Ab initio extended Hubbard model of short polyenes for efficient quantum computing
Abstract
We propose introducing an extended Hubbard Hamiltonian derived via the ab initio downfolding method, which was originally formulated for periodic materials, towards efficient quantum computing of molecular electronic structure calculations. By utilizing this method, the first-principles Hamiltonian of chemical systems can be coarse-grained by eliminating the electronic degrees of freedom in higher energy space and reducing the number of terms of electron repulsion integral from to . Our approach is validated numerically on the vertical excitation energies and excitation characters of ethylene, butadiene, and hexatriene. The dynamical electron correlation is incorporated within the framework of the constrained random phase approximation in advance of quantum computations, and the constructed models capture the trend of experimental and high-level quantum chemical calculation results. As expected, the -norm of the fermion-to-qubit mapped model Hamiltonians is significantly lower than that of conventional ab initio Hamiltonians, suggesting improved scalability of quantum computing. Those numerical outcomes and the results of the simulation of excited-state sampling demonstrate that the ab initio extended Hubbard Hamiltonian may hold significant potential for quantum chemical calculations using quantum computers.
I Introduction
Quantum computers are expected to solve electronic structure problems of chemistry that are potentially valuable to humanity and beyond the reach of classical computers Motta and Rice (2022); Dalzell et al. (2023). Quantum phase estimation is a well-known algorithm that uses a quantum computer to estimate the eigenvalues of the chemistry Hamiltonians Abrams and Lloyd (1999); Aspuru-Guzik et al. (2005). Through estimating the quantum computational cost of the phase estimation algorithms, the potential applications of fault-tolerant quantum computers have been explored to address global challenges in chemistry, such as nitrogen fixation Reiher et al. (2017); Lee et al. (2021), carbon dioxide reduction catalysis von Burg et al. (2021), materials research for batteries Rubin et al. (2023), and drug discovery Blunt et al. (2022).
One of the traditional and most attractive themes in quantum chemistry is the computation of electronic excited states Serrano-Andrés and Merchán (2005). The evaluation of excited states is still challenging for classical computation, partly because these states are often described by a linear combination of a larger number of Slater determinants, in contrast to the ground state, which is often well-described by a single Slater determinant. This highlights the suitability of quantum computational approaches leveraging the superposition nature of a quantum state. Indeed, various quantum algorithms have recently been proposed for evaluating excited states of molecules Higgott et al. (2019); Jones et al. (2019); McClean et al. (2017); Nakanishi et al. (2019); Parrish et al. (2019); Ollitrault et al. (2020); Asthana et al. (2023); Yalouz et al. (2021); Omiya et al. (2022); Yoshikura et al. (2023); Yeter-Aydeniz et al. (2020); Tsuchimochi et al. (2023a, b); Bauman et al. (2021a); Stair et al. (2020); Cortes and Gray (2022).
In principle, quantitative quantum chemical calculations on a quantum computer require a large number of quantum bits (qubits). The increase in the number of qubits is directly related to the increase in the number of molecular orbitals: in Jordan–Wigner (JW) mapping, a typical fermion-to-qubit mapping, the number of qubits representing the mapped Hamiltonian is equal to the number of spin orbitals. The use of many orbitals can provide a quantitative description of the electron correlation, especially the dynamical electron correlation, but more qubits require more quantum computational cost Gonthier et al. (2022).
The complexity of the molecular electronic structure Hamiltonian, elaborately modeled using many orbitals, triggers the fatal problem of the required number of quantum gates to implement it becoming exceedingly large. The second-quantized electronic structure Hamiltonian of a chemical system has an electron repulsion integral tensor, where is the number of spin orbitals. It induces a single Trotter step of the time-evolution operator to become circuit depth with the naive implementation Wecker et al. (2014). Such an inherent complexity poses challenges for fault-tolerant quantum computations as well as quantum simulations using near-term quantum devices; inevitably, the Coulomb operator of electronic structure Hamiltonians is factorized or sparsified to reduce the non-Clifford gate counts Berry et al. (2019); von Burg et al. (2021); Lee et al. (2021); Motta et al. (2021); Matsuzawa and Kurashige (2020); Rubin et al. (2022); Rocca et al. (2024). Simplification of electronic structure Hamiltonians in chemistry is crucial to avoid the excessively complicated quantum circuit operations.
A prospective way to save the number of qubits is to construct an effective Hamiltonian of the active space, consisting of chemically essential orbitals, prior to quantum computation. Such approaches to effectively reducing the number of electronic degrees of freedom are known as ‘downfolding’ approaches, and several downfolding methods for quantum computation have been proposed recently Bauman et al. (2019); Metcalf et al. (2020); Bauman et al. (2021b); Huang et al. (2023); Bauman and Kowalski (2022); Le and Tran (2023); Motta et al. (2020); McArdle and Tew (2020); Vorwerk et al. (2022). Downfolding approaches are very powerful because they incorporate the dynamical electron correlation related to the huge exterior space into the effective Hamiltonian in a relatively small active space. However, the complexity of ab initio electronic Hamiltonians still remains in the previous studies.
Considering those above, it is desired to construct an effective subspace model that possesses sufficient capability for discussing the electronic properties of molecules while reducing the complexity of the electron-electron interaction operators. In the context of condensed matter physics, the ab initio downfolding method was developed to make a low-energy model for periodic materials Aryasetiawan et al. (2004); Nakamura et al. (2021). This method can parameterize an extended Hubbard model by incorporating the contributions from the higher energy degrees of freedom into the electronic interactions of the lower energy degrees of freedom near the Fermi level based on constrained random phase approximation (cRPA). The ab initio downfolding approach has been widely applied to strongly correlated electronic phenomena in materials, such as the superconductivity of iron-based materials Misawa and Imada (2014) and the quantum spin liquid behavior of a molecular solid Misawa et al. (2020); Yoshimi et al. (2021). Extending this approach to quantum computing, the quasi-one-dimensional CuBr2 material as an analog of cuprate was modeled to test Amsler et al. (2023). Nonetheless, there has been little quantitative discussion about the advantage of introducing such an ab initio extended Hubbard model in quantum computing.
In this paper, we propose to use an extended Hubbard model of an isolated chemical system based on the ab initio downfolding method for efficient quantum computing. We numerically test this approach by constructing models of polyenes of short conjugation length: ethylene, butadiene, and hexatriene. Such polyene molecules, especially their excited states, have been traditionally studied as benchmarks for quantum chemistry methods Serrano‐Andrés et al. (1993); Nakayama et al. (1998); Kurashige et al. (2004); Schreiber et al. (2008); Watson and Chan (2012); Daday et al. (2012); Chien et al. (2018); Manna et al. (2020) as well as spectral experiments Mulliken (1977); Doering and McDiarmid (1980); Gavin, Jr. et al. (1973); Flicker et al. (1977); Fujii et al. (1985). The -norm of our model Hamiltonians after the fermion-to-qubit mapping is analyzed to show that the low-dimensionality and sparsity of the effective Hamiltonian can contribute to efficient quantum computing. Finally, we simulate the sampling calculations of the excited states of our models using quantum circuits optimized via the variational quantum deflation (VQD) algorithm Higgott et al. (2019).
Moreover, we examine how the number of bands of the referential first-principles electronic structure calculation affects the quality of our models. The construction of ab initio downfolded models for isolated chemical systems is still largely unexplored, and it is vital to investigate how to construct a reasonable model. The constructed models are validated by comparing the excitation energies with experimental and several quantum chemical calculation results.
The rest of this paper is organized as follows. In Sec. II, we briefly review the extended Hubbard model Hamiltonian and its construction via the ab initio downfolding method. Next, we explain the proposed approach that leverages the model Hamiltonian for quantum computation and briefly summarize the relationship between the -norm of the fermion-to-qubit mapped Hamiltonian and quantum algorithmic scaling. The computational details are explained in Sec. III. In Section IV, we present a comprehensive analysis of our models. This includes examining the dependency of the model Hamiltonian on the number of bands, validating the models through excitation energy comparisons, -norm analysis, and discussing the implications of classical simulations for quantum computation. Finally, the conclusion of this study is given in Sec. V.
II Computational Methods
II.1 Ab initio extended Hubbard Hamiltonian for molecules
We briefly review the extended Hubbard Hamiltonian of the ab initio downfolding method Aryasetiawan et al. (2004); Nakamura et al. (2021). For explanation, the effective Hamiltonian for molecules is explicitly described, with minor modifications not including the summations of inter-unit cell interactions.
We employ an extended Hubbard Hamiltonian to describe isolated molecules, which is given by,
(1) |
Here, and consists of one-body and two-body operators, respectively. The one-body term is expressed as:
(2) |
where denote the annihilation (creation) operators of the -th orbital with spin . Here, is a matrix element of the effective one-electron operator defined by
(3) |
where is the -th Wannier orbital. This integral is executed over the crystal volume, and is the Kohn-Sham (KS) Hamiltonian.
The term is introduced to prevent double counting (DC) of two-electron integrals from the one-electron integral . It can be defined as
(4) |
following the approach in Ref. Kawamura et al. (2017). is a parameter ranging for tuning the correction. The DC of the electron-electron interactions arises from the exchange-correlation functional of KS–density functional theory (DFT), and it should be noted that it cannot be eliminated completely. The one-body density matrix can be defined in the KS orbital basis as:
(5) |
denotes a screened Coulomb integral represented as:
(6) |
In this formulation, the inverse operator of the bare Coulomb interaction is replaced with the frequency-dependent screened Coulomb interaction , where is frequency. The value can be derived from cRPA Aryasetiawan et al. (2004); Nakamura et al. (2021). For our purposes, we utilize the static limit of the frequency-dependent Coulomb integrals:
(7) |
For the two-body electron-electron interaction component, , we compare the following terms: and . These terms are expressed as follows:
(8) | ||||
(9) |
Here, is a screened exchange integral represented as:
(10) |
where we also utilize the static limit for exchange integrals and refer to it as . The term discards the exchange interactions of the term.
Note that the ab initio downfolding method has scarcely been applied to isolated chemical systems; very recently, Chang et al. reported to apply to vanadocene Chang et al. (2023). Other related studies include the assessment of screened Coulomb interaction based on random phase approximation (RPA) for NbxCo () clusters Peters et al. (2018) and benzene van Loon et al. (2021). Scott et al. recently developed the moment-constrained RPA to evaluate static effective interaction rather than relying on the static limit of cRPA Scott and Booth (2024).
II.2 Ab initio extended Hubbard model approach for quantum computation
Our approach utilizes the extended Hubbard model for an isolated molecule constructed via the ab initio downfolding method in quantum computation. The conceptual figure comparing our approach with the conventional one is shown in Figure 1.

In this approach, the simplified model Hamiltonian is expected to perform quantum computations with a relatively small and shallow quantum circuit. The two-electron operator of the extended Hubbard model Hamiltonian consists of the second-order tensors and , whose indices belong to a smaller but physically essential space. It becomes a much sparse and compact representation compared to the ab initio Hamiltonian, the fourth-order tensor , whose indices belong to the entire space of orbitals.
It should be noted that our research direction is closely related to that of dynamical self-energy mapping (DSEM) Dhawan et al. (2021); Daniel et al. (2021). The DSEM procedure parametrizes a sparse Hamiltonian to reproduce the dynamical self-energy of the original molecular Hamiltonian. The sparse Hamiltonian contains at most two-body interaction terms, and it makes quantum circuits shallower and increases the feasibility of quantum computation for molecular systems.
II.3 -norm of fermion-to-qubit mapped Hamiltonian and computational costs in quantum computing
A second quantized Hamiltonian can be transformed by fermion-to-qubit mappings, such as Jordan-Wigner mapping, into the linear combination of the Pauli operators as
(11) |
where is the -th Pauli operator and is its coefficient. The summation runs over the number of the Pauli operators, denoted as .
-norm of the coefficient vector of the fermion-to-qubit mapped Hamiltonian is defined as the sum of the absolute values of the coefficients
(12) |
Given the crucial role that the parameter of a given Hamiltonian plays in determining the scalability of various quantum algorithms Lee et al. (2021); Koridon et al. (2021), evaluating the value of is essential for demonstrating the feasibility of quantum computing. For example, the scaling of the phase estimation algorithms based on the qubitization method using the single factorization and the tensor hypercontraction of the Coulomb operator is Berry et al. (2019) and Lee et al. (2021), respectively. Here, is the number of spin orbitals, and is the target precision. The qDRIFT algorithm Campbell (2019), a randomized compiler for Hamiltonian simulation, has the scaling.
Measurement is one of the most troublesome processes on variational quantum eigensolver (VQE) Peruzzo et al. (2014); Tilly et al. (2022), and the number of measurements of VQE also depends on the -norm of the Hamiltonian Wecker et al. (2015); Rubin et al. (2018). The total number of measurements is given by the sum of each number of measurements for each Pauli operator of the Hamiltonian as
(13) |
The optimal is proportional to , which is derived through the method of Lagrange multipliers Rubin et al. (2018). The optimal number of measurements is represented as
(14) |
where is the intrinsic standard deviation of the Pauli operator . Here, is the sampling error for the expectation value of the Hamiltonian. The right inequality of Eq. (14) is derived from the condition of the intrinsic variance . Hence, the total number of measurements of VQE is bounded using the -norm .
III Computational details
Here, we explain the computational details in a form that aids understanding of our calculation process.
In preparation for model construction, the first-principles electronic structure calculations were performed by Quantum ESPRESSO version 7.2 Giannozzi et al. (2009, 2017, 2020); PP . We employed a simple cubic lattice with a lattice constant , and set the plane-wave cutoff of the wave function . We used the optimized norm-conserving Vanderbilt pseudopotential Hamann (2013). For our usage, calculations were confined only to the point. The number of bands computed in the band structure calculations is an essential parameter involving the following computation of the screened Coulomb and exchange integrals, and a sufficiently large number of is necessary. In this study, we construct models varying and confirm that the sufficiently large value of . The details of increasing and the maximum number of bands for each molecule are discussed in Sec. IV.2. The molecular structure of the polyenes were all-trans isomers and optimized by DFT calculations at the B3LYP functional Becke (1993); Lee et al. (1988) and 6-31G basis-set level using Gaussian 16 software Revision C Frisch et al. (2016).
The extended Hubbard models of the polyene molecules were constructed with RESPACK-20200113 Nakamura et al. (2021); Fujiwara et al. (2003); Nohara et al. (2009); Nakamura et al. (2008, 2009, 2016). The maximally localized Wannier functions and the effective one-electron integrals were obtained to reproduce the target band energies of the -orbitals perpendicular to the carbon plane. The screened Coulomb and exchange integrals are evaluated via cRPA. The plane-wave cutoff of the polarization function was set to one-tenth of . The parameter was set to 1. We used VESTA version 3 Momma and Izumi (2011) for visualization of the Wannier functions.
The constructed models are handled using PySCF version 2.4.0 Sun et al. (2020, 2018) and OpenFermion version 1.6.0 McClean et al. (2020) or their earlier versions. OpenFermion-PySCF Sun et al. (2020, 2018); McClean et al. (2020) was also used. Jordan-Wigner mapping was employed for the fermion-to-qubit mapping, and the fermion-to-qubit mapped Hamiltonians were diagonalized to obtain their eigenstates if otherwise specified. When characterizing the eigenstates, the basis of the Hamiltonian was rotated by self-consistent field (SCF) calculation using PySCF, discussed further in Appendix A. Quantum chemical calculations for comparison were performed using the following program packages: The program version 1.5.11 Folkestad et al. (2020) was used to perform CC3 Paul et al. (2021) calculations. The OpenMolcas program version 22.06 Li Manni et al. (2023); Aquilante et al. (2020); Fdez. Galván et al. (2019) was used to perform CASCI and CASPT2 calculations. The Gaussian 16 software Revision C Frisch et al. (2016) was used to perform time-dependent density functional theory (TD-DFT) calculations, and the functional was B3LYP. Edmiston-Ruedenberg (ER) localization was performed using fcidump_rotation.f90 in the NECI program package Guther et al. (2020).
VQD calculations and the following sampling estimations were classically simulated by employing Qiskit version 0.43.2, Qiskit-aer version 0.12.1 Qiskit contributors (2023), and Qiskit-Nature version 0.6.2 The Qiskit Nature developers and contributors (2023). For the ansatz, we used a quantum circuit arranged with the particle-conserving A-gates Gard et al. (2020) in a brick-wall pattern.
IV Results and discussion
IV.1 Extended Hubbard model of short polyenes
The Wannier functions of our models are shown in Figure 2.

The modeled active space is represented as (e, o), where is the number of electrons and orbitals. The shape of these Wannier functions is reasonable because they have the -orbital character, which is a key to understanding the lower electron excitations of the polyenes. In the (4e, 4o) and (6e, 6o) cases of hexatriene, the shape of the Wannier functions is reasonable but different. The difference is that the former uses two fewer -type KS orbitals for localization.
In particular, the Wannier functions of ethylene (2e, 2o), butadiene (4e, 4o), and hexatriene (6e, 6o) have shapes reminiscent of a -orbital. This characteristic suggests that the model Hamiltonian with such a basis closely resembles the second-quantized Pariser-Parr-Pople (PPP) Hamiltonian Schulten and Karplus (1972); Tavan and Schulten (1986), which is also the extended Hubbard-type (analogous to the model with ) and composed of the set of the orthonormal -orbital basis.
Note that the PPP model is a semi-empirical molecular orbital method for -conjugated systems Pariser and Parr (1953a, b); Pople (1953). The PPP-multireference double excitation configuration interaction (PPP-MRD-CI) calculation, a configuration interaction (CI) calculation based on PPP, had achieved qualitative success for polyenes of relatively long conjugation length, even though the molecular integrals were given empirically. The PPP-MRD-CI calculations provided a qualitatively reasonable explanation for spectroscopic findings for the energetic order of the optically allowed single excitation and the forbidden double excitation Hudson and Kohler (1972, 1984), although the evaluated state order of butadiene is now known to be reversed Watson and Chan (2012).
The major difference between the second-quantized PPP and our extended Hubbard models is how the molecular integrals are determined. Whereas the former is given empirically, the latter is derived by the ab initio downfolding approach. The integral parameters of our models are summarized in Table 1.
Molecule | Active space | /eV | /eV | /eV | ||
---|---|---|---|---|---|---|
Ethylene | (2e, 2o) | 3.820 | 10.442 | 1.000 | ||
2.874 | 6.376 | 0.161 | 0.948 | |||
3.820 | 10.442 | 1.000 | ||||
Butadiene | (4e, 4o) | 3.663 | 8.298 | 0.965 | ||
2.423 | 5.651 | 0.240 | 0.459 | |||
2.749 | 5.817 | 0.211 | 0.862 | |||
0.129 | 4.440 | 0.066 | 0.056 | |||
3.663 | 8.298 | 0.965 | ||||
0.129 | 4.440 | 0.066 | 0.056 | |||
2.749 | 5.817 | 0.211 | 0.862 | |||
3.924 | 9.248 | 1.035 | ||||
0.282 | 3.610 | 0.023 | 0.387 | |||
3.924 | 9.248 | 1.035 | ||||
Hexatriene | (4e, 4o) | 3.707 | 6.977 | 1.011 | ||
2.354 | 5.466 | 0.115 | 0.887 | |||
1.887 | 4.738 | 0.074 | 0.388 | |||
0.067 | 3.889 | 0.173 | 0.055 | |||
3.583 | 8.281 | 0.989 | ||||
0.067 | 3.889 | 0.173 | 0.055 | |||
0.076 | 2.720 | 0.015 | 0.292 | |||
3.707 | 6.977 | 1.011 | ||||
2.354 | 5.466 | 0.115 | 0.887 | |||
3.583 | 8.281 | 0.989 |
Molecule | Active space | /eV | /eV | /eV | ||
---|---|---|---|---|---|---|
Hexatriene | (6e, 6o) | 3.535 | 7.658 | 0.944 | ||
2.340 | 5.528 | 0.216 | 0.473 | |||
2.726 | 5.310 | 0.312 | 0.754 | |||
0.046 | 4.267 | 0.059 | 0.026 | |||
0.322 | 4.314 | 0.077 | 0.082 | |||
0.270 | 3.410 | 0.026 | 0.292 | |||
4.001 | 8.766 | 1.081 | ||||
0.322 | 4.314 | 0.077 | 0.082 | |||
2.747 | 5.951 | 0.179 | 0.851 | |||
0.268 | 3.475 | 0.029 | 0.097 | |||
0.127 | 2.852 | 0.008 | 0.060 | |||
3.535 | 7.658 | 0.944 | ||||
0.270 | 3.410 | 0.026 | 0.292 | |||
2.340 | 5.528 | 0.216 | 0.473 | |||
0.046 | 4.267 | 0.059 | 0.026 | |||
3.746 | 9.281 | 0.975 | ||||
0.127 | 2.852 | 0.008 | 0.060 | |||
0.052 | 2.485 | 0.002 | 0.178 | |||
4.001 | 8.766 | 1.081 | ||||
2.747 | 5.951 | 0.179 | 0.851 | |||
3.746 | 9.281 | 0.975 |
These and parameters include the electron correlation effects from outside of the active space. In the next section, we examine the incorporation of the electron correlation effects by increasing the number of bands and extrapolating the continuum limit.
IV.2 Number of bands dependency
We investigate the excitation energies of the models varying the number of bands . The values of excitation energy are extrapolated to confirm the reliability of our models via estimating the continuum limit. For extrapolation, the following equation is used:
(15) |
where , , and are real fitting parameters. corresponds to the extrapolated value of the excitation energy. Fitting is performed based on the eigenvalues of the model Hamiltonian with obtained by H Kawamura et al. (2017); Ido et al. (2023). In this section, we focus on the Hamiltonian employing rather than both and . As shown in Table 1, the magnitude of is smaller than , and it is considered that the Hamiltonian with exhibits a similar trend to that with .
For further confirmation of the convergence behavior, we also vary the cutoff parameter set by comparing the condition with the condition, where the units of the lattice constant and cutoff parameters are Å and Ry, respectively. Due to our computational resources, it is difficult to increase the value of cutoff parameters while maintaining the lattice constant at 17.0 Å. However, it is important to note that this lattice constant is sufficiently large to treat the molecule as isolated, ensuring the relevance and accuracy of our results within this computational framework.
Figure 3 shows the dependency of our systems and the result of extrapolation. The state characterization is conducted in the way shown in Appendix A.

As a result, the excitation energy data varying on are successfully fitted using Eq. (15), as shown in Figure 3. Comparing the two conditions, the shapes, the intersection point, and the convergence behavior to the large limit of the curves fitted to the and states are similar. It suggests that our excitation energy results do not vary significantly with the setting of these hyperparameters.
It is found that there is a qualitative difference in the dependence among the molecules. In the ethylene (2e, 2o) case, as shown in Figure 3 (a), the excitation energies to the and states decrease in a similar way as increases. Conversely, in the other cases, the excitation energy to the state decreases, but that to the state increases as increases, as shown in Figure 3 (b)–(d). In the small region, lacking the dynamical correlation, the state is lower than the state. As increases, the order is reversed, and the excitation energies become saturated eventually. It is crucial to take a sufficiently large to reasonably discuss the energetic order of the and states and its gap width within the framework of cRPA.
In Table 2, we compare the extrapolated values with the values of the models, which employ and .
Molecule | Active space | State | |||
---|---|---|---|---|---|
Ethylene | (2e, 2o) | 572 | 7.81 | 7.76 | |
12.19 | 12.17 | ||||
Butadiene | (4e, 4o) | 1514 | 5.47 | 5.48 | |
5.64 | 5.65 | ||||
Hexatriene | (4e, 4o) | 1396 | 4.67 | 4.68 | |
5.05 | 5.08 | ||||
(6e, 6o) | 1396 | 4.32 | 4.31 | ||
4.65 | 4.63 |
These values are similar in each molecule, and it is considered that the models at the points are constructed by incorporating the electron correlation effects almost the same as the effects in the continuum limit. The values of and at the points are shown in Table 1.
IV.3 Vertical excitation energies
We compare the excitation energies of our models with the results of other quantum chemical calculations and experiments shown in Table 3.
Model or methods | Ethylene | Butadiene | Hexatriene | |||
---|---|---|---|---|---|---|
Model 1 | 7.81 | 12.19 | 5.47 | 5.64 | 4.67 (4.32) | 5.05 (4.65) |
Model 2 | 8.13 | 12.19 | 5.82 | 5.98 | 4.81 (4.65) | 5.12 (4.96) |
CC3/aug-cc-pVDZ | 7.95 | 6.25 | 6.68 | 5.35 | 5.73 | |
TD-DFT/aug-cc-pVDZ | 7.42 | 5.59 | 6.53 | 4.61 | 5.64 | |
CASCI/cc-pVDZ | 10.17 | 15.82 | 8.15 | 7.06 | 6.72 (6.93) | 6.88 (5.84) |
CASPT2/cc-pVDZ | 8.33 | 14.10 | 6.26 | 6.57 | 5.09 (5.76) | 4.74 (5.24) |
Experiment | 7.66a | 5.92b | 4.93c, 4.95d | 5.2e | ||
PPP-CIf | 5.83 | 5.34 | (5.05) | (4.36) | ||
SHCI with Extrp.g | 8.05 | 6.45 | 6.58 | (5.59) | (5.58) |
-
a
Band maximum from Ref. Mulliken (1977).
-
b
Spectral peak from Ref. Doering and McDiarmid (1980).
-
c
Band center from Ref. Gavin, Jr. et al. (1973).
-
d
A band peak reported in Ref. Flicker et al. (1977), while the intensity maximum is an adjacent peak at 5.13 eV.
- e
-
f
Ref. Tavan and Schulten (1986).
-
g
Ref. Chien et al. (2018). The excitation energy of ethylene was obtained without extrapolation.
Hereafter, the models that employ and together with the parameters in Table 1 are referred to as Model 1 and Model 2, respectively.
Overall, Model 2 practically yields a better result than Model 1. Model 1 tends to underestimate the excitation energies compared to Model 2, which is considered to be due to the exchange interactions in Model 1.
The excitation energies of Models 1 and 2 become smaller as the conjugation length becomes longer. This tendency is similar to the results of quantum chemical calculations and experiments. In particular, the excitation energies of Model 2 are in good agreement with the experimental values in either molecule. We employ the (4e, 4o) and (6e, 6o) active spaces for hexatriene, and the former model has the excitation energies closer to the experimental values. This suggests that a larger active space does not necessarily lead to a better result in the construction of the ab initio downfolding model.
The magnitude of the excitation energies to the lower excited states of polyenes has been an important research topic in traditional quantum chemistry Schulten and Karplus (1972); Tavan and Schulten (1986); Nakayama et al. (1998); Kurashige et al. (2004). In butadiene and hexatriene, it is known that the singly excited and doubly excited states are energetically close to each other. It can be said that the gap of our models is small — the excitation energy gaps of Model 2 of butadiene and hexatriene are 0.16 eV and 0.31 eV, respectively. These narrow gaps are attributed to the inclusion of dynamical electron correlation from outside of the active space. Comparison between CASCI and CASPT2 results also indicates that the dynamical electron correlation reduces the energy gap between and .
We compare the results of our models with those of CC3 and CASPT2 calculations. In our CC3 calculations, the excitation energies get close to the models’ and experimental results in each molecule. The states of CC3 are characterized using the oscillator strengths and the molecular orbital shapes related to the major single and double amplitudes. The excitation energy gaps of butadiene and hexatriene are 0.43 eV and 0.38 eV, comparable to those in Model 2. In our CASPT2 calculations, the excitation energies to the state capture the trend of the results of Model 2, experiment, and CC3. The excitation energy gaps of butadiene and hexatriene are 0.31 eV and 0.35 eV in CASPT2(4e, 4o) — The excitation energy to the state is lower than that of the state in hexatriene.
Highly precise quantum chemical calculations have provided insight into the energetic order of the and states. It has been known that the state is lower in butadiene Watson and Chan (2012); Chien et al. (2018). Our models of butadiene reproduce the order of the and states with a reasonably small energy gap.
In hexatriene, high-level quantum chemical calculations show that the state is much closer to in energy and slightly lower Nakayama et al. (1998); Kurashige et al. (2004). For instance, the result of the semi-stochastic heat-bath CI (SHCI) reported by Chien et al. Chien et al. (2018) is shown in Table 3, and the state is slightly lower than and accidentally degenerates to the state in hexatriene. The state order of our models is reversed to the SHCI result, yet the energy gap remains narrow. The order of the CC3 hexatriene result is also reversed in the order of these states. Note that the state order from experiments is similar to the results of our models and CC3. There seems to be room for a more detailed discussion about the energy gap in hexatriene. For example, the geometry optimization calculation of the ground state at a much higher level might refine our discussion.
We also compare the result of our models with that of the CI calculation using the PPP model, termed PPP-CI Tavan and Schulten (1986). The single excitation energy of PPP-CI for butadiene closely matches the experimental value, whereas the double excitation energy is smaller than the single excitation energy. On the other hand, our models reproduce the relationship between and in butadiene well. For hexatriene, PPP-CI qualitatively reproduced the order that the state is lower than the , but the gap seems large (0.69 eV). In our models, the state is lower than the state, albeit with a small gap.
Finally, we discuss the TD-DFT result. TD-DFT reproduces the excitation energy to the state well, but in the butadiene and hexatriene cases, the excitation energy to the state is about 1 eV larger than that to the state. This feature differs from our models and other established quantum chemical calculations. Conventional TD-DFT is known to encounter challenges in describing double electron excitations. While it can reasonably predict the single excitation to the state, TD-DFT struggles with representing the double excitation to the state, as highlighted in Ref. Cave et al. (2004).
IV.4 -norm values of Hamiltonians
Here, we analyze the extent to which the extended Hubbard model Hamiltonian can contribute to efficient quantum computation. The -norm and the number of terms of the fermion-to-qubit mapped Hamiltonians are shown in Table 4.
Molecule | Space | Hamiltonian | Basis | ||
Ethylene | (16e, 82o) | Full space | CMO | 7076 | |
(2e, 2o) | Active space | CMO | 0.8 | 15 | |
(2e, 2o) | Model 1 | Wannier | 1.3 | 19 | |
(2e, 2o) | Model 2 | Wannier | 1.3 | 15 | |
Butadiene | (30e, 146o) | Full space | CMO | 40392a | - |
(4e, 4o) | Active space | CMO | 3.1 | 185 | |
(4e, 4o) | Active space | ER | 3.6 | 361 | |
(4e, 4o) | Model 1 | Wannier | 3.3 | 85 | |
(4e, 4o) | Model 2 | Wannier | 3.2 | 61 | |
Hexatriene | (44e, 210o) | Full space | CMO | 101802a | - |
(4e, 4o) | Active space | CMO | 2.7 | 185 | |
(4e, 4o) | Active space | ER | 3.1 | 361 | |
(4e, 4o) | Model 1 | Wannier | 2.9 | 85 | |
(4e, 4o) | Model 2 | Wannier | 2.8 | 61 | |
(6e, 6o) | Active space | CMO | 7.1 | 919 | |
(6e, 6o) | Active space | ER | 7.0 | 1819 | |
(6e, 6o) | Model 1 | Wannier | 5.8 | 199 | |
(6e, 6o) | Model 2 | Wannier | 5.7 | 139 |
- a
Our model Hamiltonians have the -norm value comparable to the active space Hamiltonian in each molecule. Since Model 2 does not include the exchange interaction terms, the values of Model 2 are reduced compared to Model 1, yet their differences are negligible. On the contrary, the first-principles Hamiltonians of the full space have a huge -norm value. This contrast clearly shows that the quantum computing cost of our extended Hubbard Hamiltonian is similar to that of the active space Hamiltonian and significantly cheaper than that of the full space first-principles Hamiltonian. Recalling that our models include dynamical electron correlations from outside of the active space, as discussed in Section IV.3, they have a significant advantage in the efficiency of quantum computation compared to the active space Hamiltonians, which do not include such correlations.
Since the required active space is very small for these targeted molecules, we could not show numerically that the extended Hubbard models have the advantage regarding the -norm value against the corresponding active space Hamiltonian. The -norm values are, however, expected to decrease in much larger active space cases because the extended Hubbard and active space Hamiltonians have the square and quartic number of the electron-electron interaction terms, respectively. This effect may be seen in the hexatriene (6e, 6o) case and is expected to be more remarkable in a chemical system requiring a larger active space, e.g., larger -conjugated systems or transition metal complexes.
Table 4 also shows the results of ER localization of active space Hamiltonians in the butadiene and hexatriene cases. The maximally localized Wannier function is localized orbital rather than canonical molecular orbital (CMO). Compared to the canonical orbital cases, the -norm values in the ER-localized basis have slightly increased. Koridon et al. have previously reported that localized orbitals can decrease the -norm values Koridon et al. (2021), but our numerical investigation does not clearly show this trend. It can be just because the necessary active space for our system is very small.
Regarding , Model 2 is the smallest within the Hamiltonians of the same active space in either molecule. The effect of neglecting the exchange terms appears in comparing Model 1 and Model 2, even though our systems are small. The maximally localized Wannier functions do not reflect the high symmetry of the molecules. However, the extended Hubbard model Hamiltonians have a significant advantage in , whose second-quantized representation does not have the electron-electron interaction terms of the three and four kinds of distinct indices. It is expected to facilitate the execution of quantum computations.
On the contrary, in the cases of active space Hamiltonians in an ER-localized basis, the number of terms has almost doubled compared to that in a CMO basis. Since the all-trans polyenes are highly symmetrized, the fermion-to-qubit mapped active space Hamiltonian in a CMO basis is sparse, and the number of terms is reduced. Orbital localization leads to the lower symmetry of the orbitals, and it can be considered that the number of the terms increases.
The small number of terms in our models is considered to have advantages in quantum computations. For example, it can be robust against errors from the noises of the noisy intermediate-scale quantum devices Preskill (2018) because the number of measurements to evaluate the expectation values of Pauli operators is expected to become smaller. For quantum algorithms such as quantum phase estimation, the quantum circuits become smaller and shallower.
IV.5 Application: Estimation of excitation energies from VQD calculation and sampling
We further apply our extended Hubbard models to the classical simulation of quantum computations. We estimate the expectation values of the models and the ER-localized active space Hamiltonians by sampling. The quantum circuits were prepared for the ground and excited states using VQD calculations.
The VQD algorithm is an extension of VQE for calculating excited states, proposed by Higgott et al. Higgott et al. (2019). VQD’s cost function is defined as
(16) |
where is the -th target excited state that can be obtained by optimizing the quantum circuit parameters . The first term of the cost function corresponds to the expectation value of the energy of the -th excited state. The second term is the penalty term that is introduced as the sum of the overlap between the -th state and -th ansatz states (). The parameter is the weight factor of each overlap term.
Table LABEL:tab:smpl shows the estimated energy differences and the errors of the (4e, 4o) models and ER-localized active space Hamiltonians of butadiene and hexatriene.
Molecule | State | Model 1 | Model 2 | ER-localized active space Hamiltonian | ||||||
---|---|---|---|---|---|---|---|---|---|---|
Abelian grouping | No grouping | Exact | Abelian grouping | No grouping | Exact | Abelian grouping | No grouping | Exact | ||
Butadiene | S0 | 0.00 | 0.00 | 0.00 | ||||||
S1 | 5.47 | 5.82 | 6.96 | |||||||
S2 | 5.64 | 5.98 | 7.71 | |||||||
Hexatriene | S0 | 0.00 | 0.00 | 0.00 | ||||||
S1 | 4.67 | 4.81 | 6.08 | |||||||
S2 | 5.05 | 5.12 | 7.05 |
The energy difference is between the energy estimated by sampling and the exact ground-state energy. The energy is estimated by repeating the shots energy calculation times.
The estimated energy differences using our models are in good agreement with those of exact diagonalization. Those using the ER-localized active space Hamiltonians are similarly estimated well but overestimate the excitation energies due to the absence of electron correlation related to the outside of the active space. The errors are similar for both our models and ER-localized active space Hamiltonians. Hence, we validate the applicability of the ab initio downfolded model to compute excited states using quantum devices through numerical calculations using our models.
V Conclusions
We propose introducing extended Hubbard models for isolated molecules based on the ab initio downfolding method Aryasetiawan et al. (2004); Nakamura et al. (2021) toward efficient quantum computing. To demonstrate the effectiveness of this approach, we constructed models for short polyenes: ethylene, butadiene, and hexatriene. The constructed models are intended to describe the lower-lying excited states of these conjugated systems. They exhibit two key features: First, screened Coulomb and exchange integrals effectively incorporate the electron-correlation effect from outside of the active space through cRPA. Second, the extended Hubbard model has fewer terms than the usual active space Hamiltonian because the electron repulsion integral tensor is expressed as second-order rather than fourth-order. In summary, the proposed approach depicted in Figure 1 efficiently uses the ab initio downfolding method to construct an extended Hubbard model for a molecule, reducing the need for numerous qubits and quantum operations compared to the conventional approach.
It found that our models, particularly Model 2, demonstrate reasonable excitation energies compared to experimental values and high-level quantum chemical calculation results, offering practicality under limited quantum resources. The impact of introducing an approximation such as cRPA has been quantitatively analyzed: The incorporation of the dynamical electron correlation effect from outside of the active space is examined through the comparison with the continuum limit. Besides, one may say that this study has redetermined the parameters of the second-quantized PPP model Hamiltonian for short polyenes in an ab initio manner.
We also contribute to showing that our models maintain a small -norm and a reduced number of terms, making them advantageous in quantum computing. In simulations of sampling estimation, our models successfully estimate the excitation energies in molecules like butadiene and hexatriene, and they include doubly excited states, which are challenging for conventional TD-DFT. It suggests a promising direction for quantum chemical computation on quantum computers.
It is worthwhile to evaluate the performance of the ab initio extended Hubbard models for molecules with a real quantum computer. This approach can be beneficial for fault-tolerant quantum computers as well as near-term quantum devices with restricted numbers of reliable qubits and quantum gate operations Yoshioka et al. (2022); Ichikawa et al. (2023).
Updating the framework of ab initio downfolding also seems a meaningful research direction. For example, cRPA is known to overestimate the screening effect on two-body interactions Shinaoka et al. (2015); Honerkamp et al. (2018), and Scott et al. address this issue Scott and Booth (2024). Considering the future application to larger molecular systems, the fast evaluation of the effective two-body integrals, whose bottleneck is the parallelized calculation of the polarization function, as Nakamura et al. reported Nakamura et al. (2021), is also considered important.
Acknowledgements
We would like to thank Hideaki Hakoshima, Hiroshi Shinaoka, and Takahiro Misawa for fruitful discussion. This project has been supported by funding from the MEXT Quantum Leap Flagship Program (MEXTQLEAP) through Grant No. JPMXS0120319794, and the JST COI-NEXT Program through Grant No. JPMJPF2014. YY wishes to thank JSPS KAKENHI Grant No. JP21K20536. NT wishes to thank JSPS KAKENHI Grant Nos. JP19H05817, JP19H05820, JST PRESTO Grant No. JPMJPR23F6. WM wishes to thank JSPS KAKENHI Grant Nos. JP23H03819 and JP21K18933. We thank the Supercomputer Center, the Institute for Solid State Physics, the University of Tokyo for the use of the facilities. YY thank Research Center for Computational Science, Okazaki, Japan (Project: 23-IMS-C183).
References
- Motta and Rice (2022) M. Motta and J. E. Rice, WIREs Comput. Mol. Sci. 12, e1580 (2022).
- Dalzell et al. (2023) A. M. Dalzell, S. McArdle, M. Berta, P. Bienias, C.-F. Chen, A. Gilyén, C. T. Hann, M. J. Kastoryano, E. T. Khabiboulline, A. Kubica, G. Salton, S. Wang, and F. G. S. L. Brandão, arXiv (2023), arXiv:2310.03011 [quant-ph] .
- Abrams and Lloyd (1999) D. S. Abrams and S. Lloyd, Phys. Rev. Lett. 83, 5162 (1999).
- Aspuru-Guzik et al. (2005) A. Aspuru-Guzik, A. D. Dutoi, P. J. Love, and M. Head-Gordon, Science 309, 1704 (2005).
- Reiher et al. (2017) M. Reiher, N. Wiebe, K. M. Svore, D. Wecker, and M. Troyer, Proc. Natl. Acad. Sci. U.S.A. 114, 7555 (2017).
- Lee et al. (2021) J. Lee, D. W. Berry, C. Gidney, W. J. Huggins, J. R. McClean, N. Wiebe, and R. Babbush, PRX Quantum 2, 030305 (2021).
- von Burg et al. (2021) V. von Burg, G. H. Low, T. Häner, D. S. Steiger, M. Reiher, M. Roetteler, and M. Troyer, Phys. Rev. Res. 3, 033055 (2021).
- Rubin et al. (2023) N. C. Rubin, D. W. Berry, F. D. Malone, A. F. White, T. Khattar, A. E. DePrince, S. Sicolo, M. Küehn, M. Kaicher, J. Lee, and R. Babbush, PRX Quantum 4, 040303 (2023).
- Blunt et al. (2022) N. S. Blunt, J. Camps, O. Crawford, R. Izsák, S. Leontica, A. Mirani, A. E. Moylett, S. A. Scivier, C. Sünderhauf, P. Schopf, J. M. Taylor, and N. Holzmann, J. Chem. Theory Comput. 18, 7001 (2022).
- Serrano-Andrés and Merchán (2005) L. Serrano-Andrés and M. Merchán, J. Mol. Struct.: THEOCHEM 729, 99 (2005).
- Higgott et al. (2019) O. Higgott, D. Wang, and S. Brierley, Quantum 3, 156 (2019).
- Jones et al. (2019) T. Jones, S. Endo, S. McArdle, X. Yuan, and S. C. Benjamin, Phys. Rev. A 99, 062304 (2019).
- McClean et al. (2017) J. R. McClean, M. E. Kimchi-Schwartz, J. Carter, and W. A. de Jong, Phys. Rev. A 95, 042308 (2017).
- Nakanishi et al. (2019) K. M. Nakanishi, K. Mitarai, and K. Fujii, Phys. Rev. Res. 1, 033062 (2019).
- Parrish et al. (2019) R. M. Parrish, E. G. Hohenstein, P. L. McMahon, and T. J. Martínez, Phys. Rev. Lett. 122, 230401 (2019).
- Ollitrault et al. (2020) P. J. Ollitrault, A. Kandala, C.-F. Chen, P. K. Barkoutsos, A. Mezzacapo, M. Pistoia, S. Sheldon, S. Woerner, J. M. Gambetta, and I. Tavernelli, Phys. Rev. Res. 2, 043140 (2020).
- Asthana et al. (2023) A. Asthana, A. Kumar, V. Abraham, H. Grimsley, Y. Zhang, L. Cincio, S. Tretiak, P. A. Dub, S. E. Economou, E. Barnes, and N. J. Mayhall, Chem. Sci. 14, 2405 (2023).
- Yalouz et al. (2021) S. Yalouz, B. Senjean, J. Günther, F. Buda, T. E. O’Brien, and L. Visscher, Quantum Sci. Technol. 6, 024004 (2021).
- Omiya et al. (2022) K. Omiya, Y. O. Nakagawa, S. Koh, W. Mizukami, Q. Gao, and T. Kobayashi, J. Chem. Theory Comput. 18, 741 (2022).
- Yoshikura et al. (2023) T. Yoshikura, S. L. Ten-no, and T. Tsuchimochi, J. Phys. Chem. A 127, 6577 (2023).
- Yeter-Aydeniz et al. (2020) K. Yeter-Aydeniz, R. C. Pooser, and G. Siopsis, Npj Quantum Inf. 6, 63 (2020).
- Tsuchimochi et al. (2023a) T. Tsuchimochi, Y. Ryo, S. L. Ten-no, and K. Sasasako, J. Chem. Theory Comput. 19, 503 (2023a).
- Tsuchimochi et al. (2023b) T. Tsuchimochi, Y. Ryo, S. C. Tsang, and S. L. Ten-no, Npj Quantum Inf. 9, 113 (2023b).
- Bauman et al. (2021a) N. P. Bauman, H. Liu, E. J. Bylaska, S. Krishnamoorthy, G. H. Low, C. E. Granade, N. Wiebe, N. A. Baker, B. Peng, M. Roetteler, M. Troyer, and K. Kowalski, J. Chem. Theory Comput. 17, 201 (2021a).
- Stair et al. (2020) N. H. Stair, R. Huang, and F. A. Evangelista, J. Chem. Theory Comput. 16, 2236 (2020).
- Cortes and Gray (2022) C. L. Cortes and S. K. Gray, Phys. Rev. A 105, 022417 (2022).
- Gonthier et al. (2022) J. F. Gonthier, M. D. Radin, C. Buda, E. J. Doskocil, C. M. Abuan, and J. Romero, Phys. Rev. Res. 4, 033154 (2022).
- Wecker et al. (2014) D. Wecker, B. Bauer, B. K. Clark, M. B. Hastings, and M. Troyer, Phys. Rev. A 90, 022305 (2014).
- Berry et al. (2019) D. W. Berry, C. Gidney, M. Motta, J. R. McClean, and R. Babbush, Quantum 3, 208 (2019).
- Motta et al. (2021) M. Motta, E. Ye, J. R. McClean, Z. Li, A. J. Minnich, R. Babbush, and G. K.-L. Chan, npj Quantum Information 7, 83 (2021).
- Matsuzawa and Kurashige (2020) Y. Matsuzawa and Y. Kurashige, J. Chem. Theory Comput. 16, 944 (2020).
- Rubin et al. (2022) N. C. Rubin, J. Lee, and R. Babbush, J. Chem. Theory Comput. 18, 1480 (2022).
- Rocca et al. (2024) D. Rocca, C. L. Cortes, J. Gonthier, P. J. Ollitrault, R. M. Parrish, G.-L. Anselmetti, M. Degroote, N. Moll, R. Santagati, and M. Streif, arXiv (2024), arXiv:2403.03502 [quant-ph] .
- Bauman et al. (2019) N. P. Bauman, E. J. Bylaska, S. Krishnamoorthy, G. H. Low, N. Wiebe, C. E. Granade, M. Roetteler, M. Troyer, and K. Kowalski, J. Chem. Phys. 151, 014107 (2019).
- Metcalf et al. (2020) M. Metcalf, N. P. Bauman, K. Kowalski, and W. A. de Jong, J. Chem. Theory Comput. 16, 6165 (2020).
- Bauman et al. (2021b) N. P. Bauman, C. Jaroslav, V. Libor, P. Jiří, and K. Kowalski, Quantum Sci. Technol. 6, 034008 (2021b).
- Huang et al. (2023) R. Huang, C. Li, and F. A. Evangelista, PRX Quantum 4, 020313 (2023).
- Bauman and Kowalski (2022) N. P. Bauman and K. Kowalski, Mater. Theory 6, 17 (2022).
- Le and Tran (2023) N. T. Le and L. N. Tran, J. Phys. Chem. A 127, 5222 (2023).
- Motta et al. (2020) M. Motta, T. P. Gujarati, J. E. Rice, A. Kumar, C. Masteran, J. A. Latone, E. Lee, E. F. Valeev, and T. Y. Takeshita, Phys. Chem. Chem. Phys. 22, 24270 (2020).
- McArdle and Tew (2020) S. McArdle and D. P. Tew, arXiv (2020), arXiv:2006.11181 [quant-ph] .
- Vorwerk et al. (2022) C. Vorwerk, N. Sheng, M. Govoni, B. Huang, and G. Galli, Nat. Comput. Sci. 2, 424 (2022).
- Aryasetiawan et al. (2004) F. Aryasetiawan, M. Imada, A. Georges, G. Kotliar, S. Biermann, and A. I. Lichtenstein, Phys. Rev. B 70, 195104 (2004).
- Nakamura et al. (2021) K. Nakamura, Y. Yoshimoto, Y. Nomura, T. Tadano, M. Kawamura, T. Kosugi, K. Yoshimi, T. Misawa, and Y. Motoyama, Comput. Phys. Commun. 261, 107781 (2021).
- Misawa and Imada (2014) T. Misawa and M. Imada, Nat. Commun. 5, 5738 (2014).
- Misawa et al. (2020) T. Misawa, K. Yoshimi, and T. Tsumuraya, Phys. Rev. Res. 2, 032072 (2020).
- Yoshimi et al. (2021) K. Yoshimi, T. Tsumuraya, and T. Misawa, Phys. Rev. Res. 3, 043224 (2021).
- Amsler et al. (2023) M. Amsler, P. Deglmann, M. Degroote, M. P. Kaicher, M. Kiser, M. Kühn, C. Kumar, A. Maier, G. Samsonidze, A. Schroeder, M. Streif, D. Vodola, and C. Wever, arXiv (2023), arXiv:2301.11838 [quant-ph] .
- Serrano‐Andrés et al. (1993) L. Serrano‐Andrés, M. Merchán, I. Nebot‐Gil, R. Lindh, and B. O. Roos, J. Chem. Phys. 98, 3151 (1993).
- Nakayama et al. (1998) K. Nakayama, H. Nakano, and K. Hirao, Int. J. Quantum Chem. 66, 157 (1998).
- Kurashige et al. (2004) Y. Kurashige, H. Nakano, Y. Nakao, and K. Hirao, Chem. Phys. Lett. 400, 425 (2004).
- Schreiber et al. (2008) M. Schreiber, M. R. Silva-Junior, S. P. A. Sauer, and W. Thiel, J. Chem. Phys. 128, 134110 (2008).
- Watson and Chan (2012) M. A. Watson and G. K.-L. Chan, J. Chem. Theory Comput. 8, 4013 (2012).
- Daday et al. (2012) C. Daday, S. Smart, G. H. Booth, A. Alavi, and C. Filippi, J. Chem. Theory Comput. 8, 4441 (2012).
- Chien et al. (2018) A. D. Chien, A. A. Holmes, M. Otten, C. J. Umrigar, S. Sharma, and P. M. Zimmerman, J. Phys. Chem. A 122, 2714 (2018).
- Manna et al. (2020) S. Manna, R. K. Chaudhuri, and S. Chattopadhyay, J. Chem. Phys. 152, 244105 (2020).
- Mulliken (1977) R. S. Mulliken, J. Chem. Phys. 66, 2448 (1977).
- Doering and McDiarmid (1980) J. P. Doering and R. McDiarmid, The Journal of Chemical Physics 73, 3617 (1980).
- Gavin, Jr. et al. (1973) R. M. Gavin, Jr., S. Risemberg, and S. A. Rice, J. Chem. Phys. 58, 3160 (1973).
- Flicker et al. (1977) W. M. Flicker, O. A. Mosher, and A. Kuppermann, Chem. Phys. Lett. 45, 492 (1977).
- Fujii et al. (1985) T. Fujii, A. Kamata, M. Shimizu, Y. Adachi, and S. Maeda, Chem. Phys. Lett. 115, 369 (1985).
- Kawamura et al. (2017) M. Kawamura, K. Yoshimi, T. Misawa, Y. Yamaji, S. Todo, and N. Kawashima, Comput. Phys. Commun. 217, 180 (2017).
- Chang et al. (2023) Y. Chang, E. G. C. P. van Loon, B. Eskridge, B. Busemeyer, M. A. Morales, C. E. Dreyer, A. J. Millis, S. Zhang, T. O. Wehling, L. K. Wagner, and M. Rösner, arXiv (2023), arXiv:2311.05987 [cond-mat.str-el] .
- Peters et al. (2018) L. Peters, E. Şaşıoğlu, I. Mertig, and M. I. Katsnelson, Phys. Rev. B 97, 045121 (2018).
- van Loon et al. (2021) E. G. C. P. van Loon, M. Rösner, M. I. Katsnelson, and T. O. Wehling, Phys. Rev. B 104, 045134 (2021).
- Scott and Booth (2024) C. J. C. Scott and G. H. Booth, Phys. Rev. Lett. 132, 076401 (2024).
- Dhawan et al. (2021) D. Dhawan, M. Metcalf, and D. Zgid, J. Chem. Theory Comput. 17, 7622 (2021).
- Daniel et al. (2021) C. Daniel, D. Dhawan, D. Zgid, and J. K. Freericks, Eur. Phys. J. Spec. Top. 230, 1067 (2021).
- Koridon et al. (2021) E. Koridon, S. Yalouz, B. Senjean, F. Buda, T. E. O’Brien, and L. Visscher, Phys. Rev. Res. 3, 033127 (2021).
- Campbell (2019) E. Campbell, Phys. Rev. Lett. 123, 070503 (2019).
- Peruzzo et al. (2014) A. Peruzzo, J. McClean, P. Shadbolt, M.-H. Yung, X.-Q. Zhou, P. J. Love, A. Aspuru-Guzik, and J. L. O’Brien, Nat. Commun. 5, 4213 (2014).
- Tilly et al. (2022) J. Tilly, H. Chen, S. Cao, D. Picozzi, K. Setia, Y. Li, E. Grant, L. Wossnig, I. Rungger, G. H. Booth, and J. Tennyson, Phys. Rep. 986, 1 (2022).
- Wecker et al. (2015) D. Wecker, M. B. Hastings, and M. Troyer, Phys. Rev. A 92, 042303 (2015).
- Rubin et al. (2018) N. C. Rubin, R. Babbush, and J. McClean, New J. Phys. 20, 053020 (2018).
- Giannozzi et al. (2009) P. Giannozzi, S. Baroni, N. Bonini, M. Calandra, R. Car, C. Cavazzoni, D. Ceresoli, G. L. Chiarotti, M. Cococcioni, I. Dabo, A. D. Corso, S. de Gironcoli, S. Fabris, G. Fratesi, R. Gebauer, U. Gerstmann, C. Gougoussis, A. Kokalj, M. Lazzeri, L. Martin-Samos, N. Marzari, F. Mauri, R. Mazzarello, S. Paolini, A. Pasquarello, L. Paulatto, C. Sbraccia, S. Scandolo, G. Sclauzero, A. P. Seitsonen, A. Smogunov, P. Umari, and R. M. Wentzcovitch, J. Phys.: Condens. Matter 21, 395502 (2009).
- Giannozzi et al. (2017) P. Giannozzi, O. Andreussi, T. Brumme, O. Bunau, M. B. Nardelli, M. Calandra, R. Car, C. Cavazzoni, D. Ceresoli, M. Cococcioni, N. Colonna, I. Carnimeo, A. D. Corso, S. de Gironcoli, P. Delugas, R. A. DiStasio, A. Ferretti, A. Floris, G. Fratesi, G. Fugallo, R. Gebauer, U. Gerstmann, F. Giustino, T. Gorni, J. Jia, M. Kawamura, H.-Y. Ko, A. Kokalj, E. Küçükbenli, M. Lazzeri, M. Marsili, N. Marzari, F. Mauri, N. L. Nguyen, H.-V. Nguyen, A. O. de-la Roza, L. Paulatto, S. Poncé, D. Rocca, R. Sabatini, B. Santra, M. Schlipf, A. P. Seitsonen, A. Smogunov, I. Timrov, T. Thonhauser, P. Umari, N. Vast, X. Wu, and S. Baroni, J. Phys.: Condens. Matter 29, 465901 (2017).
- Giannozzi et al. (2020) P. Giannozzi, O. Baseggio, P. Bonfà, D. Brunato, R. Car, I. Carnimeo, C. Cavazzoni, S. de Gironcoli, P. Delugas, F. Ferrari Ruffino, A. Ferretti, N. Marzari, I. Timrov, A. Urru, and S. Baroni, J. Chem. Phys. 152, 154105 (2020).
- (78) http://www.quantum-espresso.org.
- Hamann (2013) D. R. Hamann, Phys. Rev. B 88, 085117 (2013).
- Becke (1993) A. D. Becke, J. Chem. Phys. 98, 5648 (1993).
- Lee et al. (1988) C. Lee, W. Yang, and R. G. Parr, Phys. Rev. B 37, 785 (1988).
- Frisch et al. (2016) M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, G. A. Petersson, H. Nakatsuji, X. Li, M. Caricato, A. V. Marenich, J. Bloino, B. G. Janesko, R. Gomperts, B. Mennucci, H. P. Hratchian, J. V. Ortiz, A. F. Izmaylov, J. L. Sonnenberg, D. Williams-Young, F. Ding, F. Lipparini, F. Egidi, J. Goings, B. Peng, A. Petrone, T. Henderson, D. Ranasinghe, V. G. Zakrzewski, J. Gao, N. Rega, G. Zheng, W. Liang, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, K. Throssell, J. A. Montgomery, Jr., J. E. Peralta, F. Ogliaro, M. J. Bearpark, J. J. Heyd, E. N. Brothers, K. N. Kudin, V. N. Staroverov, T. A. Keith, R. Kobayashi, J. Normand, K. Raghavachari, A. P. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, J. M. Millam, M. Klene, C. Adamo, R. Cammi, J. W. Ochterski, R. L. Martin, K. Morokuma, O. Farkas, J. B. Foresman, and D. J. Fox, “Gaussian 16 Revision C.01,” (2016), Gaussian Inc. Wallingford CT.
- Fujiwara et al. (2003) T. Fujiwara, S. Yamamoto, and Y. Ishii, J. Phys. Soc. Jpn. 72, 777 (2003).
- Nohara et al. (2009) Y. Nohara, S. Yamamoto, and T. Fujiwara, Phys. Rev. B 79, 195110 (2009).
- Nakamura et al. (2008) K. Nakamura, R. Arita, and M. Imada, J. Phys. Soc. Jpn. 77, 093711 (2008).
- Nakamura et al. (2009) K. Nakamura, Y. Yoshimoto, T. Kosugi, R. Arita, and M. Imada, J. Phys. Soc. Jpn. 78, 083710 (2009).
- Nakamura et al. (2016) K. Nakamura, Y. Nohara, Y. Yosimoto, and Y. Nomura, Phys. Rev. B 93, 085124 (2016).
- Momma and Izumi (2011) K. Momma and F. Izumi, J. Appl. Cryst. 44, 1272 (2011).
- Sun et al. (2020) Q. Sun, X. Zhang, S. Banerjee, P. Bao, M. Barbry, N. S. Blunt, N. A. Bogdanov, G. H. Booth, J. Chen, Z.-H. Cui, J. J. Eriksen, Y. Gao, S. Guo, J. Hermann, M. R. Hermes, K. Koh, P. Koval, S. Lehtola, Z. Li, J. Liu, N. Mardirossian, J. D. McClain, M. Motta, B. Mussard, H. Q. Pham, A. Pulkin, W. Purwanto, P. J. Robinson, E. Ronca, E. R. Sayfutyarova, M. Scheurer, H. F. Schurkus, J. E. T. Smith, C. Sun, S.-N. Sun, S. Upadhyay, L. K. Wagner, X. Wang, A. White, J. D. Whitfield, M. J. Williamson, S. Wouters, J. Yang, J. M. Yu, T. Zhu, T. C. Berkelbach, S. Sharma, A. Y. Sokolov, and G. K.-L. Chan, J. Chem. Phys. 153, 024109 (2020).
- Sun et al. (2018) Q. Sun, T. C. Berkelbach, N. S. Blunt, G. H. Booth, S. Guo, Z. Li, J. Liu, J. D. McClain, E. R. Sayfutyarova, S. Sharma, S. Wouters, and G. K.-L. Chan, WIREs Comput. Mol. Sci. 8, e1340 (2018).
- McClean et al. (2020) J. R. McClean, N. C. Rubin, K. J. Sung, I. D. Kivlichan, X. Bonet-Monroig, Y. Cao, C. Dai, E. S. Fried, C. Gidney, B. Gimby, P. Gokhale, T. Häner, T. Hardikar, V. Havlíček, O. Higgott, C. Huang, J. Izaac, Z. Jiang, X. Liu, S. McArdle, M. Neeley, T. O’Brien, B. O’Gorman, I. Ozfidan, M. D. Radin, J. Romero, N. P. D. Sawaya, B. Senjean, K. Setia, S. Sim, D. S. Steiger, M. Steudtner, Q. Sun, W. Sun, D. Wang, F. Zhang, and R. Babbush, Quantum Sci. Technol. 5, 034014 (2020).
- Folkestad et al. (2020) S. D. Folkestad, E. F. Kjønstad, R. H. Myhre, J. H. Andersen, A. Balbi, S. Coriani, T. Giovannini, L. Goletto, T. S. Haugland, A. Hutcheson, I.-M. Høyvik, T. Moitra, A. C. Paul, M. Scavino, A. S. Skeidsvoll, Å. H. Tveten, and H. Koch, J. Chem. Phys. 152, 184103 (2020).
- Paul et al. (2021) A. C. Paul, R. H. Myhre, and H. Koch, J. Chem. Theory Comput. 17, 117 (2021).
- Li Manni et al. (2023) G. Li Manni, I. Fdez. Galván, A. Alavi, F. Aleotti, F. Aquilante, J. Autschbach, D. Avagliano, A. Baiardi, J. J. Bao, S. Battaglia, L. Birnoschi, A. Blanco-González, S. I. Bokarev, R. Broer, R. Cacciari, P. B. Calio, R. K. Carlson, R. Carvalho Couto, L. Cerdán, L. F. Chibotaru, N. F. Chilton, J. R. Church, I. Conti, S. Coriani, J. Cuéllar-Zuquin, R. E. Daoud, N. Dattani, P. Decleva, C. de Graaf, M. G. Delcey, L. De Vico, W. Dobrautz, S. S. Dong, R. Feng, N. Ferré, M. Filatov(Gulak), L. Gagliardi, M. Garavelli, L. González, Y. Guan, M. Guo, M. R. Hennefarth, M. R. Hermes, C. E. Hoyer, M. Huix-Rotllant, V. K. Jaiswal, A. Kaiser, D. S. Kaliakin, M. Khamesian, D. S. King, V. Kochetov, M. Krośnicki, A. A. Kumaar, E. D. Larsson, S. Lehtola, M.-B. Lepetit, H. Lischka, P. López Ríos, M. Lundberg, D. Ma, S. Mai, P. Marquetand, I. C. D. Merritt, F. Montorsi, M. Mörchen, A. Nenov, V. H. A. Nguyen, Y. Nishimoto, M. S. Oakley, M. Olivucci, M. Oppel, D. Padula, R. Pandharkar, Q. M. Phung, F. Plasser, G. Raggi, E. Rebolini, M. Reiher, I. Rivalta, D. Roca-Sanjuán, T. Romig, A. A. Safari, A. Sánchez-Mansilla, A. M. Sand, I. Schapiro, T. R. Scott, J. Segarra-Martí, F. Segatta, D.-C. Sergentu, P. Sharma, R. Shepard, Y. Shu, J. K. Staab, T. P. Straatsma, L. K. Sørensen, B. N. C. Tenorio, D. G. Truhlar, L. Ungur, M. Vacher, V. Veryazov, T. A. Voß, O. Weser, D. Wu, X. Yang, D. Yarkony, C. Zhou, J. P. Zobel, and R. Lindh, J. Chem. Theory Comput. 19, 6933 (2023).
- Aquilante et al. (2020) F. Aquilante, J. Autschbach, A. Baiardi, S. Battaglia, V. A. Borin, L. F. Chibotaru, I. Conti, L. De Vico, M. Delcey, I. Fdez. Galván, N. Ferré, L. Freitag, M. Garavelli, X. Gong, S. Knecht, E. D. Larsson, R. Lindh, M. Lundberg, P. Å. Malmqvist, A. Nenov, J. Norell, M. Odelius, M. Olivucci, T. B. Pedersen, L. Pedraza-González, Q. M. Phung, K. Pierloot, M. Reiher, I. Schapiro, J. Segarra-Martí, F. Segatta, L. Seijo, S. Sen, D.-C. Sergentu, C. J. Stein, L. Ungur, M. Vacher, A. Valentini, and V. Veryazov, J. Chem. Phys. 152, 214117 (2020).
- Fdez. Galván et al. (2019) I. Fdez. Galván, M. Vacher, A. Alavi, C. Angeli, F. Aquilante, J. Autschbach, J. J. Bao, S. I. Bokarev, N. A. Bogdanov, R. K. Carlson, L. F. Chibotaru, J. Creutzberg, N. Dattani, M. G. Delcey, S. S. Dong, A. Dreuw, L. Freitag, L. M. Frutos, L. Gagliardi, F. Gendron, A. Giussani, L. González, G. Grell, M. Guo, C. E. Hoyer, M. Johansson, S. Keller, S. Knecht, G. Kovačević, E. Källman, G. Li Manni, M. Lundberg, Y. Ma, S. Mai, J. P. Malhado, P. Å. Malmqvist, P. Marquetand, S. A. Mewes, J. Norell, M. Olivucci, M. Oppel, Q. M. Phung, K. Pierloot, F. Plasser, M. Reiher, A. M. Sand, I. Schapiro, P. Sharma, C. J. Stein, L. K. Sørensen, D. G. Truhlar, M. Ugandi, L. Ungur, A. Valentini, S. Vancoillie, V. Veryazov, O. Weser, T. A. Wesołowski, P.-O. Widmark, S. Wouters, A. Zech, J. P. Zobel, and R. Lindh, J. Chem. Theory Comput. 15, 5925 (2019).
- Guther et al. (2020) K. Guther, R. J. Anderson, N. S. Blunt, N. A. Bogdanov, D. Cleland, N. Dattani, W. Dobrautz, K. Ghanem, P. Jeszenszki, N. Liebermann, G. Li Manni, A. Y. Lozovoi, H. Luo, D. Ma, F. Merz, C. Overy, M. Rampp, P. K. Samanta, L. R. Schwarz, J. J. Shepherd, S. D. Smart, E. Vitale, O. Weser, G. H. Booth, and A. Alavi, J. Chem. Phys. 153, 034107 (2020).
- Qiskit contributors (2023) Qiskit contributors, “Qiskit: An open-source framework for quantum computing,” (2023).
- The Qiskit Nature developers and contributors (2023) The Qiskit Nature developers and contributors, “Qiskit nature 0.6.0,” (2023), Qiskit Nature has some code that is included under other licensing. These files have been removed from the zip repository provided here and are only available via Github. See https://github.com/Qiskit/qiskit-nature#license for more details.
- Gard et al. (2020) B. T. Gard, L. Zhu, G. S. Barron, N. J. Mayhall, S. E. Economou, and E. Barnes, npj Quantum Inf. 6, 10 (2020).
- Schulten and Karplus (1972) K. Schulten and M. Karplus, Chem. Phys. Lett. 14, 305 (1972).
- Tavan and Schulten (1986) P. Tavan and K. Schulten, J. Chem. Phys. 85, 6602 (1986).
- Pariser and Parr (1953a) R. Pariser and R. G. Parr, J. Chem. Phys. 21, 466 (1953a).
- Pariser and Parr (1953b) R. Pariser and R. G. Parr, J. Chem. Phys. 21, 767 (1953b).
- Pople (1953) J. A. Pople, Trans. Faraday Soc. 49, 1375 (1953).
- Hudson and Kohler (1972) B. Hudson and B. Kohler, Chem. Phys. Lett. 14, 299 (1972).
- Hudson and Kohler (1984) B. Hudson and B. Kohler, Synth. Met. 9, 241 (1984).
- Ido et al. (2023) K. Ido, M. Kawamura, Y. Motoyama, K. Yoshimi, Y. Yamaji, S. Todo, N. Kawashima, and T. Misawa, arXiv (2023), arXiv:2307.13222 [cond-mat.str-el] .
- Rohatgi (2022) A. Rohatgi, “Webplotdigitizer: Version 4.6,” (2022).
- Cave et al. (2004) R. J. Cave, F. Zhang, N. T. Maitra, and K. Burke, Chem. Phys. Lett. 389, 39 (2004).
- (111) https://github.com/Emieeel/1-Norm_calculations.git.
- Preskill (2018) J. Preskill, Quantum 2, 79 (2018).
- Yoshioka et al. (2022) N. Yoshioka, T. Okubo, Y. Suzuki, Y. Koizumi, and W. Mizukami, (2022), arXiv:2210.14109 [quant-ph] .
- Ichikawa et al. (2023) T. Ichikawa, H. Hakoshima, K. Inui, K. Ito, R. Matsuda, K. Mitarai, K. Miyamoto, W. Mizukami, K. Mizuta, T. Mori, Y. Nakano, A. Nakayama, K. N. Okada, T. Sugimoto, S. Takahira, N. Takemori, S. Tsukano, H. Ueda, R. Watanabe, Y. Yoshida, and K. Fujii, (2023), arXiv:2307.16130 [quant-ph] .
- Shinaoka et al. (2015) H. Shinaoka, M. Troyer, and P. Werner, Phys. Rev. B 91, 245156 (2015).
- Honerkamp et al. (2018) C. Honerkamp, H. Shinaoka, F. F. Assaad, and P. Werner, Phys. Rev. B 98, 235151 (2018).
Appendix A State characterization
We describe the characterization of the 1 ground state, the singly excited 1 state, and the doubly excited 2 state of our models. In this Appendix, the model parameters specified in Table 1 are used.
First, we perform a basis transformation by SCF calculation. The resulting canonical molecular orbital is represented by the linear combination of the Wannier functions as
(17) |
where is the molecular orbital coefficient and is the number of the Wannier functions.
Next, the model Hamiltonian in the transformed basis is diagonalized. The eigenstate of the model Hamiltonian is represented by the superposition of the computational basis states :
(18) |
where is the CI coefficient of the -th basis state in the state . The tilde represents the use of the canonical orbital basis.
The computational basis state can be written as the occupation number vector defined as follows:
(19) |
where is the occupation number of the spin-orbital consisting of the -th canonical orbital with spin in the -th basis state, and thus . For example, the Hartree-Fock (HF) state in the (4e, 4o) models corresponds to .
Table 6 shows the CI coefficients and the computational basis states of the eigenstates of our models. The definitions of Model 1 and Model 2 are the same as that defined in Sec. IV.3, employing and as the electron-electron interaction part, respectively. Note that this table only shows the coefficients of the large absolute values required for the characterization. For explanation, we use for the -th computational basis state for each state instead of . This means that the different basis states are defined for each state .
Molecule, | Model | |||||||||||
---|---|---|---|---|---|---|---|---|---|---|---|---|
Space | ||||||||||||
Ethylene, | Model 1 | 1 | 0.986 | +0.169 | ||||||||
(2e, 2o) | 1 | 0.707 | +0.707 | |||||||||
2 | +0.986 | +0.169 | ||||||||||
Model 2 | 1 | 0.986 | +0.169 | |||||||||
1 | 0.707 | +0.707 | ||||||||||
2 | +0.986 | +0.169 | ||||||||||
Butadiene, | Model 1 | 1 | 0.982 | +0.139 | ||||||||
(4e, 4o) | 1 | 0.695 | +0.695 | 0.110 | +0.110 | |||||||
2 | +0.478 | +0.410 | 0.410 | +0.403 | 0.403 | |||||||
Model 2 | 1 | 0.982 | +0.135 | |||||||||
1 | 0.697 | +0.697 | 0.105 | +0.105 | ||||||||
2 | +0.489 | +0.410 | 0.410 | +0.399 | 0.399 | |||||||
Hexatriene, | Model 1 | 1 | +0.982 | 0.144 | ||||||||
(4e, 4o) | 1 | 0.694 | +0.694 | +0.092 | 0.092 | |||||||
2 | 0.561 | +0.490 | 0.490 | +0.224 | 0.224 | |||||||
Model 2 | 1 | +0.984 | 0.128 | |||||||||
1 | 0.695 | +0.695 | +0.093 | 0.093 | ||||||||
2 | 0.565 | +0.475 | 0.475 | +0.237 | 0.237 | |||||||
Hexatriene, | Model 1 | 1 | +0.974 | 0.134 | ||||||||
(6e, 6o) | 1 | +0.682 | 0.682 | +0.128 | 0.128 | |||||||
2 | 0.557 | 0.388 | +0.388 | 0.324 | +0.324 | |||||||
Model 2 | 1 | 0.974 | +0.127 | |||||||||
1 | +0.685 | 0.685 | +0.111 | 0.111 | ||||||||
2 | +0.564 | +0.389 | 0.389 | +0.319 | 0.319 |
Table 6 shows that the eigenstates are represented as the linear combination of the spin-adapted configurations. In all cases, the HF state is dominant in the 1 ground state, and the 1 and 2 states are reasonably characterized as the excited states of the one- and two-electron excitations from the highest occupied to the lowest unoccupied molecular orbitals.