.
Ferrón
,
T.
Choi
,
J.
Fernández-Rossier
,
A. J.
Heinrich
, and
C. P.
Lutz
,
Phys. Rev. Lett.
119
,
227206
(
2017
).
35.
K.
Yang
,
W.
Paul
,
F. D.
Natterer
,
J. L.
Lado
,
Y.
Bae
,
P.
Willke
,
T.
Choi
,
A.
Ferrón
,
J.
Fernández-Rossier
,
A. J.
Heinrich
, and
C. P.
Lutz
,
Phys. Rev. Lett.
122
,
227203
(
2019
).
36.
X.
Zhang
,
C.
Wolf
,
Y.
Wang
,
H.
Aubin
,
T.
Bilgeri
,
P.
Willke
,
A. J.
Heinrich
, and
T.
Choi
,
Nat. Chem.
14
,
59
(
2022
).
37.
Y.
Chen
,
Y.
Bae
, and
A. J.
Heinrich
,
Adv. Mater.
35
,
2107534
(
2022
).
38.
L.
Bi
,
K.
Liang
,
G.
Czap
,
H.
Wang
,
K.
Yang
, and
S.
Li
,
Prog. Surf. Sci.
98
,
100696
(
2023
).
39.
K.
Yang
,
W.
Paul
,
S.-H.
Phark
,
P.
Willke
,
Y.
Bae
,
T.
Choi
,
T.
Esat
,
A.
Ardavan
,
A. J.
Heinrich
, and
C. P.
Lutz
,
Science
366
,
509
(
2019
).
40.
L. M.
Veldman
,
L.
Farinacci
,
R.
Rejali
,
R.
Broekhoven
,
J.
Gobeil
,
D.
Coffey
,
M.
Ternes
, and
A. F.
Otte
,
Science
372
,
964
(
2021
).
41.
Y.
Tanimura
and
R.
Kubo
,
J. Phys. Soc. Jpn.
58
,
101
(
1989
).
42.
Y.
Tanimura
,
Phys. Rev. A
41
,
6676
(
1990
).
43.
Y.-A.
Yan
,
F.
Yang
,
Y.
Liu
, and
J.
Shao
,
Chem. Phys. Lett.
395
,
216
(
2004
).
44.
Y.
Tanimura
,
J. Chem. Phys.
153
,
020901
(
2020
).
45.
J.
Jin
,
X.
Zheng
, and
Y.
Yan
,
J. Chem. Phys.
128
,
234703
(
2008
).
46.
Z.
Li
,
N.
Tong
,
X.
Zheng
,
D.
Hou
,
J.
Wei
,
J.
Hu
, and
Y.
Yan
,
Phys. Rev. Lett.
109
,
266403
(
2012
).
47.
L.
Ye
,
X.
Wang
,
D.
Hou
,
R.-X.
Xu
,
X.
Zheng
, and
Y.
Yan
,
Wiley Interdiscip. Rev.: Comput. Mol. Sci.
6
,
608
(
2016
).
48.
D.
Zhang
,
L.
Ye
,
J.
Cao
,
Y.
Wang
,
R.-X.
Xu
,
X.
Zheng
, and
Y.
Yan
,
WIREs Comput. Mol. Sci.
14
,
e1727
(
2024
).
49.
B. M.
Garraway
,
Phys. Rev. A
55
,
4636
(
1997
).
50.
D.
Tamascelli
,
A.
Smirne
,
S. F.
Huelga
, and
M. B.
Plenio
,
Phys. Rev. Lett.
120
,
030402
(
2018
).
51.
D.
Tamascelli
,
A.
Smirne
,
J.
Lim
,
S. F.
Huelga
, and
M. B.
Plenio
,
Phys. Rev. Lett.
123
,
090402
(
2019
).
52.
N.
Lambert
,
S.
Ahmed
,
M.
Cirio
, and
F.
Nori
,
Nat. Commun.
10
,
3721
(
2019
).
53.
M.
Cirio
,
S.
Luo
,
P.
Liang
,
F.
Nori
, and
N.
Lambert
,
Phys. Rev. Res.
6
,
033083
(
2024
).
54.
N.
Lambert
,
M.
Cirio
,
J.-d.
Lin
,
P.
Menczel
,
P.
Liang
, and
F.
Nori
,
Phys. Rev. Res.
6
,
043229
(
2023
).
55.
N.
Zhou
,
Z.
Huang
,
J.
Zhu
,
V.
Chernyak
, and
Y.
Zhao
,
J. Chem. Phys.
143
,
014113
(
2015
).
56.
L.
Wang
,
L.
Chen
,
N.
Zhou
, and
Y.
Zhao
,
J. Chem. Phys.
144
,
024101
(
2016
).
57.
L.
Wang
,
Y.
Fujihashi
,
L.
Chen
, and
Y.
Zhao
,
J. Chem. Phys.
146
,
124127
(
2017
).
58.
Y.
Zhao
,
K.
Sun
,
L.
Chen
, and
M.
Gelin
,
Wiley Interdiscip. Rev.: Comput. Mol. Sci.
12
,
e1589
(
2022
).
59.
Y.
Zhao
,
J. Chem. Phys.
158
,
080901
(
2023
).
60.
K.
Shen
,
K.
Sun
,
M. F.
Gelin
, and
Y.
Zhao
,
J. Phys. Chem. Lett.
15
,
447
(
2024
).
61.
J. T.
Stockburger
and
H.
Grabert
,
Phys. Rev. Lett.
88
,
170407
(
2002
).
62.
J. T.
Stockburger
and
C. H.
Mak
,
Phys. Rev. Lett.
80
,
2657
(
1998
).
63.
J.
Shao
,
J. Chem. Phys.
120
,
5053
(
2004
).
64.
Y.
Zhou
,
Y.
Yan
, and
J.
Shao
,
Europhys. Lett.
72
,
334
(
2005
).
65.
J. M.
Moix
and
J.
Cao
,
J. Chem. Phys.
139
,
134106
(
2013
).
66.
L.
Zhu
,
H.
Liu
, and
Q.
Shi
,
New J. Phys.
15
,
095020
(
2013
).
67.
Y.-A.
Yan
and
J.
Shao
,
Phys. Rev. A
97
,
042126
(
2018
).
68.
L.
Han
,
V.
Chernyak
,
Y.-A.
Yan
,
X.
Zheng
, and
Y.
Yan
,
Phys. Rev. Lett.
123
,
050601
(
2019
).
69.
L.
Han
,
A.
Ullah
,
Y.-A.
Yan
,
X.
Zheng
,
Y.
Yan
, and
V.
Chernyak
,
J. Chem. Phys.
152
,
204105
(
2020
).
70.
A.
Ullah
,
L.
Han
,
Y.-A.
Yan
,
X.
Zheng
,
Y.
Yan
, and
V.
Chernyak
,
J. Chem. Phys.
152
,
204106
(
2020
).
71.
W. T.
Strunz
,
L.
Diósi
, and
N.
Gisin
,
Phys. Rev. Lett.
82
,
1801
(
1999
).
72.
D.
Suess
,
A.
Eisfeld
, and
W.
Strunz
,
Phys. Rev. Lett.
113
,
150403
(
2014
).
73.
G.
Cohen
,
E.
Gull
,
D. R.
Reichman
, and
A. J.
Millis
,
Phys. Rev. Lett.
115
,
266802
(
2015
).
74.
W.
Kohn
,
Rev. Mod. Phys.
71
,
1253
(
1999
).
75.
A.
Klümper
,
A.
Schadschneider
, and
J.
Zittartz
,
Europhys. Lett.
24
,
293
(
1993
).
76.
S. R.
White
and
A. E.
Feiguin
,
Phys. Rev. Lett.
93
,
076401
(
2004
).
77.
C.
Büsser
and
F.
Heidrich-Meisner
,
Phys. Rev. Lett.
111
,
246807
(
2013
).
78.
H.-D.
Meyer
,
U.
Manthe
, and
L. S.
Cederbaum
,
Chem. Phys. Lett.
165
,
73
(
1990
).
79.
H.
Wang
,
M.
Thoss
, and
W. H.
Miller
,
J. Chem. Phys.
115
,
2979
(
2001
).
80.
A.
Strathearn
,
P.
Kirton
,
D.
Kilda
,
J.
Keeling
, and
B. W.
Lovett
,
Nat. Commun.
9
,
3322
(
2018
).
81.
M. R.
Jørgensen
and
F. A.
Pollock
,
Phys. Rev. Lett.
123
,
240602
(
2019
).
82.
F.
Verstraete
,
V.
Murg
, and
J. I.
Cirac
,
Adv. Phys.
57
,
143
(
2008
).
83.
U.
Schollwöck
,
Ann. Phys.
326
,
96
(
2011
).
84.
R.
Orús
,
Ann. Phys.
349
,
117
(
2014
).
85.
R.
Borrelli
and
M. F.
Gelin
,
J. Chem. Phys.
145
,
224101
(
2016
).
86.
S. M.
Greene
and
V. S.
Batista
,
J. Chem. Theory Comput.
13
,
4034
(
2017
).
87.
Q.
Shi
,
Y.
Xu
,
Y.
Yan
, and
M.
Xu
,
J. Chem. Phys.
148
,
174102
(
2018
).
88.
Y.
Yan
,
T.
Xing
, and
Q.
Shi
,
J. Chem. Phys.
153
,
204109
(
2020
).
89.
R.
Borrelli
and
M. F.
Gelin
,
Wiley Interdiscip. Rev.: Comput. Mol. Sci.
11
,
e1539
(
2021
).
90.
Y.
Ke
,
R.
Borrelli
, and
M.
Thoss
,
J. Chem. Phys.
156
,
194102
(
2022
).
91.
Y.
Ke
,
J. Chem. Phys.
158
,
211102
(
2023
).
92.
S.
Zhang
,
Y.
Chen
, and
Q.
Shi
,
J. Chem. Phys.
160
,
054101
(
2024
).
93.
W.
Li
,
J.
Ren
,
H.
Yang
,
H.
Wang
, and
Z.
Shuai
,
J. Chem. Phys.
161
,
054116
(
2024
).
94.
M.
Cygorek
and
E. M.
Gauger
,
J. Chem. Phys.
161
,
074111
(
2024
).
95.
S.
Flannigan
,
F.
Damanet
, and
A. J.
Daley
,
Phys. Rev. Lett.
128
,
063601
(
2022
).
96.
G. E.
Fux
,
E. P.
Butler
,
P. R.
Eastham
,
B. W.
Lovett
, and
J.
Keeling
,
Phys. Rev. Lett.
126
,
200401
(
2021
).
97.
M.
Richter
and
S.
Hughes
,
Phys. Rev. Lett.
128
,
167403
(
2022
).
98.
D.
Gribben
,
D. M.
Rouse
,
J.
Iles-Smith
,
A.
Strathearn
,
H.
Maguire
,
P.
Kirton
,
A.
Nazir
,
E. M.
Gauger
, and
B. W.
Lovett
,
PRX Quantum
3
,
010321
(
2022
).
99.
M.
Popovic
,
M. T.
Mitchison
,
A.
Strathearn
,
B. W.
Lovett
,
J.
Goold
, and
P. R.
Eastham
,
PRX Quantum
2
,
020338
(
2021
).
100.
G.
Carleo
and
M.
Troyer
,
Science
355
,
602
(
2017
).
101.
G.
Torlai
and
R. G.
Melko
,
Phys. Rev. Lett.
120
,
240503
(
2018
).
102.
I. A.
Luchnikov
,
S. V.
Vintskevich
,
H.
Ouerdane
, and
S. N.
Filippov
,
Phys. Rev. Lett.
122
,
160401
(
2019
).
103.
A.
Nagy
and
V.
Savona
,
Phys. Rev. Lett.
122
,
250501
(
2019
).
104.
M. J.
Hartmann
and
G.
Carleo
,
Phys. Rev. Lett.
122
,
250502
(
2019
).
105.
F.
Vicentini
,
A.
Biella
,
N.
Regnault
, and
C.
Ciuti
,
Phys. Rev. Lett.
122
,
250503
(
2019
).
106.
J.
Jumper
,
R.
Evans
,
A.
Pritzel
,
T.
Green
,
M.
Figurnov
,
O.
Ronneberger
,
K.
Tunyasuvunakool
,
R.
Bates
,
A.
Žídek
,
A.
Potapenko
et al,
Nature
596
,
583
(
2021
).
107.
H.-Y.
Huang
,
R.
Kueng
,
G.
Torlai
,
V. V.
Albert
, and
J.
Preskill
,
Science
377
,
eabk3333
(
2022
).
108.
G.
Chen
,
F.
Song
, and
J. L.
Lado
,
Phys. Rev. Lett.
130
,
100401
(
2023
).
109.
H.
Ma
,
H.
Shang
, and
J.
Yang
,
npj Comput. Mater.
10
,
220
(
2024
).
110.
L. M.
Vandersypen
and
I. L.
Chuang
,
Rev. Mod. Phys.
76
,
1037
(
2005
).
111.
T. D.
Ladd
,
F.
Jelezko
,
R.
Laflamme
,
Y.
Nakamura
,
C.
Monroe
, and
J. L.
O’Brien
,
Nature
464
,
45
(
2010
).
112.
A. J.
Daley
,
I.
Bloch
,
C.
Kokail
,
S.
Flannigan
,
N.
Pearson
,
M.
Troyer
, and
P.
Zoller
,
Nature
607
,
667
(
2022
).
113.
J.
Biamonte
,
P.
Wittek
,
N.
Pancotti
,
P.
Rebentrost
,
N.
Wiebe
, and
S.
Lloyd
,
Nature
549
,
195
(
2017
).
114.
C.
Gross
and
I.
Bloch
,
Science
357
,
995
(
2017
).
115.
S.
McArdle
,
S.
Endo
,
A.
Aspuru-Guzik
,
S. C.
Benjamin
, and
X.
Yuan
,
Rev. Mod. Phys.
92
,
015003
(
2020
).
116.
M.
Kliesch
,
T.
Barthel
,
C.
Gogolin
,
M.
Kastoryano
, and
J.
Eisert
,
Phys. Rev. Lett.
107
,
120501
(
2011
).
117.
J.
Liu
,
Y.
Fan
,
Z.
Li
, and
J.
Yang
,
Chem. Soc. Rev.
51
,
3263
(
2022
).
118.
L. H.
Delgado-Granados
,
T. J.
Krogmeier
,
L. M.
Sager-Smith
,
I.
Avdic
,
Z.
Hu
,
M.
Sajjan
,
M.
Abbasi
,
S. E.
Smart
,
P.
Narang
,
S.
Kais
, et al, arXiv:2406.05219 (
2024
).
119.
F. A.
Schröder
,
D. H.
Turban
,
A. J.
Musser
,
N. D.
Hine
, and
A. W.
Chin
,
Nat. Commun.
10
,
1062
(
2019
).
120.
M.
Schuld
,
I.
Sinayskiy
, and
F.
Petruccione
,
Physics
12
,
74
(
2019
).
121.
J.
Chen
,
S.
Cheng
,
H.
Xie
,
L.
Wang
, and
T.
Xiang
,
Phys. Rev. B
97
,
085104
(
2018
).
122.
R. P.
Feynman
,
Int. J. Theor. Phys.
21
,
467
(
1982
).
123.
124.
M.
Greiner
,
O.
Mandel
,
T.
Esslinger
,
T. W.
Hänsch
, and
I.
Bloch
,
Nature
415
,
39
(
2002
).
125.
M.
Hartmann
,
F.
Brandão
, and
M.
Plenio
,
Nat. Phys.
2
,
849
(
2006
).
126.
R.
Gerritsma
,
G.
Kirchmair
,
F.
Zähringer
,
E.
Solano
,
R.
Blatt
, and
C.
Roos
,
Nature
463
,
68
(
2010
).
127.
S. P.
Jordan
,
K. S.
Lee
, and
J.
Preskill
,
Science
336
,
1130
(
2012
).
128.
H.
Labuhn
,
D.
Barredo
,
S.
Ravets
,
S.
De Léséleuc
,
T.
Macrì
,
T.
Lahaye
, and
A.
Browaeys
,
Nature
534
,
667
(
2016
).
129.
H.
Bernien
,
S.
Schwartz
,
A.
Keesling
,
H.
Levine
,
A.
Omran
,
H.
Pichler
,
S.
Choi
,
A. S.
Zibrov
,
M.
Endres
,
M.
Greiner
et al,
Nature
551
,
579
(
2017
).
130.
P.
Roushan
,
C.
Neill
,
J.
Tangpanitanon
,
V. M.
Bastidas
,
A.
Megrant
,
R.
Barends
,
Y.
Chen
,
Z.
Chen
,
B.
Chiaro
,
A.
Dunsworth
et al,
Science
358
,
1175
(
2017
).
131.
C.
Song
,
J.
Cui
,
H.
Wang
,
J.
Hao
,
H.
Feng
, and
Y.
Li
,
Sci. Adv.
5
,
eaaw5686
(
2019
).
132.
A.
Kandala
,
K.
Temme
,
A. D.
Córcoles
,
A.
Mezzacapo
,
J. M.
Chow
, and
J. M.
Gambetta
,
Nature
567
,
491
(
2019
).
133.
S.
Cao
,
B.
Wu
,
F.
Chen
,
M.
Gong
,
Y.
Wu
,
Y.
Ye
,
C.
Zha
,
H.
Qian
,
C.
Ying
,
S.
Guo
et al,
Nature
619
,
738
(
2023
).
134.
Y.
Cao
,
J.
Romero
,
J. P.
Olson
,
M.
Degroote
,
P. D.
Johnson
,
M.
Kieferova
,
I. D.
Kivlichan
,
T.
Menke
,
B.
Peropadre
,
N. P. D.
Sawaya
,
S.
Sim
,
L.
Veis
, and
A.
Aspuru-Guzik
,
Chem. Rev.
119
,
10856
(
2019
).
135.
A.
Miessen
,
P. J.
Ollitrault
,
F.
Tacchino
, and
I.
Tavernelli
,
Nat. Comput. Sci.
3
,
25
(
2023
).
136.
N.
Yoshioka
and
R.
Hamazaki
,
Phys. Rev. B
99
,
214306
(
2019
).
137.
A.
Lemmer
,
C.
Cormick
,
D.
Tamascelli
,
T.
Schaetz
,
S. F.
Huelga
, and
M. B.
Plenio
,
New J. Phys.
20
,
073002
(
2018
).
138.
M.
Tolunay
,
I.
Liepuoniute
,
M.
Vyushkova
, and
B. A.
Jones
,
Phys. Chem. Chem. Phys.
25
,
15115
(
2023
).
139.
C. K.
Lee
,
P.
Patil
,
S.
Zhang
, and
C. Y.
Hsieh
,
Phys. Rev. Res.
3
,
023095
(
2021
).
140.
J. D.
Guimarães
,
J.
Lim
,
M. I.
Vasilevskiy
,
S. F.
Huelga
, and
M. B.
Plenio
,
PRX Quantum
4
,
040329
(
2023
).
141.
S.
Sun
,
L.-C.
Shih
, and
Y.-C.
Cheng
,
Phys. Scr.
99
,
035101
(
2024
).
142.
T. M.
Watad
and
N. H.
Lindner
,
Quantum Sci. Technol.
9
,
025015
(
2024
).
143.
J.
Hu
,
M.
Luo
,
F.
Jiang
,
R.-X.
Xu
, and
Y.
Yan
,
J. Chem. Phys.
134
,
244106
(
2011
).
144.
L.
Ye
,
H.-D.
Zhang
,
Y.
Wang
,
X.
Zheng
, and
Y.
Yan
,
J. Chem. Phys.
147
,
074111
(
2017
).
145.
L.
Cui
,
H.-D.
Zhang
,
X.
Zheng
,
R.-X.
Xu
, and
Y.
Yan
,
J. Chem. Phys.
151
,
024110
(
2019
).
146.
H.-D.
Zhang
,
L.
Cui
,
H.
Gong
,
R.-X.
Xu
,
X.
Zheng
, and
Y.
Yan
,
J. Chem. Phys.
152
,
064107
(
2020
).
147.
D.
Zhang
,
L.
Zuo
,
L.
Ye
,
Z.-H.
Chen
,
Y.
Wang
,
R.-X.
Xu
,
X.
Zheng
, and
Y.
Yan
,
J. Chem. Phys.
158
,
014106
(
2023
).
148.
Z.-H.
Chen
,
Y.
Wang
,
X.
Zheng
,
R.-X.
Xu
, and
Y.
Yan
,
J. Chem. Phys.
156
,
221102
(
2022
).
149.
M.
Xu
,
Y.
Yan
,
Q.
Shi
,
J.
Ankerhold
, and
J.
Stockburger
,
Phys. Rev. Lett.
129
,
230601
(
2022
).
150.
H.
Takahashi
,
S.
Rudge
,
C.
Kaspar
,
M.
Thoss
, and
R.
Borrelli
,
J. Chem. Phys.
160
,
204105
(
2024
).
151.
L.
Han
,
H.-D.
Zhang
,
X.
Zheng
, and
Y.
Yan
,
J. Chem. Phys.
148
,
234108
(
2018
).
152.
R. P.
Feynman
and
F. L.
Vernon
,
Ann. Phys.
24
,
118
(
1963
).
153.
Y.
Tanimura
and
P. G.
Wolynes
,
Phys. Rev. A
43
,
4131
(
1991
).
154.
A.
Ishizaki
and
Y.
Tanimura
,
J. Phys. Soc. Jpn.
74
,
3131
(
2005
).
155.
R.-X.
Xu
,
P.
Cui
,
X.-Q.
Li
,
Y.
Mo
, and
Y.
Yan
,
J. Chem. Phys.
122
,
041103
(
2005
).
156.
R.-X.
Xu
and
Y.
Yan
,
Phys. Rev. E
75
,
031107
(
2007
).
157.
Y.
Yan
,
J.
Jin
,
R.-X.
Xu
, and
X.
Zheng
,
Front. Phys.
11
,
110306
(
2016
).
158.
Y.
Wang
and
Y.
Yan
,
J. Chem. Phys.
157
,
170901
(
2022
).
159.
R.
Hartmann
and
W. T.
Strunz
,
J. Chem. Theory Comput.
13
,
5834
(
2017
).
160.
P.
Mitrić
,
V.
Janković
,
N.
Vukmirović
, and
D.
Tanasković
,
Phys. Rev. Lett.
129
,
096401
(
2022
).
161.
V.
Janković
,
P.
Mitrić
,
D.
Tanasković
, and
N.
Vukmirović
,
Phys. Rev. B
109
,
214312
(
2024
).
162.
Y. J.
Yan
,
J. Chem. Phys.
140
,
054105
(
2014
).
163.
H.
Gong
,
Y.
Wang
,
H.-D.
Zhang
,
Q.
Qiao
,
R.-X.
Xu
,
X.
Zheng
, and
Y.
Yan
,
J. Chem. Phys.
153
,
154111
(
2020
).
164.
H.
Gong
,
Y.
Wang
,
X.
Zheng
,
R.-X.
Xu
, and
Y.
Yan
,
J. Chem. Phys.
157
,
054109
(
2022
).
165.
X.
Li
,
Y.
Su
,
Z.-H.
Chen
,
Y.
Wang
,
R.-X.
Xu
,
X.
Zheng
, and
Y.
Yan
,
J. Chem. Phys.
158
,
214110
(
2023
).
166.
X.
Li
,
S.-X.
Lyu
,
Y.
Wang
,
R.-X.
Xu
,
X.
Zheng
, and
Y.
Yan
,
Phys. Rev. A
110
,
032620
(
2024
).
167.
M.
Cirio
,
N.
Lambert
,
P.
Liang
,
P.-C.
Kuo
,
Y.-N.
Chen
,
P.
Menczel
,
K.
Funo
, and
F.
Nori
,
Phys. Rev. Res.
5
,
033011
(
2023
).
168.
A.
Bose
,
J. Chem. Phys.
158
,
204113
(
2023
).
169.
N.
Lambert
,
T.
Raheja
,
S.
Cross
,
P.
Menczel
,
S.
Ahmed
,
A.
Pitchford
,
D.
Burgarth
, and
F.
Nori
,
Phys. Rev. Res.
5
,
013181
(
2023
).
170.
T.
Lacroix
,
B.
Le Dé
,
A.
Riva
,
A. J.
Dunnett
, and
A. W.
Chin
,
J. Chem. Phys.
161
,
084116
(
2024
).
171.
W.
Guan
,
P.
Bao
,
J.
Peng
,
Z.
Lan
, and
Q.
Shi
,
J. Chem. Phys.
161
,
122501
(
2024
).
172.
G. E.
Fux
,
P.
Fowler-Wright
,
J.
Beckles
,
E. P.
Butler
,
P. R.
Eastham
,
D.
Gribben
,
J.
Keeling
,
D.
Kilda
,
P.
Kirton
,
E. D. C.
Lawrence
,
B. W.
Lovett
,
E.
O’Neill
,
A.
Strathearn
, and
R.
de Wit
,
J. Chem. Phys.
161
,
124108
(
2024
).
173.
Y.
Xu
,
C.
Liu
, and
H.
Ma
,
J. Chem. Phys.
161
,
052501
(
2024
).
174.
J. R.
Johansson
,
P. D.
Nation
, and
F.
Nori
,
Comput. Phys. Commun.
183
,
1760
(
2012
).
175.
J. R.
Johansson
,
P. D.
Nation
, and
F.
Nori
,
Comput. Phys. Commun.
184
,
1234
(
2013
).
176.
N.
Shammah
,
S.
Ahmed
,
N.
Lambert
,
S.
De Liberato
, and
F.
Nori
,
Phys. Rev. A
98
,
063815
(
2018
).
177.
B.
Li
,
S.
Ahmed
,
S.
Saraogi
,
N.
Lambert
,
F.
Nori
,
A.
Pitchford
, and
N.
Shammah
,
Quantum
6
,
630
(
2022
).
178.
D.
Zhang
,
X.
Ding
,
H.-D.
Zhang
,
X.
Zheng
, and
Y.
Yan
,
Chin. J. Chem. Phys.
34
,
905
(
2021
).
179.
X.
Zheng
,
Y.
Yan
, and
M.
Di Ventra
,
Phys. Rev. Lett.
111
,
086601
(
2013
).
180.
Y.
Wang
,
X.
Zheng
,
B.
Li
, and
J.
Yang
,
J. Chem. Phys.
141
,
084713
(
2014
).
181.
X.
Wang
,
D.
Hou
,
X.
Zheng
, and
Y.
Yan
,
J. Chem. Phys.
144
,
034101
(
2016
).
182.
Y.
Wang
,
X.
Zheng
, and
J.
Yang
,
Phys. Rev. B
93
,
125114
(
2016
).
183.
X.
Li
,
L.
Zhu
,
B.
Li
,
J.
Li
,
P.
Gao
,
L.
Yang
,
A.
Zhao
,
Y.
Luo
,
J.
Hou
,
X.
Zheng
et al,
Nat. Commun.
11
,
2566
(
2020
).
184.
X.
Li
,
H.
Gong
,
Q.
Zhuang
,
B.
Wang
,
X.
Zheng
, and
J.
Yang
,
J. Phys. Chem. C
125
,
21488
(
2021
).
185.
X.
Ding
,
D.
Zhang
,
L.
Ye
,
X.
Zheng
, and
Y.
Yan
,
J. Chem. Phys.
157
,
224107
(
2022
).
186.
Q.
Zhuang
,
X.
Wang
,
L.
Ye
,
Y.
Yan
, and
X.
Zheng
,
J. Phys. Chem. Lett.
13
,
2094
(
2022
).
187.
Q.
Zhuang
,
L.
Zuo
,
X.
Li
,
L.
Ye
,
X.
Zheng
, and
Y.
Yan
,
Chem. Phys. Lett.
830
,
140813
(
2023
).
188.
X.
Wang
,
L.
Yang
,
L.
Ye
,
X.
Zheng
, and
Y.
Yan
,
J. Phys. Chem. Lett.
9
,
2418
(
2018
).
189.
L.
Zuo
,
Q.
Zhuang
,
L.
Ye
,
Y.
Yan
, and
X.
Zheng
,
J. Phys. Chem. Lett.
13
,
11262
(
2022
).
190.
J.
Cerrillo
and
J.
Cao
,
Phys. Rev. Lett.
112
,
110401
(
2014
).
191.
L.
Banchi
,
E.
Grant
,
A.
Rocchetto
, and
S.
Severini
,
New J. Phys.
20
,
123030
(
2018
).
192.
S.
Bandyopadhyay
,
Z.
Huang
,
K.
Sun
, and
Y.
Zhao
,
Chem. Phys.
515
,
272
(
2018
).
193.
K.
Lin
,
J.
Peng
,
F. L.
Gu
, and
Z.
Lan
,
J. Phys. Chem. Lett.
12
,
10225
(
2021
).
194.
J.
Peng
,
Y.
Xie
,
D.
Hu
, and
Z.
Lan
,
J. Chem. Phys.
154
,
094122
(
2021
).
195.
K.
Lin
,
J.
Peng
,
C.
Xu
,
F. L.
Gu
, and
Z.
Lan
,
J. Chem. Theory Comput.
18
,
5837
(
2022
).
196.
K.
Lin
,
J.
Peng
,
C.
Xu
,
F. L.
Gu
, and
Z.
Lan
,
J. Phys. Chem. Lett.
13
,
11678
(
2022
).
197.
P. O.
Dral
,
Quantum Chemistry in the Age of Machine Learning
(
Elsevier
,
2022
).
198.
D.
Wu
,
Z.
Hu
,
J.
Li
, and
X.
Sun
,
J. Chem. Phys.
155
,
224104
(
2021
).
199.
L. E.
Herrera Rodríguez
and
A. A.
Kananenka
,
J. Phys. Chem. Lett.
12
,
2476
(
2021
).
200.
J.
Nelson
,
L.
Coopmans
,
G.
Kells
, and
S.
Sanvito
,
Phys. Rev. B
106
,
045402
(
2022
).
201.
I. A.
Luchnikov
,
E. O.
Kiktenko
,
M. A.
Gavreev
,
H.
Ouerdane
,
S. N.
Filippov
, and
A. K.
Fedorov
,
Phys. Rev. Res.
4
,
043002
(
2022
).
202.
G.
Cemin
,
F.
Carnazza
,
S.
Andergassen
,
G.
Martius
,
F.
Carollo
, and
I.
Lesanovsky
,
Phys. Rev. Appl.
21
,
L041001
(
2024
).
203.
A. J. I.
Dominic
,
S.
Cao
,
A.
Montoya-Castillo
, and
X.
Huang
,
J. Am. Chem. Soc.
145
,
9916
(
2023
).
204.
J.
Hermann
,
J.
Spencer
,
K.
Choo
,
A.
Mezzacapo
,
W. M. C.
Foulkes
,
D.
Pfau
,
G.
Carleo
, and
F.
Noé
,
Nat. Rev. Chem
7
,
692
(
2023
).
205.
J.
Zhang
,
C. L.
Benavides-Riveros
, and
L.
Chen
,
J. Phys. Chem. Lett.
15
,
3603
(
2024
).
206.
J.
Zhang
and
L.
Chen
,
Phys. Chem. Chem. Phys.
27
,
182
(
2025
).
207.
M.
Reh
,
M.
Schmitt
, and
M.
Gärttner
,
Phys. Rev. Lett.
127
,
230501
(
2021
).
208.
D.
Luo
,
Z.
Chen
,
J.
Carrasquilla
, and
B. K.
Clark
,
Phys. Rev. Lett.
128
,
090501
(
2022
).
209.
J.
Nys
,
Z.
Denis
, and
G.
Carleo
,
Phys. Rev. B
109
,
235120
(
2024
).
210.
L.
Cao
,
L.
Ge
,
D.
Zhang
,
X.
Li
,
Y.
Wang
,
R.-X.
Xu
,
Y.
Yan
, and
X.
Zheng
, arXiv:2404.11093 (
2024
).
211.
A. M.
Childs
and
T.
Li
,
Quantum Inf. Comput.
17
,
901
(
2017
).
212.
S.
Patsch
,
S.
Maniscalco
, and
C. P.
Koch
,
Phys. Rev. Res.
2
,
023133
(
2020
).
213.
L.
Del Re
,
B.
Rost
,
A.
Kemper
, and
J.
Freericks
,
Phys. Rev. B
102
,
125112
(
2020
).
214.
A. W.
Schlimgen
,
K.
Head-Marsden
,
L. M.
Sager
,
P.
Narang
, and
D. A.
Mazziotti
,
Phys. Rev. Lett.
127
,
270503
(
2021
).
215.
G.
García-Pérez
,
M. A.
Rossi
, and
S.
Maniscalco
,
npj Quantum Inf.
6
,
1
(
2020
).
216.
A. W.
Schlimgen
,
K.
Head-Marsden
,
L. M.
Sager
,
P.
Narang
, and
D. A.
Mazziotti
,
Phys. Rev. Res.
4
,
023216
(
2022
).
217.
X.
Li
and
C.
Wang
,
Commun. Math. Phys.
401
,
147
(
2023
).
218.
Y.
Wang
,
E.
Mulvihill
,
Z.
Hu
,
N.
Lyu
,
S.
Shivpuje
,
Y.
Liu
,
M. B.
Soley
,
E.
Geva
,
V. S.
Batista
, and
S.
Kais
,
J. Chem. Theory Comput.
19
,
4851
(
2023
).
219.
S.
McArdle
,
T.
Jones
,
S.
Endo
,
Y.
Li
,
S. C.
Benjamin
, and
X.
Yuan
,
npj Quantum Inf.
5
,
75
(
2019
).
220.
H.
Kamakari
,
S.-N.
Sun
,
M.
Motta
, and
A. J.
Minnich
,
PRX Quantum
3
,
010320
(
2022
).
221.
M.
Motta
,
C.
Sun
,
A. T.
Tan
,
M. J.
O’Rourke
,
E.
Ye
,
A. J.
Minnich
,
F. G.
Brandao
, and
G. K.-L.
Chan
,
Nat. Phys.
16
,
205
(
2020
).
222.
X.
Yuan
,
S.
Endo
,
Q.
Zhao
,
Y.
Li
, and
S. C.
Benjamin
,
Quantum
3
,
191
(
2019
).
223.
S.-J.
Wei
,
D.
Ruan
, and
G.-L.
Long
,
Sci. Rep.
6
,
30727
(
2016
).
224.
C.
Zheng
,
Sci. Rep.
11
,
3960
(
2021
).
225.
Z.
Hu
,
R.
Xia
, and
S.
Kais
,
Sci. Rep.
10
,
3301
(
2020
).
226.
Z.
Hu
,
K.
Head-Marsden
,
D. A.
Mazziotti
,
P.
Narang
, and
S.
Kais
,
Quantum
6
,
726
(
2022
).
227.
K.
Head-Marsden
,
S.
Krastanov
,
D. A.
Mazziotti
, and
P.
Narang
,
Phys. Rev. Res.
3
,
013182
(
2021
).
228.
N.
Lyu
,
P.
Khazaei
,
E.
Geva
, and
V. S.
Batista
,
J. Phys. Chem. Lett.
15
,
9535
(
2024
).
229.
Qiskit: An open-source framework for quantum computing (
2021
).
230.
C.
Long
,
L.
Cao
,
L.
Ge
,
Q.-X.
Li
,
Y.
Yan
,
R.-X.
Xu
,
Y.
Wang
, and
X.
Zheng
,
J. Chem. Phys.
161
,
084105
(
2024
).
231.
S.
Bravyi
,
O.
Dial
,
J. M.
Gambetta
,
D.
Gil
, and
Z.
Nazario
,
J. Appl. Phys.
132
,
160902
(
2022
).
232.
A. D.
Córcoles
,
A.
Kandala
,
A.
Javadi-Abhari
,
D. T.
McClure
,
A. W.
Cross
, and
K.
Temme
,
Proc. IEEE
108
,
1338
(
2020
).
233.
N.
Moll
,
P.
Barkoutsos
,
L. S.
Bishop
,
J. M.
Chow
,
A.
Cross
,
D. J.
Egger
,
S.
Filipp
,
A.
Fuhrer
,
J. M.
Gambetta
,
M.
Ganzhorn
,
A.
Kandala
,
A.
Mezzacapo
,
P.
Müller
,
W.
Riess
,
G.
Salis
,
J.
Smolin
,
I.
Tavernelli
, and
K.
Temme
,
Quantum Sci. Technol.
3
,
030503
(
2018
).