Abstract
We study the performance of our previously proposed projective quantum eigensolver (PQE) on IBM’s quantum hardware in conjunction with error mitigation techniques. For a single qubit model of H2, we find that we are able to obtain energies within 4 millihartree (2.5 kcal/mol) of the exact energy along the entire potential energy curve, with the accuracy limited by both the stochastic error and the inconsistent performance of the IBM devices. We find that an optimization algorithm using direct inversion of the iterative subspace can converge swiftly, even to excited states, but stochastic noise can prompt large parameter updates. For the 4-site transverse-field Ising model at its critical point, PQE with an appropriate application of qubit tapering can recover 99% of the correlation energy, even after discarding several parameters. The large number of CNOT gates needed for the additional parameters introduces a concomitant error that, on the IBM devices, results in a loss of accuracy despite the increased expressivity of the trial state. Error extrapolation techniques and tapering or postselection are recommended to mitigate errors in PQE hardware experiments.
1. Introduction
The rapid evolution of quantum hardware has opened a venue for addressing challenging problems in quantum chemistry with quantum computation.1−10 However, quantum computing is currently not a practical tool for electronic structure theory. The heart of the problem is that existing implementations of digital quantum computers are noisy, intermediate-scale quantum (NISQ) devices9,11,12 with a limited number of qubits and short coherence times and are subject to several sources of error. Even highly accurate state-of-the-art quantum algorithms require sophisticated error mitigation techniques to tame the noise inherent to NISQ devices.13−16 The ultimate test for quantum algorithms is, therefore, based on their utility in producing highly accurate results on hardware. Within the NISQ context, one of the most promising classes of quantum algorithms for near-term applications is based on the variational quantum eigensolver (VQE),17−19 a hybrid quantum-classical approach that variationally optimizes a trial state based on a quantum circuit. In recent years, there has been a growing interest in finding better alternatives to VQE for NISQ devices.20−42 However, most new quantum algorithms are usually assessed via emulators, and hardware benchmarks are scarce.
In this work, we implement and examine the performance of the projective quantum eigensolver (PQE)20 on quantum hardware. PQE is a NISQ-focused hybrid quantum-classical algorithm that seeks to optimize a parameterized trial state by satisfying projections of the Schrödinger equation onto a many-body basis. PQE has been studied exclusively in the context of method development. Previous work included selected variants of PQE,20 which are part of the family of adaptive methods.20,43−46 Two follow-up studies have focused on reducing the number of CNOT gates in unitary coupled cluster trial states combined with PQE.47,48 More recently, postiterative corrections to PQE have been explored.49 Additionally, a formula used in PQE for asymmetric expectation values was employed by Filip in the context of projective quantum Monte Carlo50 and by Angelakis et al. for an MP2 variant.24 Maitra and co-workers proposed a “two-phase” PQE.51 While the first phase is standard PQE, the second phase computes the values of the “small parameters” by machine learning from the first phase and, thereby, requires fewer quantum measurements. Of greater relevance to this article, a recent publication from Maitra et al. (which appeared while this manuscript was being finalized) applied pre-existing error mitigation techniques to PQE on noisy emulators.52
Although classical numerical simulations of quantum algorithms are useful, actual hardware may deviate from simulations, even for device-specific models of the computer,53,54 thus warranting its own discussion. In the literature, the closest reference points are VQE computations, as hardware results on PQE have not yet been reported. Some papers have reported VQE ground-state energies or properties.17,41,54−93 Other studies have used VQE as an ingredient of methods to compute excited-state energies and properties,41,58,60,69,70,76,77,87,94−96 linear response properties,57,97 molecular dynamics,98 and vibrational eigenstates,99,100 while yet more studies have used VQE as an active space solver for a dynamical correlation or embedding method.96,101−108 Of particular interest to this paper are those VQE computations that report the effectiveness of different error mitigation schemes,56,59,60,63,64,69,73,75,79,82,83,85,88,93,94,103,105 or that are performed on IBM devices.41,54,55,58−60,62,65,67,69,73,74,76,78,79,82,85,87−89,92,93,95,97−105,107 Since our computations will utilize IBM devices and performance is device-specific,76,93,109 comparisons with results from similar hardware are more apt.
Non-VQE methods have been applied to systems of chemical interest on hardware as well. Quantum devices have been used for ground-state methods including quantum Monte Carlo,21 quantum phase estimation,22 perturbation theory,23,24 and the Hermitian and anti-Hermitian contracted Schrödinger equations.25−28 Quantum annealing techniques have been used for ground states and other problems.29−34 Reduced density matrices from non-VQE methods may be passed into other, larger algorithms.25,110 Hardware experiments with variants of quantum imaginary time evolution (QITE) have been used to query zero-temperature and finite-temperature properties on hardware.35−42 Hardware has also been used to study the dynamics of both closed111−130 and open131−139 quantum systems. Other algorithms without VQE have also been applied to hardware to study vibrational and molecular dynamics,140,141 response properties,142 eigenstates and electronic spectra,143−150 and even problems on the frontiers of interest to computational chemistry.151−153
The present article ties these two threads of PQE and hardware together, reporting the first realization of PQE on NISQ devices to assess the effectiveness of techniques to mitigate noise in the measurements and whether the algorithmic differences between PQE and VQE matter on hardware. The paper is organized as follows: Section 2 provides a self-contained introduction to the theory of the PQE and a brief overview of the error mitigation techniques that we employ. The results of our study are reported in Section 3. We begin with a careful analysis of H2 in a minimal basis set, as described in Section 3.1. After summarizing the encoding of the problem onto a quantum computer, we evaluate the noise in PQE at a fixed geometry due to finite measurements, compute an entire dissociation curve including a model of the quantum noise, and then compute a dissociation curve using IBM quantum devices. We then studied the transverse-field Ising model (TFIM) in Section 3.2. We begin by validating our procedure, including a custom tapering scheme, on a noiseless simulator. We proceed to study the 4-site problem, first with noise models and then with results from PQE simulations on quantum hardware. Afterward, we compute TFIM with up to seven sites using noise models on quantum emulators to compare PQE and VQE. We conclude with remarks on the prospects of PQE in Section 4.
2. Methods
2.1. Projective Quantum Eigensolver
Here,
we summarize the PQE algorithm as introduced previously.20 The ground-state energy of a quantum mechanical
system is the lowest eigenvalue of its Hamiltonian, H. Let |Φ0⟩ be a reference state in the domain
of H, usually chosen as simultaneously an unentangled
qubit register and close to an exact eigenstate of H, and let U be a unitary operator on the domain
of H. Then, U|Φ0⟩ is an eigenvector of H if and only if |Φ0⟩ is an eigenvector of . This occurs when the projection of
onto any vector |Φ⊥0⟩ orthogonal to
|Φ0⟩ is zero, i.e.,
. If we construct a basis for the orthogonal
complement of |Φ0⟩, it suffices to satisfy
a residual equation for each basis element.
![]() |
1 |
When the residual equations are satisfied, the energy eigenvalue E can then be read as
![]() |
2 |
In practice, we can neither enforce all residual equations nor vary across all possible unitaries U. Instead, we parameterize our unitary U as a product of N operators
![]() |
3 |
for real parameters θμ and anti-Hermitian operators κμ. We then define the orthogonal complement basis as the set of states
![]() |
4 |
As long as the |Φμ⟩ vectors are orthogonal to each other and to |Φ0⟩, we then enforce the residual equation for each |Φμ⟩. For Fermionic systems, these κμ operators are generators of rotations between occupied (hole) and unoccupied (particle) orbital clusters in second quantization.154 If we denote occupied (unoccupied) orbitals with the indices i, j, ... (a, b, ...), then the κμ operators are of the form a†aai–a†iaa, a†aa†bajai–a†ia†jabaa, etc., and are in one-to-one correspondence with excited Slater determinants. Permuting two orbitals within a Slater determinant produces the same determinant but with a phase of −1.154 We resolve the phase ambiguity by putting orbitals in lexicographical order, which means that eq 4 requires a negative sign for some μ. This is carried through all equations involving |Φμ⟩. We neglect the sign for clarity of exposition. For spin systems, these κμ operators are imaginary multiples of Pauli strings and are associated with a specific spin state on each site.
To ensure that the residuals are zero, we first measure them. In our previous paper,20 we proposed a simple way to measure residual elements. Observe that for our choice of κμ, we have κμ|Φμ⟩ = −|Φ0⟩. Then, exp(θμκμ)|Φ0⟩ = cos(θμ)|Φ0⟩ + sin(θμ)|Φμ⟩. We then showed that the real component of the residual element may be evaluated as
![]() |
5 |
using the expectation value
![]() |
6 |
It is thus possible to measure a single residual element with two quantum measurements beyond what is needed for the ground-state energy E = Eμ(0), in analogy to the parameter-shift rule.155−157 Unlike the parameter-shift rule, this formula requires three energies. We recently derived a residual formula more similar to the parameter-shift rule, which expresses the real component of the residual using only two energies:
![]() |
7 |
We refer to eq 7 as the reference-shift rule. While these formulas are formally equivalent, it does not follow that they are equally performant on NISQ devices. Our initial numerical tests were not able to determine a statistically significant difference between eqs 5 and 7 that was robust to changes in the number of shots and the noise model. Therefore, for aesthetic reasons, our computations will use eq 7.
In PQE, the parameters are then varied
so that the measured residuals
are zero, within some numerical tolerance. In general, the zeros will
not be unique: Not only can multiple θ⃗ produce the same , but different
can approximate different eigenstates of H. Both of these hold for VQE as well. For the optimization
algorithm, we employ a quasi-Newton step accelerated with the direct
inversion of the iterative subspace (DIIS) method,158 as proposed in our previous publication.20 The use of machine learning to predict the values of small
parameters from large parameters has been studied elsewhere.51
2.2. Error Mitigation Techniques
We now sketch the error mitigation techniques that we employ. For details, see ref (9) and references therein. The technical details of how these techniques were applied to individual systems are found in each system’s respective subsection within our results.
2.2.1. Readout Error Mitigation
A major source of error in NISQ devices arises in the measurement process itself, i.e., readout error. For example, a qubit in the |1⟩ state is mistakenly identified as being in the |0⟩ state and vice versa. We correct for it by using X gates to prepare the quantum register in a given bitstring and then measuring the occupation of all qubits. By repeating this operation for all bitstrings, we construct a linear transformation from actual bitstring probabilities to experimentally observed bitstring probabilities. When we obtain any experimental measurement probabilities, we use this matrix to estimate the bitstring probabilities in the absence of readout error via Qiskit’s159 least-squares procedure, which ensures that measurement probabilities stay between 0 and 1.
2.2.2. Qubit Tapering
Qubit tapering is typically considered a technique to reduce the number of qubits and not the error. However, tapering may reduce the resources required, preventing errors from those resources. For example, in Section 3.2, a generator that entangles four qubits is replaced with one that entangles three qubits, and three parameters no longer require CNOT gates. It is therefore legitimate to consider it an error mitigation technique.
We now discuss how to perform qubit tapering.160,161 First, a Pauli string is identified that commutes with the molecular Hamiltonian. When these exist, it is normally a consequence of some physically meaningful symmetry, e.g., total spin or point group symmetry. It is possible to perform a similarity transformation so that the string becomes a single Pauli gate on qubit n. Upon performing this transformation, all operators in the Hamiltonian will take an I or Z in that qubit, and the eigenstate can be chosen to consist only of bitstrings with a 0 for that qubit or only of bitstrings with a 1 for that qubit. The experimenter can then decide which choice is correct for the eigenstate of interest to them. The qubit is then eliminated from the computation, and the I and Z are replaced with their corresponding +1 or −1 depending on what that gate would return on qubit n. Although specific forms of this similarity transformation have been prescribed previously, those prescriptions are not essential to the method.
2.2.3. Postselection
When symmetry is present, some bitstrings should have no counts. However, errors can introduce counts, and it is precisely this that postselection eliminates.154,162 In particular, suppose a state is an eigenfunction of the Pauli string P. If it is possible to simultaneously measure P and other Pauli strings (such as those in the Hamiltonian), then when we measure the Pauli strings simultaneously, we can discard all measurements that yield incorrect expectation values for P.
We mention one technical limitation. Although quantum mechanics permits us to simultaneously measure any operators that mutually commute, current quantum hardware permits us to simultaneously measure only any operators that commute for every qubit in the Pauli string, i.e., that mutually qubitwise commute. To apply postselection, we must therefore perform a unitary transformation so all operators of interest qubitwise commute, as has been studied previously.163−166
2.2.4. Error Extrapolation
Error extrapolation167−169 is flexible in what error it corrects for, only requiring that it can be scaled. The error strength is represented by ϵ, and we seek the exact energy, which is obtained at ϵ = 0. We do this by assuming that the dependence of the energy on ϵ follows some user-provided functional form and fitting its parameters according to measurements at increased values of ϵ. The “error-free” expectation value can then be read as the value at ϵ = 0. Much like the complete basis set extrapolations of quantum chemistry,170 this method is inherently heuristic but can be quite reliable. However, error extrapolation ceases to be reliable for time-dependent noise.171
A standard source of error targeted with error extrapolation is that coming from the CNOT gates. This is justified when CNOT gates are the only qubit-entangling gates, and gates that act only on a single qubit have much smaller errors. In this case, we scale the noise by a factor of 2n + 1 by adding n CNOT pairs, which are just the identity in an errorless circuit, to each CNOT gate. Because these pairs are the identity, codes such as Qiskit will remove them during a result processing step called transpilation. To prevent this, we converted transpiled circuits to OpenQASM form,172 added the CNOT pairs, and then restored them back to implementable Qiskit circuits. After the job was completed, we retrieved the executed circuits using Qiskit’s API to confirm that the CNOT pairs were inserted as expected.
The functional form for error extrapolation remains to be determined. Although the original proposals of error extrapolation used a linear function,167,168 exponential extrapolation was also found to achieve good accuracy.169 We shall study both but advise the reader that we modified the exponential extrapolation procedure for numerical stability. Details are provided in Appendix A. We also point out that Maitra and co-workers52 studied error extrapolation, but exclusively on simulators.
3. Results and Discussion
Unless otherwise indicated, all simulations on quantum hardware used 8192 shots for each measurement.
3.1. H2
3.1.1. Simulation Details
We simulated H2 with a minimal STO-6G atomic basis, which includes one basis function per atom. We first parameterize our state. From the atomic basis, two spatial orbitals can be formed: a bonding orbital ϕg of σ+g symmetry and an antibonding orbital ϕu of σ–u symmetry. The ground state possesses 1Σ+g symmetry, so by standard symmetry arguments, we can write it as
![]() |
8 |
We encode the two-dimensional space of all possible symmetry-allowed states onto a single qubit by mapping |ϕgαϕgβ⟩ to |0⟩ and |ϕuαϕuβ⟩ to |1⟩, so
![]() |
9 |
The same ansatz may be derived by beginning with Fermionic second quantized operators173 and applying Z2 symmetry reduction to the Jordan–Wigner encoding.160,161 We immediately see that U(θ) = Ry(θ). This may be efficiently implemented with no multiqubit gates. With the encoding complete, we can then construct the Hamiltonian in the basis of our two states using integrals obtained from the Psi4 package.174
We compute a dissociation curve from 0.4 to 6.0 Å, and for each bond length, we perform ten hardware experiments. A hardware experiment consists of both the optimization of the amplitude to within a residual of 0.03 and the measurement of the energy at the optimized amplitude (we find that tighter convergence metrics significantly increase the number of iterations required with no appreciable increase in accuracy). As the bond stretched, in some experiments, we observed qualitatively incorrect energies. When these occurred, the energy points were excluded from all further consideration, and a new batch of 10 experiments was recomputed. Data from problematic batches were not considered in the subsequent statistical analysis. We discuss the origin of this effect below. We used a single readout error calibration matrix for each set of ten energy computations. All computations were performed within 1 h of computing the readout error calibration matrix. Our initial guess was always θ = 0.
3.1.2. Results
We show our H2 dissociation curve in Figure 1, computed on IBM’s Lima device. We observe that for all points, the averaged measured energy (shown as a red horizontal line) agrees with the observed energy to within four mEh, a good agreement considering the convergence threshold used.
Figure 1.
Hardware experiments on the singlet ground state of the H2 molecule in a minimal basis set. Top: average energies measured in 10 experiments on a quantum computer (blue circles) compared to exact full configuration interaction energies (black curve). Bottom: errors from full configuration interaction for individual experiments (stacked red circles) and the average of 10 experiments (blue circles). Error bars show standard deviations from the average of 10 experiments. All hardware experiments were performed on the IBM Lima device with 8192 measurements (shots) per experiment, readout error mitigation, and a 1-qubit encoding.
Both the standard deviation of the experiments and the average error in the curve are smallest for geometries closest to dissociation and largest when the atoms are closest together, especially beyond the equilibrium geometry. Our group has observed a similar phenomenon previously.110 We attribute this to the simple fact that the error in a term in the energy is directly proportional to a prefactor derived from matrix elements. The energy may be expressed in terms of Hamiltonian matrix elements (Hij) and the nuclear repulsion energy (Enuc) as
![]() |
10 |
While the prefactor for the X string varies from 0.17 at compressed geometries to 0.34 at stretched geometries, the prefactor for the ⟨Z⟩ string varies from −0.99 to 0.00 along the same curve, being larger in magnitude when the atoms are near each other. We therefore attribute both the larger variance and larger error at compressed geometries to the geometry dependence of the prefactor of the Z term (the magnitude of this effect, however, is larger than we would have expected solely by comparing coefficients of ⟨Z⟩). This stochastic error is the dominant effect for stretched geometries.
However, statistical error alone cannot account for the persistent fact that for compressed geometries, the energy frequently falls below full configuration interaction. From this and the fact that PQE is a variational upper bound to the energy, we can immediately conclude that hardware noise is distorting our energy measurement. As we will discuss below, for this system, the PQE residual is equivalent to the VQE gradient, so we can regard this as a VQE optimization. We observe that noise effectively distorts the angle used for measuring X relative to the angle used to measure Z. VQE uses this to drive the energy below the full configuration interaction.
We caution the reader that due to the larger error
in the three
most compressed geometry points, we needed to recompute these data
points several times, discarding the qualitatively incorrect computations.
We observed variations in energy of over 4 mEh even when no parameters were changed. In the final
data we present, we magnified the number of shots used for readout
calibration by a factor of 10, which should improve the accuracy of
measuring the readout calibration matrix by a factor of .175 After making
this change (and no others), our energies improved by several mEh.
Because PQE is an iterative algorithm, a complete cost estimate must consider the number of iterations required to converge. We show this for two representative geometries in Figure 2. When the bond distance is 0.75 Å, close to equilibrium and where the initial guess is quite accurate, energy convergence is rapid, while amplitude convergence takes place after three iterations. At 2.95 Å, where the initial guess has only 54% overlap with the exact ground state, convergence is much more erratic but relatively swift, being achieved within eight iterations. The reader may also note a large spread in the energies after two rounds of parameter optimization, with some energies being worse than the initial guess. This is an artifact of our optimization algorithm. A pure Jacobi step will cause a large amplitude update only when the residual is large or the Hessian is nearly singular. Neither are the case here. When we optimize with DIIS, the optimization is sensitive to stochastic noise in the measurements and can cause very large extrapolations. Despite the energy having a period of 2π in the amplitude, we have observed cases where the DIIS optimizer sends parameter values far away from the initial guess of 0, into the neighborhood of ±1000. We have observed similar behavior on a noiseless simulator but still found the performance of DIIS to be superior to the performance of a pure Jacobi step. We have found that taking the DIIS amplitude modulo the period eliminates this behavior and leads to swifter convergence on noisy simulators but have not tested this on quantum hardware. Other modifications to the DIIS update, such as a trust radius, can also be imagined.
Figure 2.
Hardware experiments on the singlet ground state of the H2 molecule in a minimal basis set computed at H–H distances of 0.75 (red circles) and 2.95 (blue triangles) Å. Convergence of the energy error with respect to full configuration interaction after n iterations of the PQE algorithm for ten runs. All hardware experiments were performed on the IBM Lima device with 8192 shots, readout error mitigation, and a 1-qubit encoding.
We noted in Section 3.1.1 that for stretched geometries, the PQE algorithm returned qualitatively incorrect energies in error by over 0.5 Eh. We were able to reproduce this behavior on a quantum simulator with no noise model and shall henceforth discuss results on the simulator, where we could run PQE repeatedly. We found that the erroneous energies were within 5 mEh of the energy of the excited state at that geometry. To understand this, recall that we initialize the wave function in the state |0⟩. As the bond stretches, the overlap of |0⟩ with the ground state diminishes from over 99 to 50%. As our optimizer tries to zero projections, which can be done by targeting any eigenstate, it is unsurprising that, in this case, stochastic errors direct the optimization toward the excited state.
We view this result positively: In PQE, with an appropriate choice of amplitudes, it is possible to target an excited state (we refer the interested reader to other quantum algorithms40,95,148,176−180 to compute excited states). Indeed, by changing our initial guess, we were able to target the excited-state dissociation curve on the quantum simulator. The utility of this feature of PQE depends on how easy it is to obtain guess amplitudes and how accurately higher solutions of an approximate unitary ansatz can represent an excited state. This is a promising avenue for future work.
Lastly, we compared the performance of PQE to that of VQE after each optimization iteration at the minimum energy geometry. The results are shown in Figure 3. The numerical performances are nearly identical. In fact, it may be shown that for the special case of one-parameter systems, the PQE residual is the VQE gradient (this is most easily shown by comparing the two-measurement formulas for the residual to the VQE shift rule). At least in this case, all differences between the two algorithms are due to numerical noise and do not indicate a meaningful performance difference in the algorithms.
Figure 3.
Simulation of hardware experiments on the singlet ground state of the H2 molecule in a minimal basis set at the minimum energy geometry. Comparison of the energy convergence averaged over ten runs of the PQE (red circles) and VQE (blue triangles) algorithms. All simulations utilized the error model for the IBM Lima device, employing 8192 shots and a 1-qubit encoding.
3.2. Transverse-Field Ising Model
3.2.1. Simulation Details
The Hamiltonian for the TFIM on N sites is given by
![]() |
11 |
which describes a 1-D chain of spin sites subject to two kinds of interaction. First, adjacent spin sites have a coalignment-penalizing interaction with magnitude controlled by J. Second, an external electric or magnetic field interacts with all sites, with the magnitude controlled by h, rewarding sites that align with the field. All Pauli strings in H commute with the Pauli string ∏Ni=1Zi, which may be physically interpreted as reversing the direction of the field and mathematically interpreted as requiring that for any eigenstate; the bitstrings all have the same number of |1⟩ qubits, modulo two, i.e., parity symmetry. In the context of the present work, this symmetry enables postselection and the tapering of one qubit.
We choose h = J = 1 and, aside from the last subsection, we focus on the case N = 4. With this choice of parameters, we modeled a strongly correlated system. The choice of J/h = 1 puts this system at the critical point where it transitions between ferromagnetic and antiferromagnetic phases in the thermodynamic limit.181,182 In the uncorrelated basis of {|↑⟩,|↓⟩}⊗N, the overlap of any basis state with the correlated wave function is at most 81.2%, and the correlation energy (defined as the difference between the TFIM solution with J = 0 and the exact energy) is 15.9% of the total energy.
We have chosen our wave function ansatz as the so-called “quantum combinatorial ansatz” of ref (183). The wave operator is a product of exponentiated imaginary multiples of Pauli strings. Each Pauli string is a product of a single Y and X gate, which always acts on the largest index of the nontrivial indices for that string. This ansatz is known to represent any state with real wave function coefficients (for complex coefficients, strings without a Y gate must also be allowed). We loop over the largest qubit index, from smallest to largest, but the operator order is otherwise arbitrary. In some particular cases, we have chosen to neglect operators that are unlikely to have large amplitudes. We shall explicitly say when we do this. An advantage of this choice of ansatz is that it is an exact ansatz for any state, tapered or not. By symmetry, we may set to zero the coefficients of all operators in the untapered ansatz with an odd number of nontrivial operators. We choose our starting state to be |0⟩⊗N, and this choice of operators preserves the parity symmetry.
Let us discuss how symmetry applies to the use of postselection. Here, there are two sets of mutually commuting operators to measure. The first set consists of Z operators on various qubits, all of which already commute qubitwise with ∏Ni=1Zi. The second set consists of XiXi+1, which does not commute qubitwise with ∏Ni=1Zi. To transform these operators into ones that can be simultaneously measured, we prepend a CNOT staircase, a sequence of CNOT gates where qubit i controls qubit i + 1 for all qubits. This transforms XiXi+1 into Xi+1, allowing for the simultaneous measurement of all such operators.36 Our use of tapering is more involved and is described in the next section.
3.2.2. Exact Results
Before analyzing the robustness of PQE to noise, we first studied our algorithm on a noiseless simulator. With the procedure deployed above, we can reproduce the exact energy, as determined by Hamiltonian diagonalization, to within machine precision. This confirms the basic correctness of our ansatz and of PQE when noise and noise mitigation are neglected. Looking at the corresponding parameters, we observe that they fall into two classes. One consists of all coefficients of XiYi+1 Pauli strings, which all have a magnitude of at least 0.45. The second class of parameters has a smaller value, with a magnitude less than 0.12. Such a partitioning of PQE parameters was central to ref (51) and could also have been anticipated from the perturbative arguments of ref (183). If these three large parameters are optimized while all other parameters are set to zero, then we are able to retain 97% of the correlation energy. The remainder of this section is concerned with how this division can be exploited in conjunction with qubit tapering.
As discussed in Section 2.2.2, standard qubit tapering requires a choice of qubit to be tapered out. When we attempt all four qubit taperings from the standard procedure, we are able to reproduce the exact energy but find that only tapering qubits one or two provides an ansatz that retains at least 97% of the correlation energy when only three parameters are optimized. The other choices of qubits to taper lead to only 86% of the correlation energy when only the three largest parameters are retained. The exact solutions for the other two taperings blur the separation into large and small parameters. Tapering qubit three yields one parameter that increases to magnitude 0.35, and tapering qubit four yields two parameters that increase to magnitudes 0.23 and 0.27, respectively. However, even the two more accurate taperings are less than satisfactory. For one, the resulting Hamiltonians include a term that acts on every qubit, which costs the locality of the Hamiltonian. Additionally, the parameters that are important still retain at least one generator that acts on multiple qubits and are therefore more likely to introduce additional noise in the experiments. It would be best if the large parameters were associated with action on only one qubit.
We identified a nonstandard tapering procedure that achieves this goal. Any parity-allowed basis state can be reached by applying a sequence of XiXi+1 operations to the starting state, |0⟩⊗4. Starting from the first qubit, XiXi+1 is applied if needed to achieve the target value on qubit i, and parity is used to enforce that the final qubit is correct. We map each basis state to a state where index i is 1 if XiXi+1 was needed and 0 otherwise, producing an overall state of 3 qubits. For example, |0101⟩ = (X0X1)(X1X2)|0000⟩ maps to |011⟩ = X0X1|000⟩. We can then construct an operator in this reduced space that produces the same expectation values as in the old space after accounting for the transformation. Notice that the operator XiXi+1 will always transform to Xi in this alternative tapering scheme.
We confirmed that this tapering scheme reproduces the exact energy. As expected, there are three large parameters with a value of at least 0.45, corresponding to the Xi, while all other parameters are of a magnitude below 0.12. Restricting the ansatz to the “large” parameters leads to recovering 97% of the correlation energy as desired. Further, all terms in the Hamiltonian involve either one qubit or two adjacent qubits, meaning that locality is restored.
Before proceeding, let us assess the quality of the wave functions produced by eliminating small parameters. We can do this by measuring the spin–spin correlation functions, ⟨σiσj⟩, which measure the alignment of spin in the direction σ ∈ {X, Y, and Z} on qubits i and j. In Figure 4, we see that many of the correlation functions computed with the exact and 3-parameter standard ansatze are quite close, but those correlation functions between nonadjacent qubits, whose generators have been set to zero, can be far off. The same trend continues for the YY and ZZ correlation functions, albeit less strongly. We see similar behavior for our custom tapering scheme, which tends to be more accurate for XX and ZZ correlations but slightly less accurate for the YY correlation. We caution that we need to transform the σiσj operators into the new basis, which we do not discuss here. Nonetheless, we see results of similar quality. This confirms that our representation of the wave function has not been impaired by tapering.
Figure 4.
Spin–spin correlation functions of the 4-site TFIM. The left, center, and right columns measure correlation in the X, Y, and Z directions, respectively. The top row uses full configuration interaction, the middle row uses the three largest parameters in the standard ansatz, and the bottom row uses the three largest parameters in our tapered ansatz.
3.2.3. Simulator Results
We begin our exploration of the TFIM by studying combinations of the error mitigation techniques described in Section 2.2 using a simulator of the noisy Lima device of IBM. One instance of our simulations is given in Table 1. We caution the reader that due to both the stochastic nature of quantum sampling and, more importantly, drift in the error model describing the device, repeating this experiment leads to different percentages, with a greater difference from 100% leading to greater variability. For example, the correlation energy recovered can vary between 67 and 78% when postselection is used without extrapolation. Drift in the noise model leads us to recover between −68 and −28% of the correlation energy when only readout error is mitigated.
Table 1. Percentage Correlation Energy Recovered for the TFIM on a Simulator Modeling Noise from IBM’s Lima Devicea.
extrapolation
method |
|||
---|---|---|---|
symmetry use | none (%) | linear (%) | exponential (%) |
none | –68 | 36 | 97 |
tapering | 19 | 82 | 98 |
postselection | 67 | 98 | 104 |
Readout error mitigation was used for all computations. Results are averaged over 100 runs.
What is qualitatively clear is that for the 4-site TFIM, the extrapolation techniques are ordered, as are the symmetry-exploiting techniques. Postselection is more accurate than qubit tapering, which is more accurate than merely removing the symmetry-breaking generators. Exponential extrapolation is more accurate than linear extrapolation, which is more accurate than no extrapolation. Although it is useful to learn that eliminating erroneous measurements via postselection is a better way to eliminate error than eliminating extra qubits via tapering, all other comparisons of methods may have been easily anticipated.
3.2.4. Hardware Data
We now turn to experiments on the quantum hardware. We found that the noise on IBM devices fluctuated in intensity to the point that even with error mitigation, we would observe qualitatively different results from experiments submitted just hours apart. Significant hardware drift over time has also been reported previously.82 In one instance, results that ran and were completed within 1 h of each other differed by −0.2 Eh due to IBM recalibrating their hardware between the experiments. This severely limited the rigor with which we could perform experiments. We also observed that when we attempted error extrapolation, for some circuits, the error upon adding additional CNOT gates led the count distribution to be further away from pure noise, violating the assumptions of the exponential error extrapolation model discussed in Appendix A. Accordingly, exponential error extrapolation could not be performed reliably. Degraded performance of error extrapolation on hardware compared to that of a noise model on an emulator has been reported previously,184 and appropriate forms of error mitigation have been known to be hardware-dependent.83 For other uses of extrapolation schemes on hardware, see refs (60, 185, and 186).
One particularly successful experiment used our custom tapering for the 4-site TFIM and included only the three nonentangling generators, i.e., using an approximate ansatz that can be evaluated more accurately on the quantum hardware. We repeated this ten times and measured a final energy of −4.77 ± 0.03 Eh. The exact energy is −4.76 Eh. Within the precision possible on NISQ devices, the accuracy cannot reasonably be improved. An extension of this tapering scheme to molecular systems is unclear. Regardless, these findings make it clear that for qubit tapering, some similarity transformations vastly outperform others.
We report additional experiments that are encouraging for PQE: we combined our tapering scheme, with all generators except the three-qubit entangler, with exponential extrapolation. On the Lima device, this reproducibly gave 120% of the correlation energy. By combining linear extrapolation and qubit tapering on the IBM Perth device, as well as removing the three-site generator, we were able to measure an energy of −4.70 Eh, i.e., we recovered 92% of the correlation energy. Given this data point and the similarity of the PQE and VQE equations, we suspect that the difficulties we observed measuring data from PQE are not specific to the PQE algorithm but reflect the challenges of obtaining reliable measurements from NISQ devices with all but the simplest quantum circuits. We add that although we attempted simulations with postselection, the results were inferior to those of exponential extrapolation and tapering.
3.2.5. PQE and VQE Comparison
While the hardware results of the preceding section indicate strong noise effects, they do not conclusively establish whether the poor performance is unique to PQE. To investigate that question, we return to the simulators and perform PQE and VQE computations on the TFIM with increasing sites. We employ IBM’s noise model of their Perth device since the Lima device used earlier in the paper has only five qubits, but our experiments require seven. Numerical experiments on the Lima device, for those systems where it was possible, show the same PQE and VQE comparison as those from the Perth device but with a shift in the amount of correlation energy recovered that increases with the number of qubits. We employ linear and exponential error mitigation, with both tapering and postselection, as Section 3.2.3 shows these strategies to be advantageous. To minimize stochastic finite measurement error, we magnified our standard 8192 shots by a factor of 40. We performed VQE using analytic gradients augmented by the shift rule, accelerated by DIIS. For an additional discussion on combining VQE and DIIS, see Appendix B.
The results using the Perth device noise model are shown in the top panel of Figure 5. PQE and VQE results are extremely close in all cases. They disagree the most, by percentage, when exponential extrapolation is combined with tapering for the 7-site problem. With this combination of error mitigation techniques, although individual computations converged, different simulations would lead to different energies, and we had to repeat the experiment many times to obtain converged statistics. We suspect that performing more computations would yield closer agreement. Nonetheless, even in this case, it is clear that the main reason for the poor quality of our hardware results is hardware noise and not PQE or the optimization algorithm employed. Rather, the noise is too intense for either PQE or VQE using analytic residuals or gradients, respectively. From Figure 5, we can also observe that error steadily increases as the number of qubits increases, from 4 to 7. For the 7-qubit simulations, combining exponential extrapolation and postselection does not yield converging energies. This matter is discussed in Appendix A. Combining exponential extrapolation with tapering is the most robust approach for the larger systems, but even in this best-case scenario, the error is already prohibitive at 7 qubits.
Figure 5.
Simulation of hardware experiments on the TFIM with 4 to 7 sites. Correlation energy measured employing all combinations of postselection or tapering, with linear or exponential error extrapolation. Results in the top and bottom panels utilized error models for the IBM Perth and Brisbane devices, employing 327,680 shots.
While this paper was in preparation, IBM replaced the devices we had been using with newer ones. As an initial exploration of the quality of the new devices that were made available, we repeated our tests on TFIMs on the Brisbane device. The results are presented in the lower panel of Figure 5, which shows a radically different description of the device quality. The percentage correlation energy captured is now consistent across the extrapolation procedure and number of qubits. The only significant factor is whether postselection or qubit tapering was used, recovering about 90 and 70% of the correlation energy, respectively. These results are very encouraging for the future use of such devices.
4. Conclusions
In this article, we have studied the performance of the PQE on quantum hardware to assess whether it is a viable and competitive algorithm to solve for molecular electronic structure on NISQ devices. While we encountered difficulties in applying PQE, these can be attributed to the inherent challenges of using NISQ devices, even with error mitigation. We applied PQE to H2 in a minimal basis set with qubit tapering and mapped the state to a one-qubit problem. In this case, the PQE residual condition is formally identical to the zeroing of the gradients of VQE computed via the parameter-shift rule. For this problem, the PQE converges swiftly near the equilibrium geometry. However, when the initial guess is not as accurate, PQE may either converge to an excited state or become more sensitive to errors, leading to a large spread of energies. For the 4-site TFIM, PQE performs excellently on a noisy simulator, recovering over 99% of the correlation energy using a combination of exponential error extrapolation and postselection. On physical hardware, errors complicate the picture. While we managed to recover over 90% of the correlation energy for the TFIM, device errors prevented us from doing this consistently. We found more success by using a custom qubit tapering that allows for 99% accuracy to be achieved with generators that require no CNOT gates. More reliable hardware than that used in this study is required before the PQE and VQE can be definitively compared. An initial study using data from newer IBM devices, which replaced the ones we used while this manuscript was being finalized, suggested that PQE and VQE are comparable and can recover 90% of TFIM correlation energy with proper error mitigation and the full ansatz. A study of PQE on this newrer hardware would be highly informative.
We now identify directions for future development of the PQE until devices become more robust to error. We have identified new formulas for the residuals that need to be measured in the theory. Although not discussed in great detail here, the similarities to the shift rule allow for more fruitful incorporation of ideas from the adaptive VQE into the selected PQE. Second, we identified potential for improvement in the optimization of PQE. During the course of our studies, providing a suitable initial guess and exploiting periodicity in the parameters were both found to improve the convergence of DIIS. These areas merit further investigation, as do alternative convergence algorithms. Work in this direction has already begun.51 Lastly, and perhaps most intriguingly, our results on H2 suggest that PQE can be used to target excited states by identifying other roots of the PQE residual conditions and leveraging the fact that PQE does not use direct minimization. We plan to further explore these excited-state PQE solutions in our subsequent research.
Acknowledgments
This work is supported by the U.S. Department of Energy under award no. DE-SC0019374 and the NSF under grant no. CHE-2038019. We acknowledge helpful discussions with Ilias Magoulas, Renke Huang, Harper Grimsley, Shi-ning Sun, and Austin Minnich. We acknowledge the use of IBM Quantum services for this work. The views expressed are those of the authors and do not reflect the official policy or position of IBM or the IBM Quantum team. This article is dedicated to Gustavo Scuseria and his many contributions to elucidating the physics of many-body systems.
Appendix A
Exponential Error Extrapolation
In exponential error extrapolation, the dependence of the expectation value of a Pauli string (other than the identity) is modeled as an exponential function of the error strength, E(ϵ) = λ exp(κϵ), where λ and κ are adjustable parameters. It follows that E(0) = λ, which can be computed from the values of E(ϵ) for two particular ϵ. The formula specifically involves the ratio of the measured energies. Therefore, the exponential model inherently assumes.
The expectation value does not flip the sign.
The ratio of the expectation values is numerically stable.
Random error, e.g., from sampling, is negligible in comparison to the systematic error being corrected.
The expectation value decreases in magnitude as error increases, going to zero in the infinite noise strength limit.
The first two of these assumptions will fail when the measured expectation values are near-zero, which will happen for many Pauli strings. The third assumption will fail when systematic noise only weakly affects a particular expectation value, which the fourth assumption predicts when the true expectation value is near zero. Hence, in cases where we have small expectation values, exponential error extrapolation is unreliable due to sampling noise, but the fourth assumption predicts that it should have a small impact. PQE likely has more of these small expectation values than VQE without analytic gradients because PQE must measure residuals involving multiple wave functions that are not localized in Hilbert space.
Accordingly, for each Pauli string, we measured the expectation values for both noise strengths as if we would perform exponential extrapolation and then checked that the expectation values had the same sign and decreased as error increased, that the high-error expectation was no smaller than 0.0001, and that the ratio of the errors was no greater than 2500. If any of those conditions were false, we did not perform exponential extrapolation and assumed that the low-noise value was sufficiently accurate: failure of these assumptions is likely when the true expectation value is already small. Otherwise, we performed exponential error extrapolation. Naively performing exponential error extrapolation in all cases led to numerical instabilities.
The two numerical thresholds mentioned in the preceding paragraph
were chosen arbitrarily based on numerical experience with the 4-site
TFIM. We found these thresholds to not be robust for different systems.
For the 7-site TFIM, we frequently observed some cases where the error
ratio was on the order of 1000 due to a large measurement on the first
extrapolation point. This led to the already large measurement being
amplified by a factor of . This is fundamentally the reason we observe
the nonconvergence reported in Figure 5. While tighter thresholds would prevent error magnification,
they fail to correct the error in the first data point.
Appendix B
DIIS Acceleration of VQE
Reference (20) reported that VQE required 2–3 times as many gradient evaluations as PQE to obtain a given energy accuracy, at least for the system under consideration. We were not able to reproduce this in our experiments with PQE and VQE on TFIMs. While we found that PQE did converge faster than VQE in the absence of noise, this happened by one to two iterations.
To understand this difference, note that ref (20) compared PQE using DIIS to VQE using the BFGS algorithm.187−190 Our experiments used DIIS for both systems. Upon repeating the experiments of ref (20) with DIIS for both algorithms, we found that PQE required an iteration or two fewer than VQE, as in our results. We thus conclude that the reports of swifter convergence of PQE than VQE from ref (20) were incomplete. The convergence speed depends on the optimization algorithm. The two algorithms perform comparably with DIIS, with PQE having a slight edge. At least for minimal amounts of noise, VQE optimization with DIIS requires fewer gradients than with BFGS. We make no statement regarding other optimization algorithms or performance under more realistic noise models. Shortcomings of DIIS optimization are noted in Section 3.1.2, and the problem of PQE optimization was studied in ref (51).
Author Present Address
† Department of Chemistry, Virginia Tech, Blacksburg, Virginia 24061, United States
The authors declare no competing financial interest.
Special Issue
Published as part of The Journal of Physical Chemistry A virtual special issue “Gustavo Scuseria Festschrift”.
References
- Cao Y.; Romero J.; Olson J. P.; Degroote M.; Johnson P. D.; Kieferová M.; Kivlichan I. D.; Menke T.; Peropadre B.; Sawaya N. P. D.; Sim S.; Veis L.; Aspuru-Guzik A. Quantum Chemistry in the Age of Quantum Computing. Chem. Rev. 2019, 119, 10856. 10.1021/acs.chemrev.8b00803. [DOI] [PubMed] [Google Scholar]
- Bauer B.; Bravyi S.; Motta M.; Chan G. K.-L. Quantum Algorithms for Quantum Chemistry and Quantum Materials Science. Chem. Rev. 2020, 120, 12685–12717. 10.1021/acs.chemrev.9b00829. [DOI] [PubMed] [Google Scholar]
- Cheng H.-P.; Deumens E.; Freericks J. K.; Li C.; Sanders B. A. Application of Quantum Computing to Biochemical Systems: A Look to the Future. Front. Chem. 2020, 8, 2296. 10.3389/fchem.2020.587143. [DOI] [PMC free article] [PubMed] [Google Scholar]
- McArdle S.; Endo S.; Aspuru-Guzik A.; Benjamin S. C.; Yuan X. Quantum computational chemistry. Rev. Mod. Phys. 2020, 92, 015003. 10.1103/revmodphys.92.015003. [DOI] [Google Scholar]
- Head-Marsden K.; Flick J.; Ciccarino C. J.; Narang P. Quantum Information and Algorithms for Correlated Quantum Matter. Chem. Rev. 2021, 121, 3061–3120. 10.1021/acs.chemrev.0c00620. [DOI] [PubMed] [Google Scholar]
- Ollitrault P. J.; Miessen A.; Tavernelli I. Molecular Quantum Dynamics: A Quantum Computing Perspective. Acc. Chem. Res. 2021, 54, 4229–4238. 10.1021/acs.accounts.1c00514. [DOI] [PubMed] [Google Scholar]
- Bharti K.; Cervera-Lierta A.; Kyaw T. H.; Haug T.; Alperin-Lea S.; Anand A.; Degroote M.; Heimonen H.; Kottmann J. S.; Menke T.; Mok W.-K.; Sim S.; Kwek L.-C.; Aspuru-Guzik A. Noisy intermediate-scale quantum algorithms. Rev. Mod. Phys. 2022, 94, 015004. 10.1103/revmodphys.94.015004. [DOI] [Google Scholar]
- Miessen A.; Ollitrault P. J.; Tacchino F.; Tavernelli I. Quantum algorithms for quantum dynamics. Nat. Comput. Sci. 2022, 3, 25–37. 10.1038/s43588-022-00374-2. [DOI] [PubMed] [Google Scholar]
- Motta M.; Rice J. E. Emerging quantum computing algorithms for quantum chemistry. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2022, 12, e1580 10.1002/wcms.1580. [DOI] [Google Scholar]
- Ayral T.; Besserve P.; Lacroix D.; Guzman E. A. R. Quantum computing with and for many-body physics. arXiv 2023, arXiv:2303.04850.[quant-ph], submitted 09-27-2023 10.48550/arXiv.2303.04850. [DOI] [Google Scholar]
- Preskill J. Quantum Computing in the NISQ era and beyond. Quantum 2018, 2, 79. 10.22331/q-2018-08-06-79. [DOI] [Google Scholar]
- Lau J. W. Z.; Lim K. H.; Shrotriya H.; Kwek L. C. NISQ computing: where are we and where do we go?. AAPPS Bull. 2022, 32, 27. 10.1007/s43673-022-00058-z. [DOI] [Google Scholar]
- Cai Z.; Babbush R.; Benjamin S. C.; Endo S.; Huggins W. J.; Li Y.; McClean J. R.; O’Brien T. E. Quantum Error Mitigation. arXiv 2023, arXiv:2210.00921.[quant-ph], submitted 12-28-2023 10.48550/arXiv.2210.00921. [DOI] [Google Scholar]
- Endo S.; Cai Z.; Benjamin S. C.; Yuan X. Hybrid Quantum-Classical Algorithms and Quantum Error Mitigation. J. Phys. Soc. Jpn. 2021, 90, 032001. 10.7566/jpsj.90.032001. [DOI] [Google Scholar]
- Qin D.; Xu X.; Li Y. An overview of quantum error mitigation formulas. Chin. Phys. B 2022, 31, 090306. 10.1088/1674-1056/ac7b1e. [DOI] [Google Scholar]
- Huang H.-L.; Xu X.-Y.; Guo C.; Tian G.; Wei S.-J.; Sun X.; Bao W.-S.; Long G.-L. Near-term quantum computing techniques: Variational quantum algorithms, error mitigation, circuit compilation, benchmarking and classical simulation. Sci. China Phys. Mech. Astron. 2023, 66, 250302. 10.1007/s11433-022-2057-y. [DOI] [Google Scholar]
- Peruzzo A.; McClean J.; Shadbolt P.; Yung M.-H.; Zhou X.-Q.; Love P. J.; Aspuru-Guzik A.; O’Brien J. L. A variational eigenvalue solver on a photonic quantum processor. Nat. Commun. 2014, 5, 4213. 10.1038/ncomms5213. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tilly J.; Chen H.; Cao S.; Picozzi D.; Setia K.; Li Y.; Grant E.; Wossnig L.; Rungger I.; Booth G. H.; Tennyson J. The Variational Quantum Eigensolver: A review of methods and best practices. Phys. Rep. 2022, 986, 1–128. 10.1016/j.physrep.2022.08.003. [DOI] [Google Scholar]
- Fedorov D. A.; Peng B.; Govind N.; Alexeev Y. VQE method: a short survey and recent developments. Mater. Theory 2022, 6, 2. 10.1186/s41313-021-00032-6. [DOI] [Google Scholar]
- Stair N. H.; Evangelista F. A. Simulating Many-Body Systems with a Projective Quantum Eigensolver. PRX Quantum 2021, 2, 030301. 10.1103/prxquantum.2.030301. [DOI] [Google Scholar]
- Huggins W. J.; O’Gorman B. A.; Rubin N. C.; Reichman D. R.; Babbush R.; Lee J. Unbiasing fermionic quantum Monte Carlo with a quantum computer. Nature 2022, 603, 416–420. 10.1038/s41586-021-04351-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang Y.; Dolde F.; Biamonte J.; Babbush R.; Bergholm V.; Yang S.; Jakobi I.; Neumann P.; Aspuru-Guzik A.; Whitfield J. D.; Wrachtrup J. Quantum Simulation of Helium Hydride Cation in a Solid-State Spin Register. ACS Nano 2015, 9, 7769–7774. 10.1021/acsnano.5b01651. [DOI] [PubMed] [Google Scholar]
- Li J.; Gao X.; Sajjan M.; Su J.-H.; Li Z.-K.; Kais S. Møller-Plesset Perturbation Theory Calculations on Quantum Devices. arXiv 2023, arXiv:2308.01559.[quant-ph], submitted 08-03-2023 10.48550/arXiv.2308.01559. [DOI] [Google Scholar]
- Chee C. H.; Mak A. M.; Leykam D.; Kl Barkoutsos P.; Angelakis D. G. Computing Electronic Correlation Energies using Linear Depth Quantum Circuits. Quantum Sci. Technol. 2024, 9, 025003. 10.1088/2058-9565/ad1f3a. [DOI] [Google Scholar]
- Boyn J.-N.; Lykhin A. O.; Smart S. E.; Gagliardi L.; Mazziotti D. A. Quantum-classical hybrid algorithm for the simulation of all-electron correlation. J. Chem. Phys. 2021, 155, 244106. 10.1063/5.0074842. [DOI] [PubMed] [Google Scholar]
- Smart S. E.; Boyn J.-N.; Mazziotti D. A. Resolving correlated states of benzyne with an error-mitigated contracted quantum eigensolver. Phys. Rev. A 2022, 105, 022405. 10.1103/physreva.105.022405. [DOI] [Google Scholar]
- Smart S. E.; Mazziotti D. A. Quantum Solver of Contracted Eigenvalue Equations for Scalable Molecular Simulations on Quantum Computing Devices. Phys. Rev. Lett. 2021, 126, 070504. 10.1103/physrevlett.126.070504. [DOI] [PubMed] [Google Scholar]
- Smart S. E.; Mazziotti D. A. Verifiably Exact Solution of the Electronic Schrödinger Equation on Quantum Devices. arXiv 2023, arXiv:2303.00758.[quant-ph], submitted 03-01-2023 10.48550/arXiv.2303.00758. [DOI] [Google Scholar]
- Kumar V.; Baskaran N.; Prasannaa V. S.; Sugisaki K.; Mukherjee D.; Dyall K. G.; Das B. P. Accurate Computation of Relativistic Excitation Energies Using Quantum Annealing. arXiv 2023, arXiv:2212.01801.[quant-ph], submitted 05-22-2023 10.48550/arXiv.2212.01801. [DOI] [Google Scholar]
- Imoto T.; Susa Y.; Miyazaki R.; Kadowaki T.; Matsuzaki Y. Demonstration of the excited-state search on the D-wave quantum annealer. arXiv 2023, arXiv:2305.15974.[quant-ph], submitted 05-25-2023 10.48550/arXiv.2305.15974. [DOI] [Google Scholar]
- Teplukhin A.; Kendrick B. K.; Tretiak S.; Dub P. A. Electronic structure with direct diagonalization on a D-wave quantum annealer. Sci. Rep. 2020, 10, 20753. 10.1038/s41598-020-77315-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Teplukhin A.; Kendrick B. K.; Mniszewski S. M.; Zhang Y.; Kumar A.; Negre C. F. A.; Anisimov P. M.; Tretiak S.; Dub P. A. Computing molecular excited states on a D-Wave quantum annealer. Sci. Rep. 2021, 11, 18796. 10.1038/s41598-021-98331-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Teplukhin A.; Kendrick B. K.; Babikov D. Calculation of Molecular Vibrational Spectra on a Quantum Annealer. J. Chem. Theory Comput. 2019, 15, 4555–4563. 10.1021/acs.jctc.9b00402. [DOI] [PubMed] [Google Scholar]
- Genin S. N.; Ryabinkin I. G.; Izmaylov A. F. Quantum chemistry on quantum annealers. arXiv 2019, arXiv:1901.04715.[quant-ph], submitted 01-15-2019 10.48550/arXiv.1901.04715. [DOI] [Google Scholar]
- Tazhigulov R. N.; Sun S.-N.; Haghshenas R.; Zhai H.; Tan A. T.; Rubin N. C.; Babbush R.; Minnich A. J.; Chan G. K.-L. Simulating Models of Challenging Correlated Molecules and Materials on the Sycamore Quantum Processor. PRX Quantum 2022, 3, 040318. 10.1103/prxquantum.3.040318. [DOI] [Google Scholar]
- Sun S.-N.; Motta M.; Tazhigulov R. N.; Tan A. T.; Chan G. K.-L.; Minnich A. J. Quantum Computation of Finite-Temperature Static and Dynamical Properties of Spin Systems Using Quantum Imaginary Time Evolution. PRX Quantum 2021, 2, 010317. 10.1103/prxquantum.2.010317. [DOI] [Google Scholar]
- Motta M.; Sun C.; Tan A. T. K.; O’Rourke M. J.; Ye E.; Minnich A. J.; Brandão F. G. S. L.; Chan G. K.-L. Determining eigenstates and thermal states on a quantum computer using quantum imaginary time evolution. Nat. Phys. 2019, 16, 205–210. 10.1038/s41567-019-0704-4. [DOI] [Google Scholar]
- Sokolov I. O.; Dobrautz W.; Luo H.; Alavi A.; Tavernelli I. Orders of magnitude increased accuracy for quantum many-body problems on quantum computers via an exact transcorrelated method. Phys. Rev. Res. 2023, 5, 023174. 10.1103/physrevresearch.5.023174. [DOI] [Google Scholar]
- Lin S.-H.; Dilip R.; Green A. G.; Smith A.; Pollmann F. Real- and Imaginary-Time Evolution with Compressed Quantum Circuits. PRX Quantum 2021, 2, 010342. 10.1103/prxquantum.2.010342. [DOI] [Google Scholar]
- Yeter-Aydeniz K.; Pooser R. C.; Siopsis G. Practical quantum computation of chemical and nuclear energy levels using quantum imaginary time evolution and Lanczos algorithms. NPJ Quantum Inf. 2020, 6, 63. 10.1038/s41534-020-00290-1. [DOI] [Google Scholar]
- Barison S.; Galli D. E.; Motta M. Quantum simulations of molecular systems with intrinsic atomic orbitals. Phys. Rev. A 2022, 106, 022404. 10.1103/physreva.106.022404. [DOI] [Google Scholar]
- Bassman Oftelie L.; Klymko K.; Liu D.; Tubman N. M.; de Jong W. A. Computing Free Energies with Fluctuation Relations on Quantum Computers. Phys. Rev. Lett. 2022, 129, 130603. 10.1103/physrevlett.129.130603. [DOI] [PubMed] [Google Scholar]
- Grimsley H. R.; Economou S. E.; Barnes E.; Mayhall N. J. An adaptive variational algorithm for exact molecular simulations on a quantum computer. Nat. Commun. 2019, 10, 3007. 10.1038/s41467-019-10988-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ryabinkin I. G.; Lang R. A.; Genin S. N.; Izmaylov A. F. Iterative Qubit Coupled Cluster Approach with Efficient Screening of Generators. J. Chem. Theory Comput. 2020, 16, 1055–1063. 10.1021/acs.jctc.9b01084. [DOI] [PubMed] [Google Scholar]
- Yordanov Y. S.; Armaos V.; Barnes C. H. W.; Arvidsson-Shukur D. R. M. Qubit-excitation-based adaptive variational quantum eigensolver. Commun. Phys. 2021, 4, 228. 10.1038/s42005-021-00730-0. [DOI] [Google Scholar]
- Tang H. L.; Shkolnikov V.; Barron G. S.; Grimsley H. R.; Mayhall N. J.; Barnes E.; Economou S. E. Qubit-ADAPT-VQE: An Adaptive Algorithm for Constructing Hardware-Efficient Ansätze on a Quantum Processor. PRX Quantum 2021, 2, 020310. 10.1103/prxquantum.2.020310. [DOI] [Google Scholar]
- Magoulas I.; Evangelista F. A. CNOT-Efficient Circuits for Arbitrary Rank Many-Body Fermionic and Qubit Excitations. J. Chem. Theory Comput. 2023, 19, 822–836. 10.1021/acs.jctc.2c01016. [DOI] [PubMed] [Google Scholar]
- Magoulas I.; Evangelista F. A. Linear-Scaling Quantum Circuits for Computational Chemistry. J. Chem. Theory Comput. 2023, 19, 4815–4821. 10.1021/acs.jctc.3c00376. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Magoulas I.; Evangelista F. A. Unitary Coupled Cluster: Seizing the Quantum Moment. J. Phys. Chem. A 2023, 127, 6567–6576. 10.1021/acs.jpca.3c02781. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Filip M.-A. Fighting noise with noise: a stochastic projective quantum eigensolver. arXiv 2023, arXiv:2306.14540.[quant-ph], submitted 08-04-2023 10.48550/arXiv.2306.14540. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Halder S.; Patra C.; Mondal D.; Maitra R. Machine learning aided dimensionality reduction toward a resource efficient projective quantum eigensolver: Formal development and pilot applications. J. Chem. Phys. 2023, 158, 244101. 10.1063/5.0155009. [DOI] [PubMed] [Google Scholar]
- Halder S.; Shrikhande C.; Maitra R. Development of zero-noise extrapolated projective quantum algorithm for accurate evaluation of molecular energetics in noisy quantum devices. J. Chem. Phys. 2023, 159, 114115. 10.1063/5.0166433. [DOI] [PubMed] [Google Scholar]
- Russo V.; Mari A.; Shammah N.; LaRose R.; Zeng W. J. Testing Platform-Independent Quantum Error Mitigation on Noisy Quantum Computers. IEEE Trans. Quantum Eng. 2023, 4, 1–18. 10.1109/tqe.2023.3305232. [DOI] [Google Scholar]
- Anselme Martin B.; Simon P.; Rančić M. J. Simulating strongly interacting Hubbard chains with the variational Hamiltonian ansatz on a quantum computer. Phys. Rev. Res. 2022, 4, 023190. 10.1103/physrevresearch.4.023190. [DOI] [Google Scholar]
- Kumar S.; Singh R. P.; Behera B. K.; Panigrahi P. K. Quantum simulation of negative hydrogen ion using variational quantum eigensolver on IBM quantum computer. arXiv 2019, arXiv:1903.03454.[quant-ph], submitted 05-30-2019 10.48550/arXiv.1903.03454. [DOI] [Google Scholar]
- Arute F.; Arya K.; Babbush R.; Bacon D.; Bardin J. C.; Barends R.; Boixo S.; Broughton M.; Buckley B. B.; Buell D. A.; Burkett B.; Bushnell N.; Chen Y.; Chen Z.; Chiaro B.; Collins R.; Courtney W.; Demura S.; Dunsworth A.; Farhi E.; Fowler A.; Foxen B.; Gidney C.; Giustina M.; Graff R.; Habegger S.; Harrigan M. P.; Ho A.; Hong S.; Huang T.; Huggins W. J.; Ioffe L.; Isakov S. V.; Jeffrey E.; Jiang Z.; Jones C.; Kafri D.; Kechedzhi K.; Kelly J.; Kim S.; Klimov P. V.; Korotkov A.; Kostritsa F.; Landhuis D.; Laptev P.; Lindmark M.; Lucero E.; Martin O.; Martinis J. M.; McClean J. R.; McEwen M.; et al. Hartree-Fock on a superconducting qubit quantum computer. Science 2020, 369, 1084–1089. 10.1126/science.abb9811. [DOI] [PubMed] [Google Scholar]
- Huang K.; Cai X.; Li H.; Ge Z.-Y.; Hou R.; Li H.; Liu T.; Shi Y.; Chen C.; Zheng D.; Xu K.; Liu Z.-B.; Li Z.; Fan H.; Fang W.-H. Variational Quantum Computation of Molecular Linear Response Properties on a Superconducting Quantum Processor. J. Phys. Chem. Lett. 2022, 13, 9114–9121. 10.1021/acs.jpclett.2c02381. [DOI] [PubMed] [Google Scholar]
- Huang B.; Sheng N.; Govoni M.; Galli G. Quantum Simulations of Fermionic Hamiltonians with Efficient Encoding and Ansatz Schemes. J. Chem. Theory Comput. 2023, 19, 1487–1498. 10.1021/acs.jctc.2c01119. [DOI] [PubMed] [Google Scholar]
- Zhang S.; Hill C. D.; Usman M. Comparative analysis of error mitigation techniques for variational quantum eigensolver implementations on IBM quantum system. arXiv 2022, arXiv:2206.07907.[quant-ph], submitted 06-16-2022 10.48550/arXiv.2206.07907. [DOI] [Google Scholar]
- Huang B.; Govoni M.; Galli G. Simulating the Electronic Structure of Spin Defects on Quantum Computers. PRX Quantum 2022, 3, 010339. 10.1103/prxquantum.3.010339. [DOI] [Google Scholar]
- Kandala A.; Temme K.; Córcoles A. D.; Mezzacapo A.; Chow J. M.; Gambetta J. M. Error mitigation extends the computational reach of a noisy quantum processor. Nature 2019, 567, 491–495. 10.1038/s41586-019-1040-7. [DOI] [PubMed] [Google Scholar]
- Yu H.; Zhao Y.; Wei T.-C. Simulating large-size quantum spin chains on cloud-based superconducting quantum computers. Phys. Rev. Res. 2023, 5, 013183. 10.1103/physrevresearch.5.013183. [DOI] [Google Scholar]
- Gao Q.; Nakamura H.; Gujarati T. P.; Jones G. O.; Rice J. E.; Wood S. P.; Pistoia M.; Garcia J. M.; Yamamoto N. Computational Investigations of the Lithium Superoxide Dimer Rearrangement on Noisy Quantum Devices. J. Phys. Chem. A 2021, 125, 1827–1836. 10.1021/acs.jpca.0c09530. [DOI] [PubMed] [Google Scholar]
- Lolur P.; Skogh M.; Dobrautz W.; Warren C.; Biznárová J.; Osman A.; Tancredi G.; Wendin G.; Bylander J.; Rahm M. Reference-State Error Mitigation: A Strategy for High Accuracy Quantum Computation of Chemistry. J. Chem. Theory Comput. 2023, 19, 783–789. 10.1021/acs.jctc.2c00807. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Suchsland P.; Barkoutsos P. K.; Tavernelli I.; Fischer M. H.; Neupert T. Simulating a ring-like Hubbard system with a quantum computer. Phys. Rev. Res. 2022, 4, 013165. 10.1103/physrevresearch.4.013165. [DOI] [Google Scholar]
- Sennane W.; Piquemal J.-P.; Rančić M. J. Calculating the ground-state energy of benzene under spatial deformations with noisy quantum computing. Phys. Rev. A 2023, 107, 012416. 10.1103/physreva.107.012416. [DOI] [Google Scholar]
- Rice J. E.; Gujarati T. P.; Motta M.; Takeshita T. Y.; Lee E.; Latone J. A.; Garcia J. M. Quantum computation of dominant products in lithium–sulfur batteries. J. Chem. Phys. 2021, 154, 134115. 10.1063/5.0044068. [DOI] [PubMed] [Google Scholar]
- Kandala A.; Mezzacapo A.; Temme K.; Takita M.; Brink M.; Chow J. M.; Gambetta J. M. Hardware-efficient variational quantum eigensolver for small molecules and quantum magnets. Nature 2017, 549, 242–246. 10.1038/nature23879. [DOI] [PubMed] [Google Scholar]
- Gao Q.; Jones G. O.; Motta M.; Sugawara M.; Watanabe H. C.; Kobayashi T.; Watanabe E.; Ohnishi Y. y.; Nakamura H.; Yamamoto N. Applications of quantum computing for investigations of electronic transitions in phenylsulfonyl-carbazole TADF emitters. NPJ Comput. Mater. 2021, 7, 70. 10.1038/s41524-021-00540-6. [DOI] [Google Scholar]
- Colless J. I.; Ramasesh V. V.; Dahlen D.; Blok M. S.; Kimchi-Schwartz M. E.; McClean J. R.; Carter J.; de Jong W. A.; Siddiqi I. Computation of Molecular Spectra on a Quantum Processor with an Error-Resilient Algorithm. Phys. Rev. X 2018, 8, 011021. 10.1103/physrevx.8.011021. [DOI] [Google Scholar]
- Hempel C.; Maier C.; Romero J.; McClean J.; Monz T.; Shen H.; Jurcevic P.; Lanyon B. P.; Love P.; Babbush R.; Aspuru-Guzik A.; Blatt R.; Roos C. F. Quantum Chemistry Calculations on a Trapped-Ion Quantum Simulator. Phys. Rev. X 2018, 8, 031022. 10.1103/physrevx.8.031022. [DOI] [Google Scholar]
- O’Malley P. J. J.; Babbush R.; Kivlichan I. D.; Romero J.; McClean J. R.; Barends R.; Kelly J.; Roushan P.; Tranter A.; Ding N.; Campbell B.; Chen Y.; Chen Z.; Chiaro B.; Dunsworth A.; Fowler A. G.; Jeffrey E.; Lucero E.; Megrant A.; Mutus J. Y.; Neeley M.; Neill C.; Quintana C.; Sank D.; Vainsencher A.; Wenner J.; White T. C.; Coveney P. V.; Love P. J.; Neven H.; Aspuru-Guzik A.; Martinis J. M. Scalable Quantum Simulation of Molecular Energies. Phys. Rev. X 2016, 6, 031007. 10.1103/physrevx.6.031007. [DOI] [Google Scholar]
- Egger D. J.; Capecci C.; Pokharel B.; Barkoutsos P. K.; Fischer L. E.; Guidoni L.; Tavernelli I. Pulse variational quantum eigensolver on cross-resonance-based hardware. Phys. Rev. Res. 2023, 5, 033159. 10.1103/physrevresearch.5.033159. [DOI] [Google Scholar]
- Crippa L.; Tacchino F.; Chizzini M.; Aita A.; Grossi M.; Chiesa A.; Santini P.; Tavernelli I.; Carretta S. Simulating Static and Dynamic Properties of Magnetic Molecules with Prototype Quantum Computers. Magnetochemistry 2021, 7, 117. 10.3390/magnetochemistry7080117. [DOI] [Google Scholar]
- O’Brien T. E.; Anselmetti G.; Gkritsis F.; Elfving V. E.; Polla S.; Huggins W. J.; Oumarou O.; Kechedzhi K.; Abanin D.; Acharya R.; Aleiner I.; Allen R.; Andersen T. I.; Anderson K.; Ansmann M.; Arute F.; Arya K.; Asfaw A.; Atalaya J.; Bardin J. C.; Bengtsson A.; Bortoli G.; Bourassa A.; Bovaird J.; Brill L.; Broughton M.; Buckley B.; Buell D. A.; Burger T.; Burkett B.; Bushnell N.; Campero J.; Chen Z.; Chiaro B.; Chik D.; Cogan J.; Collins R.; Conner P.; Courtney W.; Crook A. L.; Curtin B.; Debroy D. M.; Demura S.; Drozdov I.; Dunsworth A.; Erickson C.; Faoro L.; Farhi E.; Fatemi R.; Ferreira V. S.; Flores Burgos L.; et al. Purification-based quantum error mitigation of pair-correlated electron simulations. Nat. Phys. 2023, 19, 1787–1792. 10.1038/s41567-023-02240-y. [DOI] [Google Scholar]
- Tilly J.; Jones G.; Chen H.; Wossnig L.; Grant E. Computation of molecular excited states on IBM quantum computers using a discriminative variational quantum eigensolver. Phys. Rev. A 2020, 102, 062425. 10.1103/physreva.102.062425. [DOI] [Google Scholar]
- Yamamoto K.; Manrique D. Z.; Khan I. T.; Sawada H.; Ramo D. M. n. Quantum hardware calculations of periodic systems with partition-measurement symmetry verification: Simplified models of hydrogen chain and iron crystals. Phys. Rev. Res. 2022, 4, 033110. 10.1103/physrevresearch.4.033110. [DOI] [Google Scholar]
- Zhang Y.; Cincio L.; Negre C. F. A.; Czarnik P.; Coles P. J.; Anisimov P. M.; Mniszewski S. M.; Tretiak S.; Dub P. A. Variational quantum eigensolver with reduced circuit complexity. NPJ Quantum Inf. 2022, 8, 96. 10.1038/s41534-022-00599-z. [DOI] [Google Scholar]
- McCaskey A. J.; Parks Z. P.; Jakowski J.; Moore S. V.; Morris T. D.; Humble T. S.; Pooser R. C. Quantum chemistry as a benchmark for near-term quantum computers. NPJ Quantum Inf. 2019, 5, 99. 10.1038/s41534-019-0209-0. [DOI] [Google Scholar]
- Nam Y.; Chen J.-S.; Pisenti N. C.; Wright K.; Delaney C.; Maslov D.; Brown K. R.; Allen S.; Amini J. M.; Apisdorf J.; Beck K. M.; Blinov A.; Chaplin V.; Chmielewski M.; Collins C.; Debnath S.; Hudek K. M.; Ducore A. M.; Keesan M.; Kreikemeier S. M.; Mizrahi J.; Solomon P.; Williams M.; Wong-Campos J. D.; Moehring D.; Monroe C.; Kim J. Ground-state energy estimation of the water molecule on a trapped-ion quantum computer. NPJ Quantum Inf. 2020, 6, 33. 10.1038/s41534-020-0259-3. [DOI] [Google Scholar]
- Sagastizabal R.; Bonet-Monroig X.; Singh M.; Rol M. A.; Bultink C. C.; Fu X.; Price C. H.; Ostroukh V. P.; Muthusubramanian N.; Bruno A.; Beekman M.; Haider N.; O’Brien T. E.; DiCarlo L. Experimental error mitigation via symmetry verification in a variational quantum eigensolver. Phys. Rev. A 2019, 100, 010302. 10.1103/physreva.100.010302. [DOI] [Google Scholar]
- Czarnik P.; McKerns M.; Sornborger A. T.; Cincio L. Improving the efficiency of learning-based error mitigation. arXiv 2022, arXiv:2204.07109.[quant-ph], submitted 04-14-2022 10.48550/arXiv.2204.07109. [DOI] [Google Scholar]
- Maupin O. G.; Burch A. D.; Yale C. G.; Ruzic B.; Russo A.; Lobser D. S.; Revelle M. C.; Chow M. N.; Clark S. M.; Landahl A. J.; Love P. J. Error mitigation, optimization, and extrapolation on a trapped ion testbed. arXiv 2023, arXiv:2307.07027.[quant-ph], submitted 11-03-2023 10.48550/arXiv.2307.07027. [DOI] [Google Scholar]
- Zhao L.; Goings J.; Shin K.; Kyoung W.; Fuks J. I.; Kevin Rhee J. K.; Rhee Y. M.; Wright K.; Nguyen J.; Kim J.; Johri S. Orbital-optimized pair-correlated electron simulations on trapped-ion quantum computers. NPJ Quantum Inf. 2023, 9, 60. 10.1038/s41534-023-00730-8. [DOI] [Google Scholar]
- Urbanek M.; Nachman B.; de Jong W. A. Error detection on quantum computers improving the accuracy of chemical calculations. Phys. Rev. A 2020, 102, 022427. 10.1103/physreva.102.022427. [DOI] [Google Scholar]
- Goings J.; Zhao L.; Jakowski J.; Morris T.; Pooser R. Molecular Symmetry in VQE: A Dual Approach for Trapped-Ion Simulations of Benzene. arXiv 2023, arXiv:2308.00667.[quant-ph], submitted 08-01-2023 10.48550/arXiv.2308.00667. [DOI] [Google Scholar]
- Ganzhorn M.; Egger D.; Barkoutsos P.; Ollitrault P.; Salis G.; Moll N.; Roth M.; Fuhrer A.; Mueller P.; Woerner S.; Tavernelli I.; Filipp S. Gate-Efficient Simulation of Molecular Eigenstates on a Quantum Computer. Phys. Rev. Appl. 2019, 11, 044092. 10.1103/physrevapplied.11.044092. [DOI] [Google Scholar]
- Huo M.; Li Y. Dual-state purification for practical quantum error mitigation. Phys. Rev. A 2022, 105, 022427. 10.1103/physreva.105.022427. [DOI] [Google Scholar]
- Stenger J. P. T.; Hellberg C. S.; Gunlycke D. Implementing Jastrow-Gutzwiller operators on a quantum computer using the cascaded variational quantum eigensolver algorithm. Phys. Rev. A 2023, 107, 062606. 10.1103/physreva.107.062606. [DOI] [Google Scholar]
- Shen Y.; Zhang X.; Zhang S.; Zhang J.-N.; Yung M.-H.; Kim K. Quantum implementation of the unitary coupled cluster for simulating molecular electronic structure. Phys. Rev. A 2017, 95, 020501. 10.1103/physreva.95.020501. [DOI] [Google Scholar]
- Feniou C.; Claudon B.; Hassan M.; Courtat A.; Adjoua O.; Maday Y.; Piquemal J.-P. Greedy Gradient-free Adaptive Variational Quantum Algorithms on a Noisy Intermediate Scale Quantum Computer. arXiv 2023, arXiv:2306.17159.[quant-ph], submitted 09-11-2023 10.48550/arXiv.2306.17159. [DOI] [Google Scholar]
- Mazzola G.; Ollitrault P. J.; Barkoutsos P. K.; Tavernelli I. Nonunitary Operations for Ground-State Calculations in Near-Term Quantum Computers. Phys. Rev. Lett. 2019, 123, 130501. 10.1103/physrevlett.123.130501. [DOI] [PubMed] [Google Scholar]
- Bentellis A.; Matic-Flierl A.; Mendl C. B.; Lorenz J. M.. Benchmarking the Variational Quantum Eigensolver using different quantum hardware. 2023 IEEE International Conference on Quantum Computing and Engineering (QCE), 2023.
- Khan I. T.; Tudorovskaya M.; Kirsopp J. J. M.; Muñoz Ramo D.; Warrier P.; Papanastasiou D. K.; Singh R. Chemically aware unitary coupled cluster with ab initio calculations on an ion trap quantum computer: A refrigerant chemicals’ application. J. Chem. Phys. 2023, 158, 214114. 10.1063/5.0144680. [DOI] [PubMed] [Google Scholar]
- Ollitrault P. J.; Kandala A.; Chen C.-F.; Barkoutsos P. K.; Mezzacapo A.; Pistoia M.; Sheldon S.; Woerner S.; Gambetta J. M.; Tavernelli I. Quantum equation of motion for computing molecular excitation energies on a noisy quantum processor. Phys. Rev. Res. 2020, 2, 043140. 10.1103/physrevresearch.2.043140. [DOI] [Google Scholar]
- Ma H.; Govoni M.; Galli G. Quantum simulations of materials on near-term quantum computers. NPJ Comput. Mater. 2020, 6, 85. 10.1038/s41524-020-00353-z. [DOI] [Google Scholar]
- Gocho S.; Nakamura H.; Kanno S.; Gao Q.; Kobayashi T.; Inagaki T.; Hatanaka M. Excited state calculations using variational quantum eigensolver with spin-restricted ansätze and automatically-adjusted constraints. NPJ Comput. Mater. 2023, 9, 13. 10.1038/s41524-023-00965-1. [DOI] [Google Scholar]
- Sokolov I. O.; Barkoutsos P. K.; Moeller L.; Suchsland P.; Mazzola G.; Tavernelli I. Microcanonical and finite-temperature ab initio molecular dynamics simulations on quantum computers. Phys. Rev. Res. 2021, 3, 013125. 10.1103/physrevresearch.3.013125. [DOI] [Google Scholar]
- Lötstedt E.; Yamanouchi K.; Tsuchiya T.; Tachikawa Y. Calculation of vibrational eigenenergies on a quantum computer: Application to the Fermi resonance in CO2. Phys. Rev. A 2021, 103, 062609. 10.1103/physreva.103.062609. [DOI] [Google Scholar]
- Lötstedt E.; Yamanouchi K.; Tachikawa Y. Evaluation of vibrational energies and wave functions of CO2 on a quantum computer. AVS Quantum Sci. 2022, 4, 036801. 10.1116/5.0091144. [DOI] [Google Scholar]
- Krompiec M.; Ramo D. M. Strongly Contracted N-Electron Valence State Perturbation Theory Using Reduced Density Matrices from a Quantum Computer. arXiv 2022, arXiv:2210.05702.[quant-ph], submitted 10-11-2023 10.48550/arXiv.2210.05702. [DOI] [Google Scholar]
- Kirsopp J. J. M.; Di Paola C.; Manrique D. Z.; Krompiec M.; Greene-Diniz G.; Guba W.; Meyder A.; Wolf D.; Strahm M.; Muñoz Ramo D. Quantum computational quantification of protein–ligand interactions. Int. J. Quantum Chem. 2022, 122, e26975 10.1002/qua.26975. [DOI] [Google Scholar]
- Tammaro A.; Galli D. E.; Rice J. E.; Motta M. N-Electron Valence Perturbation Theory with Reference Wave Functions from Quantum Computing: Application to the Relative Stability of Hydroxide Anion and Hydroxyl Radical. J. Phys. Chem. A 2023, 127, 817–827. 10.1021/acs.jpca.2c07653. [DOI] [PubMed] [Google Scholar]
- Rossmannek M.; Pavošević F.; Rubio A.; Tavernelli I. Quantum Embedding Method for the Simulation of Strongly Correlated Systems on Quantum Computers. J. Phys. Chem. Lett. 2023, 14, 3491–3497. 10.1021/acs.jpclett.3c00330. [DOI] [PubMed] [Google Scholar]
- Tilly J.; Sriluckshmy P. V.; Patel A.; Fontana E.; Rungger I.; Grant E.; Anderson R.; Tennyson J.; Booth G. H. Reduced density matrix sampling: Self-consistent embedding and multiscale electronic structure on current generation quantum computers. Phys. Rev. Res. 2021, 3, 033230. 10.1103/physrevresearch.3.033230. [DOI] [Google Scholar]
- Rungger I.; Fitzpatrick N.; Chen H.; Alderete C. H.; Apel H.; Cowtan A.; Patterson A.; Ramo D. M.; Zhu Y.; Nguyen N. H.; Grant E.; Chretien S.; Wossnig L.; Linke N. M.; Duncan R. Dynamical mean field theory algorithm and experiment on quantum computers. arXiv 2020, arXiv:1910.04735.[quant-ph], submitted 01-08-2020 10.48550/arXiv.1910.04735. [DOI] [Google Scholar]
- Backes S.; Murakami Y.; Sakai S.; Arita R. Dynamical mean-field theory for the Hubbard-Holstein model on a quantum device. Phys. Rev. B 2023, 107, 165155. 10.1103/physrevb.107.165155. [DOI] [Google Scholar]
- Kawashima Y.; Lloyd E.; Coons M. P.; Nam Y.; Matsuura S.; Garza A. J.; Johri S.; Huntington L.; Senicourt V.; Maksymov A. O.; Nguyen J. H. V.; Kim J.; Alidoust N.; Zaribafiyan A.; Yamazaki T. Optimizing electronic structure simulations on a trapped-ion quantum computer using problem decomposition. Commun. Phys. 2021, 4, 245. 10.1038/s42005-021-00751-9. [DOI] [Google Scholar]
- Kordzanganeh M.; Buchberger M.; Kyriacou B.; Povolotskii M.; Fischer W.; Kurkin A.; Somogyi W.; Sagingalieva A.; Pflitsch M.; Melnikov A. Benchmarking Simulated and Physical Quantum Processing Units Using Quantum and Hybrid Algorithms. Adv. Quantum Technol. 2023, 6, 2300043. 10.1002/qute.202300043. [DOI] [Google Scholar]
- Huang R.; Li C.; Evangelista F. A. Leveraging Small-Scale Quantum Computers with Unitarily Downfolded Hamiltonians. PRX Quantum 2023, 4, 020313. 10.1103/prxquantum.4.020313. [DOI] [Google Scholar]
- Maruyoshi K.; Okuda T.; Pedersen J. W.; Suzuki R.; Yamazaki M.; Yoshida Y. Conserved charges in the quantum simulation of integrable spin chains. J. Phys. A: Math. Theor. 2023, 56, 165301. 10.1088/1751-8121/acc369. [DOI] [Google Scholar]
- Chen W.; Zhang S.; Zhang J.; Su X.; Lu Y.; Zhang K.; Qiao M.; Li Y.; Zhang J.-N.; Kim K. Error-mitigated quantum simulation of interacting fermions with trapped ions. NPJ Quantum Inf. 2023, 9, 122. 10.1038/s41534-023-00784-8. [DOI] [Google Scholar]
- Sopena A.; Gordon M. H.; Sierra G.; López E. Simulating quench dynamics on a digital quantum computer with data-driven error mitigation. Quantum Sci. Technol. 2021, 6, 045003. 10.1088/2058-9565/ac0e7a. [DOI] [Google Scholar]
- Lee C.-K.; Zhong Lau J. W.; Shi L.; Kwek L. C. Simulating Energy Transfer in Molecular Systems with Digital Quantum Computers. J. Chem. Theory Comput. 2022, 18, 1347–1358. 10.1021/acs.jctc.1c01296. [DOI] [PubMed] [Google Scholar]
- Lee Y. Symmetric Trotterization in digital quantum simulation of quantum spin dynamics. J. Korean Phys. Soc. 2023, 82, 479–485. 10.1007/s40042-023-00722-z. [DOI] [Google Scholar]
- Arute F.; Arya K.; Babbush R.; Bacon D.; Bardin J. C.; Barends R.; Bengtsson A.; Boixo S.; Broughton M.; Buckley B. B.; Buell D. A.; Burkett B.; Bushnell N.; Chen Y.; Chen Z.; Chen Y.-A.; Chiaro B.; Collins R.; Cotton S. J.; Courtney W.; Demura S.; Derk A.; Dunsworth A.; Eppens D.; Eckl T.; Erickson C.; Farhi E.; Fowler A.; Foxen B.; Gidney C.; Giustina M.; Graff R.; Gross J. A.; Habegger S.; Harrigan M. P.; Ho A.; Hong S.; Huang T.; Huggins W.; Ioffe L. B.; Isakov S. V.; Jeffrey E.; Jiang Z.; Jones C.; Kafri D.; Kechedzhi K.; Kelly J.; Kim S.; Klimov P. V.; Korotkov A. N.; Kostritsa F.; et al. Observation of separated dynamics of charge and spin in the Fermi-Hubbard model. arXiv 2020, arXiv:2010.07965.[quant-ph], submitted 10-15-2020 10.48550/arXiv.2010.07965. [DOI] [Google Scholar]
- Zhukov A. A.; Remizov S. V.; Pogosov W. V.; Lozovik Y. E. Algorithmic simulation of far-from-equilibrium dynamics using quantum computer. Quantum Inf. Process. 2018, 17, 223. 10.1007/s11128-018-2002-y. [DOI] [Google Scholar]
- Babukhin D. V.; Zhukov A. A.; Pogosov W. V. Hybrid digital-analog simulation of many-body dynamics with superconducting qubits. Phys. Rev. A 2020, 101, 052337. 10.1103/physreva.101.052337. [DOI] [Google Scholar]
- Rost B.; Re L. D.; Earnest N.; Kemper A. F.; Jones B.; Freericks J. K. Demonstrating robust simulation of driven-dissipative problems on near-term quantum computers. arXiv 2021, arXiv:2108.01183.[quant-ph], submitted 08-02-2021 10.48550/arXiv.2108.01183. [DOI] [Google Scholar]
- Smith A.; Kim M. S.; Pollmann F.; Knolle J. Simulating quantum many-body dynamics on a current digital quantum computer. NPJ Quantum Inf. 2019, 5, 106. 10.1038/s41534-019-0217-0. [DOI] [Google Scholar]
- Barends R.; Lamata L.; Kelly J.; García-Álvarez L.; Fowler A. G.; Megrant A.; Jeffrey E.; White T. C.; Sank D.; Mutus J. Y.; Campbell B.; Chen Y.; Chen Z.; Chiaro B.; Dunsworth A.; Hoi I.-C.; Neill C.; O’Malley P. J. J.; Quintana C.; Roushan P.; Vainsencher A.; Wenner J.; Solano E.; Martinis J. M. Digital quantum simulation of fermionic models with a superconducting circuit. Nat. Commun. 2015, 6, 7654. 10.1038/ncomms8654. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gibbs J.; Gili K.; Holmes Z.; Commeau B.; Arrasmith A.; Cincio L.; Coles P. J.; Sornborger A. Long-time simulations for fixed input states on quantum hardware. NPJ Quantum Inf. 2022, 8, 135. 10.1038/s41534-022-00625-0. [DOI] [Google Scholar]
- Martin B. A.; Ayral T.; Jamet F.; Rančić M. J.; Simon P. Combining Matrix Product States and Noisy Quantum Computers for Quantum Simulation. arXiv 2023, arXiv:2305.19231.[quant-ph], submitted 01-08-2024 10.48550/arXiv.2305.19231. [DOI] [Google Scholar]
- Doronin S. I.; Fel’dman E. B.; Kuznetsova E. I.; Zenchuk A. I. Simulation of Multiple-Quantum NMR Dynamics of Spin Dimer on Quantum Computer. Appl. Magn. Reson. 2021, 53, 1121–1131. 10.1007/s00723-021-01435-x. [DOI] [Google Scholar]
- Bassman Oftelie L.; Liu K.; Krishnamoorthy A.; Linker T.; Geng Y.; Shebib D.; Fukushima S.; Shimojo F.; Kalia R. K.; Nakano A.; Vashishta P. Towards simulation of the dynamics of materials on quantum computers. Phys. Rev. B 2020, 101, 184305. 10.1103/physrevb.101.184305. [DOI] [Google Scholar]
- Tornow S.; Gehrke W.; Helmbrecht U. Non-equilibrium dynamics of a dissipative two-site Hubbard model simulated on IBM quantum computers. J. Phys. A: Math. Theor. 2022, 55, 245302. 10.1088/1751-8121/ac6bd0. [DOI] [Google Scholar]
- Lötstedt E.; Wang L.; Yoshida R.; Zhang Y.; Yamanouchi K. Error-mitigated quantum computing of Heisenberg spin chain dynamics. Phys. Scr. 2023, 98, 035111. 10.1088/1402-4896/acbcac. [DOI] [Google Scholar]
- Lamm H.; Lawrence S. Simulation of Nonequilibrium Dynamics on a Quantum Computer. Phys. Rev. Lett. 2018, 121, 170501. 10.1103/physrevlett.121.170501. [DOI] [PubMed] [Google Scholar]
- Urbanek M.; Nachman B.; Pascuzzi V. R.; He A.; Bauer C. W.; de Jong W. A. Mitigating Depolarizing Noise on Quantum Computers with Noise-Estimation Circuits. Phys. Rev. Lett. 2021, 127, 270502. 10.1103/physrevlett.127.270502. [DOI] [PubMed] [Google Scholar]
- Kim Y.; Eddins A.; Anand S.; Wei K. X.; van den Berg E.; Rosenblatt S.; Nayfeh H.; Wu Y.; Zaletel M.; Temme K.; Kandala A. Evidence for the utility of quantum computing before fault tolerance. Nature 2023, 618, 500–505. 10.1038/s41586-023-06096-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Guimarães J. D.; Lim J.; Vasilevskiy M. I.; Huelga S. F.; Plenio M. B. Noise-Assisted Digital Quantum Simulation of Open Systems Using Partial Probabilistic Error Cancellation. PRX Quantum 2023, 4, 040329. 10.1103/prxquantum.4.040329. [DOI] [Google Scholar]
- Guimarães J. D.; Vasilevskiy M. I.; Barbosa L. S. Efficient method to simulate non-perturbative dynamics of an open quantum system using a quantum computer. arXiv 2022, arXiv:2203.14653.[quant-ph], submitted 11-09-2022 10.48550/arXiv.2203.14653. [DOI] [Google Scholar]
- Peetz J.; Smart S. E.; Tserkis S.; Narang P. Simulation of Open Quantum Systems via Low-Depth Convex Unitary Evolutions. arXiv 2023, arXiv:2307.14325.[quant-ph], submitted 07-27-2023 10.48550/arXiv.2307.14325. [DOI] [Google Scholar]
- Head-Marsden K.; Krastanov S.; Mazziotti D. A.; Narang P. Capturing non-Markovian dynamics on near-term quantum computers. Phys. Rev. Res. 2021, 3, 013182. 10.1103/physrevresearch.3.013182. [DOI] [Google Scholar]
- Gaikwad A.; Arvind; Dorai K. Simulating open quantum dynamics on an NMR quantum processor using the Sz.-Nagy dilation algorithm. Phys. Rev. A 2022, 106, 022424. 10.1103/physreva.106.022424. [DOI] [Google Scholar]
- Hu Z.; Xia R.; Kais S. A quantum algorithm for evolving open quantum dynamics on quantum computing devices. Sci. Rep. 2020, 10, 3301. 10.1038/s41598-020-60321-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Han J.; Cai W.; Hu L.; Mu X.; Ma Y.; Xu Y.; Wang W.; Wang H.; Song Y. P.; Zou C.-L.; Sun L. Experimental Simulation of Open Quantum System Dynamics via Trotterization. Phys. Rev. Lett. 2021, 127, 020504. 10.1103/physrevlett.127.020504. [DOI] [PubMed] [Google Scholar]
- Kamakari H.; Sun S.-N.; Motta M.; Minnich A. J. Digital Quantum Simulation of Open Quantum Systems Using Quantum Imaginary–Time Evolution. PRX Quantum 2022, 3, 010320. 10.1103/prxquantum.3.010320. [DOI] [Google Scholar]
- Schlimgen A. W.; Head-Marsden K.; Sager L. M.; Narang P.; Mazziotti D. A. Quantum Simulation of Open Quantum Systems Using a Unitary Decomposition of Operators. Phys. Rev. Lett. 2021, 127, 270503. 10.1103/physrevlett.127.270503. [DOI] [PubMed] [Google Scholar]
- Richerme P.; Revelle M. C.; Yale C. G.; Lobser D.; Burch A. D.; Clark S. M.; Saha D.; Lopez-Ruiz M. A.; Dwivedi A.; Smith J. M.; Norrell S. A.; Sabry A.; Iyengar S. S. Quantum Computation of Hydrogen Bond Dynamics and Vibrational Spectra. J. Phys. Chem. Lett. 2023, 14, 7256–7263. 10.1021/acs.jpclett.3c01601. [DOI] [PubMed] [Google Scholar]
- Gaidai I.; Babikov D.; Teplukhin A.; Kendrick B. K.; Mniszewski S. M.; Zhang Y.; Tretiak S.; Dub P. A. Molecular dynamics on quantum annealers. Sci. Rep. 2022, 12, 16824. 10.1038/s41598-022-21163-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sun S.-N.; Marinelli B.; Koh J. M.; Kim Y.; Nguyen L. B.; Chen L.; Kreikebaum J. M.; Santiago D. I.; Siddiqi I.; Minnich A. J. Quantum Computation of Frequency-Domain Molecular Response Properties Using a Three-Qubit iToffoli Gate. arXiv 2023, arXiv:2302.04271.[quant-ph], submitted 02-08-2023 10.48550/arXiv.2302.04271. [DOI] [Google Scholar]
- Neill C.; McCourt T.; Mi X.; Jiang Z.; Niu M. Y.; Mruczkiewicz W.; Aleiner I.; Arute F.; Arya K.; Atalaya J.; Babbush R.; Bardin J. C.; Barends R.; Bengtsson A.; Bourassa A.; Broughton M.; Buckley B. B.; Buell D. A.; Burkett B.; Bushnell N.; Campero J.; Chen Z.; Chiaro B.; Collins R.; Courtney W.; Demura S.; Derk A. R.; Dunsworth A.; Eppens D.; Erickson C.; Farhi E.; Fowler A. G.; Foxen B.; Gidney C.; Giustina M.; Gross J. A.; Harrigan M. P.; Harrington S. D.; Hilton J.; Ho A.; Hong S.; Huang T.; Huggins W. J.; Isakov S. V.; Jacob-Mitos M.; Jeffrey E.; Jones C.; Kafri D.; Kechedzhi K.; Kelly J.; Kim S.; et al. Accurately computing the electronic properties of a quantum ring. Nature 2021, 594, 508–512. 10.1038/s41586-021-03576-2. [DOI] [PubMed] [Google Scholar]
- Zhu J.; Nakagawa Y. O.; Zhang Y.-S.; Li C.-F.; Guo G.-C. Calculating the Green’s function of two-site fermionic Hubbard model in a photonic system. New J. Phys. 2022, 24, 043030. 10.1088/1367-2630/ac61d1. [DOI] [Google Scholar]
- Motta M.; Jones G. O.; Rice J. E.; Gujarati T. P.; Sakuma R.; Liepuoniute I.; Garcia J. M.; Ohnishi Y. y. Quantum chemistry simulation of ground- and excited-state properties of the sulfonium cation on a superconducting quantum processor. Chem. Sci. 2023, 14, 2915–2927. 10.1039/d2sc06019a. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhu D.; Johri S.; Nguyen N. H.; Alderete C. H.; Landsman K. A.; Linke N. M.; Monroe C.; Matsuura A. Y. Probing many-body localization on a noisy quantum computer. Phys. Rev. A 2021, 103, 032606. 10.1103/physreva.103.032606. [DOI] [Google Scholar]
- Santagati R.; Wang J.; Gentile A. A.; Paesani S.; Wiebe N.; McClean J. R.; Morley-Short S.; Shadbolt P. J.; Bonneau D.; Silverstone J. W.; Tew D. P.; Zhou X.; O’Brien J. L.; Thompson M. G. Witnessing eigenstates for quantum simulation of Hamiltonian spectra. Sci. Adv. 2018, 4, eaap9646 10.1126/sciadv.aap9646. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Libbi F.; Rizzo J.; Tacchino F.; Marzari N.; Tavernelli I. Effective calculation of the Green’s function in the time domain on near-term quantum processors. Phys. Rev. Res. 2022, 4, 043038. 10.1103/physrevresearch.4.043038. [DOI] [Google Scholar]
- Cruz D.; Magano D. Superresolution of Green’s functions on noisy quantum computers. Phys. Rev. A 2023, 108, 012618. 10.1103/physreva.108.012618. [DOI] [Google Scholar]
- Li Z.; Liu X.; Wang H.; Ashhab S.; Cui J.; Chen H.; Peng X.; Du J. Quantum Simulation of Resonant Transitions for Solving the Eigenproblem of an Effective Water Hamiltonian. Phys. Rev. Lett. 2019, 122, 090504. 10.1103/physrevlett.122.090504. [DOI] [PubMed] [Google Scholar]
- O’Brien T. E.; Ioffe L. B.; Su Y.; Fushman D.; Neven H.; Babbush R.; Smelyanskiy V. Quantum Computation of Molecular Structure Using Data from Challenging-To-Classically-Simulate Nuclear Magnetic Resonance Experiments. PRX Quantum 2022, 3, 030345. 10.1103/prxquantum.3.030345. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yoshida R.; Lötstedt E.; Yamanouchi K. Quantum computing of Hückel molecular orbitals of π-electron systems. J. Chem. Phys. 2022, 156, 184117. 10.1063/5.0086489. [DOI] [PubMed] [Google Scholar]
- Smart S. E.; Mazziotti D. A. Quantum-classical hybrid algorithm using an error-mitigating N-representability condition to compute the Mott metal-insulator transition. Phys. Rev. A 2019, 100, 022517. 10.1103/physreva.100.022517. [DOI] [Google Scholar]
- McArdle S.; Yuan X.; Benjamin S. Error-Mitigated Digital Quantum Simulation. Phys. Rev. Lett. 2019, 122, 180501. 10.1103/physrevlett.122.180501. [DOI] [PubMed] [Google Scholar]
- Schuld M.; Bergholm V.; Gogolin C.; Izaac J.; Killoran N. Evaluating analytic gradients on quantum hardware. Phys. Rev. A 2019, 99, 032331. 10.1103/physreva.99.032331. [DOI] [Google Scholar]
- Vidal J. G.; Theis D. O. Calculus on parameterized quantum circuits. arXiv 2018, arXiv:1812.06323.[quant-ph], submitted 12-31-2018 10.48550/arXiv.1812.06323. [DOI] [Google Scholar]
- Kottmann J. S.; Anand A.; Aspuru-Guzik A. A feasible approach for automatically differentiable unitary coupled-cluster on quantum computers. Chem. Sci. 2021, 12, 3497–3508. 10.1039/d0sc06627c. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pulay P. Convergence acceleration of iterative sequences. the case of scf iteration. Chem. Phys. Lett. 1980, 73, 393–398. 10.1016/0009-2614(80)80396-4. [DOI] [Google Scholar]
- Anis M. S.; Abraham H.; AduOffei; Agarwal R.; Agliardi G.; Aharoni M.; Akhalwaya I. Y.; Aleksandrowicz G.; Alexander T.; Amy M.; Anagolum S.; Arbel E.; Asfaw A.; Athalye A.; Avkhadiev A.; Azaustre C.; Bhole P.; Banerjee A.; Banerjee S.; Bang W.; Bansal A.; Barkoutsos P.; Barnawal A.; Barron G.; Barron G. S.; Bello L.; Ben-Haim Y.; Bennett M. C.; Bevenius D.; Bhatnagar D.; Bhobe A.; Bianchini P.; Bishop L. S.; Blank C.; Bolos S.; Bopardikar S.; Bosch S.; Brandhofer S.; Brandon; Bravyi S.; Bronn N.; Bucher D.; Burov A.; Cabrera F.; Calpin P.; Capelluto L.; Carballo J.; Carrascal G.; Carriker A.; Carvalho I.. et al. Qiskit: An Open-source Framework for Quantum Computing, 2021.
- Bravyi S.; Gambetta J. M.; Mezzacapo A.; Temme K. Tapering off qubits to simulate fermionic Hamiltonians. arXiv 2017, arXiv:1701.08213.[quant-ph], submitted 01-27-2017 10.48550/arXiv.1701.08213. [DOI] [Google Scholar]
- Setia K.; Chen R.; Rice J. E.; Mezzacapo A.; Pistoia M.; Whitfield J. D. Reducing Qubit Requirements for Quantum Simulations Using Molecular Point Group Symmetries. J. Chem. Theory Comput. 2020, 16, 6091–6097. 10.1021/acs.jctc.0c00113. [DOI] [PubMed] [Google Scholar]
- Bonet-Monroig X.; Sagastizabal R.; Singh M.; O’Brien T. E. Low-cost error mitigation by symmetry verification. Phys. Rev. A 2018, 98, 062339. 10.1103/physreva.98.062339. [DOI] [Google Scholar]
- Gokhale P.; Angiuli O.; Ding Y.; Gui K.; Tomesh T.; Suchara M.; Martonosi M.; Chong F. T. Minimizing State Preparations in Variational Quantum Eigensolver by Partitioning into Commuting Families. arXiv 2019, arXiv:1907.13623.[quant-ph], submitted 07-31-2019 10.48550/arXiv.1907.13623. [DOI] [Google Scholar]
- Crawford O.; Straaten B. v.; Wang D.; Parks T.; Campbell E.; Brierley S. Efficient quantum measurement of Pauli operators in the presence of finite sampling error. Quantum 2021, 5, 385. 10.22331/q-2021-01-20-385. [DOI] [Google Scholar]
- Yen T.-C.; Verteletskyi V.; Izmaylov A. F. Measuring All Compatible Operators in One Series of Single-Qubit Measurements Using Unitary Transformations. J. Chem. Theory Comput. 2020, 16, 2400–2409. 10.1021/acs.jctc.0c00008. [DOI] [PubMed] [Google Scholar]
- Hamamura I.; Imamichi T. Efficient evaluation of quantum observables using entangled measurements. NPJ Quantum Inf. 2020, 6, 56. 10.1038/s41534-020-0284-2. [DOI] [Google Scholar]
- Li Y.; Benjamin S. C. Efficient Variational Quantum Simulator Incorporating Active Error Minimization. Phys. Rev. X 2017, 7, 021050. 10.1103/physrevx.7.021050. [DOI] [Google Scholar]
- Temme K.; Bravyi S.; Gambetta J. M. Error Mitigation for Short-Depth Quantum Circuits. Phys. Rev. Lett. 2017, 119, 180509. 10.1103/physrevlett.119.180509. [DOI] [PubMed] [Google Scholar]
- Endo S.; Benjamin S. C.; Li Y. Practical Quantum Error Mitigation for Near-Future Applications. Phys. Rev. X 2018, 8, 031027. 10.1103/physrevx.8.031027. [DOI] [Google Scholar]
- Feller D.; Peterson K. A.; Grant Hill J. On the effectiveness of CCSD(T) complete basis set extrapolations for atomization energies. J. Chem. Phys. 2011, 135, 044102. 10.1063/1.3613639. [DOI] [PubMed] [Google Scholar]
- Schultz K.; LaRose R.; Mari A.; Quiroz G.; Shammah N.; Clader B. D.; Zeng W. J. Impact of time-correlated noise on zero-noise extrapolation. Phys. Rev. A 2022, 106, 052406. 10.1103/physreva.106.052406. [DOI] [Google Scholar]
- Cross A. W.; Bishop L. S.; Smolin J. A.; Gambetta J. M. Open Quantum Assembly Language. arXiv 2017, arXiv:1707.03429.[quant-ph], submitted 07-13-2017 10.48550/arXiv.1707.03429. [DOI] [Google Scholar]
- Helgaker T.; Jørgensen P.; Olsen J.. Molecular Electronic-Structure Theory; John Wiley & Sons, Ltd, 2000. [Google Scholar]
- Smith D. G. A.; Burns L. A.; Simmonett A. C.; Parrish R. M.; Schieber M. C.; Galvelis R.; Kraus P.; Kruse H.; Di Remigio R.; Alenaizan A.; James A. M.; Lehtola S.; Misiewicz J. P.; Scheurer M.; Shaw R. A.; Schriber J. B.; Xie Y.; Glick Z. L.; Sirianni D. A.; O’Brien J. S.; Waldrop J. M.; Kumar A.; Hohenstein E. G.; Pritchard B. P.; Brooks B. R.; Schaefer H. F.; Sokolov A. Y.; Patkowski K.; DePrince A. E.; Bozkaya U.; King R. A.; Evangelista F. A.; Turney J. M.; Crawford T. D.; Sherrill C. D. PSI4 1.4: Open-source software for high-throughput quantum chemistry. J. Chem. Phys. 2020, 152, 184108. 10.1063/5.0006002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- McClean J. R.; Romero J.; Babbush R.; Aspuru-Guzik A. The theory of variational hybrid quantum-classical algorithms. New J. Phys. 2016, 18, 023023. 10.1088/1367-2630/18/2/023023. [DOI] [Google Scholar]
- Mizuta K.; Fujii M.; Fujii S.; Ichikawa K.; Imamura Y.; Okuno Y.; Nakagawa Y. O. Deep variational quantum eigensolver for excited states and its application to quantum chemistry calculation of periodic materials. Phys. Rev. Res. 2021, 3, 043121. 10.1103/physrevresearch.3.043121. [DOI] [Google Scholar]
- Jones T.; Endo S.; McArdle S.; Yuan X.; Benjamin S. C. Variational quantum algorithms for discovering Hamiltonian spectra. Phys. Rev. A 2019, 99, 062304. 10.1103/physreva.99.062304. [DOI] [Google Scholar]
- Parrish R. M.; Hohenstein E. G.; McMahon P. L.; Martínez T. J. Quantum Computation of Electronic Transitions Using a Variational Quantum Eigensolver. Phys. Rev. Lett. 2019, 122, 230401. 10.1103/physrevlett.122.230401. [DOI] [PubMed] [Google Scholar]
- Higgott O.; Wang D.; Brierley S. Variational Quantum Computation of Excited States. Quantum 2019, 3, 156. 10.22331/q-2019-07-01-156. [DOI] [Google Scholar]
- Tsuchimochi T.; Ryo Y.; Ten-no S. L.; Sasasako K. Improved Algorithms of Quantum Imaginary Time Evolution for Ground and Excited States of Molecular Systems. J. Chem. Theory Comput. 2023, 19, 503–513. 10.1021/acs.jctc.2c00906. [DOI] [PubMed] [Google Scholar]
- He Y.; Guo H. The boundary effects of transverse field Ising model. J. Stat. Mech. 2017, 2017, 093101. 10.1088/1742-5468/aa85b0. [DOI] [Google Scholar]
- Lieb E.; Schultz T.; Mattis D. Two soluble models of an antiferromagnetic chain. Ann. Phys. 1961, 16, 407–466. 10.1016/0003-4916(61)90115-4. [DOI] [Google Scholar]
- Herasymenko Y.; O’Brien T. A diagrammatic approach to variational quantum ansatz construction. Quantum 2021, 5, 596. 10.22331/q-2021-12-02-596. [DOI] [Google Scholar]
- Roggero A.; Gu C.; Baroni A.; Papenbrock T. Preparation of excited states for nuclear dynamics on a quantum computer. Phys. Rev. C 2020, 102, 064624. 10.1103/physrevc.102.064624. [DOI] [Google Scholar]
- McArdle S. Learning from Physics Experiments with Quantum Computers: Applications in Muon Spectroscopy. PRX Quantum 2021, 2, 020349. 10.1103/prxquantum.2.020349. [DOI] [Google Scholar]
- Giurgica-Tiron T.; Hindy Y.; LaRose R.; Mari A.; Zeng W. J.. Digital zero noise extrapolation for quantum error mitigation. 2020 IEEE International Conference on Quantum Computing and Engineering (QCE). 2020; p 306.
- Broyden C. G. The Convergence of a Class of Double-rank Minimization Algorithms: 2. The New Algorithm. IMA J. Appl. Math. 1970, 6, 222–231. 10.1093/imamat/6.3.222. [DOI] [Google Scholar]
- Fletcher R. A new approach to variable metric algorithms. Comput. J. 1970, 13, 317–322. 10.1093/comjnl/13.3.317. [DOI] [Google Scholar]
- Goldfarb D. A family of variable-metric methods derived by variational means. Math. Comput. 1970, 24, 23. 10.2307/2004873. [DOI] [Google Scholar]
- Shanno D. F. Conditioning of quasi-Newton methods for function minimization. Math. Comput. 1970, 24, 647. 10.2307/2004840. [DOI] [Google Scholar]