Jong-Kwon Haa and
Ryan J. MacDonell
*ab
aDepartment of Chemistry, Dalhousie University, 6243 Alumni Cres, Halifax, NS B3H 4R2, Canada. E-mail: rymac@dal.ca
bDepartment of Physics and Atmospheric Science, Dalhousie University, 1453 Lord Dalhousie Dr, Halifax, NS B3H 4R2, Canada
First published on 16th September 2025
Quantum computing has the potential to reduce the computational cost required for quantum dynamics simulations. However, existing quantum algorithms for coupled electron-nuclear dynamics simulation either require fault-tolerant devices, or involve the Born–Oppenheimer (BO) approximation and pre-calculation of electronic states on classical computers. We present the first quantum simulation approach for molecular vibronic dynamics in a pre-BO framework with an analog mapping of nuclear degrees of freedom, i.e. without the separation of electrons and nuclei, by mapping the molecular Hamiltonian to a device with coupled qubits and bosonic modes. We perform a proof-of-principle emulation of our ansatz using a single-mode model system which represents vibronic dynamics of chemical systems, such as nonadiabatic charge transfer involving polarization of the medium, and propose an implementation of our approach on a trapped-ion device. We show that our approach has exponential savings in resource and computational costs compared to the equivalent classical algorithms. Furthermore, our approach has a much smaller resource and implementation scaling than the existing pre-BO quantum algorithms for chemical dynamics. The low cost of our approach will enable an exact treatment of electron-nuclear dynamics on near-term quantum devices.
The Born–Oppenheimer (BO) approximation neglects the coupling between electronic and nuclear degrees of freedom, known as nonadiabatic coupling (NAC).14 While it is useful for properties of the ground electronic state, it fails to describe light-induced chemical reactions where the molecule can reach strong NAC regions after photo-excitation of the molecular electronic state.14 The BO approximation can be extended to the group BO approximation (GBOA), which takes into account NACs between a group of BO states considered to be relevant to the dynamics, while NACs to states outside of the group are neglected.14 The molecular dynamics are then described as nuclear dynamics on multiple BO potential energy surfaces coupled via NACs. The majority of existing simulation methods make use of the GBOA since it extends the scope of molecular dynamics simulation while using conventional electronic structure methods.15–17 Ideally, the couplings neglected by the GBOA are minimal, but choosing the number of BO electronic states requires some intuition. Furthermore, since existing simulation methods with the GBOA require an accurate calculation of BO states and their gradients (including NACs), the computational cost for accurate simulations can become intractable.
Pre-Born–Oppenheimer (pre-BO) methods, on the other hand, naturally include nonadiabatic effects since they treat nuclei and electrons without separation.18–24 There have been many studies on pre-BO theory with both first quantization18–21 and second quantization22–24 of the molecular Hamiltonian. One approach using the pre-BO framework is the nuclear-electronic orbital (NEO) method, which expands certain nuclei in a basis of nuclear orbitals and those the nuclei as quantum particles.22 Recently, an approach was developed based on the multi-configurational time-dependent Hartree method using a second quantization representation of electronic degrees of freedom, which enables the description of coupled electron-nuclear dynamics without electronic potential energy surfaces.24 Nevertheless, the development of methods in the pre-BO framework is still far behind the methods in the BO framework since the computational cost of a pre-BO treatment of molecular dynamics is much greater.
Quantum computing can significantly reduce the computational cost of the simulation of quantum mechanical systems by exploiting the intrinsic quantum nature of the computational device.25–28 Most near-term quantum computing research in the field of quantum chemistry has focused on obtaining electronic properties at fixed nuclear configurations based on the variational quantum eigensolver (VQE) method29–33 using the BO approximation, with proposed extensions to pre-BO eigenvalues using NEO.34 For chemical dynamics, although there are several proposed quantum algorithms without the BO approximation for fault-tolerant devices,35–37 most methods still adopt the GBOA for near-term applications with variational quantum algorithms38–40 or analog mappings.41–44 The intractability of quantum dynamics simulations with a pre-BO wavefunction on classical computers suggests that a pre-BO quantum simulation could show an earlier quantum advantage than methods using the BO framework.
In this work, we propose the first quantum simulation method for molecular vibronic dynamics in the pre-BO framework with an analog mapping of nuclear degrees of freedom. Our approach maps the second quantized pre-BO representation of the molecular vibronic Hamiltonian onto an analog quantum device, and thus treats the electron-nuclear interactions exactly. Specifically, we map the nuclear vibrational motions to bosonic modes of a device and use a fermion-qubit mapping for electronic degrees of freedom. We show that our method can efficiently simulate the exact molecular vibronic dynamics within a given single-particle basis set, and therefore it is suitable for accurate near-term quantum simulations of coupled electron-nuclear dynamics. As an example, we show how a single-mode vibronic model system of nonadiabatic charge transfer45,46 could be simulated on currently available trapped ion quantum computers with existing experimental techniques and noise. Finally, we show how our approach has the potential to out-perform all previous approaches in terms of scaling of hardware resources, implementation cost, and classical pre-calculation.
The full molecular vibronic Hamiltonian is given by
![]() | (1) |
To reduce the electronic basis size, we adopt a basis of orthonormal spin orbitals that depend parametrically on the nuclear positions, ϕp(x; Q) = φp(r; Q)σp(s), where x is a vector of electronic spatial coordinates r and a spin coordinate s, i.e. x = {r, s}, and φp and σp are spatial and spin functions of the spin orbital, respectively.
We express the molecular vibronic wavefunction using Slater determinants of the position-dependent spin orbitals for the electronic degrees of freedom, and Hartree products of harmonic oscillator eigenstates of each normal mode for the vibrational degrees of freedom. In the second quantization representation, the Slater determinants and the Hartree product can be expressed as occupation number vectors (ONVs). An electronic ONV for a Slater determinant is written as |n〉e = |n1⋯nNo〉e, where nj is the occupation of the j-th spin orbital and No is the number of spin orbitals, while a vibrational ONV for a Hartree product is given as |v〉n = |v1⋯vNmode〉n, where vν is the occupation (Fock state) of mode ν. Therefore, our wavefunction ansatz becomes
![]() | (2) |
Sasmal and Vendrell showed that the ansatz above transforms the molecular vibronic Hamiltonian to the second quantization representation:24
![]() | (3) |
For fixed Taylor expansion orders, ων and Q0 only affect the accuracy of our approach by the number of harmonic eigenstates required to describe the wavefunction. In other words, a quantum simulator with lower oscillator noise can tolerate lower-accuracy values of ων and Q0, including approximations derived from the electronic integral Taylor expansions. For practical cases in the near term, we expect that density functional theory will provide an ideal balance between accuracy and classical computational cost when finding Q0 and ων, and Kohn–Sham orbitals provide a compact spin orbital basis.48 We can reduce the number of terms in eqn (3) using an active space. The active space defines a subset of spin orbitals for which all possible electron configurations are included, while the remaining (inactive) orbitals have fixed occupations of 0 or 1.49 The terms corresponding to the inactive orbitals can be included in the electron integrals of active orbitals, and an external potential Vinact() that depends only on the nuclear configuration (i.e. Vnn(
) → Vnn(
) + Vinact(
)).
In general, the orbital vibronic coupling functions dν,pq and gν,pq can be extremely localized and non-analytical at some nuclear geometries. This requires a high-order expansion in terms of the bosonic ladder operators in eqn (3), which greatly increases the number of terms in the Hamiltonian. One way to reduce the expansion order would be a unitary transformation of orbitals for which the derivative couplings vanish, which must satisfy ∂c/∂Qν + dνc = 0, where c is the transformation matrix to the resulting set of “diabatic” orbitals, . This transformation is identical in form to the many-body BO state diabatization condition.14 However, much like multi-electron states in multi-mode systems,50 a set of strictly diabatic orbitals does not exist for an incomplete basis. Instead, because our goal is simply to achieve a low-order expansion in terms of Q, we can use “quasi-diabatic” orbitals where the above strict diabatization condition is not satisfied but all integral coefficients are smooth functions of Q. These orbitals can likewise borrow techniques from BO state diabatization.51
An illustration of our pre-BO analog simulation method in comparison with the BO framework is summarized in Fig. 1. We start by finding a nuclear position dependent spin orbital basis (Fig. 1a) and deriving the expansion coefficients in eqn (3). We then encode the pre-BO wavefunction (Fig. 1b) on a cMQB device, e.g. a trapped-ion device with ion electronic states representing electronic ONVs, and ion motional modes representing the nuclear component of the wavefunction (Fig. 1c). In contrast, existing classical/quantum algorithms in the BO framework require the pre-calculation of BO electronic states (Fig. 1d) with the spin orbital basis. The vibronic wavefunction is then propagated on a truncated BO basis (Fig. 1e).
For the electronic degrees of freedom, we use the Jordan–Wigner transformation which directly maps the occupation number of a spin orbital to the qubit state.55 However, other fermion-qubit mappings could be equivalently employed. The fermionic creation operator for spin orbital p is mapped to a tensor product of Pauli operators (k, Ŷk, Ẑk) for qubits by âp = Ẑ1 ⊗⋯⊗Ẑp−1 ⊗ (
p + iŶp), and the annihilation operator is its Hermitian conjugate. The general expression for the cMQB Hamiltonian becomes a sum of tensor products of multiple bosonic and qubit operators, i.e.
![]() | (4) |
Our electronic basis spans the entire electronic Fock space, including electron configurations with different number of electrons and spins. This implies that our approach can describe intersystem crossing if the corresponding coefficients derived from integrals of the spatial orbitals and electron angular momentum are non-zero. Our approach is also capable of describing ionization processes by adding interactions with terms that remove electrons, which has potential uses for predicting photoelectron spectra. In both cases, the number of quantum resources is unchanged for the simulation.
A schematic illustration of a circuit for pre-BO dynamics simulation with the cMQB mapping consists of initialization, time-evolution, and measurement, as shown in Fig. 2. We provide details for each step in what follows.
As with any quantum dynamics simulation, our approach requires a robust preparation of the initial state which can be generated by a unitary operator Ûinit (Fig. 2a). For most applications in photochemical dynamics we can assume that the initial state is well represented as a product state of electronic and nuclear components of the wavefunction which can be prepared by Ûinit = Ûe ⊗ Ûn, with separate operations on qubits with Ûe and motional states with Ûn. According to the Franck–Condon approximation, the equilibrium nuclear state is unperturbed by photoexcitation and the excitation is instantaneous. This means the initial state can be prepared with a ground (or coherent) nuclear state, and with qubit states corresponding to the photoexcited spin orbital occupation.
Future devices with longer coherence times and more sophisticated quantum control will allow for increasingly accurate initial state preparation. For example, controlled dissipation of the simulator could be used to prepare the vibronic ground state wavefunction, with the fidelity approaching 1 for longer experimental times. The dynamics could then be initialized by a purely electronic (Franck–Condon) or vibronic transition dipole operator.
The direct implementation of our approach outlined above requires the simultaneous control of all degrees of freedom in the simulation. However, our approach uses a fermion-qubit mapping for the electronic degrees of freedom, which involves a change of basis between non-commuting electronic terms. Therefore, we must use Trotterization59 of the time-evolution operator. Trotterization approximates the time-evolution operator as a product of operators, each of which corresponds to a term ĤI in eqn (4). The Trotterized cMQB time-evolution operator is given by where Nop is the number of divided operators and Nt is the number of the Trotter steps (Fig. 2b).
The Trotterized cMQB time-evolution operator for each term with an electronic component in our Hamiltonian (eqn (4)) is given by
![]() | (5) |
This approach can be considered a digital-analog quantum simulation57,58 because the gates involved in generating qubit entanglement take a digital form. Using a coupled fermionic-bosonic operation would eliminate the need for digital gates I for a fully analog simulation. Even with the fermion-qubit mapping, a more compact form of the time-evolution operator could be achieved using time- and spin-dependent squeezing/displacement operators on the collective ionic motions.60 For the remainder of this paper we consider the digital-analog approach.
Trapped ions and cQED typically use bosonic modes to achieve entangling gates in digital quantum algorithms. The quantum gates necessary to implement I are thus readily achieved with these architectures. However, since the entangling gates require a bosonic mode, some bosonic modes (the “bus modes”) must be reserved from those used for simulating nuclear vibrational degrees of freedom. The total number of bosonic modes in the simulator must therefore be at least one greater than the total number of molecular vibrational modes.
Our approach encodes the electronic degrees of freedom of the vibronic wavefunction as an ONV, and nuclear degrees of freedom are mapped to the bosonic modes. This difference in encoding means that all nuclear observables can be measured directly from the simulator, whereas some electronic and vibronic observables will depend on the electronic orbital basis functions. In general, the expectation values of observables can be obtained with Hadamard tests61 for the corresponding unitary operator Ô as shown in Fig. 2a.
For example, a one-electron property 〈Ô1e〉 ≡ 〈Ψ|Ô1e|Ψ〉 is obtained by the sum of the measured elements of the one-electron reduced density matrix (1RDM), 〈â†pâq〉, weighted by the integrals 〈ϕp|Ô1e|ϕq〉 for spin orbitals p and q. The 1RDM is calculated from the quantum simulator by measuring expectation values of Pauli strings, whereas the integrals are obtained on a classical computer. Conversely, the expectation value of a unitary nuclear operator that can be expressed as an exponential, Ôn = exp(Ân), can be implemented directly on the analog simulator with an effective Hamiltonian Ĥeffn = iÂn/τ by evolving for a time τ.
Real-space density functions are important observables for characterizing the dynamics of a molecule. They are expressed in terms of the real-space projection operator, |, Q〉〈
, Q| where
= {x1,…, xNe}. For example, the projection operator for the joint nuclear normal modes and electronic density is
, which yields
![]() | (6) |
Electron and nuclear densities can be obtained by integrating out other degrees of freedom, given by and
respectively. Measurement of the nuclear density, which provides the basis for the analysis of chemical reactions such as isomerization and bond dissociation, simplifies to
![]() | (7) |
As an analog simulation approach, our approach is subject to environmental noise over the course of the simulation, which is a source of error in addition to the Trotter error. The noise effect on an open quantum system can be described by the Lindblad master equation:63
![]() | (8) |
The noise does not equally affect electronic and vibrational degrees of freedom in cMQB simulations, since they are encoded in different resources of the simulator. In addition, because cMQB simulators use motional modes to achieve entangling gates, the electronic degrees of freedom are subject to noise on both qubit states and motional modes.
In ion-trap cMQB devices, the dominant source of noise is the decoherence of the motional modes, with a rate orders of magnitude greater than the other sources of noise.64 Therefore, in what follows, we only consider the noise effect from motional decoherence (ν =
†ν
ν) in an ion-trap cMQB device. While the motional modes representing the nuclear degrees of freedom are subject to the decoherence for the entire experiment time, each qubit is subject to the motional noise only when an entanglement operator is applied on the qubit. Thus, the native rates for the motional decoherence and the indirect qubit noise scale differently for vibrational and electronic degrees of freedom of the molecular state. The native decoherence rate γdmot of motional modes for a cMQB device is scaled to the molecular vibrational decoherence rate γvibmol as
![]() | (9) |
![]() | (10) |
![]() | (11) |
Without spin–orbit coupling, the maximum number of BO electronic states is equal to the number of configuration state functions (CSFs, i.e. spin-adapted linear combinations of Slater determinants) with a fixed electron count and spin, NCSF, which we will discuss in more detail in Section 6. The summation over the BO states in eqn (11) is exact within the given orbital basis set only when NBO = NCSF. The relation between the Born–Huang expansion and our ansatz (eqn (2)) becomes clear via the relation , where
. However, NCSF scales rapidly with the orbital basis set size whereas NBO must be kept small for practical simulations. Therefore, the GBOA is employed (NBO < NCSF) to reduce the computational cost of the BO state calculations and time-propagation of the molecular vibronic state (Fig. 1e) based on prior knowledge on the system and its dynamics.
We construct the electronic Fock space with four spin orbitals {|ϕi〉} using the spin-up and spin-down configurations of two diabatic spatial orbitals, ηa and ηb in Fig. 3b and c respectively, i.e. φ1 = φ3 = ηa, and φ2 = φ4 = ηb. As a result, a total of four ions are needed in the experimental trapped ion setup for the simulation, with only two of the 12 available motional modes needed for the simulation. The diabatic orbitals were obtained by numerically integrating two adiabatic orbitals from the one-electron Shin–Metiu Hamiltonian45 subject to the diabatization condition. The resulting orbitals are localized around the left and right fixed ions, but delocalize slightly as the moving ion approaches the fixed ion (Fig. 3b and c). We replaced the Coulomb potential for Vnn with a harmonic potential to simplify the Hamiltonian and to confine the spatial extent of the wavefunction. The one- and two-electron integrals are truncated at first order in the nuclear position: v() ≈ v0 + v1
, where
with ladder operators
† and
of our single mode with frequency ω of the harmonic nuclear repulsion potential. With these approximations, the cMQB Hamiltonian can be written as
![]() | (12) |
We consider the initial molecular state (Fig. 3d), where the ground state wavefunction of the harmonic oscillator for bosonic degrees of freedom is displaced by R0 = 0.1 a.u. in real space, and electrons are in the closed shell configuration of the orbital ηa (Fig. 3b). To achieve the initial state experimentally on an ion trap, we would first prepare the ground state of two bosonic modes and of each of the qubits, |0000〉q. Then, we would apply a digital quantum operation
1
3, and subsequently apply the displacement operator using a spin-dependent force on the ions with a laser.67 During the simulation, an additional bosonic mode is required to implement qubit entanglement operations.
For the time-evolution, we assume that the base Hamiltonian of the ion trap (the first term in eqn (12)) is always present during the simulation and rescale the Trotter steps to compensate for it.41 To achieve the digital-analog time evolution, we would apply a series of digital quantum gates on the trapped ions with single-ion addressing. The second bosonic mode would be used to apply the entangling gates.67
In Fig. 3, we show the density functions ρ(r, R, t) obtained from exact time-evolution, Trotterized cMQB time-evolution, and the GBOA. Experimental determination of the density function would require an ancilla qubit which controls the operator for the density measurement. Controlled Pauli strings correspond to CNOT and single-qubit gates, and controlled displacement operators can be achieved with a spin-dependent force on the trapped ions.62,67 The sum of all orbital pairs with Fourier transforms over a grid of displacements yields the density, as previously shown in eqn (6). For this example, only 12 sets of measurements are required: two Pauli strings with four on-diagonal and two non-zero off-diagonal pairs. This number of measurements can be halved because the densities of the two spin components are equal. We can obtain a nuclear density resolution of 0.02 a.u. using a grid spacing of 1.26 a.u. in the momentum space with 250 points.
NCSF = 3 is the size of complete electronic basis of singlet electronic states with two electrons in two spatial orbitals. We have excluded the ground BO state from the dynamics with the GBOA (NBO = 2), which has relatively small NACs with the first excited state (see SI for densities of electronic states in the BO framework). Within a short simulation time (56.1 a.u.), the GBOA density shows a spurious electron transfer from the left fixed ion to the right (Fig. 3g) where the subsequent density deviates even further (Fig. 3j). The electron transfer at t = 56.1 a.u. does not occur in the exact simulation (Fig. 3e) due to vibronic coupling to the ground BO state. The inaccuracy of the GBOA in this example is much greater than one would expect for realistic molecules, due to the minimal nature of the model. On the other hand, the Trotterized cMQB time-evolution reproduces the exact density with a converged Trotter step of Δt = 5.6 a.u (Fig. 3f and i) with the fidelity |〈Ψ|Ψexact〉|2 > 0.95, where the convergence of fidelity with respect to the Trotter step is reported in the SI. A more detailed discussion on the comparison of our approach with the BO framework is given in Discussion.
We perform open quantum system dynamics simulations using eqn (8) with Trotterization to estimate the noise effect on the dynamics of our model system, using a Rabi rate for ion–laser interaction of Ω = 2π × 1.0 MHz (ref. 68) and the native vibrational decoherence rate of γdmot = 30 s−1 (ref. 64) based on existing ion trap cMQB devices. We report the time evolution of the fractional occupation numbers (FONs) of ϕ1 and the fidelities in Fig. 4 with vibrational decoherence and the indirect qubit noise effect on the molecular dynamics from the motional decoherence of an ion-trap, in comparison with the result without noise. The indirect qubit noise effect is negligible compared to the direct vibrational decoherence effect on nuclear degrees of freedom in FON, while the fidelity is more sensitive (see SI).
Although the noise effect leads to a deviation from the exact dynamics, the FONs are qualitatively well reproduced (Fig. 4a). Furthermore, the actual noise effect can be minimized by circuit optimization and experimental techniques to reduce the experiment time required for cMQB simulations. For example, the cMQB Hamiltonian of our model system consists of many two-qubit Pauli strings without bosonic terms, where their time-evolution operators can each be realized by a single native two-qubit entanglement operator (e.g. using ) instead of nesting a single qubit exponential operator with CNOT gates.
Fig. 4b confirms the trade-off between the Trotter error and the noise effect. The fidelity with Δt = 9.6 a.u. is greater than the fidelity with Δt = 16.8 a.u. due to the smaller Trotter error. However, the fidelity decreases again when the Trotter step is further reduced to Δt = 5.6 a.u. due to the contribution from the digital part in the molecular scaling factor (eqn (9) and (10)) for the noise. Therefore, the Trotter step should be chosen carefully considering both the Trotter error and the noise.
![]() | ||
Fig. 5 Resource and computational cost comparison of our cMQB approach with other classical and quantum algorithms. (a) Resource scaling of electronic degrees of freedom with respect to No in a logarithmic scale for classical (log2(NCSF)) and quantum pre-BO simulations (log2(Nq)) with singlet electronic states. The resource scaling of our method is represented by the red line, wheras those of equivalent classical simulations is represented by the blue lines for the maximum (dashed), minimum (dotted), and No/Ne = 6 case (solid). (b) Scaling in the number of interactions/gates required to implement the time-evolution operator (log(Nop)) with respect to No for our approach (red), and existing quantum algorithms in the BO framework (yellow) labeled as “qBO” for the maximum (dashed), minimum (dotted), and No/Ne = 6 case (solid). (c) The number of qubits (Nq) required vs. the number of electrons Ne for the pre-BO quantum algorithm proposed in ref. 35 using grids where Nq = 3nNe with n = 10 (black) and our approach with Nq = 6Ne (red). |
The number of qubits required for our approach is equal to the number of spin orbitals, No. On the other hand, the equivalent number of electronic states in the BO framework corresponds to the exact Born–Huang expansion limit (NBO = NCSF). Even for cases where NBO can be kept small, the resources required for the classical calculation of the BO states equivalent to our approach (i.e. FCI, or CASCI for a subset of orbitals) scales with NCSF. Therefore, the classical resources of the BO framework equivalent to our approach scale proportionally to NCSF for a fixed spin-multiplicity, which can be calculated according to the Weyl's formula for the corresponding total spin angular momentum S:
![]() | (13) |
The simulation cost of our approach, which can be measured by the number of interaction terms required, scales as O(No4Nmodek) where k is the maximum order of the Taylor expansion for the functions of bosonic degrees of freedom fI. This cost also remains far smaller than the exponential cost of quantum dynamics simulations on classical computers:
Previous quantum algorithms, including the MQB approach, can potentially have a smaller quantum resource scaling for electronic degrees of freedom than our method: log2(NCSF) < No for the complete CSF basis38,39,41 since the CSF space is a subspace of the spin orbital Fock space. However, the simulation cost Nop would scale as O(NCSF2) for electronic degrees of freedom in existing quantum simulation approaches, whereas the cMQB mapping scales as O(No4) (Fig. 5b). For the nuclear degrees of freedom, our approach has an advantage over other digital or hybrid quantum algorithms for molecular quantum dynamics38,39 since our approach has the same resource scaling as the MQB mapping41 due to the direct mapping on the bosonic modes.
The GBOA reduces Nop to O(NBO2), and its use is justified where a limited number of BO electronic states are accessible at different nuclear geometries. Nevertheless, even if the number of BO states can be reduced, the BO states may be obtained approximately since an exact pre-calculation would require fully diagonalizing the FCI or CASCI Hamiltonian on a classical computer at a prohibitive cost (O(NCSF3)), whereas the cMQB approach requires only the pre-calculation of electron integrals that scales as O(No4). Highly accurate electronic structure methods have a cost scaling greater than our approach (O(No6–8)),69 whereas methods with a lower accuracy require careful characterization. Furthermore, the scalar coupling terms are impractical to calculate for most electronic structure methods and are often neglected, whereas the analogous terms are easily included in our pre-BO approach.
The first proposal of a quantum algorithm for pre-BO molecular dynamics used a real-space, first quantization representation which has a linear resource scaling with respect to the number of grid points per each degree of freedom: 3n(Nn + Ne), where n is the number of qubits required for each degree of freedom and Nn represents the number of nuclei. The authors suggested a minimum grid size of n = 10.35 Although their approach also scales linearly with the number of electrons, our approach has a lower resource cost for most molecules with a reasonable basis set size, where Nmode < Ne and No < 30Ne (Fig. 5c).
The noise effect on the cMQB device depends primarily on the Trotter step size and the number of digital gates used to generate multi-qubit entanglement (eqn. (9) and (10)), which has a trade-off with the Trotter error. Relative to the direct effect of motional decoherence on modes representing vibrations, the indirect qubit noise is negligible. The effect of noise is highly specific to the target problem and device; however, the scaling factors for the noise effect suggest desirable improvements in both computational and experimental aspects of the cMQB approach for a near-term quantum advantage: reducing NCNOT via circuit optimization and reducing tCNOT in cMQB devices.
From a different perspective, noise can be leveraged as an advantage in analog simulators by enabling the simulation of open quantum systems at minimal cost as demonstrated in previous work,41,64,70,71 rather than treating it as an effect that must be suppressed. For example, our model problem can be interpreted as a nonadiabatic charge transfer process occurring in solution under the influence of solvent interactions, or within a crystal at an interface with a substrate.
Considering the quantum resource requirements and noise effects, our approach can be implemented on current and near-term hardware. Early applications are well suited to molecules with a small active space, particularly rigid molecules whose nuclear dependence is well-described by low-order Taylor expansions. These molecules are also the target of efficient classical algorithms, but our approach has advantages for the accuracy by including all terms in the vibronic Hamiltonian for the choice of orbital basis. However, practical advantage should be examined carefully by assessing the achievable efficiency and accuracy in the BO framework and the effect of errors in near-term cMQB devices.
For long term applications, we note that the range of orbital energies in a molecule are typically much greater than the range of energetically accessible BO electronic states. The Trotter step size Δtmol decreases proportional to the orbital energy range. This in turn decreases the scaling factor F due to the minimum experimental time step, leading to longer simulation times and thus proportionally greater noise. To reduce the energy range, we can choose a set of active orbitals at the cost of reducing simulation accuracy. Because our approach is more strongly limited by the orbital energies rather than the number of orbitals, we expect it to accommodate larger active spaces than classical computing approaches. For this reason, our approach will be particularly advantageous for strongly correlated systems, which can have large numbers of nearly degenerate orbitals, leading to extremely large numbers of CSFs and intractable BO-based calculations.
Nuclear quantum effects, such as proton tunneling, are naturally included in our approach. Such effects are essential for describing many chemical processes, especially in biological systems.72 Our choice of model system represents two static moieties (fixed ions) coupled to two electrons and a moving ion (e.g. a proton). The time evolution resulted in a superposition of the moving ion between left and right positions, demonstrating the importance of the quantum description. Furthermore, the results for the model system demonstrate how small changes in the electronic description can result in large differences in the dynamics. These results demonstrate that our approach will enable accurate and efficient simulations of the quantum effects in photochemical dynamics and proton-coupled electron transfer.
A key approximation in our approach is the exclusion of the Coriolis (rotational-vibrational) coupling from the Hamiltonian. On the timescale of ultrafast processes (fs–ps), we expect that the effect of Coriolis coupling would be a negligible perturbation on the vibronic states since the rotational frequencies are relatively small. This approximation also comes with an advantage for simulating physical observables. Any observable quantity that depends on the molecular (Eckart) frame is averaged out over the rotational wavefunction, including electronic and nuclear densities. Experimental measurement of these observables in a lab frame results from spontaneous symmetry breaking between the molecule and its environment.73 This gives a potential advantage for interpretability of our wavefunction compared to other quantum simulation proposals.35,37 Future work could explore how rotations could be included as an environmental effect without compromising the wavefunction interpretability.
As previously mentioned, our approach can describe intersystem crossing by considering spin–orbit coupling electron integrals and ionization by including additional terms in the Hamiltonian (eqn (3)) and incorporating the corresponding operations in the cMQB simulator without any additional quantum resources. Such simulations have the potential to make our approach even more advantageous since the classical and quantum simulations in the BO framework require a greater number of basis states and the corresponding electronic structure calculations for different spin multiplicities. Based on the form of the vibronic Hamiltonian, our Hamiltonian can be considered as an extended version of electron–phonon coupling models used in the solid state physics field such as the Hubbard–Holstein model.74 Therefore, our approach can be readily extended to dynamics involving intersystem crossing and ionization, and translated onto electron-phonon coupling dynamics in solids.
For equivalent descriptions of vibronic dynamics, our approach shows clear scaling advantages over existing quantum algorithms when the accuracy, quantum resources, number of operations, and number of pre-calculations are considered. However, true advantages of our approach for realistic molecules require careful consideration of potential sources of error such as Trotterization, noise, and Taylor expansion truncation, as well as comparison to classical computing approximations. Model systems such as the example presented in this work can be readily implemented on existing quantum hardware such as trapped ion quantum computers with realistic experimental noise. With improvements in quantum hardware and quantum control, we expect that our approach will demonstrate an early quantum advantage for the simulation of quantum chemistry.
This journal is © The Royal Society of Chemistry 2025 |