This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

A Quantum Algorithm to Calculate Band Structure at the EOM Level of Theory

Yi Fan Hefei National Laboratory for Physical Sciences at the Microscale, University of Science and Technology of China, Hefei, Anhui 230026, China    Jie Liu Hefei National Laboratory for Physical Sciences at the Microscale, University of Science and Technology of China, Hefei, Anhui 230026, China    Zhenyu Li [email protected] Hefei National Laboratory for Physical Sciences at the Microscale, University of Science and Technology of China, Hefei, Anhui 230026, China    Jinlong Yang Hefei National Laboratory for Physical Sciences at the Microscale, University of Science and Technology of China, Hefei, Anhui 230026, China
Abstract

Band structure is a cornerstone to understand electronic properties of materials. Accurate band structure calculations using a high-level quantum-chemistry theory can be computationally very expensive. It is promising to speed up such calculations with a quantum computer. In this study, we present a quantum algorithm for band structure calculation based on the equation-of-motion (EOM) theory. First, we introduce a new variational quantum eigensolver algorithm named ADAPT-C for ground-state quantum simulation of solids, where the wave function is built adaptively from a complete set of anti-Hermitian operators. Then, on top of the ADAPT-C ground state, quasiparticle energies and the band structure can be calculated using the EOM theory in a quantum-subspace-expansion (QSE) style, where the projected excitation operators guarantee that the killer condition is satisfied. As a proof of principle, such an EOM-ADAPT-C protocol is used to calculate the band structures of silicon and diamond using a quantum computer simulator.

keywords:
quantum computation, electronic structure, variational quantum eigensolver

Many properties of a chemical system can be predicted by the electronic structure theory. Over the past decades, density functional theory (DFT)1, 2, 3, 4 has become one of the most widely used electronic structure method. However, its accuracy is difficult to be systematically improved since the exchange-correlation part of the exact energy functional in DFT is unknown. For band structure calculation, GW approximation5 can be used to improve the accuracy of DFT results with local functionals. However, GW calculation usually has a high computational cost and its error is also hard to assess due to different self-consistency protocols adopted in practical GW calculations6.

Wave function based methods, such as configuration interaction (CI), Møller-Plesset perturbation theory, and coupled cluster theory (CC), offer a systematic way to approach the exact solution of the Schrödinger equation. For example, although typical CC calculations are truncated at single and double excitations (CCSD), we can in principle obtain accurate wave function by including all possible excitations. Recently, equation-of-motion coupled-cluster (EOM-CC) theory has been implemented in studying excitations and quasiparticle band structures in solid systems7, 8, 9, 10, 11. Numerical calculations indicate that EOM-CCSD produces quasiparticle energies which are usually more accurate than those from the GW approximation12. It can be proved that the EOM-CCSDT Green’s function includes all diagrams contained in the GW approximation, along with many high-order vertex corrections. Although EOM-CC can reach a very high accuracy by incorporating high-order terms, it must be addressed that the computational cost also grows fast and quickly goes beyond the capability of current supercomputers.

Recent advances in quantum computing provides possibilities to obtain the exact solution to the many-electron Schrödinger equation within a polynomial time complexity. Quantum algorithms for electronic structure calculations include quantum phase estimation (QPE) and variational quantum eigensolver (VQE)13, 14, 15, 16, 17, 18, 19, 20, 21, 22, 23, 24, 25, 26, 27, 28, 29, 30, 31, 32. The former can efficiently calculate the energy spectrum of a specific Hamiltonian. However, the high gate fidelity and error-correction devices required by QPE is typically far beyond the capability of the near-term noisy intermediate-scale quantum (NISQ) computers. By embedding efficient quantum state preparation and measurement in a classical optimization process, VQE yields a much shallower quantum circuit compared to QPE and it is thus more suitable for NISQ devices. The first experiment of VQE to study chemical systems was carried out on a photonic quantum processor to study the ground state of HeH+ 20. In recent years, VQE has been demonstrated on most of the leading quantum computer architectures such as trapped-ion system21, 22, 23 and superconducting-qubit-based platform24, 25, 26.

The key ingredient of VQE is a parametric wave function ansätze. For example, one possibility is assuming that the wave function can be written in a unitary coupled-cluster (UCC) form33, 34, 35. Recently, a more efficient and accurate wave function ansätze, adaptive derivative-assembled pseudo-trotter (ADAPT)36, is proposed, which relies on an iterative construction of the wave function. ADAPT algorithm provides a very accurate solution to the ground-state wave function, making it a favorable reference state for studying excitations under the EOM framework. Although VQE has been used to extract the bands of silicon from a tight-binding Hamiltonian37, quantum algorithms for band structure calculations at a high ab-initio level is not yet available. It is thus desirable to extend the EOM theory to the ADAPT wave function and predict band structures in an EOM-ADAPT way on a quantum computer.

Notice that the development of an EOM-ADAPT quantum algorithm for band structure calculation is not straightforward. First, like most quantum algorithms used in chemical simulations, the ADAPT algorithm is designed for molecular systems. In a previous study38, we have pointed out that a direct application of the ADAPT algorithm to periodic solid systems suffers from a residual error introduced by complex wave functions. A simple way to solve this problem is transforming complex Hartree-Fock orbitals at sampling kk-points in a unit cell to real orbitals at the Γ\Gamma point in a supercell, which is termed as the K2G transformation38. However, this leads to a significant growth in the number of operator terms since the restriction of crystal momentum conservation is removed. At the same time, it is not convenient for band structure calculations especially when a non-orthogonal kk-grid is used. Another difficulty in EOM-ADAPT band structure calculation is that the Hamiltonian similarity transformation technique which is widely used in EOM-CCSD 39, 40 to satisfy the killer condition is not applicable as detailed below. Therefore, extra efforts are required to avoid deviations between different EOM functionals, which can otherwise lead to additional errors in the predicted band gap.

In this study, we present an EOM algorithm for quasiparticle band structure calculations on quantum computers. First, we introduce a modified ADAPT-VQE ansätze termed ADAPT-C for preparing the accurate ground state. In contrast to the K2G transformation scheme, the Hartree-Fock orbitals at sampling kk-points are explicitly treated and the residual errors are minimized by introducing a complementary operator pool. Then, the quasiparticle states are obtained through a quantum subspace expansion (QSE)31 style technique derived from the projected-operator EOM-IP and EOM-EA formalism39. The converged ground-state wave function obtained with ADAPT-C is served as the reference state, and the excitation operators are transformed using the ground-state projector to satisfy the killer condition. Finally, the quasiparticle bands are extracted by identifying correct quasiparticle excitations from the solved eigenstates. We name this method as EOM-ADAPT-C, which, as a demonstration, is successfully applied to simulate band structures of silicon and diamond.

ADAPT-C algorithm. Given the Hartree-Fock orbitals as linear combinations of Bloch atomic orbitals at sampling k{k} points, the Hamiltonian can be written in the second quantized form as

H^\displaystyle\hat{H} =𝒌p,𝒌qp,qhq𝒌qp𝒌pT^q𝒌qp𝒌p\displaystyle=\sum_{\bm{k}_{p},\bm{k}_{q}}^{\prime}\sum_{p,q}h_{q\bm{k}_{q}}^{p\bm{k}_{p}}\hat{T}_{q\bm{k}_{q}}^{p\bm{k}_{p}} (1)
+12𝒌p,𝒌q𝒌r,𝒌sp,qr,sgr𝒌rs𝒌sp𝒌pq𝒌qT^r𝒌rs𝒌sp𝒌pq𝒌q\displaystyle+\frac{1}{2}\sum_{\begin{subarray}{c}\bm{k}_{p},\bm{k}_{q}\\ \bm{k}_{r},\bm{k}_{s}\end{subarray}}^{\prime}{\sum_{\begin{subarray}{c}p,q\\ r,s\end{subarray}}}g_{r\bm{k}_{r}s\bm{k}_{s}}^{p\bm{k}_{p}q\bm{k}_{q}}\hat{T}_{r\bm{k}_{r}s\bm{k}_{s}}^{p\bm{k}_{p}q\bm{k}_{q}}

with

T^q𝒌qp𝒌p\displaystyle\hat{T}_{q\bm{k}_{q}}^{p\bm{k}_{p}} =a^p𝒌pa^q𝒌q\displaystyle=\hat{a}^{\dagger}_{p\bm{k}_{p}}\hat{a}_{q\bm{k}_{q}} (2)
T^r𝒌rs𝒌sp𝒌pq𝒌q\displaystyle\hat{T}_{r\bm{k}_{r}s\bm{k}_{s}}^{p\bm{k}_{p}q\bm{k}_{q}} =a^p𝒌pa^q𝒌qa^r𝒌ra^s𝒌s\displaystyle=\hat{a}_{p\bm{k}_{p}}^{\dagger}\hat{a}_{q\bm{k}_{q}}^{\dagger}\hat{a}_{r\bm{k}_{r}}\hat{a}_{s\bm{k}_{s}}

hq𝒌qp𝒌ph_{q\bm{k}_{q}}^{p\bm{k}_{p}} and gr𝒌rs𝒌sp𝒌pq𝒌qg_{r\bm{k}_{r}s\bm{k}_{s}}^{p\bm{k}_{p}q\bm{k}_{q}} are one- and two-electron integrals, respectively, which can be calculated with classical routines. Primed summations enforce crystal momentum conservation, i.e. the difference between summations of crystal momenta of creation and annihilation operators should be a reciprocal lattice vector. The creation and annihilation operators in Hamiltonian can be mapped to weighted Pauli strings and thus part of the quantum circuit using the Jordan-Wigner or Bravyi-Kitaev transformations41, 13, 42, 43.

The ground-state wave function can be obtained by solving the Schrödinger equation

H^|Ψ0=E0|Ψ0\hat{H}|\Psi_{0}\rangle=E_{0}|\Psi_{0}\rangle (3)

via a variational algorithm, where the energy expectation value is minimized with a specific wave function ansätze. The UCC ansätze is a widely used physically motivated ansätze for quantum simulations, where the wave function takes the form

|Ψ=eT^T^|Φ0|\Psi\rangle=e^{\hat{T}-\hat{T}^{\dagger}}|\Phi_{0}\rangle (4)

|Φ0|\Phi_{0}\rangle is a reference state, typically the Hartree-Fock ground state. If truncated to the second order (UCCSD), the cluster operators are

T^\displaystyle\hat{T} =𝒌i,𝒌aa,iθi𝒌ia𝒌aT^i𝒌ia𝒌a\displaystyle=\sum_{\bm{k}_{i},\bm{k}_{a}}^{\prime}\sum_{a,i}\theta_{i\bm{k}_{i}}^{a\bm{k}_{a}}\hat{T}_{i\bm{k}_{i}}^{a\bm{k}_{a}} (5)
+14𝒌a,𝒌b𝒌i,𝒌ja,bi,jθi𝒌ij𝒌ja𝒌ab𝒌bT^i𝒌ij𝒌ja𝒌ab𝒌b\displaystyle+\frac{1}{4}\sum_{\begin{subarray}{c}\bm{k}_{a},\bm{k}_{b}\\ \bm{k}_{i},\bm{k}_{j}\end{subarray}}^{\prime}{\sum_{\begin{subarray}{c}a,b\\ i,j\end{subarray}}}\theta_{i\bm{k}_{i}j\bm{k}_{j}}^{a\bm{k}_{a}b\bm{k}_{b}}\hat{T}_{i\bm{k}_{i}j\bm{k}_{j}}^{a\bm{k}_{a}b\bm{k}_{b}}

It has been suggested that the performance of UCCSD can be improved by replacing occupied and virtual orbitals by general orbitals (UCCGSD)44, 45, 46, 47.

T^\displaystyle\hat{T} =12𝒌p,𝒌qp,qθq𝒌qp𝒌pT^q𝒌p𝒌\displaystyle=\frac{1}{2}\sum_{\bm{k}_{p},\bm{k}_{q}}^{\prime}\sum_{p,q}\theta_{q\bm{k}_{q}}^{p\bm{k}_{p}}\hat{T}_{q\bm{k}}^{p\bm{k}} (6)
+14𝒌p,𝒌q𝒌r,𝒌sp,qr,sθr𝒌rs𝒌sp𝒌pq𝒌qT^r𝒌rs𝒌sp𝒌pq𝒌q\displaystyle+\frac{1}{4}\sum_{\begin{subarray}{c}\bm{k}_{p},\bm{k}_{q}\\ \bm{k}_{r},\bm{k}_{s}\end{subarray}}^{\prime}{\sum_{\begin{subarray}{c}p,q\\ r,s\end{subarray}}}\theta_{r\bm{k}_{r}s\bm{k}_{s}}^{p\bm{k}_{p}q\bm{k}_{q}}\hat{T}_{r\bm{k}_{r}s\bm{k}_{s}}^{p\bm{k}_{p}q\bm{k}_{q}}

Here, we use {i,j,k,}\{i,j,k,\dots\}, {a,b,c,}\{a,b,c,\dots\}, and {p,q,r,}\{p,q,r,\dots\} to indicate occupied, virtual, and general orbitals, respectively.

The exponential unitary operator in the UCC ansätze cannot be converted to a quantum circuit directly. In practice, it is decomposed into one- and two-qubit quantum gates using an approximated scheme named Trotter-Suzuki decomposition48, 49:

eA^+B^(eA^/keB^/k)ke^{\hat{A}+\hat{B}}\approx(e^{\hat{A}/k}e^{\hat{B}/k})^{k} (7)

The corresponding UCC wave function has an expression of

|Ψ=limNk=1NiNieθiNτ^i|Φ0|\Psi\rangle=\lim_{N\rightarrow\infty}{\prod_{k=1}^{N}{\prod_{i}^{N_{i}}{e^{\frac{\theta_{i}}{N}\hat{\tau}_{i}}|\Phi_{0}\rangle}}} (8)

with {τ^i}\{\hat{\tau}_{i}\} being anti-Hermitian operators:

τ^q𝒌qp𝒌p=T^q𝒌qp𝒌pT^p𝒌pq𝒌q\hat{\tau}_{q\bm{k}_{q}}^{p\bm{k}_{p}}=\hat{T}_{q\bm{k}_{q}}^{p\bm{k}_{p}}-\hat{T}^{q\bm{k}_{q}}_{p\bm{k}_{p}} (9)
τ^r𝒌rs𝒌sp𝒌pq𝒌q=T^r𝒌rs𝒌sp𝒌pq𝒌qT^q𝒌qp𝒌ps𝒌sr𝒌r\hat{\tau}_{r\bm{k}_{r}s\bm{k}_{s}}^{p\bm{k}_{p}q\bm{k}_{q}}=\hat{T}_{r\bm{k}_{r}s\bm{k}_{s}}^{p\bm{k}_{p}q\bm{k}_{q}}-\hat{T}^{s\bm{k}_{s}r\bm{k}_{r}}_{q\bm{k}_{q}p\bm{k}_{p}} (10)

Here, general excitation operators {τ^q𝒌qp𝒌p,τ^r𝒌rs𝒌sp𝒌pq𝒌q}\{\hat{\tau}_{q\bm{k}_{q}}^{p\bm{k}_{p}},\hat{\tau}_{r\bm{k}_{r}s\bm{k}_{s}}^{p\bm{k}_{p}q\bm{k}_{q}}\} should also satisfy the crystal momentum conservation rule. The Trotterization procedure is to be truncated at finite order and the ordering of the Trotterized cluster operators is not unique, which may introduce a considerable Trotterization error48, 49.

The recently proposed ADAPT ansätze36 is capable of bypassing the Trotterization error by generating a compact sequence of unitary operators. Instead of optimizing a large number of UCC parameters simultaneously, a pseudo-Trotter wave function with finite variational parameters is constructed

|Ψ(k+1)=eθ(k+1)τ^(k+1)|Ψ(k)|\Psi^{(k+1)}\rangle=e^{\theta^{(k+1)}\hat{\tau}^{(k+1)}}|\Psi^{(k)}\rangle (11)

where the exponential terms are taken from the operator pool 𝒪\mathcal{O} defined by anti-Hermitian operators τi^𝒪{τ^q𝒌qp𝒌p,τ^r𝒌rs𝒌sp𝒌pq𝒌q}\hat{\tau_{i}}\in\mathcal{O}\equiv\{\hat{\tau}_{q\bm{k}_{q}}^{p\bm{k}_{p}},\hat{\tau}_{r\bm{k}_{r}s\bm{k}_{s}}^{p\bm{k}_{p}q\bm{k}_{q}}\}. At the kkth iteration, the energy is minimized through the variational procedure:

E(k)=min{θ(l)}l=1k{Ψ(k)|H^|Ψ(k)}E^{(k)}=\min_{\{\vec{\theta}^{(l)}\}_{l=1}^{k}}{\{\langle\Psi^{(k)}|\hat{H}|\Psi^{(k)}\rangle\}} (12)

At the (k+1)(k+1)th iteration, among all operators in 𝒪\mathcal{O}, the one with the largest residual gradient is selected as τ^(k+1)\hat{\tau}^{(k+1)}. The residual gradient of operator τ^i\hat{\tau}_{i} is defined as

Ri(k)\displaystyle R_{i}^{(k)} =E(k+1)θ(k+1)|θ(k+1)=0,τ^(k+1)=τ^i\displaystyle=\left.\frac{\partial{E^{(k+1)}}}{\partial{\theta^{(k+1)}}}\right|_{\theta^{(k+1)}=0,\hat{\tau}^{(k+1)}=\hat{\tau}_{i}} (13)
=Ψ(k)|[H^,τ^i]|Ψ(k)\displaystyle=\langle\Psi^{(k)}|[\hat{H},\hat{\tau}_{i}]|\Psi^{(k)}\rangle

The L2L2-norm of residual gradients R(k)=(R1(k),R2(k),)\vec{R}^{(k)}=(R_{1}^{(k)},R_{2}^{(k)},\dots) is used as the convergence criterion

R(k)2=i|Ri(k)|2<ε\left\lVert\vec{R}^{(k)}\right\rVert_{2}=\sqrt{\sum_{i}{\left.|R_{i}^{(k)}|\right.^{2}}}<\varepsilon (14)

Such a convergence criterion guarantees that the real part of the anti-Hermitian contracted Schrödinger equation (ACSE)50 is satisfied. The imaginary part of ACSE is automatically satisfied only for real-valued wave functions. Therefore, there is a residual error originated from the imaginary part of ACSE when the ADAPT algorithm is straightforwardly applied to solids with complex wave function38. In the UCC case, it is suggested to use complex cluster amplitudes to avoid this problem.51

In ADAPT, it is more natural to introduce a complementary operator pool 𝒪c\mathcal{O}_{c}

𝒪c={τ~^q𝒌qp𝒌p,τ~^r𝒌rs𝒌sp𝒌pq𝒌q}\mathcal{O}_{c}=\{\hat{\tilde{\tau}}_{q\bm{k}_{q}}^{p\bm{k}_{p}},\hat{\tilde{\tau}}_{r\bm{k}_{r}s\bm{k}_{s}}^{p\bm{k}_{p}q\bm{k}_{q}}\} (15)

where τ~^q𝒌qp𝒌p\hat{\tilde{\tau}}_{q\bm{k}_{q}}^{p\bm{k}_{p}} and τ~^r𝒌rs𝒌sp𝒌pq𝒌q\hat{\tilde{\tau}}_{r\bm{k}_{r}s\bm{k}_{s}}^{p\bm{k}_{p}q\bm{k}_{q}} are anti-Hermitian operators defined as

τ~^q𝒌qp𝒌p=i(T^q𝒌qp𝒌p+T^p𝒌pq𝒌q)\hat{\tilde{\tau}}_{q\bm{k}_{q}}^{p\bm{k}_{p}}=i(\hat{T}_{q\bm{k}_{q}}^{p\bm{k}_{p}}+\hat{T}^{q\bm{k}_{q}}_{p\bm{k}_{p}}) (16)
τ~^r𝒌rs𝒌sp𝒌pq𝒌q=i(T^r𝒌rs𝒌sp𝒌pq𝒌q+T^q𝒌qp𝒌ps𝒌sr𝒌r).\hat{\tilde{\tau}}_{r\bm{k}_{r}s\bm{k}_{s}}^{p\bm{k}_{p}q\bm{k}_{q}}=i(\hat{T}_{r\bm{k}_{r}s\bm{k}_{s}}^{p\bm{k}_{p}q\bm{k}_{q}}+\hat{T}^{s\bm{k}_{s}r\bm{k}_{r}}_{q\bm{k}_{q}p\bm{k}_{p}}). (17)

The wave function is then updated using the extended operator pool

𝒪~=𝒪𝒪c={τ^q𝒌qp𝒌p,τ^r𝒌rs𝒌sp𝒌pq𝒌q,τ~^q𝒌qp𝒌p,τ~^r𝒌rs𝒌sp𝒌pq𝒌q}\tilde{\mathcal{O}}=\mathcal{O}\cup\mathcal{O}_{c}=\{\hat{\tau}_{q\bm{k}_{q}}^{p\bm{k}_{p}},\hat{\tau}_{r\bm{k}_{r}s\bm{k}_{s}}^{p\bm{k}_{p}q\bm{k}_{q}},\hat{\tilde{\tau}}_{q\bm{k}_{q}}^{p\bm{k}_{p}},\hat{\tilde{\tau}}_{r\bm{k}_{r}s\bm{k}_{s}}^{p\bm{k}_{p}q\bm{k}_{q}}\} (18)

The size of the operator pool is doubled regardless of the Brillouin zone sampling scheme. At the same time, the crystal momentum conservation law is still preserved. We name such a modified ADAPT algorithm with a complementary operator pool as ADAPT-C. In ADAPT-C, variational parameters for operators in both 𝒪\mathcal{O} and 𝒪c\mathcal{O}_{c} are kept to be real numbers.

EOM theory. The EOM formalism based on the coupled-cluster theory52, 53, 54, 55 is a well-established approach to calculate excitation energies in quantum chemistry. In EOM theory, we focus on the excitation operator ^\hat{\mathcal{R}} connecting the ground state |Ψ0|\Psi_{0}\rangle and a target excited state |Ψx|\Psi_{x}\rangle. From the definition ^=|ΨxΨ0|\hat{\mathcal{R}}=|\Psi_{x}\rangle\langle\Psi_{0}|, we can obtain an EOM-like equation for ^\hat{\mathcal{R}}

[H^,^]=ΔEx^[\hat{H},\hat{\mathcal{R}}]=\Delta E_{x}\hat{\mathcal{R}} (19)

where the excitation energy is defined as ΔEx=ExE0\Delta E_{x}=E_{x}-E_{0}. Such an equation provides us a way to access excitation energies without the knowledge about the ground and excited states explicitly.

Different types of target states can be studied in the EOM theory by using different excitation operators. Ionization potential (IP) and electron affinity (EA) can be obtained if the operator ^\hat{\mathcal{R}} includes a different number of creation and annihilation operators and thus not electron conserving. The fundamental band gap of a NN-electron neutral system is characterized as the energy difference between the IP and EA. Therefore, we can use the EOM-IP energy ΔExIP\Delta E^{IP}_{x} and the EOM-EA energy ΔExEA\Delta E^{EA}_{x} at sampling kk-points to construct the valence and conduction bands.7

Here, we truncate the EOM-IP and EOM-EA excitation operator up to the second order, which means the ionization energies are calculated by diagonalizing the Hamiltonian in the space of 1-hole (1hh) and 2-hole, 1-particle (2hh1pp) states, while electron affinities are obtained in the space of 1-particle (1pp) and 2-particle, 1-hole (2pp1hh) states. The corresponding excitation operators ^IP\hat{\mathcal{R}}_{IP} and ^EA\hat{\mathcal{R}}_{EA} are expressed as a linear combination of a set of Fermion excitation operators as

^IP(𝒌)\displaystyle\hat{\mathcal{R}}_{IP}(\bm{k}) =prp𝒌a^p𝒌\displaystyle=\sum_{p}{r_{p\bm{k}}\hat{a}_{p\bm{k}}} +𝒌p𝒌q,𝒌sp,q,srq𝒌qs𝒌sp𝒌pa^p𝒌pa^q𝒌qa^s𝒌s\displaystyle+\sum_{\begin{subarray}{c}\bm{k}_{p}\\ \bm{k}_{q},\bm{k}_{s}\end{subarray}}^{\prime}{\sum_{p,q,s}{r_{q\bm{k}_{q}s\bm{k}_{s}}^{p\bm{k}_{p}}{\hat{a}_{p\bm{k}_{p}}^{\dagger}\hat{a}_{q\bm{k}_{q}}\hat{a}_{s\bm{k}_{s}}}}} (20)
^EA(𝒌)\displaystyle\hat{\mathcal{R}}_{EA}(\bm{k}) =prp𝒌a^p𝒌\displaystyle=\sum_{p}{r^{p\bm{k}}\hat{a}^{\dagger}_{p\bm{k}}} +𝒌p,𝒌q𝒌sp,q,srs𝒌sp𝒌pq𝒌qa^p𝒌pa^q𝒌qa^s𝒌s\displaystyle+\sum_{\begin{subarray}{c}\bm{k}_{p},\bm{k}_{q}\\ \bm{k}_{s}\end{subarray}}^{\prime}{\sum_{p,q,s}{r_{s\bm{k}_{s}}^{p\bm{k}_{p}q\bm{k}_{q}}{\hat{a}_{p\bm{k}_{p}}^{\dagger}\hat{a}_{q\bm{k}_{q}}^{\dagger}\hat{a}_{s\bm{k}_{s}}}}} (21)

where the rr-coefficients of the operator basis are parameters to be determined. Again, crystal momentum conservation rule 𝒌p𝒌q𝒌s±𝒌=𝑮m\bm{k}_{p}\mp\bm{k}_{q}-\bm{k}_{s}\pm\bm{k}=\bm{G}_{m} should be satisfied7, 8, 9, 10, 11.

In practical EOM calculations, a reference state |Ψ|\Psi\rangle is chosen. Then, eigenvalues of Eq. (19) can be obtained from stationary values of the following functional

ΔEx=Ψ|^[H^,^]|ΨΨ|^^|Ψ\Delta E_{x}=\frac{\langle\Psi|\hat{\mathcal{R}}^{\dagger}[\hat{H},\hat{\mathcal{R}}]|\Psi\rangle}{\langle\Psi|\hat{\mathcal{R}}^{\dagger}\hat{\mathcal{R}}|\Psi\rangle} (22)

for any reference state |Ψ|\Psi\rangle that has a nonzero overlap with the exact ground state |Ψ0|\Psi_{0}\rangle. Notice that, if the operator basis to represent ^\hat{\mathcal{R}} is incomplete, the accuracy of the calculated excitation energy depends on the choice of reference state |Ψ|\Psi\rangle. In the Liouvillian superoperator formalism, Eq. (22) corresponds to a simple metric defined in the operator space (ρi^|ρj^)=Ψ|ρi^ρj^|Ψ(\hat{\rho_{i}}|\hat{\rho_{j}})=\langle\Psi|\hat{\rho_{i}}^{\dagger}\hat{\rho_{j}}|\Psi\rangle 39.

Sometimes, it is more convenient to choose the commutator metric (ρi^|ρj^)=Ψ|[ρi^,ρj^]+|Ψ(\hat{\rho_{i}}|\hat{\rho_{j}})=\langle\Psi|[\hat{\rho_{i}}^{\dagger},\hat{\rho_{j}}]_{+}|\Psi\rangle. Then, the corresponding EOM functional becomes

ΔEx=Ψ|[^,[H^,^]]+|ΨΨ|[^,^]+|Ψ.\Delta E_{x}=\frac{\langle\Psi|[\hat{\mathcal{R}}^{\dagger},[\hat{H},\hat{\mathcal{R}}]]_{+}|\Psi\rangle}{\langle\Psi|[\hat{\mathcal{R}}^{\dagger},\hat{\mathcal{R}}]_{+}|\Psi\rangle}. (23)

Such an expression can lower the rank of density matrices involved in classical calculations55, 39 and reduce the number of Pauli terms in quantum simulations. However, it is not equivalent to the simple metric functional Eq. (22) for a practical ansätze of ^\hat{\mathcal{R}} such as those specified in Eqs. (20) and (21). If the so-called killer condition ^|Ψ=0\hat{\mathcal{R}}^{\dagger}|\Psi\rangle=0 (it means that the ground state cannot be de-excited) is satisfied, the simple and commutator metric expressions becomes equivalent.

In band structure calculations with complex wave function involved, another important issue is if the excitation energy will be a real number. With such a consideration, a symmetric double commutator form of the EOM functional has been introduced55

ΔEx=Ψ|[^,H^,^]+|ΨΨ|[^,^]+|Ψ\Delta E_{x}=\frac{\langle\Psi|[\hat{\mathcal{R}}^{\dagger},\hat{H},\hat{\mathcal{R}}]_{+}|\Psi\rangle}{\langle\Psi|[\hat{\mathcal{R}}^{\dagger},\hat{\mathcal{R}}]_{+}|\Psi\rangle} (24)

where [^,H^,^]+=12([[^,H^],^]++[^,[H^,^]]+)[\hat{\mathcal{R}}^{\dagger},\hat{H},\hat{\mathcal{R}}]_{+}=\frac{1}{2}([[\hat{\mathcal{R}}^{\dagger},\hat{H}],\hat{\mathcal{R}}]_{+}+[\hat{\mathcal{R}}^{\dagger},[\hat{H},\hat{\mathcal{R}}]]_{+}). This EOM functional guarantees real-valued excitation energies. A similar expression has been implemented in the qEOM method to calculate the electron-excitation (EE) energies.56 If ^=|ΨxΨ0|\hat{\mathcal{R}}=|\Psi_{x}\rangle\langle\Psi_{0}|, Eq. (24) can be obtained directly from Eq. (19). For a practical ansätze of ^\hat{\mathcal{R}}, Eq. (24) can be obtained from Eq. (23) if the exact ground state |Ψ0|\Psi_{0}\rangle is used as the reference state.

Since in practical calculations ^\hat{\mathcal{R}} is typically truncated and |Ψ|\Psi\rangle is not the exact ground state, a direct use of Eq. (24) may lead to significant errors as will be demonstrated in Si and diamond band structure calculations. To solve this problem, we adopt the projected excitation operator formalism 39 for both EOM-IP and EOM-EA

~^IP(EA)=^IP(EA)|ΨΨ|\hat{\tilde{\mathcal{R}}}_{IP(EA)}=\hat{\mathcal{R}}_{IP(EA)}|\Psi\rangle\langle\Psi| (25)

Since ~^|Ψ=^|Ψ\hat{\tilde{\mathcal{R}}}|\Psi\rangle=\hat{\mathcal{R}}|\Psi\rangle, we can substituting ^\hat{\mathcal{R}} in EOM functionals with ~^\hat{\tilde{\mathcal{R}}}. Then, the killer condition becomes satisfied for any reference wavefunction. More importantly, the IP or EA energies obtained from Eq. (24) become equivalent to those from Eqs. (22) and (23).

The problem of such a projected operator formalism in quantum simulation is that the projector |ΨΨ||\Psi\rangle\langle\Psi| is difficult to be directly implemented in a quantum circuit. Therefore, a simplified working equation is obtained from either Eqs. (22), (23), or (24), which is given as

ΔEx\displaystyle\Delta E_{x} =Ψ|[~^,H^,~^]+|ΨΨ|[~^,~^]+|Ψ\displaystyle=\frac{\langle\Psi|[\hat{\tilde{\mathcal{R}}}^{\dagger},\hat{H},\hat{\tilde{\mathcal{R}}}]_{+}|\Psi\rangle}{\langle\Psi|[\hat{\tilde{\mathcal{R}}}^{\dagger},\hat{\tilde{\mathcal{R}}}]_{+}|\Psi\rangle} (26)
=Ψ|^H^^|ΨΨ|^^|ΨΨ|H^|Ψ\displaystyle=\frac{\langle\Psi|\hat{\mathcal{R}}^{\dagger}\hat{H}\hat{\mathcal{R}}|\Psi\rangle}{\langle\Psi|\hat{\mathcal{R}}^{\dagger}\hat{\mathcal{R}}|\Psi\rangle}-\langle\Psi|\hat{H}|\Psi\rangle

In contrast to Eqs. (22) and (23), this working equation always leads to real excitation energies.

The second term of the working equation (A1) is a constant and the first term can be converted into a generalized Hermitian eigenvalue problem in a QSE31 style

𝑯𝑪=𝑺𝑪𝑬\bm{H}\bm{C}=\bm{S}\bm{C}\bm{E} (27)

The matrix elements of 𝑯\bm{H} and 𝑺\bm{S} can be evaluated as

Huv=ΨC|ρ^uH^ρ^v|ΨC\displaystyle H_{uv}=\langle\Psi_{C}|\hat{\rho}_{u}^{\dagger}\hat{H}\hat{\rho}_{v}|\Psi_{C}\rangle (28)
Suv=ΨC|ρ^uρ^v|ΨC\displaystyle S_{uv}=\langle\Psi_{C}|\hat{\rho}_{u}^{\dagger}\hat{\rho}_{v}|\Psi_{C}\rangle

where ρ^i\hat{\rho}_{i} stands for Fermion excitation terms in Eqs. (20) and (21), |ΨC|\Psi_{C}\rangle is the ADAPT-C ground-state wave function. Calculation of these matrix elements can take advantage of a quantum computer, while eigenvalues and eigenvectors of Eq.(27) can be calculated on a classical computer.

Among all the eigenstates, only the quasiparticle states should be considered for band structure. Therefore, we denote R1={rp𝒌,rp𝒌}\vec{R}_{1}=\{r_{p\bm{k}},r^{p\bm{k}}\} and define the quasiparticle weight (QPWTQPWT) as

QPWT=R12QPWT=\left\lVert\vec{R}_{1}\right\rVert_{2} (29)

For example, the QPWTQPWT for EOM-IP takes the form

QPWTIP=p|rp𝒌|2.QPWT_{IP}=\sqrt{\sum_{p}{\left.|r_{p\bm{k}}|\right.^{2}}}. (30)

where {rp𝒌}\{r_{p\bm{k}}\} are coefficients of single ionization operators {a^p𝒌}\{\hat{a}_{p\bm{k}}\}. To construct the band structure, we find excitation energies with a large QPWTQPWT, which indicates the corresponding excitation is dominated by single electron addition (for EOM-EA) or removal (for EOM-IP). The protocol described above for band structure calculation is named EOM-ADAPT-C.

Notice that, in classical EOM-CC calculations, a similarity transformation technique is widely used, where a transformed Hamiltonian H~^=eT^H^eT^\hat{\tilde{H}}=e^{-\hat{T}}\hat{H}e^{\hat{T}} is considered with the Hartree-Fock wave function |ΨHF|\Psi_{HF}\rangle as the reference state54, 40. Such a scheme guarantees that the killer condition is satisfied ^|ΨHF0\hat{\mathcal{R}}^{\dagger}|\Psi_{HF}\rangle\equiv 0, although it is not due to the orthogonality between excited and ground states as it suppose to be. The similarity-transformed Hamiltonian has the same excitation energy spectrum as the bare Hamiltonian. More importantly, due to the commutation relationship [^,e±T^]0[\hat{\mathcal{R}},e^{\pm\hat{T}}]\equiv 0, we also have [H~^,^]|ΨHF=ΔEx^|ΨHF[\hat{\tilde{H}},\hat{\mathcal{R}}]|\Psi_{HF}\rangle=\Delta E_{x}\hat{\mathcal{R}}|\Psi_{HF}\rangle, which means that ^\hat{\mathcal{R}} is the same for the bare and transformed Hamiltonians. In the UCC or ADAPT case with anti-Hermitian operators {τ^i}\{\hat{\tau}_{i}\} introduced, since [^,eτ^]0[\hat{\mathcal{R}},e^{\hat{\tau}}]\neq 0, ^\hat{\mathcal{R}} for the bare and transformed Hamiltonians are different. Such a difference is consistent with the intermediate state representation57, where the so-called self-consistent excitation operators eTT^eTTe^{T-T^{\dagger}}\hat{\mathcal{R}}e^{T^{\dagger}-T} is applied on the UCC or ADAPT ground state and the similarity-transformed Hamiltonian appears in the working equation. As a result of this difference, the identification of quasiparticle state will become more difficult. At the same time, estimation of the matrix element Ψ|ρ^ueTTH^eTTρ^v|Ψ\langle\Psi|\hat{\rho}_{u}^{\dagger}e^{T^{\dagger}-T}\hat{H}e^{T-T^{\dagger}}\hat{\rho}_{v}|\Psi\rangle for the similarity-transformed Hamiltonian leads to a more complicated quantum circuit. Therefore, the similarity transformation technique is not used in this study.

Numerical results. We first use a one dimensional hydrogen chain system to test the accuracy of ADAPT-C. Then, band structures of diamond and silicon are calculated using the EOM-ADAPT-C algorithm. All calculations are performed using an in-house developed code interfaced with several open-source software packages. One- and two-electron coefficients in the Hamiltonian in Eq. (1) are calculated with PySCF58. The mapping from Fermion operators to Pauli operators are performed using the Jordan-Wigner transformation implemented in the OpenFermion code59. In ADAPT-C optimization, the limited-memory Broyden-Fletcher-Goldfarb-Shanno (L-BFGS-B) algorithm in SciPy60 is used. The GTH-SVZ basis set is used together with the GTH pseudopotential. Full configuration interaction (FCI) results are obtained by direct diagonalization of the system Hamiltonian in the qubit space. The convergence threshold of residual gradient ε\varepsilon in Eq.(14) is taken to be 1×1031\times 10^{-3} Hartree in both potential energy surface and band structure calculations.

A unit cell with two hydrogen atoms are used to build the one-dimensional hydrogen chain. A 1×1×41\times 1\times 4 kk-point grid is used for Brillouin zone sampling. We use the notation ADAPT{XX}, ADAPT-C{XX} and ADAPT-K2G{XX} to represent different types of ADAPT-VQE calculations where XX\in {SD,GSD}. For example, ADAPT-C{GSD} stands for an ADAPT-C calculation with general single and double excitation cluster operators.

Refer to caption
Refer to caption
Figure 1: (a) Ground-state potential energy surface of one-dimensional hydrogen chain calculated with different methods. (b) Absolute energy error compared with the FCI result. The chemical accuracy (1 kcal/mol) is marked by a dashed line.

As shown in Figure 1, for both ADAPT and ADAPT-C, the generalized {GSD} gives better results than {SD}, which is in consistent with previous UCC results38, 47. More importantly, the ADAPT-C ansätze significantly improves the accuracy over ADAPT by 1 to 3 orders of magnitude, since the former eliminating the residual error introduced by complex orbitals. In the whole range of H-H distance from 0.40 to 2.70 Ångstrom, the chemical accuracy is reached with ADAPT-C{GSD} and the mean absolute error (MAE) throughout the potential energy curve is only 0.06 kcal/mol, which is much smaller than that of ADAPT{GSD} (Table 1).

Table 1: Mean absolute error (MAE) and maximum absolute error (MaxAE) in ground-state potential energy surface of one-dimensional hydrogen chain in kcal/mol.
HF ADAPT {SD} ADAPT {GSD} ADAPT-C{SD} ADAPT-C{GSD} ADAPT-K2G{GSD}
MAE 51.57 2.39 2.13 0.89 0.06 0.12
MaxAE 121.11 4.28 3.09 1.70 0.23 0.37

While the accuracy of the K2G transformation technique which we proposed previously38 is at the same order of magnitude as ADAPT-C, the required computational resource is increased roughly by a factor of NkN_{k}, where NkN_{k} is the total number of kk-points. The numbers of terms in the Hamiltonian and the size of the operator pool scale as O(Nk3No4)O(N_{k}^{3}N_{o}^{4}) for ADAPT and ADAPT-C, while for the K2G transformation technique this will be increased to O(Nk4No4)O(N_{k}^{4}N_{o}^{4}), where NoN_{o} is the number of orbitals in a unit cell. The difference becomes even more significant if higher excitation operators such as triples and quadruples are included. Therefore, the ADAPT-C ansätze provides an accurate and efficient solution to periodic system ground-state energy calculation.

Band structures of diamond and silicon are calculated at the experimental lattice constants (a=5.431a=5.431 Å for silicon and a=3.567a=3.567 Å for diamond). At each specific kk-point 𝒌\bm{k}, a two-step process is executed. (1) Perform an ADAPT-C{GSD} calculation to obtain the ground-state wave function |ΨC|\Psi_{C}\rangle with a 1×1×11\times 1\times 1 kk-point sampling grid centered at 𝒌\bm{k}. (2) Perform EOM-ADAPT-C calculations with ~^IP|ΨC\hat{\tilde{\mathcal{R}}}_{IP}|\Psi_{C}\rangle and ~^EA|ΨC\hat{\tilde{\mathcal{R}}}_{EA}|\Psi_{C}\rangle as the target states. Solve the generalized eigenvalue problem defined in Eq.(28) and compare QPWTQPWT to obtain the IP and EA energy spectrum at 𝒌\bm{k}.

Refer to caption
Refer to caption
Figure 2: Band structures of (a) silicon and (b) diamond calculated using different methods.

As shown in Figure 2, the band structures calculated with EOM-ADAPT-C agree well with classical EOM-CCSD results. The average deviation is within 0.05 eV. If DFT at the generalized gradient approximation PBE level61 is used to calculate the band structure with the same kk-point scheme, a significant deduction of the band gap is observed as expected. On the other hand, if Eq. (24) is used directly without adopting the projected operator Eq. (25), the so-called EOM-NP result is significantly deviated from the EOM-ADAPT-C result. For example, the band gap of Si is changed from 1.04 to 0.11 eV.

Notice that noise in quantum device should be considered for real applications. Therefore, a noise-included EOM-ADAPT-C simulation to calculate the IP and EA energies of a one-dimensional hydrogen chain is performed. As shown in Figure S1, noise can indeed introduce non-negligible errors in the EOM-ADAPT-C algorithm. These errors can also be significantly reduced by applying error mitigation techniques as for other quantum algorithms.

Relatively small kk-grids and the minimal basis set are used in this study, which are only suitable for algorithm demonstration and they are not expected to predict accurate electronic structures. Even if clever Brillouin zone sampling technique and accurate Wannier62 basis functions can be used and, at the same time, the requirement to use a high-order truncation in the wave function and excitation operator ansätze for strongly correlated systems is not considered, classical simulation of EOM-ADAPT-C algorithm is still strongly restricted by the computational resource cost which grows exponentially. Actually, running an EOM-ADAPT-C simulation on a classical computer is even more expensive than a classical EOM-CC calculation. For example, if the DZVP basis set with a 3×3×33\times 3\times 3 kk-point grid is used for C and Si band structure calculations as in the previous classical EOM-CCSD study,7 a total number of 1404 qubits will be required, leading to a memory consumption of approximately 21404×162^{1404}\times 16 bytes in a brute-force simulation, which is far beyond the capacity of any existing supercomputer.

However, if the EOM-ADAPT-C algorithm is run on a quantum computer, the expensive construction of Eq.(28) can be realized in parallel via a polynomial number of measurements on a quantum processor. Therefore, the EOM-ADAPT-C algorithm is expected to be a very useful method to obtain highly accurate band structures in the quantum computer era.

In summary, we have proposed the ADAPT-C approach for ground state quantum simulation of solid states and developed a quantum EOM algorithm for quasiparticle band structure calculations. In the ADAPT-C variational ansätze, we introduce a set of complementary operators to the original anti-Hermitian operator pool. The complementary operators are targeted at describing the imaginary part of the anti-Hermitian contracted Schrödinger equation. Optimizing the energy functional of ADAPT-C wave functions, we reach highly accurate ground-state energy as in the K2G transformation case with a significantly reduced computational cost and improved flexibility. EOM calculations can be performed using the accurate ADAPT-C ground-state wave function as the reference state. A QSE-style EOM equation is derived by adopting projected EOM-IP and EOM-EA operators to make sure that the killer condition is satisfied. As a demonstration, band structures of silicon and diamond are calculated, which is in good agreement with the classical EOM-CCSD result. The EOM-ADAPT-C method developed in this study represents an attractive way for accurate quantum simulation of band structures.

{acknowledgement}

This work was partially supported by the NSFC (21825302), by the Fundamental Research Funds for the Central Universities (WK2060000018), and by the USTC Supercomputing Center.

Appendix A Appendices

A.1 A1. Derivation of the EOM-ADAPT-C working equation

We starting from Eq. (24) in the main text

ΔEx\displaystyle\Delta E_{x} =Ψ|[~^,H^,~^]+|ΨΨ|[~^,~^]+|Ψ\displaystyle=\frac{\langle\Psi|[\hat{\tilde{\mathcal{R}}}^{\dagger},\hat{H},\hat{\tilde{\mathcal{R}}}]_{+}|\Psi\rangle}{\langle\Psi|[\hat{\tilde{\mathcal{R}}}^{\dagger},\hat{\tilde{\mathcal{R}}}]_{+}|\Psi\rangle} (A1)

The numerator NN and the denominator DD of Eq. (A1) are

N=Ψ|[~^,H^,~^]+|Ψ\displaystyle N=\langle\Psi|[\hat{\tilde{\mathcal{R}}}^{\dagger},\hat{H},\hat{\tilde{\mathcal{R}}}]_{+}|\Psi\rangle (A2)
D=Ψ|[~^,~^]+|Ψ\displaystyle D=\langle\Psi|[\hat{\tilde{\mathcal{R}}}^{\dagger},\hat{\tilde{\mathcal{R}}}]_{+}|\Psi\rangle

Expand the double commutator for the numerator NN:

2N\displaystyle 2N =2Ψ|[~^,H^,~^]+|Ψ\displaystyle=2\langle\Psi|[\hat{\tilde{\mathcal{R}}}^{\dagger},\hat{H},\hat{\tilde{\mathcal{R}}}]_{+}|\Psi\rangle (A3)
=Ψ|[~^,H^]~^+~^[~^,H^]|Ψ+Ψ|~^[H^,~^]+[H^,~^]~^|Ψ\displaystyle=\langle\Psi|[\hat{\tilde{\mathcal{R}}}^{\dagger},\hat{H}]\hat{\tilde{\mathcal{R}}}+\hat{\tilde{\mathcal{R}}}[\hat{\tilde{\mathcal{R}}}^{\dagger},\hat{H}]|\Psi\rangle+\langle\Psi|\hat{\tilde{\mathcal{R}}}^{\dagger}[\hat{H},\hat{\tilde{\mathcal{R}}}]+[\hat{H},\hat{\tilde{\mathcal{R}}}]\hat{\tilde{\mathcal{R}}}^{\dagger}|\Psi\rangle
=Ψ|[~^,H^]~^|Ψ+Ψ|~^[H^,~^]|Ψ\displaystyle=\langle\Psi|[\hat{\tilde{\mathcal{R}}}^{\dagger},\hat{H}]\hat{\tilde{\mathcal{R}}}|\Psi\rangle+\langle\Psi|\hat{\tilde{\mathcal{R}}}^{\dagger}[\hat{H},\hat{\tilde{\mathcal{R}}}]|\Psi\rangle
=2Ψ|~^H^~^|ΨΨ|~^~^H^|ΨΨ|H^~^~^|Ψ\displaystyle=2\langle\Psi|\hat{\tilde{\mathcal{R}}}^{\dagger}\hat{H}\hat{\tilde{\mathcal{R}}}|\Psi\rangle-\langle\Psi|\hat{\tilde{\mathcal{R}}}^{\dagger}\hat{\tilde{\mathcal{R}}}\hat{H}|\Psi\rangle-\langle\Psi|\hat{H}\hat{\tilde{\mathcal{R}}}^{\dagger}\hat{\tilde{\mathcal{R}}}|\Psi\rangle

where, the killer condition that ~^|Ψ=Ψ|~^=0\hat{\tilde{\mathcal{R}}}^{\dagger}|\Psi\rangle=\langle\Psi|\hat{\tilde{\mathcal{R}}}=0 is applied in the third line. Now substitute ~^=^|ΨΨ\hat{\tilde{\mathcal{R}}}=\hat{\mathcal{R}}|\Psi\rangle\langle\Psi into the equality

Ψ|~^H^~^|Ψ\displaystyle\langle\Psi|\hat{\tilde{\mathcal{R}}}^{\dagger}\hat{H}\hat{\tilde{\mathcal{R}}}|\Psi\rangle =Ψ|ΨΨ|^H^^|ΨΨ|Ψ\displaystyle=\langle\Psi|\Psi\rangle\langle\Psi|\hat{\mathcal{R}}^{\dagger}\hat{H}\hat{\mathcal{R}}|\Psi\rangle\langle\Psi|\Psi\rangle (A4)
=Ψ|^H^^|Ψ\displaystyle=\langle\Psi|\hat{\mathcal{R}}^{\dagger}\hat{H}\hat{\mathcal{R}}|\Psi\rangle
Ψ|~^~^H^|Ψ\displaystyle\langle\Psi|\hat{\tilde{\mathcal{R}}}^{\dagger}\hat{\tilde{\mathcal{R}}}\hat{H}|\Psi\rangle =Ψ|ΨΨ|^^|ΨΨ|H^|Ψ\displaystyle=\langle\Psi|\Psi\rangle\langle\Psi|\hat{\mathcal{R}}^{\dagger}\hat{\mathcal{R}}|\Psi\rangle\langle\Psi|\hat{H}|\Psi\rangle (A5)
=Ψ|H^|ΨΨ|^^|Ψ\displaystyle=\langle\Psi|\hat{H}|\Psi\rangle\langle\Psi|\hat{\mathcal{R}}^{\dagger}\hat{\mathcal{R}}|\Psi\rangle
Ψ|H^~^~^|Ψ\displaystyle\langle\Psi|\hat{H}\hat{\tilde{\mathcal{R}}}^{\dagger}\hat{\tilde{\mathcal{R}}}|\Psi\rangle =Ψ|H^|ΨΨ|^^|ΨΨ|Ψ\displaystyle=\langle\Psi|\hat{H}|\Psi\rangle\langle\Psi|\hat{\mathcal{R}}^{\dagger}\hat{\mathcal{R}}|\Psi\rangle\langle\Psi|\Psi\rangle (A6)
=Ψ|H^|ΨΨ|^^|Ψ\displaystyle=\langle\Psi|\hat{H}|\Psi\rangle\langle\Psi|\hat{\mathcal{R}}^{\dagger}\hat{\mathcal{R}}|\Psi\rangle

The numerator NN finally becomes

N\displaystyle N =Ψ|[~^,H^,~^]+|Ψ\displaystyle=\langle\Psi|[\hat{\tilde{\mathcal{R}}}^{\dagger},\hat{H},\hat{\tilde{\mathcal{R}}}]_{+}|\Psi\rangle (A7)
=Ψ|^H^^|ΨΨ|H^|ΨΨ|^^|Ψ\displaystyle=\langle\Psi|\hat{\mathcal{R}}^{\dagger}\hat{H}\hat{\mathcal{R}}|\Psi\rangle-\langle\Psi|\hat{H}|\Psi\rangle\langle\Psi|\hat{\mathcal{R}}^{\dagger}\hat{\mathcal{R}}|\Psi\rangle

Similarly, the denominator DD can be reduced to

D\displaystyle D =Ψ|[~^,~^]+|Ψ\displaystyle=\langle\Psi|[\hat{\tilde{\mathcal{R}}}^{\dagger},\hat{\tilde{\mathcal{R}}}]_{+}|\Psi\rangle (A8)
=Ψ|^^|Ψ\displaystyle=\langle\Psi|\hat{\mathcal{R}}^{\dagger}\hat{\mathcal{R}}|\Psi\rangle

Therefore, the final working equation becomes

ΔEx\displaystyle\Delta E_{x} =ND=Ψ|^H^^|ΨΨ|H^|ΨΨ|^^|ΨΨ|^^|Ψ\displaystyle=\frac{N}{D}=\frac{\langle\Psi|\hat{\mathcal{R}}^{\dagger}\hat{H}\hat{\mathcal{R}}|\Psi\rangle-\langle\Psi|\hat{H}|\Psi\rangle\langle\Psi|\hat{\mathcal{R}}^{\dagger}\hat{\mathcal{R}}|\Psi\rangle}{\langle\Psi|\hat{\mathcal{R}}^{\dagger}\hat{\mathcal{R}}|\Psi\rangle} (A9)
=Ψ|^H^^|ΨΨ|^^|ΨΨ|H^|Ψ\displaystyle=\frac{\langle\Psi|\hat{\mathcal{R}}^{\dagger}\hat{H}\hat{\mathcal{R}}|\Psi\rangle}{\langle\Psi|\hat{\mathcal{R}}^{\dagger}\hat{\mathcal{R}}|\Psi\rangle}-\langle\Psi|\hat{H}|\Psi\rangle

which is the second or third equality of Eq. (26) in the main text.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure A1: (a), (b) Ground state potential energy curve and error with respect to FCI results of the one-dimensional hydrogen chain. (c), (d) The lowest IP calculated using EOM-ADAPT-C. The results are compared with classical EOM-IP-CCSD method. (e), (f) Same as (c) or (d), but for the lowest EA.

A.2 A2. Noise-included simulation of EOM-ADAPT-C

Simulation in the presence of noise is performed for the one-dimensional hydrogen chain. The ground-state potential energy curve is calculated using ADAPT-C, and the lowest IP (ionization potential) and EA (electron affinity) states are obtained by subsequent EOM-ADAPT-C calculations. The kk-point grid used in this test is reduced to 1×1×11\times 1\times 1, leading to a 4-qubit quantum circuit. The noise model is implemented by including depolarizing gate errors for all qubits participating in the gates, with the depolarizing parameter λ=0.001\lambda=0.001. Error mitigation is performed using the ZNE (zero-noise extrapolation)63, 64 technique, with scaling factors [1.0,1.25,1.50][1.0,1.25,1.50] and linear extrapolation method. Expectation values are averaged over 2172^{17} trials, and the experiment is repeatedly carried out for 16 times. All simulations are performed using the Qiskit toolkit65 with the built-in QASM simulator and the Mitiq package66.

The results of the calculated ground-state and excitation energies are summarized in Figure.(SA1). For the ground state, one double excitation cluster operator is selected by the ADAPT algorithm. The noise introduced by depolarizing gate errors leads to remarkable errors in both the ground-state and the IP or EA energies. A linear extrapolation model is able to reduce the errors significantly.

References

  • Hohenberg and Kohn 1964 Hohenberg, P.; Kohn, W. Inhomogeneous Electron Gas. Phys. Rev. 1964, 136, B864–B871
  • Kohn and Sham 1965 Kohn, W.; Sham, L. J. Self-Consistent Equations Including Exchange and Correlation Effects. Phys. Rev. 1965, 140, A1133–A1138
  • Kohn et al. 1996 Kohn, W.; Becke, A. D.; Parr, R. G. Density Functional Theory of Electronic Structure. J. Phys. Chem. 1996, 100, 12974–12980
  • Cohen et al. 2012 Cohen, A. J.; Mori-Sanchez, P.; Yang, W. Challenges for density functional theory. Chem. Rev. 2012, 112, 289–320
  • Klimeš et al. 2014 Klimeš, J.; Kaltak, M.; Kresse, G. Predictive GWGW calculations using plane waves and pseudopotentials. Phys. Rev. B 2014, 90, 075125
  • Stan et al. 2009 Stan, A.; Dahlen, N. E.; van Leeuwen, R. Levels of self-consistency in the GW approximation. J. Phys. Chem. 2009, 130, 114105
  • McClain et al. 2017 McClain, J.; Sun, Q.; Chan, G. K.-L.; Berkelbach, T. C. Gaussian-Based Coupled-Cluster Theory for the Ground-State and Band Structure of Solids. J. Chem. Theory Comput. 2017, 13, 1209–1218
  • Dittmer et al. 2019 Dittmer, A.; Izsák, R.; Neese, F.; Maganas, D. Accurate Band Gap Predictions of Semiconductors in the Framework of the Similarity Transformed Equation of Motion Coupled Cluster Theory. Inorg. Chem. 2019, 58, 9303–9315
  • Wang and Berkelbach 2020 Wang, X.; Berkelbach, T. C. Excitons in Solids from Periodic Equation-of-Motion Coupled-Cluster Theory. J. Chem. Theory Comput. 2020, 16, 3095–3103
  • Gallo et al. 2021 Gallo, A.; Hummel, F.; Irmler, A.; Grüneis, A. A periodic equation-of-motion coupled-cluster implementation applied to F-centers in alkaline earth oxides. J. Chem. Phys. 2021, 154, 064106
  • McClain et al. 2016 McClain, J.; Lischner, J.; Watson, T.; Matthews, D. A.; Ronca, E.; Louie, S. G.; Berkelbach, T. C.; Chan, G. K.-L. Spectral functions of the uniform electron gas via coupled-cluster theory and comparison to the GW and related approximations. Phys. Rev. B 2016, 93, 235139
  • Lange and Berkelbach 2018 Lange, M. F.; Berkelbach, T. C. On the Relation between Equation-of-Motion Coupled-Cluster Theory and the GW Approximation. J. Chem. Theory Comput. 2018, 14, 4224–4236
  • Bravyi and Kitaev 2002 Bravyi, S. B.; Kitaev, A. Y. Fermionic Quantum Computation. Ann. Phys. 2002, 298, 210–226
  • McArdle et al. 2020 McArdle, S.; Endo, S.; Aspuru-Guzik, A.; Benjamin, S. C.; Yuan, X. Quantum computational chemistry. Rev. Mod. Phys. 2020, 92, 015003
  • Cao et al. 2019 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, 10856–10915
  • Preskill 2018 Preskill, J. Quantum Computing in the NISQ era and beyond. Quantum 2018, 2, 79
  • Georgescu et al. 2014 Georgescu, I. M.; Ashhab, S.; Nori, F. Quantum simulation. Rev. Mod. Phys. 2014, 86, 153–185
  • Aspuru-Guzik et al. 2005 Aspuru-Guzik, A.; Dutoi, A. D.; Love, P. J.; Head-Gordon, M. Simulated Quantum Computation of Molecular Energies. Science 2005, 309, 1704–1707
  • Wang et al. 2008 Wang, H.; Kais, S.; Aspuru-Guzik, A.; Hoffmann, M. R. Quantum algorithm for obtaining the energy spectrum of molecular systems. Phys. Chem. Chem. Phys. 2008, 10, 5388–5393
  • Peruzzo et al. 2014 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
  • Hempel et al. 2018 Hempel, C.; Maier, C.; Romero, J.; McClean, J.; Monz, T.; Shen, H.; Jurcevic, P.; Lanyon, B. P.; Love, P.; Babbush, R.; Aspuru-Guzik, A.; Blatt, R.; Roos, C. F. Quantum Chemistry Calculations on a Trapped-Ion Quantum Simulator. Phys. Rev. X 2018, 8, 031022
  • Nam et al. 2020 Nam, Y. et al. Ground-state energy estimation of the water molecule on a trapped-ion quantum computer. NPJ Quantum Inf. 2020, 6, 33
  • Shen et al. 2017 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: At., Mol., Opt. Phys. 2017, 95, 020501
  • O’ Malley et al. 2016 O’ Malley, P. J. J. et al. Scalable Quantum Simulation of Molecular Energies. Phys. Rev. X 2016, 6, 031007
  • Kandala et al. 2017 Kandala, A.; Mezzacapo, A.; Temme, K.; Takita, M.; Brink, M.; Chow, J. M.; Gambetta, J. M. Hardware-efficient variational quantum eigensolver for small molecules and quantum magnets. Nature 2017, 549, 242–246
  • Colless et al. 2018 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
  • McClean et al. 2016 McClean, J. R.; Romero, J.; Babbush, R.; Aspuru-Guzik, A. The theory of variational hybrid quantum-classical algorithms. New J. Phys. 2016, 18, 023023
  • Lanyon et al. 2010 Lanyon, B. P.; Whitfield, J. D.; Gillett, G. G.; Goggin, M. E.; Almeida, M. P.; Kassal, I.; Biamonte, J. D.; Mohseni, M.; Powell, B. J.; Barbieri, M.; Aspuru-Guzik, A.; White, A. G. Towards quantum chemistry on a quantum computer. Nat. Chem. 2010, 2, 106–111
  • Romero et al. 2018 Romero, J.; Babbush, R.; McClean, J. R.; Hempel, C.; Love, P. J.; Aspuru-Guzik, A. Strategies for quantum computing molecular energies using the unitary coupled cluster ansatz. Quantum Sci. Technol. 2018, 4, 014008
  • Higgott et al. 2019 Higgott, O.; Wang, D.; Brierley, S. Variational Quantum Computation of Excited States. Quantum 2019, 3, 156
  • McClean et al. 2017 McClean, J. R.; Kimchi-Schwartz, M. E.; Carter, J.; de Jong, W. A. Hybrid quantum-classical hierarchy for mitigation of decoherence and determination of excited states. Phys. Rev. A 2017, 95, 042308
  • Yung et al. 2014 Yung, M.-H.; Casanova, J.; Mezzacapo, A.; McClean, J.; Lamata, L.; Aspuru-Guzik, A.; Solano, E. From transistor to trapped-ion computers for quantum chemistry. Scientific Reports 2014, 4, 3589
  • Kutzelnigg 1982 Kutzelnigg, W. Quantum chemistry in Fock space. I. The universal wave and energy operators. J. Chem. Phys. 1982, 77, 3081–3097
  • Bartlett et al. 1989 Bartlett, R. J.; Kucharski, S. A.; Noga, J. Alternative coupled-cluster ansätze II. The unitary coupled-cluster method. Chem. Phys. Lett. 1989, 155, 133–140
  • Taube and Bartlett 2006 Taube, A. G.; Bartlett, R. J. New perspectives on unitary coupled-cluster theory. Int. J. Quantum Chem. 2006, 106, 3393–3401
  • Grimsley et al. 2019 Grimsley, H. R.; Economou, S. E.; Barnes, E.; Mayhall, N. J. An adaptive variational algorithm for exact molecular simulations on a quantum computer. Nat. Commun. 2019, 10, 3007
  • Cerasoli et al. 2020 Cerasoli, F. T.; Sherbert, K.; Sławińska, J.; Nardelli, M. B. Quantum computation of silicon electronic band structure. Phys. Chem. Chem. Phys. 2020, 22, 21816–21822
  • Liu et al. 2020 Liu, J.; Wan, L.; Li, Z.; Yang, J. Simulating Periodic Systems on a Quantum Computer Using Molecular Orbitals. J. Chem. Theory Comput. 2020, 16, 6904–6914
  • Szekeres et al. 2001 Szekeres, Z.; Ágnes Szabados,; Kállay, M.; Surján, P. R. On the “killer condition” in the equation-of-motion method: ionization potentials from multi-reference wave functions. Physical Chemistry Chemical Physics 2001, 3, 696–701
  • Levchenko and Krylov 2004 Levchenko, S. V.; Krylov, A. I. Equation-of-motion spin-flip coupled-cluster model with single and double substitutions: Theory and application to cyclobutadiene. J. Chem. Phys. 2004, 120, 175–185
  • Jordan and Wigner 1928 Jordan, P.; Wigner, E. Über das Paulische Äquivalenzverbot. Zeitschrift für Physik 1928, 47, 631–651
  • Seeley et al. 2012 Seeley, J. T.; Richard, M. J.; Love, P. J. The Bravyi-Kitaev transformation for quantum computation of electronic structure. J. Chem. Phys. 2012, 137, 224109
  • Tranter et al. 2018 Tranter, A.; Love, P. J.; Mintert, F.; Coveney, P. V. A Comparison of the Bravyi–Kitaev and Jordan–Wigner Transformations for the Quantum Simulation of Quantum Chemistry. J. Chem. Theory Comput. 2018, 14, 5617–5630
  • Mazziotti 2004 Mazziotti, D. A. Exactness of wave functions from two-body exponential transformations in many-body quantum theory. Phys. Rev. A: At., Mol., Opt. Phys. 2004, 69, 012507
  • Mukherjee and Kutzelnigg 2004 Mukherjee, D.; Kutzelnigg, W. Some comments on the coupled cluster with generalized singles and doubles (CCGSD) ansatz. Chem. Phys. Lett. 2004, 397, 174–179
  • Ronen 2003 Ronen, S. Can the Eigenstates of a Many-Body Hamiltonian Be Represented Exactly Using a General Two-Body Cluster Expansion? Phys. Rev. Lett. 2003, 91, 123002
  • Lee et al. 2019 Lee, J.; Huggins, W. J.; Head-Gordon, M.; Whaley, K. B. Generalized Unitary Coupled Cluster Wave functions for Quantum Computation. J. Chem. Theory Comput. 2019, 15, 311–324
  • Grimsley et al. 2020 Grimsley, H. R.; Claudino, D.; Economou, S. E.; Barnes, E.; Mayhall, N. J. Is the Trotterized UCCSD Ansatz Chemically Well-Defined? J. Chem. Theory Comput. 2020, 16, 1–6
  • Babbush et al. 2015 Babbush, R.; McClean, J.; Wecker, D.; Aspuru-Guzik, A.; Wiebe, N. Chemical basis of Trotter-Suzuki errors in quantum chemistry simulation. Phys. Rev. A 2015, 91, 022311
  • Mazziotti 2007 Mazziotti, D. A. Anti-Hermitian part of the contracted Schrödinger equation for the direct calculation of two-electron reduced density matrices. Phys. Rev. A 2007, 75, 022505
  • Manrique et al. 2021 Manrique, D. Z.; Khan, I. T.; Yamamoto, K.; Wichitwechkarn, V.; Ramo, D. M. Momentum-Space Unitary Coupled Cluster and Translational Quantum Subspace Expansion for Periodic Systems on Quantum Computers. arXiv:quant-ph 2021, arXiv:2008.08694
  • ROWE 1968 ROWE, D. J. Equations-of-Motion Method and the Extended Shell Model. Rev. Mod. Phys. 1968, 40, 153–166
  • Stanton and Bartlett 1993 Stanton, J. F.; Bartlett, R. J. The equation of motion coupled-cluster method. A systematic biorthogonal approach to molecular excitation energies, transition probabilities, and excited state properties. J. Chem. Phys. 1993, 98, 7029–7039
  • Krylov 2008 Krylov, A. I. Equation-of-Motion Coupled-Cluster Methods for Open-Shell and Electronically Excited Species: The Hitchhiker’s Guide to Fock Space. Annu. Rev. Phys. Chem. 2008, 59, 433–462
  • McWeeny 1992 McWeeny, R. Methods of molecular quantum mechanics; Academic Press, 1992
  • Ollitrault et al. 2020 Ollitrault, P. J.; Kandala, A.; Chen, C.-F.; Barkoutsos, P. K.; Mezzacapo, A.; Pistoia, M.; Sheldon, S.; Woerner, S.; Gambetta, J.; Tavernelli, I. Quantum equation of motion for computing molecular excitation energies on a noisy quantum processor. Phys. Rev. Res. 2020, 2, 043140
  • Hodecker and Dreuw 2020 Hodecker, M.; Dreuw, A. Unitary coupled cluster ground- and excited-state molecular properties. J. Chem. Phys 2020, 153, 084112
  • Sun et al. 2018 Sun, Q.; Berkelbach, T. C.; Blunt, N. S.; Booth, G. H.; Guo, S.; Li, Z.; Liu, J.; McClain, J. D.; Sayfutyarova, E. R.; Sharma, S.; Wouters, S.; Chan, G. K.-L. PySCF: the Python-based simulations of chemistry framework. Wiley Interdisciplinary Reviews: Computational Molecular Science 2018, 8, e1340
  • McClean et al. 2017 McClean, J. R. et al. OpenFermion: The Electronic Structure Package for Quantum Computers. arXiv:quant-ph 2017, arXiv:1710.07629
  • Virtanen et al. 2020 Virtanen, P.; Gommers, R.; Oliphant, T. E.; Haberland, M.; Reddy, T.; Cournapeau, D.; Burovski, E.; Peterson, P.; Weckesser, W.; Bright, J.; van der Walt, S. J.; Brett, M.; Wilson, J.; Jarrod Millman, K., et al. SciPy 1.0: Fundamental Algorithms for Scientific Computing in Python. Nature Methods 2020, 17, 261–272
  • Perdew et al. 1996 Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868
  • Marzari et al. 2012 Marzari, N.; Mostofi, A. A.; Yates, J. R.; Souza, I.; Vanderbilt, D. Maximally localized Wannier functions: Theory and applications. Rev. Mod. Phys. 2012, 84, 1419–1475
  • Temme et al. 2017 Temme, K.; Bravyi, S.; Gambetta, J. M. Error Mitigation for Short-Depth Quantum Circuits. Phys. Rev. Lett. 2017, 119, 180509
  • Kandala et al. 2019 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
  • Aleksandrowicz et al. 2019 Aleksandrowicz, G. et al. Qiskit: An Open-source Framework for Quantum Computing. 2019,
  • LaRose et al. 2020 LaRose, R.; Mari, A.; Kaiser, S.; Karalekas, P. J.; Alves, A. A.; Czarnik, P.; Mandouh, M. E.; Gordon, M. H.; Hindy, Y.; Robertson, A.; Thakre, P.; Shammah, N.; Zeng, W. J. Mitiq: A software package for error mitigation on noisy quantum computers. arXiv:quant-ph 2020, arXiv:2009.04417