You are on page 1of 38

Spin coupling is all you need:

Encoding strong electron correlation on quantum computers


Daniel Marti-Dafcik,1, ∗ Hugh G. A. Burton,2 , David P. Tew1
1
Physical and Theoretical Chemistry Laboratory, University of Oxford, South Parks Road, Oxford, OX1 3QZ, U.K.
2
Yusuf Hamied Department of Chemistry, University of Cambridge, Lensfield Road, Cambridge, CB2 1EW, U.K.
(Dated: April 30, 2024)
The performance of quantum algorithms for eigenvalue problems, such as computing Hamiltonian
spectra, depends strongly on the overlap of the initial wavefunction and the target eigenvector. In
a basis of Slater determinants, the representation of energy eigenstates of systems with N strongly
correlated electrons requires a number of determinants that scales exponentially with N . On classi-
cal processors, this restricts simulations to systems where N is small. Here, we show that quantum
computers can efficiently simulate strongly correlated molecular systems by directly encoding the
arXiv:2404.18878v1 [quant-ph] 29 Apr 2024

dominant entanglement structure in the form of spin-coupled initial states. This avoids resorting
to expensive classical or quantum state preparation heuristics and instead exploits symmetries in
the wavefunction. We provide
 quantum circuits for deterministic preparation of a family of spin
N
eigenfunctions with N/2 Slater determinants with depth O(N ) and O(N 2 ) local gates. Their use
as highly entangled initial states in quantum algorithms reduces the total runtime of quantum phase
estimation and related fault-tolerant methods by orders of magnitude. Furthermore, we assess the
application of spin-coupled wavefunctions as initial states for a range of heuristic quantum algo-
rithms, namely the variational quantum eigensolver, adiabatic state preparation, and different ver-
sions of quantum subspace diagonalization (QSD) including QSD based on real-time-evolved states.
We also propose a novel QSD algorithm that exploits states obtained through adaptive quantum
eigensolvers. For all algorithms, we demonstrate that using spin-coupled initial states drastically
reduces the quantum resources required to simulate strongly correlated ground and excited states.
Our work paves the way towards scalable quantum simulation of electronic structure for classically
challenging systems.

I. INTRODUCTION exact eigenstate.[5, 6] Therefore, the success of quan-


tum algorithms relies on finding an initial state which
Computing the low-lying eigenstates and energies of approximately contains the structure of the true eigen-
electronic Hamiltonians remains a fundamental challenge state. This can be achieved by computing approxi-
in the physical sciences and scientific computing. De- mate wavefunctions through classical heuristics, such
spite the success of classical computational approaches as Hartree–Fock, configuration interaction,[7] or density
for quantum chemistry in simulating a wide range of matrix renormalization group (DMRG) approaches,[8]
molecules, their application to systems exhibiting strong and loading such states on quantum hardware.[9] On
electron correlation is limited due to the difficulty of effi- fault-tolerant quantum hardware, quantum heuristics—
ciently representing highly entangled quantum states.[1] such as the variational quantum eigensolver (VQE),[10–
Quantum computers can generate and transform vec- 27] adiabatic state preparation (ASP),[28–35] and quan-
tors whose dimension scales exponentially in the num- tum subspace diagonalization (QSD)[14, 36–48] —could
ber of qubits with remarkable efficiency.[2] This suggests be used to prepare initial states for quantum phase esti-
that they might provide a solution to the curse of di- mation, a method for which the runtime rigorously de-
mensionality. The reality is more nuanced. Despite the pends on the overlap between the initial state and the
plethora of algorithms developed for computing energy target eigenstate.[49–51]
eigenvalues or preparing eigenstates of many-body sys- The challenge is that, for systems where such heuris-
tems, complexity-theoretic results rule out exponential tics are accurate, the problem can often be solved entirely
quantum speedups for worst-case versions of the elec- using classical algorithms (to sufficient accuracy). This
tronic Schrödinger equation.[3, 4] Although this does not casts doubts on the advantage of using quantum comput-
prohibit practical speedups for realistic physical systems, ers over classical machines for quantum chemistry.[35] To
a fundamental problem lies at the core of quantum com- obtain quantum speedups, the approximate initial state
puting for electronic structure: the performance of nearly found through heuristics must be of insufficient accuracy
all quantum algorithms for eigenvalue problems strongly for chemistry applications. At the same time, its struc-
depends on the accuracy of the initial state. ture must be such that further refining the wavefunction
A random vector in the many-electron Hilbert space is through classical algorithms is hard, while encoding it
expected to have exponentially small overlap with any as an initial state for quantum algorithms is easy. In
this work, we tackle the strong correlation problem in
quantum chemistry by leveraging a recently-developed
∗ classical heuristic that satisfies these requirements and
dmartidafcik@gmail.com
2

is ideally suited for initial state preparation on quantum few spin-coupled states. These spin-coupled states can be
computers.[52] deduced from chemical intuition and spatial-spin sym-
To understand what features make an initial state metry arguments, and their representation is given di-
preparation method advantageous for quantum compu- rectly from the standard Clebsch–Gordan coefficients.[62]
tation, consider the success and limitations of scalable This challenges the common notion that strongly corre-
classical algorithms. All polynomially-scaling methods lated states are inevitably complex. While they might be
for quantum chemistry rely on identifying qualitatively highly-dimensional when expressed as vectors expanded
accurate, yet approximate, initial states with a simple, in a single-particle basis, they encode a relatively small
compact wavefunction description. Restricted Hartree– amount of information.
Fock theory provides good initial states for systems Here, we enable their efficient application in quantum
with weak electron correlation, because it accurately en- algorithms by providing quantum circuits  that prepare a
N
codes the mean-field character of delocalized molecular family of spin-coupled states with N/2 determinants in
wavefunctions.[53] The many-body Hartree–Fock state is depth O(N ) and using O(N 2 ) gates. We achieve this
simply an antisymmetrized product of delocalized molec- by exploiting the symmetry structure in spin-coupled
ular orbitals, which corresponds to a single Slater deter- wavefunctions and connecting them to Dicke states, a
minant. Since this state has a compact representation in well-known family of entangled states.[63] This approach
the single-particle basis of Hartree–Fock orbitals, one can avoids the exponential scaling of generic, black-box state
expand around it in a controlled manner, e.g. with meth- preparation methods.[9, 58]
ods based on many-body perturbation theory, to improve We numerically assess the application of this state
the state without incurring exponential cost.[54] preparation method in VQE, ASP, and QSD, where
For strongly correlated systems, the Hartree–Fock the quantum algorithms are initialized with spin-coupled
state is inaccurate because the molecular orbital picture states. We also propose a new QSD algorithm, ADAPT-
underpinning it breaks down. Even a qualitatively ac- QSD, which is of interest in its own right. It builds a
curate wavefunction for such systems requires a super- subspace from states obtained through adaptive quan-
position of Slater determinants that scales exponentially tum eigensolvers such as ADAPT-VQE.[12] For all al-
with the number of strongly correlated electrons, regard- gorithms, we demonstrate that the use of spin-coupled
less of the choice of single-particle basis. In such scenar- states greatly reduces the quantum resources (circuit
ios, no efficient classical heuristics exist, and one resorts depth and gate counts) and the number of degrees of
to brute-force algorithms which in general scale expo- freedom (variational parameters) required to achieve a
nentially as they require a linear combination of all the given accuracy, compared to using the Hartree–Fock ref-
relevant Slater determinants.[55–57] erence. This confirms our spin-coupled framework as
The high dimension of such states also poses a chal- a low-cost approach for improving the performance of
lenge for quantum algorithms due to the initial state de- heuristic quantum algorithms.
pendency. Suppose that an approximate wavefunction We highlight QSD as a class of algorithms that can
is found through a state-of-the-art classical algorithm, most strongly benefit from access to different reference
such as selective configuration interaction or DMRG. states, in particular for multireference systems i.e. those
Even then, its use in quantum algorithms is inefficient, for which multiple initial states are required for an ac-
because preparing a superposition with arbitrary coef- curate wavefunction description. These are generally
ficients through a quantum circuit requires a number hard to tackle using classical algorithms and therefore a
of steps (gates) proportional to the number of basis promising target for quantum computation. Combining
states (determinants).[9, 58] Quantum heuristics could QSD with spin-coupled states also allows computation of
in principle offer an alternative, but often also suffer excited state energies at low cost.
in strongly correlated regimes.[19] Therefore, there ap- Finally, we analyze the state preparation question in
pears to be a trade-off between the accuracy of the initial the context of fault-tolerant computation of electronic
state used in quantum algorithms and the cost of state structure based on quantum phase estimation. We con-
preparation.[59–61] Fundamentally, the challenge is that sider the number of non-Clifford gates required to pre-
quantum computers need to exploit structure in the prob- pare initial states with high ground state overlap for
lem, and while such structure is present in Hartree–Fock systems with many (up to 35) spin-coupled electrons
states, it is washed away in strongly correlated wavefunc- such as FeMoCo.[31, 35, 64] Our extrapolated gate count
tions due to the brute-force nature of algorithms that are estimates suggest that one could greatly reduce the
used for their approximation. state preparation cost by directly encoding the entan-
In this work, we present a heuristic state preparation glement due to spin coupling through a circuit similar
method that solves the initial state problem for a range of to the ones presented here, compared to preparation of
strongly correlated systems. This relies on our recently- states obtained from classical black-box algorithms such
developed generalized molecular orbital theory,[52] where as DMRG or selected CI. This would enable efficient
we found that molecular wavefunctions that arise during fault-tolerant quantum simulation of such strongly corre-
the stretching of chemical bonds are often highly struc- lated molecules—precisely the systems for which classical
tured and can be approximated to high accuracy through methods are most likely to remain insufficient.
3

Overall, our work provides a scalable framework with Since this notation only specifies the occupation of each
the necessary ingredients required to unlock the unique orbital, we must specify the nature and ordering of the
power of quantum computers for challenging chemical orbitals separately. We will do so through an ordered list
systems. which we define by curly brackets {ϕ1 , ϕ2 , ...}.
In Sec. II, we introduce our wavefunction notation
and encoding and the background on spin-coupled states.
Most of the content appeared in Ref. 52 in a language tai- B. Discovering and describing strong electron
lored to chemists; here, we present these concepts to a correlation through symmetry
general quantum science audience. In Sec. III we present
an overview of our main results. In Sec. IV, we provide Common examples of strongly correlated states in-
the quantum circuits for preparing spin eigenfunctions. clude electronic wavefunctions of molecules at dissoci-
In Sec. V, we discuss the numerical results obtained from ation. Here, we consider the nitrogen molecule, which
classical simulations of quantum algorithms, and intro- illustrates the challenges of both classical[55, 66] and
duce the ADAPT-QSD algorithm. In Sec. VI, we analyze quantum[19, 33, 67] algorithms.
the initial state preparation task within the longer-term, In Ref. 52, we showed that the ground state wavefunc-
fault-tolerant quantum computing context. tion can be represented in terms of a few states with both
mean-field and strongly-correlated character. Each state
encodes a unique entanglement pattern that corresponds
II. BACKGROUND AND CONCEPTUAL to a clear physical motif. Bonding is manifest by elec-
FRAMEWORK trons that doubly occupy delocalized molecular orbitals,
and strong correlation corresponds to electrons occupying
A. Encoding and notation for qubit states open-shell, localized orbitals that couple through their
spin angular momentum. A superposition of four refer-
We work with the second-quantized representation, ence states yields an accurate wavefunction approxima-
where basis states are antisymmetric product states tion at any nuclear geometry (Fig. 1), without requiring
(Slater determinants) in a fermionic Fock space and the a large number of variational parameters. Below, we dis-
antisymmetry is manifest through anticommuting oper- cuss the derivation of the four reference states.
ator algebra. Using the Jordan-Wigner mapping,[65] an In a minimal STO-3G basis, the Re-
electronic wavefunction of 2M spin-orbitals (M spatial stricted Hartree–Fock method yields the fol-
orbitals) can be mapped to a system of 2M qubits, where lowing set of delocalized molecular orbitals:
each qubit represents the occupation of a spin-orbital (0 {1σg , 1σu , 2σg , 2σu , 3σg , 1πu,x , 1πu,y , 1πg,x , 1πg,y , 3σu }.
means unoccupied, 1 means occupied). Every computa- We use a Hamiltonian that freezes the lowest four
tional basis state |j⟩ uniquely encodes a Slater determi- orbitals and therefore only work with the 1π-orbitals
nant represented as and the 3σ valence orbitals.[68] This yields a system of
six electrons in six orbitals that can be mapped to twelve
M 
O  qubits. We drop the shell indices from the specification
|j⟩ = |fiα ⟩ ⊗ |fiβ ⟩ , (1) of the orbitals and therefore the delocalized orbitals
i=1 considered here are: {σg , πu,x , πu,y , πg,x , πg,y , σu }. In
this basis, the RHF state doubly occupies the three
where fiα , fiβ ∈ {0, 1} represent the occupation of the lowest orbitals (Fig. 1a, blue), given by
α (spin-up) and β (spin-down) spin-orbital correspond-
ing to the i-th spatial orbital, and the Hamming weight |ΦRHF ⟩ = |222000⟩ (3)
(number of 1s) is the number of electrons. We define
the qubit ordering such that a spin-up orbital is followed using the compact notation introduced in Eq. (2).
by the corresponding spin-down orbital. This convention Although a rigorous definition of electron correlation
allows us to introduce a more compact notation: remains an open research question,[69] one natural defi-
nition is to take the view that any product state in second
M
O quantization is uncorrelated. In this sense, the RHF state
|j⟩ = |gi ⟩ , (2) in Eq. (3) is uncorrelated. This is consistent with it being
i=1 a fermionic mean-field state, and with the widely adopted
definition of the correlation energy Ecorr = E − ERHF . In
where gi ∈ {0, 2, α, β} is the occupation of a spatial
this paradigm, correlation is therefore a measure propor-
orbital: 0 if unoccupied, 2 if doubly occupied, α or
tional to the deviation of an entangled quantum state
β if singly occupied by a spin-up or spin-down elec-
from a product state, using the Hartree–Fock orbitals as
tron. We will typically drop the tensor product i.e.
the prescribed single-particle basis.
|gi ⟩ ⊗ |gj ⟩ =: |gi gj ⟩, and often write products of K iden-
K Around the equilibrium bond length, the RHF state is
tical single-qubit states as |gi ⟩ . The compact notation a relatively accurate wavefunction approximation (Fig. 1,
(Eq. (2)) relates to the standard notation (Eq. (1)) as blue). This reflects the success of molecular orbital the-
follows: |0⟩ → |00⟩, |α⟩ → |10⟩, |β⟩ → |01⟩, |2⟩ → |11⟩. ory in chemistry, which describes bonding as consisting of
4

1.0

0.8

Contribution
0.6

0.4

0.2

0.0
1.0 1.5 2.0 2.5 3.0 3.5 4.0
Bond length [Å]

(b) Coefficients of each CSF in ΦLC .


1.0

0.8

Squared overlap
0.6

0.4

0.2

0.0
1.0 1.5 2.0 2.5 3.0 3.5 4.0
Bond length [Å]

(a) Energies and configurations of the valence CSFs. (c) Squared overlaps |⟨Ψ|Φi ⟩|2 .

FIG. 1: The dissociation of N2 in a STO-3G basis in the valence space of 6 electrons in 6 orbitals (12 qubits) using
four CSFs as predicted by generalised MO theory. Reproduced with permission from Ref. 52.

electrons occupying delocalized orbitals, a useful model calized on the atom, the wavefunction for the atom is
for many molecules at equilibrium. therefore a quartet (total spin s = 3/2) that ferromag-
However, this model breaks down at dissociation. The netically couples three spin-1/2 particles (Fig. 1a, red).
electronic wavefunction exhibits strong correlation ef- This spin-coupled state can be determined directly from
fects, as reflected in the poor energy of the RHF state the Clebsch–Gordan coefficients for angular momentum
(Fig. 1a) and its low overlap with the exact ground state coupling:
(Fig. 1c). The exact singlet ground state in the RHF ba- 3 3
sis is a superposition of 44 Slater determinants of similar 
 | 2 , 2 ⟩ = |ααα⟩
3 1 √1

| 2, 2⟩ = |ααβ⟩ + |αβα⟩ + |βαα⟩

weights. Classical methods encounter challenges in such 3 
3
situations since the wavefunction is no longer compact in |3, ⟩ = 3 1 1

2 | 2 , − 2 ⟩ = √3 |βαα⟩ + |βαβ⟩ + |ββα⟩

the RHF basis. Quantum algorithms inherit these chal- 
3 3
| 2 , − 2 ⟩ = |βββ⟩ ,
lenges if the RHF state is used as the initial state.
Fortunately, many of the low-energy eigenstates that (4)
appear in nature are highly structured, even if they can where on the left we have used the notation |n, s⟩ for
be highly entangled. Approximations thereof can be ob- the particle and spin quantum number, in the center we
tained from simple symmetry considerations. have used |s, ms ⟩ for the total spin and spin projection
Let us now consider the wavefunction for N2 at disso- quantum numbers s and ms , and on the right we use α
ciation. The wavefunction for the dissociated state is not and β to represent the state of each site as either spin-up
the direct product of the atomic states, because the lo- or spin-down.
cal spins of the dissociated atoms are coupled and the The latter representation is equivalent to electronic
wavefunction is entangled across the two atomic sub- Slater determinants with singly-occupied spatial orbitals
systems. The ground state of each isolated nitrogen using the compact notation for fermions/qubits from
atom is well-understood from basic chemical principles: Eq. (2), and assuming a single-particle basis of localized
the three valence electrons repel each other and there- orbitals. These localized orbitals can be obtained from
fore form the maximally antisymmetric spatial wavefunc- the RHF orbitals for N2 through a trivial unitary trans-
tion, which maximizes the average distance between elec- formation of the molecular orbitals: zL = √12 (σg + σu ),
trons. Due to the overall antisymmetry requirement for zR = √12 (σg − σu ), xL = √12 (πu,x + πg,x ), xR = √12 (πu,x −
fermionic wavefunctions, the spin wavefunction must be πg,x ), yL = √12 (πu,y + πg,y ), yR = √12 (πg,y − πg,y ). Here,
fully symmetric. Expressed in a basis of spin-orbitals lo- the delocalized RHF orbitals have been transformed to a
5

set of atomic-like, mutually orthogonal molecular orbitals nant is in general an eigenfunction of Ŝz but not of Ŝ 2 .
x, y, z localized on the left (L) or right (R) atom. Exceptions include fully closed-shell determinants with
In the dissociation limit, the N2 ground state can be S, MS = 0, 0, such as the state in Eq. (3).
any state that couples the two atomic quartets with Configuration state functions (CSFs) are eigenfunc-
proper quantum numbers. However, as the interatomic tions of both Ŝ 2 and Ŝz with quantum numbers
distance R decreases, singlet-coupled electrons can de- formed by a symmetry-adapted linear combination of
localise across the bond, which reduces the kinetic en- determinants.[7] All the reference states considered in
ergy. This one-body delocalization lifts the degener- this work are CSFs. These are products of closed-shell
acy between eigenstates and the ground state requires delocalized orbitals as well as open-shell spin-coupled or-
that the two three-electron subsystems couple to a six- bitals. The spin-coupled part is defined by the pattern
electron singlet. This state, which we will refer to as in which the single electrons/orbitals are coupled to form
6,1
|Φ6 ⟩ = |O0,0 ⟩, can also be determined from the Clebsch– the many-body wavefunction. We define the following
Gordan coefficients and corresponds to a superposition first-quantized representation for the CSFs used in this
of 20 Slater determinants in the localized orbital basis work (Eq. 6 in Ref. 52):
(Eq. (A1c) in Appendix A). In contrast to the mean-field
N,i
RHF state, this spin-coupled wavefunction is exact at |Φ⟩ = N Â |ϕ2c · · ·⟩ OS,M S
(ϕo · · · ), (5)
dissociation but very inaccurate at shorter bond lengths
(Fig. 1). Precisely due to its symmetry origin, the state where S, MS are the spin quantum numbers and N is
at dissociation is predictable despite being strongly cor- the number of spin-coupled electrons (and therefore the
related. It has low Kolmogorov complexity in the sense number of spin-coupled, localized spatial orbitals). Here,
that few steps are required to specify it: one must simply we have included the antisymmetrization operator  and
define the orbitals involved and their spin coupling pat- normalization constant N , but we omit these in the
tern. From this, the coefficients of each basis state can remaining text. The term |ϕ2c · · ·⟩ is a product state
be computed through the standard formula for Clebsch– with doubly occupied (closed-shell) orbitals {ϕc , ...}. The
N,i
Gordan coefficients. term OS,M S
(ϕo · · · ) is a state with singly-occupied (open-
As shown in Ref. 52, the same is true for the dissocia- shell) orbitals {ϕo , ...} that are spin-coupled. The index
tion of a range of molecules including first- and second- i specifies the coupling pattern and the amplitudes of
row diatomics, water, and hydrogen clusters: in the limit the expansion of the CSF in a product basis are simply
of long bond length, they can all be described by a sin- the Clebsch–Gordan coefficients; a list of relevant spin-
gle spin-coupled wavefunction, whose general form we coupled states is provided in Appendix A.
will define below. For N strongly correlated electrons, In the compact, second-quantized qubit representation
N,1 N
the spin-coupled state |O0,0 ⟩ is a superposition of N/2 (Eq. (2)), the closed-shell state maps as |ϕ2c · · ·⟩ → |2 · · ·⟩,
N,i
Slater determinants (where N = 6 for N2 ). This expo- the open-shell state is |OS,M S
(ϕo · · · )⟩, and we explicitly
nential scaling in the number of electrons means that it is include the occupation of the unoccupied (virtual) or-
hard to exploit such states as reference states in classical bitals {ϕv , ...} as |0 · · ·⟩. Therefore, the qubit representa-
algorithms. tion of any CSF with orbitals {ϕc , ..., ϕo , ..., ϕv , ...} takes
At intermediate bond lengths, neither the RHF state the form
6,1
nor the fully spin-coupled state |O0,0 ⟩ are an accurate N,i
description of the ground state. Instead, multiple refer- |Φ⟩ = |2 · · ·⟩ |OS,M S
(ϕo · · · )⟩ |0 · · ·⟩ , (6)
ence states have significant weights in the wavefunction
where the indices c, o and v run over the closed-shell,
expansion. Such situations are referred to as multirefer-
open-shell, and virtual orbitals.
ence systems in quantum chemistry.[70, 71]
In the following, we introduce the notation used for the
general configuration state functions that describe each D. Multireference quantum chemistry
of the contributing states relevant to strongly correlated
regimes where bonds are partially broken.
To describe situations where multiple correlation ef-
fects occur simultaneously, such as bond breaking in the
C. Configuration State Functions intermediate region of the binding curve, we form hy-
brid, partially entangled states. By localizing only the
RHF orbitals involved in the πx -bond to obtain the ba-
The strong correlation corresponding to localized spin-
sis {σg , πu,y , xL , xR , πg,y , σu }, we form a state with one
coupled orbitals requires using entangled states whose
stretched bond (two spin-coupled electrons):
coefficients are determined from symmetry. The Hamil-
tonian operator commutes with both the total spin oper- 2,1 1
ator, Ŝ 2 , and the spin projection operator, Ŝz . Therefore, |Φ2x ⟩ = |σ2g π2u,y ⟩ O0,0 |π2g,y σ2u ⟩ → |22⟩ √ (|αβ⟩−|βα⟩) |00⟩ .
2
energy eigenstates must simultaneously be eigenstates of (7)
the spin operators with corresponding total spin and spin This is traditionally referred to as a diradical state. Due
projection quantum numbers S, MS . A Slater determi- to the cylindrical symmetry of the molecule, whose bond
6

points along the z-axis, the state has an exactly degener- key result from Ref. 52 is that a qualitatively accurate
ate counterpart with spin coupling along the y direction. wavefunction can be built for systems with increasingly
Using orbitals {σg , πu,x , yL , yR , πg,x , σu }, this state has strong electron correlation by parametrizing strong cor-
the same qubit representation relation effects due to spin coupling. Just as MO theory
predicts the Hartree–Fock state to be the dominant elec-
1
|Φ2y ⟩ = |σg πu,x ⟩ O0,0 |πg,x σu ⟩ → |22⟩ √ (|αβ⟩−|βα⟩) |00⟩ . tron configuration, our generalized MO theory predicts
2 2 2,1 2 2
2 the dominant entangled states for strongly spin-coupled
(8) systems from chemical intuition and symmetry consider-
In calculations, we will form a fixed, symmetric linear ations. This type of correlation is typically referred to as
combination of the two states strong or static correlation in quantum chemistry.
1 To obtain quantitatively accurate energy estimates,
|Φ2 ⟩ = √ (|Φ2x ⟩ + |Φ2y ⟩) (9) the remaining (dynamic or weak ) correlation needs to
3 be added to the spin-coupled reference states.[72] This
and treat this as a single reference state without any loss requires powerful computational methods to generate
in circuit/state expressivity. a large amount of additional Slater determinants with
We also form a state describing two stretched bonds small weights. Classical computational methods gener-
by localizing four of the orbitals. With orbitals ally struggle to provide such accuracy due to the unavoid-
{σg , xL , yL , xR , yR , σu }, the state is able memory/time complexity of simultaneously storing
and processing spin-coupled as well as delocalized mean-
|Φ4 ⟩ = |σ2g ⟩ O0,04,1 4,1
(xL , yL , xR , yR ) |σ2u ⟩ → |2⟩ |O0,0 ⟩ |0⟩ , field states, as this requires exponential wavefunction ex-
(10) pansions regardless of the choice of single-particle basis.
4,1 On the other hand, quantum information processors
where |O0,0 ⟩ is given in the Appendix (Eq. (A1b)). At
can efficiently store and transform exponentially large
arbitrary bond lengths, the wavefunction is accurately
quantum states and therefore this limitation is not
described by the optimal linear combination of all CSFs,
present in quantum models of computation.[2] Within
|ΦLC ⟩ = c1 |ΦRHF ⟩ + c2 |Φ2 ⟩ + c3 |Φ4 ⟩ + c4 |Φ6 ⟩ , (11) the digital quantum circuit model, the key requirement
to leverage such reference states is the ability to efficiently
which can be found by diagonalizing the Hamiltonian in prepare them by applying elementary gates on the initial
the subspace built from the four (nonorthogonal) states. product state of the qubits, typically |0⟩ · · · |0⟩. In Sec-
The binding curve (Fig. 1a) clearly displays the ex- tion IV, we show how this can be achieved through quan-
pected behaviour: the RHF state is accurate around the tum circuits of depth O(N ) and using O(N 2 ) gates for
equilibrium distance but poor at dissociation. The spin- systems of N spin-coupled electrons, by mapping such
coupled states |Φ2 ⟩, |Φ4 ⟩ have increasingly longer minima states to the well-studied family of Dicke states. This
and improved energies at dissociation. The fully spin- confirms that, in addition to their low Kolmogorov com-
coupled state |Φ6 ⟩ is exact at dissociation. plexity, the quantum circuit complexity of such states is
The trend seen in the energies is fully matched by the also low owing to their high degree of symmetry.
trend in the coefficients and squared overlaps (Figs. 1b This approach unlocks the power of quantum al-
and 1c). At short bond lengths, the RHF state domi- gorithms for strongly correlated electronic structure.
nates the linear combination with high overlap with the Through extensive classical simulations of quantum al-
exact ground state. As the bond is stretched, the in- gorithms, we demonstrate how the use of spin-coupled
termediate CSFs |Φ2 ⟩ and |Φ4 ⟩ become increasingly im- initial states drastically enhances the performance of
portant, and at long bond lengths the fully spin-coupled quantum algorithms for challenging electronic systems.
state |Φ6 ⟩ dominates, with 100% overlap in the infinite In VQE, it reduces the number of variational parame-
limit. The linear combination is least accurate at the ters, and therefore the number of parametrized quan-
intermediate region; nevertheless the minimum squared tum gates, required to achieve chemical accuracy (Sec-
overlap is 92%. Using the linear combination state |ΦLC ⟩ tion V A). In quantum subspace approaches, it greatly
instead of |ΦRHF ⟩ directly leads to a reduction in the lowers the circuit depth and the number of measurements
runtime of QPE of more than one order of magnitude (Sections V B and V C). In adiabatic state preparation,
at stretched geometries (see Section VI for an extended it increases the accuracy that can be achieved through a
discussion). given number of steps, effectively speeding up the adia-
batic evolution process required to reach a wavefunction
of high accuracy (Section V D).
III. THIS WORK: OVERVIEW OF RESULTS Finally, spin-coupled reference states reduce the cost
of fault-tolerant algorithms based on QPE by orders of
The ability to build a qualitatively accurate reference magnitude (Section VI). Although alternative methods
state for weakly correlated electronic system through for initial state preparation have been proposed by other
mean-field theory is what has driven the success of most authors,[9, 58, 73] fault-tolerant applications relying on
practical approaches to computational chemistry. The QPE will greatly benefit from our approach. This stems
7

from the fact that, by directly encoding the entanglement and Ref. 52).
due to spin coupling through bespoke quantum circuits, The work in Ref. 81 is similar in spirit to ours in that it
our state preparation circuits are much more efficient considers superpositions of states with a varying number
than generic, black-box state preparation techniques. of spin-coupled electrons. However, it defines the states
in terms of delocalized orbitals and relies on symmetry-
broken mean-field solutions. By choosing appropriate lo-
IV. QUANTUM CIRCUITS FOR calized orbitals and obtaining the spin-coupled states di-
PREPARATION OF SPIN EIGENFUNCTIONS rectly from the Clebsch–Gordan coefficients, our work
provides spin-pure wavefunctions derived from a clear
To exploit spin-coupled initial states in quantum al- physical picture.
gorithms, they must be efficiently loaded onto quan- In Section IV A, we analyze the general form of a fam-
tum registers. Despite their exponentially-scaling sup- ily of spin eigenfunctions that captures the entanglement
port in the computational basis, the spin-coupled states structure of molecular bonds. We discuss the connection
considered in this work have a highly symmetric and to Dicke states in Section IV B. In Section IV C, we re-
well-defined structure, and the number of distinct co- view the circuits for Dicke state preparation from Ref. 63.
efficients scales linearly with the number of electrons. We introduce a more explicit notation for operators and
Crucially, we can exploit this structure to derive quan- circuits in Section IV D, discuss the circuits for prepara-
N,1 N,1
tum circuits that prepare the spin eigenfunctions |O0,0 ⟩, tion of the states |O0,0 ⟩ in Section IV E, the mapping
N

which correspond to a superposition of N/2 determi- from the representation in the spin Hilbert space to the
Fock space in Section IV F, the total state preparation
nants, with O(N 2 ) rotation and CNOT gates and depth
cost in Section IV G, the preparation of linear combina-
O(N ). We achieve this by connecting the spin eigenfunc-
N,1 tions of CSFs in Section IV I, as well as generalizations
tions |O0,0 ⟩ to Dicke states, which form a different fam- to other spin eigenfunctions in Section IV H.
ily of entangled wavefunctions.[63] The recursive circuit
structure also reduces the cost of preparing linear com-
binations of spin-coupled states. Furthermore, we con- A. Spin eigenfunctions for bond dissociation
sider their implementation on a fault-tolerant architec-
ture, where there is an additional log(N ) overhead in the Each spin eigenfunction presented in Section II in-
non-Clifford gate count due to the need for synthesizing cludes a set of doubly occupied, delocalized orbitals (typ-
rotation gates through Clifford + T gates (Section IV G ically, Hartree–Fock orbitals), and a set of singly occu-
and Appendix C 4). pied, localized orbitals which capture the strong elec-
We compute the exact CNOT and Toffoli gate counts tron correlation (Eqs. (5) and (6)). The weakly corre-
and find that both are very low in practice for a range lated component can be trivially prepared by applying
of systems sizes relevant to quantum chemistry (Ta- Pauli-X gates on each of the qubits corresponding to
ble I in Section IV G). For example, for N = 34, only the closed-shell Hartree–Fock orbitals. This is the cir-
∼ 103 gates are required to prepare the spin-coupled cuit structure used in most quantum algorithm imple-
34,1
CSF |O0,0 ⟩, which is a superposition of L ∼ 109 Slater mentations for preparing the Hartree–Fock state. The
determinants.[74] entanglement due to spin coupling occurs in the remain-
In contrast, there exist general algorithms for prepar- ing singly-occupied orbitals, which are localized in this
ing arbitrary linear combinations of computational basis work. The state preparation task thus reduces to prepar-
states (Slater determinants).[9, 58, 73, 75–77] Such meth- N,i
ing a particular type of spin eigenfunction |OS,M ⟩ for N
ods do not exploit any particular features of the states
strongly correlated electrons in 2N spin orbitals (qubits).
that are being prepared and their cost scales at best lin-
The physical model that we propose requires a parti-
early with the number of Slater determinants. Therefore,
tioning of the molecular electronic system into subsys-
the scaling is exponential for strongly correlated systems.
tems, and considers the coupling within the subsystem
For the example of N = 34, our method is at least eight
and across the subsystems. For the systems considered
orders of magnitude more efficient than preparing the
in Ref. 52, these subsystems always contain the same
same linear combination of computational basis states
number of electrons. Therefore, for a total number of
one determinant at a time using the algorithm presented
N strongly correlated electrons in NS subsystems, each
in Ref. 73 (see Section VI B for more details on the com-
subsystem has n = N/NS electrons.
parison between our method and others).
In the case of diatomic bond breaking at dissociation,
Other works have focused on the preparation of certain
N strongly correlated electrons localize onto n = N/2
spin eigenfunctions.[78–80]. However, these either scale
subsystems, where N = 2 for H2 , and N = 6 for N2 (Sec-
exponentially[80] or only prepare trivial geminal prod-
tion II). The local coupling within a molecular fragment
ucts that couple two-electron singlets[78, 79] and there-
is ferromagnetic, which corresponds to a state of maxi-
fore do not have the structure necessary to encode the
mum spin multiplicity for each subsystem. By consider-
electronic correlation present in many molecular systems
ing the local coupling first, we can build the global wave-
such as the ones considered in this work (see Section II
function from coupling subsystem wavefunctions. We
8

and β → 0. For example |D23 ⟩ = √13 |110⟩+|101⟩+|011⟩



continue to use the example of stretching a triple bond
as in N2 for illustration. corresponds to | 23 , 12 ⟩ in Eq. (4).
The subsystem state with spin s = 3/2 state that fer- This encoding is natural for spin systems, where the
romagnetically couples three electrons is the quartet in N -qubit states exist in the 2N -dimensional Hilbert space
6,1
Eq. (4). The state |O0,0 ⟩ that describes the dissociation of N spin-1/2 sites, as might be found in site-based model
of a triple molecular bond into two three-electron quar- Hamiltonians. However, in quantum chemistry we con-
tets is the product of linear combinations of the 2s + 1 sider spin-1/2 fermions, and each site is a localized spa-
components of each quartet subsystem that forms an tial orbital that can have four possible occupations. In
overall N -electron state of spin S = 0, MS = 0: this context, the appropriate encoding requires one qubit
" # per spin-orbital (Section II A). Therefore, we first only
6,1 1 3 3 3 3 1 1 1 1 work with in the M qubits for the spin-up orbitals to
|O0,0 ⟩ = | ⟩ |− ⟩ − |− ⟩ | ⟩ − | ⟩ |− ⟩ + |− ⟩ | ⟩
2 2 2 2 2 2 2 2 2 N,1
prepare the |O0,0 ⟩ spin eigenfunction, i.e., we map spin-
1 N,1
coupled CSFs |O0,0 ⟩ to a superposition of basis states
h i
= |αααβββ⟩ − |βββααα⟩
2
1 h M
+ |αββααβ⟩ + |αββαβα⟩ + |αβββαα⟩ + |ββαβαα⟩
O
6 |j⟩ = |fiα ⟩ . (15)
+ |ββααβα⟩ + |ββαααβ⟩ + |βαββαα⟩ + |βαβαβα⟩ i=1

+ |βαβααβ⟩ − |βααββα⟩ − |βααβαβ⟩ − |βαααββ⟩ Afterwards, we map the occupation from this spin
− |αβαββα⟩ − |αβαβαβ⟩ − |αβααββ⟩ − |ααβββα⟩ space to the Fock space to recover the correct encoding
i (Eqs. (1) and (2)), as detailed in Section IV F.
− |ααββαβ⟩ − |ααβαββ⟩ . A symmetric state of n qubits |ψSn ⟩ is a state that re-
(12) mains invariant under permutations P of the symmetry
group of n qubits, Sn :
In the first row, we have used the notation |msL ⟩ |msR ⟩
for compactness to refer to the spin projection quantum |ψSn ⟩ = P |ψSn ⟩ . (16)
number of the left and right spin subsystems, i.e., we
have dropped the spin quantum number s = 3/2 and Furthermore, the n + 1 Dicke states |Dkn ⟩ of an n qubit-
only specify the spin projection quantum numbers of the system (k ∈ {0, 1, ..., n}) are orthonormal and form a
subsystems. The general form reads: complete basis for the n + 1 symmetric states of n
qubits.[83] Therefore, any symmetric state can be ex-
N,1 1 Xh i
pressed as a linear combination of Dicke states:
|O0,0 ⟩= am |m⟩ |−m⟩ + bm |−m⟩ |m⟩ (1 − δm,0 ) ,
N m
n
(13)
X
|ψSn ⟩ = ck |Dkn ⟩ . (17)
where s = n/2 = N/4 is the spin of each subsystem, k=0
m = s, s − 1, ... ≥ 0 is the absolute magnitude of√the spin
projection quantum number, m = |ms |, N = n + 1 is The states |O0,0 N,1
⟩ are constructed from the prod-
a normalizing constant, and the relative signs am , bm ∈ ucts of states of highest multiplicity, which are symmet-
{1, −1} are given from the Clebsch–Gordan coefficients. ric (Eqs. (4), (12), and (13)). Thus, the construction
N,1
of |O0,0 ⟩ is equivalent to constructing linear combina-
tions of products of Dicke states with different Hamming
B. Mapping to Dicke states and symmetric states
weights. For example, the state in Eq. (12) can be rewrit-
ten as
Although Eq. (13) contains N N
 
n = N/2 Slater deter-
minants, we can efficiently prepare this expansion on a 6,1 1 3
|O0,0 ⟩= |D3 ⟩ |D03 ⟩ − |D03 ⟩ |D33 ⟩
digital quantum computer by mapping it to Dicke states, 2  (18)
for which efficient quantum circuits are known. A Dicke − |D23 ⟩ |D13 ⟩ + |D13 ⟩ |D23 ⟩ .
state |Dkn ⟩ is defined as an equal-weight superposition of
all possible states of an n-qubit system with Hamming The general N -electron case (Eq. (13)) expressed in the
weight k[82] Dicke basis reads:
 −1/2 X 
n k n−k

N,1 1 Xh n n
|Dkn ⟩ = P |1⟩ |0⟩ , (14) |O0,0 ⟩= am |Ds+m ⟩ |Ds−m ⟩
k N m
P
i (19)
n n
where P denotes any permutation of qubits and the sum +bm |Ds−m ⟩ |Ds+m ⟩ (1 − δm,0 ) .
runs over all nk possible permutations. The open-shell
spin eigenfunctions discussed in this work can naturally We now discuss the preparation of Dicke states and sym-
be expressed in terms of Dicke states by rewriting α → 1 metric states following the approach proposed in Ref. 63.
9

C. Preparation of Dicke and symmetric states products of two types of elementary operators, Ml,l−1
and Ml,k (Lemma 2 in Ref. 63):
Dicke states have long been created on experimen-
k n
tal platforms such as ion traps,[84] but their efficient Y Y
Un,k = Ml,l−1 Ml,k . (22)
preparation through quantum circuits was a longstanding
l=2 l=k+1
challenge.[85] Recently, Bärtschi and Eidenbenz provided
a protocol for efficiently preparing any Dicke state |Dkn ⟩ For preparation of symmetric states we set k = n and
or any symmetric state using shallow circuits and without the unitary reduces to
requiring any ancilla qubits.[63] Dicke states have direct
applications in combinatorial optimization, where they n
Y
can be used as initial states of the Quantum Alterna- Un,n = Ml,l−1 := Sn . (23)
tive Operator Ansatz algorithm,[86] but their connection l=2
to spin eigenfunctions has previously not been explored.
Here, we use the Dicke state preparation method from Here, the operators Ml,l−1 can be constructed from prod-
Ref 63 as a circuit primitive to prepare spin eigenfunc- ucts of two primitive building blocks: one two-qubit gate
tions. We summarize the relevant results below. block (Fig. 2b) followed by l − 2 three-qubit gate blocks
(Fig. 2c). This enables the preparation of an n-qubit
1. Efficient preparation of Dicke states: The Dicke state Dicke state in terms of subcircuits acting on l < n qubits.
|Dkn ⟩, for k ≤ n, can be prepared deterministically by An example circuit for S4 is shown in Fig. 2a. We discuss
n−k k the decomposition of such circuits into gates consisting
applying a unitary Un,k on the input state |0⟩ |1⟩ .
of only CNOT and single-qubit rotations in Appendix C,
A decomposition of Un,k in terms of single-qubit and
to show that the number of CNOT gates required is,
CNOT gates is provided. The resulting quantum circuit
has depth O(n) and uses O(kn) gates. 5 2 9
CSn = n + n+2 (24)
2. Efficient preparation of symmetric states: Given any 2 2
input state of the same number of qubits n but with a as derived in Eq. (C1).
n−l l
smaller Hamming weight, |0⟩ |1⟩ with l ≤ k, applying One attractive feature of this protocol is the fact that
the operator Un,k onto said state outputs the Dicke state only one unitary Sn needs to be applied to generate a lin-
|Dln ⟩. Consequently, any superposition of Dicke states of ear combination of multiple Dicke states, and therefore
equal n and different l can be generated by applying the no special techniques such as Linear Combination of Uni-
unitary Un,n on the appropriate superposition of input taries [76, 87] are required for state preparation. Instead,
states: the desired superposition of Dicke states is achieved
X X n−l l by preparing a linear combination with the appropriate
|ψ⟩ = cl |Dln ⟩ = Un,n cl |0⟩ |1⟩ . (20) weights in the relatively simple input state. This is much
l l more efficient than applying state preparation unitaries
Since the n + 1 Dicke states |Dln ⟩ form an orthonor- controlled by ancilla qubits.[57, 58, 76] By slightly mod-
mal basis of the fully symmetric subspace of all n-qubit ifying the circuit for input state preparation, we can ef-
states (Eq. (17)), the circuits for unitaries Un,n can be ficiently use this circuit structure to prepare the CSFs
used to prepare any symmetric state |ψSn ⟩, provided that relevant for chemical bond breaking. For reference, the
they cost for up to n = 6 is shown in Table III in Appendix C.
P actnon the appropiate input states. The input state
l cl |ψl ⟩ can be prepared by a quantum circuit of depth
and gate count O(n). The overall complexity for prepa-
D. Notation for operators and circuits
ration of any symmetric state is thus depth O(n) and
O(n2 ) gates. These circuits can be implemented directly
on a qubit architecture with linear connectivity, which is In the above, the qubits on which Ml,k acts are spec-
beneficial e.g. for applications on superconducting quan- ified in the circuit diagrams (Fig. 2). For further clar-
tum hardware. ity, we define the following more explicit notation for all
quantum circuits below.
The key insight that enables this efficient protocol is Superscripts denote the qubits on which gates act, e.g.
the following recursive formula: X i is a Pauli-X gate on qubit i, where 1 is the top qubit
in the circuit diagram and its state is specified by the
r r i,j
n k k−1 n − k n−1 leftmost bit in the ket: |j⟩ = |j1 , j2 , ...⟩. We use Uj−i+1
|Dk ⟩ = |D ⟩ |1⟩ + |Dk ⟩ |0⟩ . (21) to denote an operator acting on (j − i + 1) qubits with
n n−1 n
indices i, i + 1, ..., j − 1, j, e.g. Sn1,n means Sn acting on
Given a unitary operator Un,k that prepares the Dicke qubits 1 through n. For two unitary operators U and V ,
n−k k
state |Dkn ⟩ when acting on the input state |0⟩ |1⟩ , one the product is U V (unlike for states where products of
can recursively decompose the unitary operator Un,k into kets are tensor products, see Section II A), therefore the
tensor product of two unitaries must be written explicitly
10

as U ⊗ V . Tensor products with identity are implied the circuit diagram. We often use the CNOT gate with
c,t
whenever the unitaries act on a subset of the qubits. CX = X c CX c,t X c .
For two-qubit controlled gates, we use the following
notation: CU c,t applies the operator U on target qubit
t controlled by qubit c, i.e. CU c,t = P0c ⊗ I t + P1c ⊗ U t , E. Preparation of singlet states with locally
where the projectors are P0 = |0⟩ ⟨0| and P1 = |1⟩ ⟨1|. On ferromagnetic coupling
circuit diagrams, this corresponds to a full black dot on
c,t
qubit i. We use CU for a gate that controls application Equipped with the circuits for Dicke state preparation,
of the operator U on target qubit t such that U is applied we can rewrite the CSF for bond breaking (Eq. (19)) in
c,t
if c is in the state |0⟩, i.e. CU = P0c ⊗ U t + P1c ⊗ I t . terms of the corresponding unitaries acting on carefully
This corresponds to a white dot on the control qubit c in chosen input states:

N,1 1 Xh n n n n
i
|OS,M ⟩= am |Ds+m ⟩ |Ds−m ⟩ + bm |Ds−m ⟩ |Ds+m ⟩ (1 − δm,0 )
N m
(25)
1 Xh s−m s+m s+m s−m s+m s−m s−m s+m
i
= Sn1,n ⊗ Snn+1,2n am |0⟩ |1⟩ |0⟩ |1⟩ + bm |0⟩ |1⟩ |0⟩ |1⟩ (1 − δm,0 )
N m

8,1
For example, to prepare the spin-coupled state |O0,0 ⟩, we only need to apply the operator S4 twice, in parallel, on an
input state which is a linear combination of five states:

8,1 1  
|O0,0 ⟩ = √ |D44 ⟩ |D04 ⟩ + |D04 ⟩ |D44 ⟩ − |D34 ⟩ |D14 ⟩ − |D14 ⟩ |D34 ⟩ + |D24 ⟩ |D24 ⟩
5
(26)
5,8 1
 
1,4
= S4 ⊗ S4 √ |1111000⟩ + |00001111⟩ − |01110001⟩ − |00010111⟩ + |00110011⟩ .
5

1,N
To prepare input states of the form We define Uin,N as the unitary for preparation of the in-
X n−l l put state for N spin- 12 particles (spin-coupled electrons):
cl |0⟩ |1⟩ , (27) 1,N N
l
|vN ⟩ = Uin,N |0⟩ . An example for N = 8 is shown in
Fig. 3 (left panel).
one could simply use the circuits in Fig. 4 of Ref. 63,
which are composed of ladders of controlled-Ry gates ap-
n
plied on the initial state |0⟩ . However, the spin eigen- F. Mapping from spin to Fock space
functions considered in this work are not symmetric and
require parallel application of two circuits for Sn , with
If one desires to prepare spin eigenfunctions for pure
input states of the form in Eq. (25).
spin systems, e.g. for quantum simulation of Heisenberg
We define the input state |vN ⟩ for preparation of the
N,1 or Ising spin models, this circuit above completes the en-
CSF |O0,0 ⟩ = Sn1,n ⊗ Snn+1,2n |vN ⟩ as |vN ⟩ = |L⟩ |R⟩ = tire state preparation procedure. However, here we are
P
i ci |Li ⟩ |Ri ⟩, where the left and right states |L⟩ and interested in systems of electrons (spin- 12 fermions). In
|R⟩ correspond to qubits {1, 2, ..n} and qubits {n + 1, n + this context, the |α⟩ and |β⟩ states in the expansion of
2, ..., 2n}. With n = N/2, the left and right qubits cor- the spin eigenfunctions (Appendix A) represent the spin
respond to dividing the system into two molecular sub- projection of an electron occupying a localized spatial or-
systems (Section II). The following circuit produces the bital (rather than of a spin orbital or qubit), where we
appropriate initial state: have restricted each orbital to be occupied by exactly
1. Flip the last n qubits by applying X-gates to obtain one electron. For a general many-electron wavefunction,
2n n n
the state ⊗2n n
n X |0⟩ = |0⟩ |1⟩ . any spatial orbital can be singly occupied, unoccupied or
Q  doubly occupied (Eq. (2)). We must therefore map the
n−1 i+1,i
2. Apply i=1 CRy (θi ) Ryn (θn ) on the first n state from the spin space of N spins, with dimension 2N ,
P
qubits to produce the state i ci |Li ⟩ |1⟩ .
n to the Fock space of N electrons occupying M spatial or-
bitals, with dimension 2M

N . For a minimal basis set and
QN −1
3. Apply a CNOT accordion i=0 CX 1+i,N −i that N spin-coupled valence electrons in the fully dissociated
flips the state of the right qubits controlled by the limit, we have M = N spatial orbitals (Section II).
occupation of the left P qubits to obtain the input To account for this, we have so far chosen to work
state |vN ⟩ = |L⟩ |R⟩ = i ci |Li ⟩ |Ri ⟩. entirely on the qubits corresponding to α spin orbitals
2 11

• Ry (✓) • • Ry (✓) • • Ry (✓) •


• Ry (✓) • • • Ry (✓) • • • n−1 • Ry (θ) • n−l
n • n−l+1
• Ry (✓) • • • • ···
• • • (a) Two-qubit
(b) Two-qubit gate. gate n
M4,3 M3,2 M2,1 1
| {z }
S
n −2 1 •} Ry (θ) • n − l • Ry (θ) •
| {z
S3 n • − l + 11: Gate •blocks used for appl
nFIG.
| {z } ···
S4 (a) Two-qubit gate preparation
n
unitaries, as derived i

(a) Circuit for S4 = S3 M4,3 = (S2 M3,2 )M4,3 = M2,1 M3,2 M4,3 . (b) Three-qubit
(c) Three-qubit gate. gate
FIG. 2: Circuit for S4 = M2,1 M3,2 M4,3 .
FIG. 1: Gate blocks used for application of Dicke state
FIG. 2: (a) Quantum circuit proposed in Ref. 63 to implement the symmetric state preparation unitary
preparation unitaries, as derived in Bartschi.
S4 = M2,1 M3,2 M4,3 (see Eq. (23)). (b), (c) Gate blocks in Dicke state preparation circuitssuchas (a), also from
q  q
1 l
Ref. 63. The angles are θ = 2 arccos n for the two-qubit gate blocks and θ = 2 arccos n for the three-qubit
gate blocks. 3

↵1 Ry (✓1 ) •
↵2 Ry (✓2 ) • •
M4,3
↵3 Ry (✓3 ) • • M3,2
↵1
M2,1
↵4 Ry (✓1 ) •
Ry (✓4 ) • •
↵2 Ry (✓2 ) • •
↵5 X M4,3
↵3 Ry (✓3 ) • • M3,2
↵6 X M2,1
↵4 Ry (✓4 ) • • M4,3
↵7 X M3,2
↵5 X M2,1
↵8 X M2,1
↵6 X M3,2
M4,3
↵7 1 X
↵8 2 X
3
1
4
2
3 5
4 6
5 7
6
8
7
1,8
8
Uin,8 S41,4 ⌦ S45,8 Umap,8
1,8
Uin,8 S41,4 ⌦ S45,8 Umap,8
8,1 8
FIG. Quantum
5: A circuit thatthatimplements the unitary
unitaryV8V8toto prepare
preparethe spin-coupled state |O8,1
0,0 i =VV 8 |00i . It consists of 8
FIG.
FIG. 3: circuit implements
the unitary Vthe the spin-coupled state . It|O 0,0 ⟩ = of 8 |00⟩
8,1 8
3: A circuit that implements 8 to prepare the spin-coupled state |O0,0 i = V8 |00i consists
three
(Eq. parts,
three(26)).
parts, It
delimited
consists
delimited byofthe
by
three the dashed
parts,
dashed
vertical
delimited
vertical
lines:
by the preparation
dashed
lines: preparation of the vertical
of (left),
thepreparation
lines:
input state
input state of(left),
preparationof the preparation
input
the spin
of the spin
state (left),
preparation of the spin eigenfunction through Dicke state preparation circuits (center, see Fig. 2), and mapping fromoccupation
eigenfunction
eigenfunction through
through Dicke Dicke
state state
preparationpreparation
circuits circuits
(center), and (center),
mapping and
from mapping
spin to from
spatial orbital spin to spatial
occupation orbital
(right).
(righy).
spin Is thespace
to Fock direction of the second
occupation U4,4 Qubits
(right). correct? labelled
Or do youαineed
andtoβreverse the input state? Check your qiskit
i correspond to the up and down spin-orbital with
code.
the same spatial orbital ϕi .
• •
• • • • • •
to prepare the superposition of Dicke states, leaving the The full operator VN for preparing the N -electron CSF
• • • • • • • • •
qubits corresponding to β spin-orbitals in the |0⟩ state. in two molecular subsystems (n = N/2) is thus
• • • • • •
After• the spin eigenfunction• is prepared through Dicke N,1 N  1,N N
circuits, we map the occupation as |α⟩ = |10⟩ → |10⟩ |O0,0 ⟩ = VN |00⟩ = Umap,N Sn1,n ⊗Snn,n+1 Uin,N |00⟩ .
and |β⟩ = |00⟩ → |01⟩, where the first qubit corresponds (28)
1,6 2,5 3,4
to an α spin-orbital andCX the second qubit represents a CXAn example CXfor N = 8 is shown in Fig. 3.
β spin-orbital. This can be achieved by a single layer of
FIG. 6: Recompilation of CNOT accordion for N = 6, namely CX 1,6 CX 2,5 CX 3,4 , into gates restricted to linear
CNOT gates that flips the state of the i-th β spin-orbital
connectivity.
if the state of the i-th α spin-orbital is |0⟩: Umap,N =
NN iα ,iβ
i=1 CX .

D( ) S( )

D( ) S( )
12

G. Total circuit cost and examples gates to rotate the single-particle basis, which can be
done in depth O(N ) and at worst O(N 2 ) gates, although
The total number of CNOTs assuming all-to-all qubit often O(N ) gates suffice. We exclude these from the
connectivity is analysis in this section since they depend on the system
considered, the choice of basis, as well as the quantum al-
all 5 2 gorithm in which they are being used. In Appendix B, we
Ctot = N + 2N + 2 (29)
4 present the circuits for basis rotations, their cost, and op-
(see Eq. (C4)). In Appendix C, we provide a detailed timal implementations for the molecules considered here.
derivation and discuss the case with linear or planar qubit
connectivity. The latter might be of interest for experi- all
N Ctot pla
Ctot lin
Ctot b T L System
ments on near-term superconducting hardware.[88] The 2 3 3 5 - 0 2 H2
depth of the circuits with all-to-all and planar connectiv- 4 14 19 63 13 49 6 H2 O
ity is O(N ). For linear connectivity, we obtain a depth 6 35 49 163 14 114 20 N2
scaling as O(N 2 ). 8 66 93 309 14 203 70
The CNOT counts are the most important metric for 10 107 151 501 14 317 252 Fe2 S2 (est.)
implementations on noisy quantum hardware without er- 12 158 223 739 15 477 924 Cr2
18 371 523 1729 15 1072 48620 Fe4 S4 (est.)
ror correction because two-qubit gates are usually the
34 1379 1939 6393 16 3989 2.3 × 109 FeMoCo (est.)
source of the largest errors.[89] In Appendix C 4, we dis-
cuss a fault-tolerant implementation of the state prepa-
ration circuit VN using a Clifford + Toffoli gate set. In TABLE I: Number of gates required for preparation of
N,1
short, the number of non-Clifford (Toffoli) gates depends the state |O0,0 ⟩ with N spin-coupled sites through our
on the number of rotations, R, and the accuracy required state preparation circuits. In the quantum chemistry
for synthesizing each continuous rotation gate through a context discussed here and in Ref. 52, N is the number
discrete gate set. To bound the error in the state prepa- of open-shell electrons occupying N localized orbitals.
ration unitary as ||VN − ṼN || ≤ ϵ, where ϵ is the target The total CNOT cost of preparing each spin
accuracy, the number of Toffoli gates required is eigenfunction is given by Ctot = Cin + 2CSN/2 + Cmap ,
  where the superscript indicates different hardware
T = R 0.2875⌈log(R/ϵ)⌉ + 4.6 . (30) connectivities. The number of determinants
Here, R = 14 N 2 is the total number of rotation gates (computational basis states) with non-zero coefficients
N,1
N,1 N in the state |O0,0 ⟩ is L. The number of Toffoli gates in
in the circuit that implements VN : |O0,0 ⟩ = VN |00⟩
a fault-tolerant implementation is given by T (see
(including Ry , CRy , and CCRy gates, see Eq. (C16)).
Appendix C 4). We also report b, the number of digits
We use log for logarithm to the base 2 throughout this
required for the binary representation of the rotation
manuscript.
angle.
Table I shows the total cost of preparing a spin eigen-
N,1
function |O0,0 ⟩, and the number of determinants (com-
N,1
putational basis states) in the expansion of |O0,0 ⟩ in a
−7
product basis. We have chosen ϵ = 10 for the accuracy H. Other spin eigenfunctions
of initial state preparation in the fault-tolerant setting.
It is evident that the preparation of these states is very We have so far discussed the preparation of CSFs for
efficient in practice, due to the O(N 2 ) scaling and the the coupling of two subsystems with maximum spin, cor-
relatively small constant factors. responding to fully symmetric spin functions, into an
H2 , H2 O, N2 , Cr2 are examples of molecules for which, overall singlet. While this spin coupling pattern repre-
in a minimal basis of N spatial orbitals, a single spin sents the dominant entanglement structure of electronic
eigenfunction describes the exact ground state at disso- eigenstates in typical covalent bonds,[52] which are ubiq-
2,1 2,2
ciation. For N = 2, |O0,0 ⟩ = |O0,0 ⟩ and we can therefore uitous in chemistry, other systems might require different
directly use the more compact circuit for preparation of types of spin eigenfunctions. For completeness, here we
N,2
|O0,0 ⟩, which only contains Clifford gates (see Section present circuits required to prepare the remaining CSFs
IV H). For the FeS systems, we hypothesize that the gate in Ref. 52.
counts are within the correct order of magnitude because The state of linear and square (cyclic) H4 at stretched
the number of open-shell electrons N is known from ex- geometries can be expressed as a superposition of two
isting literature,[90–92] and the number of determinants states,[52] each of the form in Eq. (A2):
as well as the cost of CSF preparation depends mainly on
h 1 i2
N . However, the estimates are by no means exact since 4,2
|O0,0 ⟩ = √ (|αβ⟩ − |βα⟩) . (31)
the correct spin eigenfunctions are currently unknown 2
(see Section VI B for details).
Finally, depending on the application, the total cost of The general form is a tensor product of singlets (anti-
the reference state preparation might require additional
13

symmetric functions) of N/2 two-spin systems The encoding used for the ancilla register can take
# N2 various forms that trade off circuit depth with space
(qubits). One could either work with one ancilla qubit
"
N,2 1
|O0,0 ⟩ = √ (|αβ⟩ − |βα⟩) . (32) with a higher gate cost, as shown in Ref. 9, use a one-hot
2 encoding with L ancilla qubits for L CSFs, or use a com-
pressed register of log(L) ancilla qubits as is common in
This state, also considered in Refs. 78, 79, is a
general Linear Combination of Unitaries approaches.[76]
product of Bell states and can be implemented triv-
The optimal encoding of the ancilla register depends on
ially in constant depth and using only one CNOT
h iN/2 the particular hardware and molecule considered.
gate per two-electron singlet √12 (|10⟩ − |01⟩) = The gate overhead for controlling the unitaries Vj
NN/2 2i,2i−1 2i N/2 scales linearly with N , which is lower than the cost of
i=1 CX Ry (−π/2) |00⟩ . The total CNOT
N,2
the circuits for Vj themselves (Appendix C 5). Exploit-
cost for preparing |O0,0 ⟩, including the cost of the spin- ing the recursive structure of the Dicke circuits enables an
to-Fock-space mapping Umap , is given by (Appendix C 6): additional reduction of the cost of implementing |ΦLC ⟩.
Specifically, from Eq. (23) it follows that
all 3
Ctot = N. (33)
2 Sn = Ml,l−1 Sn−1 . (34)
Since Ry (−π/2) = HX, this rotation can be imple- This implies that the circuit for Vj−2 can be reused as a
mented without any non-Clifford (e.g., Toffoli) gates. subcircuit for Vj rather than being applied twice.
On a fault-tolerant architecture, the preparation of such To summarize, the cost for preparing linear combina-
states has negligible cost (Appendix C 4). tions of CSFs is only slightly higher than the cost of
One might wonder if the circuits can be extended to preparing a single CSF, where the exact overhead de-
prepare any type of CSF. The efficient preparation of pends on the particular case considered.
arbitrary spin eigenfunctions remains an open problem,
as existing approaches with rigorous performance guar-
antees require exponentially many gates.[80] In general, V. HEURISTIC QUANTUM ALGORITHMS
highly symmetric states allow efficient state preparation
procedures, and one might expect that the complexity While the algorithmic error of fault-tolerant quantum
of the state preparation circuit increases with the num- algorithms can be bound analytically,[49, 50, 93–95] the
ber of distinct coefficients in the CSF. We hypothesize factors determining the performance of heuristic quan-
that most chemical systems contain enough structure tum algorithms such as VQE, quantum subspace diago-
that their spin coupling is given by paths that have a nalization, and adiabatic state preparation are less well
significant amount of symmetry. understood. Regardless, the numerical evidence accu-
Crucially, for applications in electronic structure, the mulated so far suggests that the accuracy and efficiency
choice of CSFs greatly depends on the choice of the of near-term algorithms strongly depends on the qual-
single-particle basis as well as the method used to con- ity of the initial state. In this section, we show that
struct the many-body CSFs.[62] In particular, rather spin-coupled reference states have the correct features to
than working in a fixed orthonormal basis, it is more ef- provide this increase in accuracy and efficiency.
ficient to use a bespoke basis for each physically-relevant
CSF. The choice of basis should appropriately reflect
the physical spin coupling pattern, e.g., by partition- A. Variational quantum eigensolver
ing the system into molecular fragments and separately
parametrizing the correlations due to local vs global cou- We first consider the VQE approach, which avoids deep
plings, as is done in this work. The extension of our state quantum circuits by combining quantum and classical
preparation protocols to more systems is an interesting processors in a hybrid algorithm. A quantum circuit with
direction for future work. parametrized gates prepares the electronic wavefunction
on a quantum computer, and a classical computer vari-
ationally optimizes the parameters. This is repeated in
I. Linear combinations of spin eigenfunctions
an iterative loop until convergence is reached.
At a fundamental level, VQE exploits the variational
When multireference initial states are required for principle to rotate the initial qubit state |0 · · · 0⟩ to
VQE or QPE, such as for molecular bonds at interme- the true electronic ground state. Naturally, the more
diate bond lengths, it is useful
P to prepare a linear combi- complex the eigenstate, the more parametrized quan-
nation of CSFs |ΦLC ⟩ = j cj |Φj ⟩ for j ∈ {0, 2, ..., N }. tum gates are required to accurately approximate it.
For example, N = 0, 2, 4, 6 for N2 (Eq. (11)). This can In practice, finding a compact and accurate ansatz is
be done by controlling the state preparation unitaries difficult.[26] In particular, strong electron correlation ex-
j,1 N
Vj : |O0,0 ⟩ = Vj |00⟩ from an ancilla register, for each acerbates the challenges of VQE, which reduces the po-
j. tential of this quantum approach to outperform classical
14

algorithms.[19, 24] For example, benchmark calculations the number of parametrized ansatz operators required
show that many ansätze accurately describe weakly- to reach quantitative accuracy.
correlated wavefunctions of molecular bonds at equilib- We demonstrate this improvement by performing clas-
rium geometries, but their accuracy deteriorates signifi- sical numerical simulations of VQE with the quantum-
cantly at strongly-correlated stretched geometries.[19] number-preserving (QNP) ansatz,[18] which uses a lay-
To reduce the computational cost, VQE implementa- ered circuit structure that can be systematically im-
tions use a classical heuristic to obtain an approximate proved to the exact result by increasing the number of
initial state, |Φ⟩, usually the Hartree–Fock state. Here, repeating layers k. Each layer includes a series of spin-
we show that using a spin-coupled initial state can greatly adapted one-body rotations and paired two-body rota-
reduce the cost required to achieve high-accuracy approx- tion operators acting between neighbouring spatial or-
imations of strongly correlated ground states. bitals, with the arrangement of these operators depicted
Formally, the quantum computer is first initialized to in Fig. 4. This choice of operators preserves the spin
the reference state quantum numbers of the initial state, which ensures that
2M
approximate wavefunctions are exact spin eigenfunctions.
|Φ⟩ = Uref |0⟩ . (35) Ref. 26 proves that this type of ansatz is universal, i.e.
it can yield exact wavefunctions within the spin- and
Here, 2M is the number of spin-orbitals (qubits) and particle-number-preserving subspace. In practice, it is
Uref is the unitary transformation required to prepare the unclear how many layers are required to achieve suffi-
reference state. To further correlate the reference state, cient accuracy.
a unitary transformation Uansatz (θ) is applied to it. This We consider two versions of VQE: a single reference
is implemented as a sequence of parametrized quantum approach in which a single unitary transformation is ap-
gates: plied on one sole reference state composed of a linear
Y combination of CSFs, and a multireference approach that
|Ψ(θ)⟩ = Uansatz (θ) |Φ⟩ = Ui (θi ) |Φ⟩ . (36) applies a different unitary transformation on each indi-
i vidual CSF.
The particular choice and ordering of elementary unitary
transformations Ui defines a variational ansatz, which de-
termines the wavefunctions that can be generated by the
quantum circuit through variation of the rotation param- D( ) S( )
eters θ, starting from the reference state |Φ⟩. The varia-
tional energy

E(θ) = ⟨Φ|Uansatz (θ) Ĥ Uansatz (θ)|Φ⟩ (37)
D( ) S( )
is estimated by measuring Hamiltonian expectation val-
ues on each qubit followed by operator averaging.[11] D( ) S( )
These estimates are then passed to a classical computer,
which uses traditional optimization algorithms to pro-
poses updates to θ that minimise E(θ). D( ) S( )
Practical VQE implementations face substantial chal-
lenges. Firstly, although compact and accurate electronic
states can in principle be prepared with short circuit D( ) S( )
depths,[12, 16, 25, 26, 96] finding these ansätze currently
relies on computationally expensive strategies that op-
timize the gate sequences.[12, 26] Secondly, the numer-
ical optimization appears to be fundamentally challeng- FIG. 4: One layer of the QNP ansatz (k = 1) includes
ing. The non-linearity of the ansatz means that E(θ) is FIG. 6: One layer (k = 1) of the QNP
an alternating series of paired two-body operators D(λ)
ans
highly non-convex and the optimization is prone to get S(spin-preserving
and ). one-body operators S(γ). These are
stuck in local minima.[20] Finally, estimating the energy applied between neighbouring spatial orbitals, each
may incur prohibitive measurement costs.[15, 22] acting on four qubits, as described in Ref. 18. The
Strong electron correlation greatly increases the dif- fermionic excitation operators S(γ) and D(λ) can be
ficulties of VQE.[19, 24] This poor performance can be decomposed into Givens rotations that act non-trivially
attributed to the use of a mean-field RHF initial state, on a two-qubit subspace.[18, 21]
which is uncorrelated and therefore burdens the ansatz
circuit Uansatz to recover a large amount of electron corre-
lation. Instead, we propose to use a spin-coupled wave-
function as the initial state, which has a greater over- S( )
lap with the true ground state. This greatly reduces

FIG. 7: Single excitation .


15

100 RHF Variable combination NO-VQE

10 1
Error from FCI [H]

10 2
1.6 mH
10 3
k=1
10 4 k=2
k=3
10 5 k=4
k=5
10 6
1.0 1.5 2.0 2.5 1.0 1.5 2.0 2.5 1.0 1.5 2.0 2.5
Bond Length [Å] Bond Length [Å] Bond Length [Å]
FIG. 5: Energy errors from the QNP ansatz for N2 (STO-3G) using either a fixed RHF reference state or a variable
linear combination of CSFs. The linear combination of CSFs significantly improves the accuracy in the
strongly-correlated dissociation limit. The NO-VQE approach applies a bespoke QNP circuit to each reference state
prior to constructing the linear combination and reaches chemical accuracy using shallower circuits with k = 2.

1. Single-reference VQE with variable linear combination reduces the number of parameters and unitary opera-
tors required to reach quantitative accuracy with a VQE
Our first approach is to apply the QNP ansatz to an ansatz.
initial state defined as the variable linear combination
L
X 2. Multireference ansatz with a nonorthogonal variational
|Φ⟩ = CI |ΦI ⟩ . (38) quantum eigensolver
I=1

We optimize both the QNP rotation angles as well as the An alternative strategy is to uniquely correlate each
linear coefficients CI in the initial state definition, since reference state. This is particularly useful for multirefer-
the latter are also parametrized in terms of single-qubit ence systems where more than one CSF dominates the
rotations through the circuits for preparing linear com- wavefunction. Since each CSF is a unique vector in
binations of CSFs (Section IV I). This allows the linear the Hilbert space, separately rotating each vector en-
combination of spin-coupled states to relax in the pres- ables multiple regions of the Hilbert space to be ex-
ence of the correlation generated by the QNP circuit. plored in a targetted manner. Although this idea has
For the strongly correlated N2 binding curve, the vari- been explored in other works e.g. using unrestricted
able linear combination of CSFs significantly reduces the Hartree–Fock determinants,[23] our spin-coupled wave-
energetic error compared to the RHF initial state at function approach has the advantage that significant en-
large bond lengths (Fig. 5), reaching chemical accuracy at tanglement is already included in the reference state.
R > 2.5 Å even with a single layer (i.e. k = 1). This result Achieving the same result through single-determinant
is expected since the spin-coupled wavefunction becomes reference states would require exponentially many ini-
exact in the dissociation limit (Section II B). Each layer tial states. Furthermore, since we strictly use CSFs,
corresponds to ten variational parameters (five for each any spin-preserving ansatz is guaranteed to produce spin-
operator in Fig. 4), and there is a fixed constant num- pure wavefunctions, without relying on approximately re-
ber of three parameters required to prepare the linear covering the correct spin state after symmetry breaking.
combination of four CSFs (one is fixed through normal- On a quantum circuit implementation, the multirefer-
ization). Even around the equilibrium region, where the ence approach avoids the preparation of a linear combi-
RHF configuration dominates the linear combination of nation of different CSFs on the quantum register and in-
CSFs (Fig. 1b), the spin-coupled wavefunction improves stead relies on the measurement of Hamiltonian and over-
the accuracy by around half an order of magnitude for lap matrix elements between the individually correlated
each value of k. These results demonstrate that using reference states. Furthermore, the increased number of
a linear combination of CSFs to define the initial state variational parameters provides a very flexible ansatz us-
16

ing limited circuit depth. accuracy requirement for each ansatz circuit, as the clas-
Formally, a unique set of unitary operations is applied sical diagonalization step gives increased variational free-
to the each reference state before constructing the linear dom. Here, we will show how using spin-coupled refer-
combination, giving ence states can greatly improve the performance of QSD
methods for both ground and excited state calculations.
L
!
X Y
|Ψ(Θ, C)⟩ = CI UIi (θIi ) |ΦI ⟩ , (39)
I=1 i 1. Background

where Θ = (θ1 , . . . , θL ). Each reference state is corre-


lated with a bespoke set of quantum gates that each have Quantum subspace diagonalization methods have
a unique variational parameter, and the gate parameters emerged as a promising class of hybrid quantum-
Θ are optimized simultaneously with the linear combi- classical algorithms for computing Hamiltonian eigenval-
nation C. The correlated basis states in this expansion ues. Starting from a reference state |Φ⟩, the quantum de-
will generally not be orthogonal, and we must therefore vice is used to generate a set of basis states {|Ψj ⟩} using
obtain the overlap matrix elements to guide the optimiza- various unitary transformations Uj , where |Ψj ⟩ = Uj |Φ⟩.
tion as in any nonorthogonal quantum eigensolver.[14, 23] The Hamiltonian is then diagonalized within the corre-
Here, we run classical simulations of this nonorthogonal sponding subspace to give the optimal linear combination
VQE (NO-VQE) algorithm. We describe details of our X
numerical implementation in Appendix D. |ψ(v)⟩ = vj |Ψj ⟩ . (40)
j
The multireference NO-VQE provides chemically ac-
curate energies across the full binding curve of N2 (STO- In general, the basis states are mutually nonorthogo-
3G) using only two layers of the QNP ansatz for each nal and the subspace expansion takes the form of a
reference state (Fig. 5, right panel). In contrast, five lay- nonorthogonal configuration interaction with the gener-
ers are required to reach an equivalent accuracy when alized eigenvalue problem
applied on a single reference state (Fig. 5, central panel),
as per the approach described in Section V A 1. This Hv = ESv, (41)
corresponds to a reduction in circuit depth by a factor
of 2 to 3, which is highly desirable on noisy quantum where Hjk = ⟨Ψj |Ĥ|Ψk ⟩ and Sjk = ⟨Ψj |Ψk ⟩. This eigen-
hardware. Therefore, although the NO-VQE approach value problem is solved on a classical computer. An
introduces more parameters, it achieves greater accuracy advantage of this subspace diagonalization is that the
with a shallower circuit than applying a single unitary to quantum device is only used to measure the matrix ele-
the variable linear combination of CSFs. The efficient ex- ments Hjk and Sjk , avoiding the significant overhead in
ploration of the Hilbert space achieved by starting from gate count and circuit depth associated with explicitly
different spin-coupled wavefunctions makes this a natural preparing the linear combination in Eq. (40) as a quan-
application of our work in multiconfigurational quantum tum circuit. Furthermore, QSD also gives direct access to
chemistry. excited state energies. Various strategies for generating
the basis states have been proposed, including:

B. Quantum subspace diagonalization through 1. fermionic excitation operators applied to a ground


real-time evolution state approximation to target excited states,[36, 37]
2. application of parametrized quantum circuits
Although VQE might be suited for noisy hardware Uj (θj ) defined through optimization[14] or pertur-
because typical ansatz circuits have low gate counts bation theories,[23]
and depths, severe implementation challenges in the
optimization and measurement make its practical use 3. application of the imaginary time-evolution opera-
questionable.[20, 22, 27] tor Uj = e−Hτj , where τj = j ∆t is total duration
In this section, we consider an approach that avoids of the evolution in imaginary time defined by in-
non-linear optimization and instead requires real-time- creasing integers j = 0, 1, 2, ...,[38]
evolution of the reference states under the Hamiltonian.
4. real-time evolution operators Uj = e−iHτj , where i
This falls within the broader category of quantum sub-
is the imaginary unit.[40–46, 48]
space diagonalization (QSD) methods (one could also
consider the nonorthogonal VQE from Section V A 2 as a Here, we focus on the real-time evolution approach,
QSD approach). QSD methods are uniquely well-suited also referred to as variational quantum phase estimation
for exploiting different reference states like the ones pre- (VQPE)[42] or quantum Krylov.[40, 46, 48] This method
sented in this work, as they can explore the Hilbert space only requires transforming the reference state through
from multiple directions in parallel, through transforma- unitary time-evolution, which is a natural operation for
tions applied separately on each reference state. The abil- quantum processors whose quantum circuit implementa-
ity to rotate each reference state individually relaxes the tion has greatly been optimized over the past decades.[29,
17

44, 50, 64, 97] Previous work has shown that the real- has NR × (1 + NT ) states. Note that each reference state
time-evolved states provide a very compact subspace for is time-evolved independently.
diagonalizing the Hamiltonian, in terms of the number We test this approach for N2 at R = 1.5 Å using noise-
of variational parameters needed.[40, 42, 44, 45] Fur- less simulations on classical hardware. We use a maxi-
thermore, numerical[42] and theoretical analysis[43, 48] mum of NT = 120 steps in a linear time grid tj = j∆t
suggest that this algorithm is robust to noise. Starting for three different time-step values ∆t ∈ {0.1, 1.0, 2.0}.
from the RHF reference state, the subspace is typically We set thresholds of 10−6 , 10−4 , and 10−2 for the singu-
built using a linear grid of NT equally-spaced time points lar values of the overlap matrix to remove the null space
τj = j∆t (j = 0, 1, ..., NT ), where the fixed time step ∆t in the generalized eigenvalue problem. Higher thresholds
must be chosen prior to the calculation. Naively, a to- might be more suited for implementation on noisy de-
tal of 2M 2 matrix elements would need to be evaluated vices, whereas using lower thresholds can help to retain
on the quantum device in order to set up the eigenvalue more expansion states and achieve faster convergence.
problem (Eq. (41)), where M is the number of expansion We observed no qualitative difference for the three dif-
states. This computation can be reduced to 2M unique ferent thresholds and therefore limit our discussion to
matrix elements if the time-evolution operator is imple- results with the threshold 10−6 . We apply a first-order
mented exactly, since the Hamiltonian and overlap matri- Trotter approximation of the time-evolution operator
P for
ces then have a Toeplitz structure, Hj,k = Hj+1,k+1 and a Hamiltonian given by a sum of terms H = k hk in
Sj,k = Sj+1,k+1 .[42] However, if the time-evolution op- the Pauli basis, given by
erator is approximated through Trotterization,[29] then Y
the Toeplitz structure of the Hamiltonian matrix is lost. e−iHtj ≈ e−ihk tj . (42)
In this scenario, and for linear time grids, the Toeplitz k
structure can still be recovered if the eigenvalue problem
is reformulated using the time-evolution operator instead Compared to a single RHF reference state, using mul-
of the Hamiltonian (see Appendix E and Refs. 39, 42). tiple spin-coupled reference states provides significantly
The main disadvantage of VQPE is that relatively faster convergence with respect to the number of time
deep quantum circuits are required to implement the steps (and thus circuit depth) for every ∆t considered
time-evolution operator. Specifically, the depth and gate (Fig. 6, top left). This implies that the multireference
count scale linearly with NT . The performance of VQPE approach should be faster and less susceptible to hard-
worsens in the presence of strong correlation, meaning ware noise, which becomes worse for deeper quantum
that more time steps are required.[40] The convergence circuits. One might wonder whether this comes at the
with respect to NT has also been shown to be slower cost of an increased number of measurements, since the
for excited-state energies.[42, 44] Furthermore, numer- increased number of reference states results in an addi-
ical results suggest that the accuracy of ground-state tional NR − 1 subspace states per time step. Encour-
calculations deteriorates when the time-evolution is im- agingly, we find that the multireference expansion also
converges more rapidly with respect to the total number
plemented with low-order Trotter approximations.[40]
of expansion states, meaning that fewer measurements
Higher-order Trotter formulas could be applied, but this
would be required overall (Fig. 6, top right). The faster
comes at great cost since the circuit depth for the k-th or-
convergence with the number of expansion states indi-
der Trotter formula scales exponentially as O(5k ). Below,
cates that starting from different reference states and in-
we show how combining QSD with spin-coupled reference
dependently time-evolving each of them might help to
states can significantly mitigate these limitations.
explore the Hilbert space more efficiently than if the time-
evolution is applied onto a single same initial state.
2. QSD with spin-coupled reference states The accuracy of VQPE strongly depends on the choice
of ∆t. We can understand this dependence through the
phase cancellation picture discussed in Ref. 42, since ∆t
Rather than using more time steps, the deep circuits
controls the phase applied to each eigenstate contained in
associated with the time-evolution operator can be re-
the reference space. Our results suggest that larger values
duced by building the subspace using multiple refer-
of ∆t enable faster cancellations and avoid the step-like
ence states.[40, 41] For example, Stair et al. used a
plateaus that occur for ∆t = 0.1, which are associated
set of reference determinants that are identified using
with near-linear dependencies in the expansion subspace.
iteratively-grown subspaces.[40] Here, we show that us-
VQPE can also provide access to excited state energies,
ing the physically-inspired spin-coupled wavefunctions
a key benefit that has received surprisingly limited at-
introduced in Ref. 52 (see Sections II, IV) to define
tention so far. When the time-evolution is Trotterized at
the reference states allows for the accurate computa-
first order, VQPE with both the RHF or multireference
tion of ground- and excited-state energies at signifi-
expansion can compute the lowest eigenstates with spa-
cantly reduced circuit depth compared to a single RHF
tial symmetry corresponding to A1g , E2g , and E4g , with
reference state. Starting from NR spin-coupled refer-
no restriction on the spin symmetry (Fig. 6, bottom left).
ence states {|Φ1 ⟩ , . . . , |ΦNR ⟩}, we construct a subspace
Like for the ground state calculation, the multireference
{e−iHτj |Φ1 ⟩ , . . . , e−iHτj |ΦNR ⟩} for j = 1, . . . , NT , which
expansion requires an order of magnitude fewer time
18

Trotterized Trotterized

Trotterized
Trotterized Exact
Exact

1.6mH

FIG. 6: When applied on spin-coupled reference states, real-time evolution accurately yields ground and excited
state energies after subspace diagonalization, requiring very few time steps. Top row: Ground state energy error as a
function of the number of time steps (left) and number of expansion states (right), where the time-evolved states are
obtained through first-order Trotter evolution (Eq. (42)). Bottom row: Energy error for the 10 lowest eigenstates
obtained from first-order Trotter evolution (left), and for the 10 lowest singlet eigenstates obtained from exact
time-evolution (right), with time-evolution step size ∆t = 2.0. Dashed lines correspond to subspaces formed from
time-evolving the RHF state; solid lines correspond to time-evolving four spin-coupled reference states. The labels
indicate the spin (superscript) and spatial symmetry of each eigenstate, where the first number is the index of the
eigenstate of a particular symmetry, in increasing energy and starting at 1 for the ground state. All results are for
N2 in the STO-3G basis at bond length R = 1.5 Å.

steps to reach the exact excited state energies compared Since the initial states are totally symmetric, this means
to the single RHF reference state. The presence of excited that we obtain excited states that reduce to Ag in D2h
states with different spin symmetry to the reference state (which correspond to the D∞h irreps A1g and Eng for
arises because the Trotter approximation of the time- even n). Consequently, although approximate energies
evolution operator breaks the spin-symmetry of two-body can be obtained with high accuracy, the corresponding
operators in the Hamiltonian. Similarly, non-abelian spa- eigenstates might not have pure spin or spatial symme-
tial symmetries in the D∞h point group are also partially try unless they are exact (within numerical accuracy).
broken since the Trotterized time-evolution operator only Contrast this with the version where we apply the ex-
conserves symmetries within the D2h abelian subgroup. act time-evolution operator (Fig. 6, bottom right). Since
19

the exact time-evolution operator commutes with all the Rather than using a fixed quantum circuit, the
symmetry operators of the Hamiltonian, the time-evolved ADAPT-VQE algorithm dynamically grows an ansatz
subspace states are guaranteed to preserve the total spin circuit tailored to the system of interest.[12] At each iter-
and the spatial symmetry of the reference state. This ation, the gradient of the energy estimate with respect to
allows us to systematically target states of a particular a pool of candidate operators is measured. The operator
symmetry. Here, we consider the lowest 1 A1g states in with the largest gradient is appended to the ansatz cir-
N2 . QSD with exact time evolution applied on multiple cuit, and all variational parameters of the ansatz unitary
CSFs yields all ten eigenenergies with only 5 time steps, are reoptimized following the usual quantum-classical
whereas QSD applied on the RHF state requires 24 time VQE loop. The ADAPT-VQE algorithm has been shown
steps. (Note that our multireference approach also yields to provide very compact representations of electronic
states of 1 E4g symmetry since the CSF |Φ2 ⟩, defined in wavefunctions.[12, 16] However, as the ansatz growth is
Eq. (9), contains spatial contaminants with azimuthal only guided by a local gradient criterion, its convergence
orbital angular momentum Lz = 4, 8, . . . .) It is remark- can sometimes be slow and it can get stuck in a local
able that our multireference approach provides such rapid minimum within the discrete operator space.[26] Further-
convergence despite the fact that the spin-coupled config- more, as with any VQE algorithm, the parameter opti-
urations are tailored explicitly for the ground state.[52] mization can converge to one of the many high-energy
This can be understood because the excited states of local minima in the continuous space.[20]
N2 correspond to configurations where electrons are pro- Here, we propose a QSD approach where a subspace
moted from bonding to antibonding orbitals, which are is defined using the sequence of wavefunctions obtained
exactly the configurations included in our CSF reference on each iteration of the ADAPT-VQE algorithm (Fig 7;
states. left). This can be done either using a single reference
In summary, defining a set of multiple reference states wavefunction such as the restricted Hartree–Fock (RHF)
using spin-coupled configurations significantly reduces state, or multiple reference states including different spin-
the number of time steps required to converge ground- coupled wavefunctions. ADAPT-QSD performs individ-
and excited-state VQPE energies compared to a single ual ADAPT-VQE calculations on each reference state,
RHF reference state. This faster convergence results in and uses the combined set of correlated states to define
shallower quantum circuits, making the VQPE algorithm the subspace within which the Hamiltonian is diagonal-
more suitable for implementation on quantum devices. ized.
We believe that this success arises because the multiref- Numerical simulations on N2 demonstrate that, even
erence expansion is conducive to exploring distant sectors in the single reference case, this approach can mitigate
of the Hilbert space. Any QSD method requires the gen- two of the main difficulties of the ADAPT-VQE algo-
eration of linearly dependent basis states,[42] but this is rithm. Firstly, ADAPT-QSD significantly accelerates the
generally hard to achieve due to the redundancies that convergence of the energy estimates with respect to the
arise when applying the time-evolution operator to a sin- number of operators, and therefore reduces the circuit
gle reference state. By independently time-evolving each depth (Fig 7, right). Secondly, when the ADAPT-VQE
reference state, we generate a basis that has fewer linear wavefunctions stagnate at a local energy mininum, sub-
dependencies and is thus more efficient at phase cancel- space diagonalization in ADAPT-QSD can escape that
lation. minimum, yielding an energy estimate that is orders of
The perturb-then-diagonalize form of QSD is naturally magnitude lower than ADAPT-VQE at comparable cir-
well-suited for multireference quantum chemistry. Just as cuit depth.
for the nonorthogonal VQE algorithm explored in V A 2, Comparing the role of the reference states, our re-
QSD with real-time-evolved states is greatly enhanced by sults again confirm the advantage of using multiple spin-
using spin-coupled initial states that contain a large part coupled reference states in QSD approaches. We find that
of the relevant electron correlation. the multireference approach to ADAPT-QSD requires
significantly fewer ADAPT-VQE iterations compared to
the approach that employs a single RHF reference state.
C. ADAPT-QSD: QSD with adaptive ansatz We achieve convergence to the exact ground state energy
using multireference ADAPT-QSD with roughly half the
Despite its simplicity, the real-time evolution approach number of operators per circuit needed in the single ref-
to QSD requires circuits that are significantly deeper erence case, further reducing the circuit depth require-
than is feasible with near-term quantum hardware, be- ments.
cause each time-evolution step requires the application ADAPT-QSD also systematically yields the low-lying
of every term in the Hamiltonian. Although the circuit excited states in the molecule, as shown in Fig. 8. This
depth can be reduced through randomized Hamiltonian enables excited state energies to be computed using the
simulation,[46] typically such circuits are still deeper than ADAPT-VQE formalism without requiring constrained
those for a VQE ansatz. To mitigate this, we propose an optimization.[13] For the results presented here, we chose
alternative QSD approach using adaptive quantum eigen- to work with the operators presented in the original
solvers. ADAPT-VQE paper,[12] which are not spin-symmetry-
20

100 R = 1.2 Å
10 1 RHF
Subspace grown using ADAPT-VQE iterations
10 2 Multireference
10 3

Error [H]
10 4
10 5
10 6
|Φ0 ⟩ −→ et1
(0) (0)
κ̂1
|Φ0 ⟩ −→ et2
(0) (0)
κ̂2
et1
(0) (0)
κ̂1
|Φ0 ⟩ −→ · · · 10 70
10 R = 1.5 Å
10 1 ADAPT-VQE
|Φ1 ⟩ −→ e
(1) (1)
t1 κ̂1
|Φ1 ⟩ −→ e
(1) (1)
t2 κ̂2
e
(1) (1)
t1 κ̂1
|Φ1 ⟩ −→ · · ·
10 2 ADAPT-QSD
10 3

Error [H]
10 4
(2) (2) (2) (2) (2) (2) 10 5
|Φ2 ⟩ −→ et1 κ̂1
|Φ2 ⟩ −→ et2 κ̂2
et1 κ̂1
|Φ2 ⟩ −→ · · · 10 6
10 70
(3) (3) (3) (3) (3) (3)
10 R = 2.5 Å
|Φ3 ⟩ −→ et1 κ̂1
|Φ3 ⟩ −→ et2 κ̂2
et1 κ̂1
|Φ3 ⟩ −→ · · · 10 1
10 2 1.6 mH
10 3

Error [H]
Iteration 1
10 4
Iteration 2 10 5
10 6
Iteration 3 10 7
0 20 40 60 80 100
ADAPT-VQE iterations
FIG. 7: ADAPT-QSD: Performing QSD in the basis of sequential ADAPT-VQE states can significantly accelerate
the convergence with respect to the number of operators and avoid stagnation, as demonstrated here for N2
(STO-3G) at R = 1.5 Å.

preserving. Furthermore, these operators also do not con- expressiveness in the ansatz or because the optimization
serve all spatial symmetries for non-abelian point groups. gets stuck in local traps.[17, 20] This becomes particu-
Therefore, like the Trotterized RT-QSD approach, the larly difficult when hardware noise is present in a real
ADAPT-QSD expansion yields excited states that trans- quantum device implementation, because the gate errors
form as the Ag irreducible representation of the D2h are usually higher than the 10−3 Eh accuracy needed for
abelian subgroup. One could use symmetry-preserving the energy estimate in chemistry applications.[27] We ex-
operators if states with well-defined symmetry were de- pect that QSD methods can mitigate this problem since
sired. even if hardware noise corrupts the measured overlap and
We have considered three algorithms that perform Hamiltonian matrix elements, high accuracy might still
subspace diagonalization: the nonorthogonal quantum be achieved through the linear QSD expansion, in partic-
eigensolver, QSD with real-time-evolved states (also ular when the diagonalization step is regularized through
known as VQPE or quantum Krylov), and our newly- thresholding the eigenvalue problem, as done in our simu-
proposed ADAPT-QSD. We have seen that the perfor- lations. There is theoretical[43, 48] and numerical[42, 47]
mance of all three algorithms significantly improves when evidence that the quantum Krylov approach discussed in
combined with spin-coupled reference states. This is par- Section V B is indeed noise resilient when combined with
ticularly useful for multireference systems as well excited thresholding. It is likely that this noise-resilience is also
states, which are precisely the situations for which clas- a feature of ADAPT-QSD, but further work is needed to
sical algorithms are known to struggle. QSD appears establish this.
particularly promising because it exploits the quantum
device to represent highly entangled states in different
bases, as well as to yield matrix elements in nonorthog- D. Adiabatic state preparation from the molecular
onal bases. For comparison, the ability to efficiently dissociation limit
provide such information through measurements is also
the core argument for potential quantum advantage with All the algorithms we have considered so far are hybrid
other quantum algorithms.[14, 98] quantum-classical algorithms that exploit the variational
There is an additional reason to prefer QSD over VQE principle to minimize the energy through iterative up-
methods. Often, VQE implementations struggle to re- dates of some trial ansatz parameters. Adiabatic state
solve the last few digits of accuracy in the energy es- preparation (ASP) offers an interesting alternative. It is
timate. Typically, this occurs either due to a lack of a purely-quantum algorithm for preparing Hamiltonian
21

evolving the state along a pathway that interpolates be-


106.0 1 1A1g 1 1E2g tween Ĥ0 and ĤF , defined as
1 5A1g 2 5A1g  
Ĥ s(t) = 1 − s(t) Ĥ0 + s(t)ĤF . (43)
1 3E2g 2 1E2g
106.5
Energy [H]

2 1A1g 3 1A1g Here, s(t) is some continuous function with s(0) = 0


1 1E4g 1 3A1g and s(τ ) = 1, where τ is the total evolution time. The
107.0 required total evolution time τ can be estimated as[32]

|⟨Ψ1 (s)| ∂s H(s) |Ψ0 (s)⟩|


τ ≫ max , (44)
107.5 s∈[0,1] |E1 (s) − E0 (s)|
2

where |Ψ0 ⟩ and |Ψ1 ⟩ are the ground and first excited
10 1 states of H(s), respectively. On a digital quantum de-
vice, the path variable s(t) can be discretized using con-
stant time steps ∆t. Therefore, slower evolution, which
Error [H]

10 3 increases the likelihood of remaining in the ground state,


requires a larger number of time steps and deeper quan-
tum circuits.
10 5
Since the ground state of Ĥ0 must be easily obtained
and prepared, the obvious choice for electronic problems
is the mean-field Fock operator, for which the (restricted)
10 7
0 5 10 15 20 25 30 Hartree–Fock wavefunction is the ground state.[30] The
ADAPT-VQE iterations accuracy of ASP is typically quantified using the squared
2
overlap |⟨Ψ(τ )|Ψ0 ⟩| between the final state |Ψ(τ )⟩ and
FIG. 8: Multireference ADAPT-QSD captures the the physical ground state |Ψ0 ⟩. Numerical studies have
low-lying excited states with the same spatial symmetry shown that the accuracy strongly depends on the ini-
as the input CSFs, as shown for N2 at R = 1.5 Å. The tial state,[30, 33–35] meaning that a mean-field refer-
lower energy states converge more rapidly as the ence is unlikely to be sufficient for strong electron cor-
subspace expands, and the states considered here are relation. Indeed, simulations of chemical bond break-
converged below chemical accuracy using 22 ing have demonstrated that much larger τ values are re-
ADAPT-VQE iterations. quired at long bond lengths due to both the inadequacy
of the Hartree–Fock state and the small energy gap be-
tween eigenstates that leads to an exact degeneracy for
ground states on digital quantum computers that does R → ∞.[30, 34]
not rely on any parametrized ansatz.[28, 32] Instead, it To overcome the limitations of the mean-field Fock
exploits the adiabatic theorem, which requires using some operator, alternative initial states and reference Hamil-
Hamiltonian whose ground state can be easily prepared tonians have been explored, including Hamiltonians in-
on a quantum computer. volving a subset of the molecular orbitals with com-
Typically, the initial state of choice is the Hartree– plete active space configuration interaction (CASCI)
Fock state.[29] As usual, this works well for weakly cor- wavefunctions,[30, 33] active space Hamiltonians ob-
related problems, but it is inefficient for strongly corre- tained from n-electron valence perturbation theory
lated systems. The CSFs considered in this work are the (NEVPT)[35], and unrestricted Slater determinants that
exact ground state of a family of molecular Hamiltoni- break Ŝ 2 symmetry.[34] The CASCI approach reduces
ans at dissociation (Section II and Ref. 52). Here, we the adiabatic evolution time, but it requires the brute-
show that by starting from the dissociation limit with force computation of the ground state wavefunction
a spin-coupled state, we can speed up adiabatic state within the active space, which is difficult to predict a
preparation of strongly correlated eigenstates. priori. Thus, active-space methods are unlikely to be ap-
plicable for systems with many strongly correlated elec-
trons. Furthermore, while the symmetry-broken states
1. Background analyzed in Ref. 34 can be easily prepared on a quantum
circuit, the spin symmetry must be restored by adding
The ASP algorithm was proposed as a method to pre- a penalty term to constrain the ⟨Ŝ 2 ⟩ expectation value,
pare initial states with high overlap with the ground state which increases the circuit cost of implementing time-
for quantum phase estimation.[29, 31, 35] Starting from evolution.
some initial Hamiltonian Ĥ0 whose ground state can be
easily prepared on a quantum computer, the ground state
of the full Hamiltonian ĤF can be generated by slowly
22

103 1.0
102 0.8
Energy error [mH]

101
0.6

Fidelity
100 = 10
= 20
= 30
0.4
10 1
= 50
= 100
10 2 = 200 0.2
= 300
10 31.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 0.01.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5
Bond length [ ] Bond length [ ]
FIG. 9: Adiabatic state preparation is greatly accelerated by using a spin-coupled wavefunction in the strong
correlation limit. (Left) Error of energy estimates and (right) squared overlap between the exact ground state for N2
and the adiabatically evolved states after full evolution (s(τ ) = 1). Solid lines correspond to the version of ASP that
6,1
starts with the state |O0,0 ⟩ and the Hamiltonian at dissociation, whereas dashed lines refer to ASP starting with the
RHF state and the Fock operator. The total evolution time is τ , and the horizontal black line denotes chemical
accuracy in the energy estimate.

2. Interpolation from the exact dissociation limit much larger than chemical accuracy. In contrast, ASP
6,1
starting from the fully open-shell CSF |O0,0 ⟩ achieves
Here, we exploit the fact that the ground state of dis- fidelities of > 90% for all geometries above R > 1.4 Å
sociated molecules can often be described using only one with a short evolution time τ ≥ 30 E−1 h . This demon-
N,1 strates that the spin-coupled CSF provides a much bet-
spin eigenfunction of the form |O0,0 ⟩.[52] This suggests
an alternative approach to apply ASP in the regime of ter starting point in the strongly correlated dissociation
strong correlation. We propose using the ground state regime. For a given τ , simulations starting from the
and the fully-interacting Hamiltonian at molecular disso- open-shell CSF perform better than those starting from
ciation as the starting point, and applying ASP to inter- the Hartree–Fock state for R > 1.7 Å, which corresponds
polate between the dissociation limit and the target ge- to the bond length at which the open-shell CSF switches
ometry to compute the entire binding curve. We numeri- with the RHF determinant as the dominant contribution
cally investigate this approach through the N2 molecule. to the ground wavefunction (Fig. 1c). This observation
The ground state for N2 at dissociation is exactly rep- provides further numerical evidence for the direct corre-
6,1
resented by one open-shell CSF |O0,0 ⟩ which spin-couples lation between the accuracy of the initial state and the
the six valence p-electrons and can be predicted a pri- cost of ASP, in line with previous research.[30, 33–35]
ori, from symmetry arguments (see Sections II). We de- Since the circuit depth and gate cost of implementing
fine the initial Hamiltonian Ĥ0 as the full interacting time-evolution on a quantum computer scale polynomi-
Hamiltonian within the (6, 6) active space at R = 4.5 Å, ally with the total time,[99] state preparation based on
which is representative of the dissociation limit since open-shell CSFs provides orders of magnitude more ef-
6,1
2 ficiency for strong correlation compared to RHF-based
1 − ⟨O0,0 |Ψ0 ⟩ ≈ 4 × 10−7 , where |Ψ0 ⟩ is the exact ASP. Furthermore, since the CSF is explicitly an eigen-
ground state. For simplicity, we only consider the linear function of the Ŝ 2 operator, we do not require additional
interpolation s(t) = t/τ . We use exact time evolution modifications of the time-evolution circuits to enforce
with a constant timestep of ∆t = 0.1 E−1 h and different
spin-pure wavefunction, in contrast to the symmetry-
total evolution times up to τ = 300 E−1 h . We compare
broken approach in Ref. 34. Finally, because the cost
CSF-based simulations with standard ASP calculations of preparing the open-shell CSFs is negligible compared
that start from the RHF reference state with Ĥ0 being to the cost of implementing the time-evolution opera-
the Fock operator. tor and we avoid the classical complexity of identifying
The accuracy of the adiabatically-evolved RHF state CASCI initial wavefunctions,[30, 57] our approach can be
for a given τ decreases as the bond length increases extended to many strongly correlated electrons.
(Fig. 9). For example, a very long evolution time of Looking forward, although we have only considered a
τ = 300 E−1 h is required to achieve over 90 % fidelity minimal basis representation of the N2 dissociation, the
(squared overlap) with the true ground state in the disso- same principles can be readily applied to larger basis sets.
ciation limit, and the corresponding energetic error is still For example, one possible protocol would be to start with
23

a minimal basis representation incorporating the domi- trolled by the ancilla register accumulates a phase
nant spin coupling and continuously evolving the state on the ancilla qubits corresponding to the eigenval-
into the larger basis set at the target geometry using ASP ues of the operator.
by turning on the matrix elements coupling the orbitals
in the minimal and large bases. A similar approach could 3. After application of the Quantum Fourier Trans-
be applied to other strongly-correlated systems, such as form on the ancilla register, the ancilla qubits are
FeS clusters,[31, 92] where the fully-interacting ground measured in the computational basis. The mea-
state could be prepared starting from a suitable model surement results encode the energy eigenvalue as a
Hamiltonian with a single CSF ground state. binary bitstring that is read out classically.
If |Φref ⟩ = |Ei ⟩, the measurement yields the energy
corresponding to the i-th eigenstate with probability 1
VI. AVOIDING THE ORTHOGONALITY
CATASTROPHE IN QUANTUM PHASE
(exactly, up to a desired numerical precision). However,
ESTIMATION if the initial state is a superposition of energy eigenstates,
the probability of projecting to the desired eigenstate and
measuring the corresponding energy eigenvalue is given
A. Quantum phase estimation
by the squared overlap |γi |2 = | ⟨Φref |Ei ⟩ |2 between the
initial state and the desired eigenstate. Therefore, to
Large-scale error-corrected quantum hardware remains obtain a sufficiently accurate energy estimate, the entire
a distant prospect and near-term applications are likely procedure must be repeated several times, with the num-
to rely on heuristic algorithms. However, in the long ber of repeats scaling as O(|γi |−2 ). Post-QPE techniques
term, the most anticipated application of quantum com- can improve this scaling to O(|γi |−1 ).[93–95] However,
puting in chemistry is the simulation of challenging sys- the overlap dependency is a fundamental feature of all
tems through quantum phase estimation.[29, 31, 35, 64, projective algorithms, and indeed the O(|γi |−1 ) scaling
97] has been found to be near-optimal.[93]
PQPE can be used to project some initial state |Φref ⟩ = Hamiltonian simulation is remarkably efficient, and the
i γi |Ei ⟩ onto one of the energy eigenstates |Ei ⟩ of the latest qubitization-based algorithms achieve quadratic
Hamiltonian.[49–51] Although the algorithm is highly scaling with the basis set size (number of orbitals).[64,
promising for quantum advantage with fault-tolerant 101] However, the QMA-hardness of electronic structure
hardware architectures, its scaling depends inversely on manifests through the initial state dependency of QPE.[4]
the initial state overlap. This is problematic for strongly The reliance on the global overlap of an approximate
correlated electrons where the Hartree–Fock wavefunc- wavefunction means that in principle, the algorithm will
tion has poor overlap with the exact state, which is also suffer from the orthogonality catastrophe,[5, 6] requiring
where classical methods are most limited. For example, a number of repetitions that scales exponentially in the
for N2 , |γ0 |2 = | ⟨ΦRHF |E0 ⟩ |2 ≈ 0.92 at equilibrium but system size. In practice, this worst-case scenario can be
decreases to |γ0 |2 ≈ 0.06 at long bond lengths (Fig. 1c). avoided by identifying approximate wavefunctions that
Here, we propose using an initial state state that con- contain the dominant entanglement structure of the ex-
sists of a linear combination of the fully spin-coupled act wavefunction.
N,1
state |O0,0 ⟩ and intermediate states with increasingly For the N2 example studied here, a linear combination
fewer correlated electrons (N − 2, N − 4, ...0), includ- of four CSFs (Eq. (11)) can enhance the squared overlap
ing the fully uncorrelated RHF state (see Eq. (11) for of the reference state and the exact ground state by a
N2 , where N = 0, 2, 4, 6). This allows us to prepare factor of 16 when compared to the Hartree–Fock state at
states with high overlap with the exact ground state us- stretched geometries (Fig. 1c). This implies an equiva-
ing O(N 2 ) Toffoli gates (Eq. (30)), mitigating the ini- lent reduction in the total runtime of QPE since the cost
tial state problem for systems where the accuracy of the of preparing the initial state is negligible compared to
Hartree–Fock state diminishes rapidly with the number the cost of Hamiltonian simulation. In Ref. 52, we pre-
of electrons. sented analogous results for more systems including most
The QPE algorithm involves three steps: second-row diatomics and the water molecule, where the
overlaps of linear combinations of CSFs are always above
1. The main register is initialized to some reference
85%. The accuracy of this linear combination state lies
state through application of a state preparation cir-
in the ability to capture spin coupling effects between
cuit Uref |0⟩ = |Φref ⟩.
localized, open-shell orbitals.
2. An invertible function of the Hamiltonian is applied This improved initial state removes the bias of
in the form of a unitary operator. This is often Hartree–Fock theory towards delocalized states and al-
the time-evolution operator U = e−iHt ,[29, 31, 50] lows for ferromagnetic coupling within a local subsystem
but it can also be a qubitized quantum walk U = of electrons and global antiferromagnetic coupling across
ei arcsin(H/λ) ,[64, 100, 101] where λ is a rescaling subsystems. We expect that diatomics with even more
parameter proportional to the norm of the Hamil- strongly correlated electrons, such as Cr2 , can also be
tonian. Evolution of the initial state under U con- treated in this way.
24

Nevertheless, all systems mentioned so far, including situation, it is clear that one would need post-Hartree–
the Cr2 ,[102] are tractable with classical methods, at Fock initial states for QPE as the overlap between the
least if only the ground state is considered. This is due to Hartree–Fock state and the ground state is very small.
the fact that a relatively small number of electrons are Indeed, Lee et al. [35] found an overlap of ∼ 10−7 be-
strongly spin-coupled (N = 12 for Cr2 ), and therefore tween the Hartree–Fock state and a matrix product state
compressed classical representations of the wavefunction obtained through the DMRG algorithm. Although re-
can still be processed on classical devices without a big cent work by Ollitrault et al. shows that the overlap of a
loss in accuracy. In other words, the exponential asymp- single CSF and the ground state can be boosted by opti-
totic scaling in N is not prohibitive when N = 12. mizing the orbitals, more sophisticated state preparation
methods are required for larger systems.[103]
As an alternative to preparing the Hartree–Fock state
B. Resource estimation for fault-tolerant or an open-shell CSF, one could consider using the best
simulation of FeS systems possible selected CI state,[73] e.g., as found through
FCIQMC,[55, 104] heat-bath CI,[56], or adaptive sam-
The long-term hope is that fault-tolerant quantum de- pling CI.[9, 57] However, quantum circuits that prepare
vices with hundreds of logical qubits will be able to com- a generic linear combination of L determinants typically
pute energies for systems that are out of reach for all require O(L) gates,[58, 105] usually with an additional
classical methods.[31, 64, 97] O(log L) [58] or O(N ) factor for N qubits,[9] as well as
An often-cited class of molecules with potential techno- the use of an ancilla
√ register. At best, the cost can be
logical applications, whose simulation is classically chal- reduced to Õ(
√ L), but this requires trading off Toffoli
lenging, are transition metal clusters such as systems gates for O( L) qubits.[58]
composed of multiple iron-sulfur centers.[90, 91] In par- The state-of-the-art algorithm for computing the
ticular, the electronic structure of FeMoCo is considered ground state energy of FeMoCo through phase estima-
a benchmark for QPE-based quantum advantage.[31, 35, tion requires 3.2 × 1010 Toffoli gates.[64] Using the clean
64, 92] It is strongly correlated, containing up to 35 open- qubit approach from Low et al.,[58] recent work gave an
shell electrons.[91] By increasing the number of FeS cen- algorithm for preparing a CI state of L determinants with
ters, it has been shown that both the Hartree–Fock state cost of
and a naive single CSF have an overlap with the best-
available matrix product state that diminishes exponen- (2 log L − 2)L + 2log L+1 + L (45)
tially with increasing system size.[35] Although the poor Toffoli gates.[73] Therefore, preparing a superposition of
Hartree–Fock overlap is expected, we hypothesize that it the L = 2.3×109 determinants that form the state |O0,034,1

is possible to obtain much higher overlaps using a single 11
would cost 1.43 × 10 Toffoli gates. This would make
CSF or a linear combination of few CSFs. This requires
the cost of initial state preparation higher than the cost
understanding the spin coupling structure and choosing
of QPE, which is problematic given the already-long ex-
a suitable set of orbitals to parametrize an appropriate
pected runtimes for evaluation of a single energy, which
CSF.
are in the order of multiple days on future superconduct-
To get a rough estimate for the cost of preparing an
ing quantum hardware.[64]
accurate initial state, first assume that a CSF like the
N,1 To reduce the state preparation cost, one could con-
state |O0,0 ⟩ is an accurate approximation of the ground sider only preparing a few of the most important de-
state, where N is the number of open-shell, spin-coupled terminants that dominate the wavefunction. Examining
electrons. Although we have not attempted to obtain the N,1
the CSF |O0,0 ⟩, while the number of determinants scales
correct CSFs for FeS systems in this work, we expect that N
 n n
the number of determinants is mostly determined by the exponentially as N/2 , the determinants |α⟩ |β⟩ and
n n
number of open-shell electrons. Therefore, the open-shell |β⟩ |α⟩ , where n = N/2, have significantly more√ weight
N,1 that the rest. Specifically, their amplitude is 1/ n + 1
CSF |O0,0 ⟩, whose structure and circuit representation
we have studied in this work, serves as a proxy for the (Eq. (13)). One could therefore build a state
cost of preparing an approximation to the eigenstates of 1 n n n n
FeS systems. |ϕ⟩ = √ (|α⟩ |β⟩ − |β⟩ |α⟩ ) (46)
2
From previous studies,[90] it is known that N = 10 and
N = 18 for Fe2 S2 and Fe4 S4 . FeMoCo is believed to have N,1
that would have squared overlap with the |O0,0 ⟩ CSF
up to 35 open-shell electrons,[91] and we thus set N = 34. scaling inverse linearly in the number of electrons:
The smaller systems have been tackled with relative suc- N,1 2
|γN |2 = | ⟨ϕ|O0,0 ⟩ | = 4/(N + 2). For example, for
cess using state-of-the-art DMRG programs,[90, 91] but N2 , this would give an initial state with squared overlap
accurately resolving the electronic structure of FeMoCo 6,1
|γ6 |2 = 0.5 with the |O0,0 ⟩ CSF (and therefore with the
remains an open problem.[35,  92] Table I contains the ground state near dissociation) at a very cheap cost.
N
number of determinants N/2 required in the expansion However, this has several downsides. First, the state
N,1 |ϕ⟩ is not a spin eigenfunction. Furthermore, |ϕ⟩ has sig-
of the state |O0,0 ⟩. For N = 34, this is 2.3 × 109 . In this
nificant overlap with eigenstates other than the ground
25

state, which is problematic for application in QPE, as it ergy estimates.[35, 90, 91, 103] Table II compares the Tof-
makes it highly probable to measure an undesired eigen- foli cost of MPS state preparation for different FeS sys-
state. In the N2 example, at R = 3.0 Å the squared tems as a function of the bond dimension, as per Eq. (47).
overlap of |ϕ⟩ with the singlet (S = 0) ground state and Clearly, preparation of a MPS of high (χ > 1000) bond
a quintet (S = 2) excited state is equal at 50%, and dimension would constitute a significant fraction of the
therefore it is equally likely to measure either state. This overall cost of one run of QPE.
scenario also occurs in more challenging systems such Formichev et al. also proposed reducing the cost of
as transition metal clusters, where the strong correla- MPS state preparation by taking a MPS with high bond
tion due to spin coupling leads to many near-degenerate dimension (e.g., χ ≈ 2000) obtained from a DMRG cal-
eigenstates of different spin, or eigenstates of the same culation, and compressing it to a MPS with lower bond
spin but different spin coupling. In such situations, it is dimension (χ ≈ 5 − 10).[73] For some systems, includ-
essential to use accurate initial states to separate the dif- ing Fe4 S4 , they found that while the energy of that MPS
ferent eigenstates, which requires a much larger number is very inaccurate, compression does not significantly af-
of determinants. fect the wavefunction quality, as there remains a strong
A further alternative would be to use a classical DMRG overlap between the compressed MPS and the MPS of
calculation to find a good initial state.[106] The tensor higher bond dimension. Assuming that such compressed
product structure of the MPS can be exploited to directly states with bond dimension in the order of χ = 10 in-
encode the DMRG state in a quantum circuit, circum- deed provide accurate ground state approximations, the
venting any explicit CI expansion. To our knowledge, state preparation cost would be relatively low compared
the only method for MPS preparation with guaranteed to the cost of QPE (Table II). Although this compression
success is the sequential algorithm proposed by Schön et approach could be promising for cheaper state prepara-
al.[107] The number of gates scales as O(M χ2 ) for a MPS tion, it requires significant classical cost to compute the
with M sites and bond dimension χ.[73] For FeMoCo, we MPS with high bond dimension. It also remains to be
can take M = N = 34, i.e. the number of sites of the seen how accurate the compression is for larger and more
MPS is the number of open-shell orbitals/spin-coupled complex systems.
electrons.
Despite the polynomial scaling in N and χ of MPS N System χ b TMPS
preparation circuits, the cost of gate-level implementa- 10 Fe2 S2 10 34 4.56 × 104
tions have large constant factors as they require im- 50 37 8.51 × 105
plementing non-trivial quantum arithmetic operations 2000 43 1.16 × 109
and the use of O(χ) ancilla qubits.[73] For example, 18 Fe4 S4 10 35 8.7 × 105
Formichev et al. [73] gave a circuit decomposition of the 50 37 1.61 × 106
algorithm by Schön et al. [107] with approximate cost 2000 44 2.2 × 109
34 FeMoCo 10 36 1.71 × 105
TMPS ≈ (M − 1)χ[32χ + (b + 1) log(4χ)] (47) 50 38 3.13 × 106
2000 44 4.26 × 109
(Appendix F). More fundamentally, the entanglement
captured by a MPS of bond dimension χ scales as
O(log(χ)). Therefore, for systems whose entanglement TABLE II: Estimated number of Toffoli gates required
grows with the system size, as opposed to Hamiltonians for preparation of matrix product states for N electrons
with area-law ground states,[108] the bond dimension can in M = N spatial orbitals for different choices of bond
scale very steeply, at worst exponentially in N to achieve dimension χ. Here, b is the number of digits (ancilla
high accuracy through a MPS. qubits) that encode each rotation gate in binary
In practice, the success of the DMRG algorithm when representation, as required to guarantee a state
applied to strongly correlated systems, at least for smaller preparation error of at most ϵ = 10−7 . For comparison,
FeS clusters,[90, 91] indicates that the worst-case ex- the number of Toffoli gates required for running
ponential scaling might not be strong enough to pro- qubitization-based quantum phase estimation circuits is
hibit classical simulations for the system sizes consid- 3.2 × 1010 for FeMoCo.[64]
ered. Another explanation could be that locality damp-
ens electron-electron interactions, e.g. due to charge In contrast, our CSF preparation circuits directly en-
screening, reducing the strength of long-range entangle- code the relevant entanglement structure into bespoke
ment between different FeS clusters. It remains to be seen quantum circuits, which we expect to maximize the state
what bond dimensions would be required to accurately preparation efficiency nearly optimally. Preparation of
N,1
resolve larger systems like FeMoCo, and what would be the |O0,0 ⟩ for N = 34 has a low cost of ≈ 4 × 103 Toffoli
the exact cost of the quantum circuits for sufficiently- gates (Table I), and does not require any ancillas, signifi-
accurate state preparation in the context of phase esti- cantly reducing the space-time volume and hardware con-
mation. State-of-the-art DMRG calculations for different nectivity requirements compared to the aforementioned
FeS systems typically involve bond dimensions of up to black-box state preparation approaches. Note that our
several thousands (2000 − 8000) to achieve accurate en- current understanding of the spin coupling structure of
26

FeMoCo is not as precise as our intuition for the bond magnitude improvement at stretched geometries. The
breaking examples in Ref. 52 (Section II). Unlike for the multireference non-orthogonal variant of this VQE algo-
bond breaking cases, where one type of spin coupling pat- rithm produces chemically accurate energies across the
tern dominates, accurate approximations for FeS eigen- whole binding curve at much reduced gate depths. Quan-
states might require superpositions of a larger number of tum subspace diagonalization through real-time evolu-
spin-coupled states. However, we anticipate that similar tion also benefits greatly from using CSF-based initial
machinery could be applied to such systems, and that states. We find that the subspace formed by evolving
the number of spin-coupled CSFs required scales at most each of the chemically relevant CSF states independently
linearly in the number of open-shell electrons. Preparing rapidly spans the relevant low-energy regions of Hilbert
linear combinations of CSFs would add a small overhead space, which provides a powerful method for computing
to the ∼ 103 gate counts for state preparation in Table I, accurate excited states, as well as accelerating conver-
but it is unlikely to be significant enough to affect our gence to the ground state. Since the CSFs are eigen-
conclusions (see Section IV I). functions of space and spin symmetry operators, this has
the advantage that the space and spin symmetry sectors
can be treated separately, which makes it possible to tar-
VII. CONCLUSIONS get specific molecular excited states. For ASP, beginning
with a single spin-coupled wavefunction at dissociation
While quantum algorithms such as QPE can in prin- enables preparation of the ground state at stretched ge-
ciple compute the exact eigenenergies of arbitrary elec- ometries with a much-reduced evolution time and there-
tronic Hamiltonians, in practice their performance de- fore lower circuit depth.
pends critically on the overlap of the target eigenstate We attribute the success of our CSF-based algorithms
with the initial state. Since the overlap between two to the accuracy of our spin-coupled molecular orbital
random vectors is inversely proportional to the size of theory for representing the dominant correlation effects
the Hilbert space, strategies for preparing initial states in molecules. This conceptual model provides both an
with the principal features of the Hamiltonian eigen- enhanced understanding of the nature of the electronic
states are an essential component of quantum simula- states, and a more compact representation of the key
tion. In quantum chemistry, the mean-field Hartree–Fock features of the eigenstate. By directly encoding the
method provides an initial state with a high overlap with dominant entanglement structure of the exact eigenstate
the exact ground eigenstate for weakly correlated elec- into the initial state, fewer variational parameters are
tronic systems, which pushes the orthogonality catastro- required for convergence. In the language of quantum
phe to larger electron numbers. Classically challenging chemistry, the initial state is constructed to contain the
problems, however, concern strongly correlated states. static, or strong correlation contributions, and the sub-
Here, the overlap of the RHF wavefunction with the ex- sequent refinement captures the remaining dynamic cor-
act eigenstate shrinks exponentially with the number of relation and orbital relaxation. Although we have re-
strongly correlated electrons, which presents a severe ini- stricted the analysis to a minimal basis set, quantita-
tial state problem. tively accurate energy estimates require a more refined
In previous work, we have demonstrated that the struc- discretization of the Hilbert space in the form of a larger
ture of strongly correlated eigenstates in chemical sys- basis set.[72] This does not present a problem since the
tems can be predicted from symmetry arguments, us- CSFs in a minimal basis can be projected into the larger
ing generalized spin-coupled molecular orbital theory.[52] basis, and orbital relaxation effects captured through the
Replacing the HF wavefunction with more general spin- VQE or QSD algorithms.
coupled orbital wavefunctions, linear combinations of A key benefit of the QSD through real-time evolu-
CSFs, provides high-quality initial states across the tion algorithm is that it can be used to compute excited
whole range of correlation regimes found in molecules. states and has less stringent requirements on the accu-
In this article, we have shown that the CSF-based initial racy of each ansatz circuit. However, implementing real-
states improve the performance of a wide range of quan- time evolution requires higher gate-depths than is feasible
tum algorithms for the simulation of quantum chemistry with near-term quantum hardware. We have therefore in-
on quantum computers, which can be applied both on troduced a new quantum algorithm, ADAPT-QSD, that
fault-tolerant quantum hardware as well as on near-term combines the advantages of QSD with the benefit of the
devices. The CSFs directly encode the strong correlation short-depth circuits of VQE. We have demonstrated that
in molecules and constitute specific patterns of entan- performing QSD in the basis of sequential ADAPT-VQE
glement. We have presented circuits for building CSFs states can significantly accelerate the convergence with
with depth linear in the number of electrons N , which respect to the number of operators and avoid stagnation,
generate entanglement exponential in N . and that ADAPT-QSD yields an energy estimate that is
For VQE with a fixed-depth quantum number preserv- orders of magnitude lower than ADAPT-VQE at com-
ing ansatz, our numerical simulations for N2 show that parable circuit depth. In view of the high popularity of
the CSF-based initial state returns significantly more ac- ADAPT-VQE for ground state computation, we believe
curate energies than a HF initial state, with orders of that ADAPT-QSD is a promising algorithm for excited
27

state calculations. Junior Research Fellowship) and Downing College, Cam-


The long-term hope is that fault-tolerant quantum de- bridge (Kim and Julianna Silverman Research Fellow-
vices with hundreds of logical qubits will be able to com- ship).
pute energies for systems that are out of reach for all clas-
sical methods. For strongly correlated molecules, such
as FeMoCo, which is considered a major target of quan- Appendix A: Spin eigenfunctions
tum simulation, our CSF-based formalism provides a sys-
tematic approach to constructing initial states with high Below are some examples of the spin-coupled state
overlap with the exact state to unlock the power of the N,1
|O0,0 ⟩. The general form is given in Eq. (13).
QPE algorithm. From approximate resource estimates,
we conclude that our spin-coupling-based approach to
initial state preparation is likely a necessary and suffi- 2,1 1
|O0,0 ⟩ = √ (|αβ⟩ − |βα⟩), (A1a)
cient requirement. This is because the CSF states can 2
be identified and prepared with much lower numbers of
Toffoli gates than expensive approximate classical wave-
4,1 1
functions such as DMRG, for a similar level of initial |O0,0 ⟩ = √ (|ααββ⟩ + |ββαα⟩)
state overlap. We anticipate that a possible fault-tolerant 3
(A1b)
workflow would be the following: 1) prepare the relevant 1
− √ (|αβαβ⟩ + |βαβα⟩ + |αββα⟩ + |βααβ⟩),
CSFs in the minimal basis, 2) optimize the orbitals, given 2 3
the CSF state, 3) apply a heuristic quantum algorithm
to compute additional correlation, e.g. through VQE or
a QSD algorithm, 4) use phase estimation to obtain the 6,1 1
|O0,0 ⟩= (|αααβββ⟩) − |βββααα⟩)
final energy. 2
The ultimate goal of computational chemistry is to in- 1
+ (|αββααβ⟩ + |αββαβα⟩ + |αβββαα⟩ + |ββαβαα⟩
crease our understanding of complex chemical systems. 6
Chemical properties are often the result of finely balanced + |ββααβα⟩ + |ββαααβ⟩ + |βαββαα⟩ + |βαβαβα⟩
competing effects and large parts of modern development + |βαβααβ⟩ − |βααββα⟩ − |βααβαβ⟩ − |βαααββ⟩
have focused on methods for obtaining tightly converged
− |αβαββα⟩ − |αβαβαβ⟩ − |αβααββ⟩ − |ααβββα⟩
energy estimates. The powerful numerical algorithms
such as DMRG and FCI-QMC fall in to this category, − |ααββαβ⟩ − |ααβαββ⟩)
as do the prevailing quantum algorithms. While these (A1c)
methods can accurately compute properties directly from
quantum mechanics, the brute-force nature of these algo-
rithms obscures the interpretation of the results. In con-
trast, our spin-coupled theory reveals configurations that 4,2 1
|O0,0 ⟩= (|αβαβ⟩ − |αββα⟩ − |βααβ⟩ + |βαβα⟩),
correspond to the dominant contributions to the ground 2
(A2)
and low-lying excited eigenstates. The CSFs correspond
to clear bonding motifs, providing direct insight into the
underlying electronic structure. The success of our CSF- 1
4,1
based quantum algorithms results from leveraging this |O1,1 ⟩= (|αααβ⟩ − |ααβα⟩ − |αβαα⟩ + |βααα⟩),
2
increased understanding. By understanding and exploit- (A3)
ing the spin coupling responsible for high levels of en-
tanglement in electronic states, our work paves the way
towards scalable quantum simulation of electronic struc- Appendix B: Quantum circuits for basis rotations
ture for classically challenging systems.
1. General fermionic basis rotations
ACKNOWLEDGMENTS
We can rotate the single-particle basis of any many-
particle state by applying exponential unitary transfor-
D.M.D. thanks Simon C. Benjamin, Nicholas Lee,
mations that in general act on the entire many-body
Katherine Klymko and Norm Tubman for help-
Hilbert space. A basis transformation of K basis func-
ful discussions. D.M.D. acknowledges the use of
tions (here, spin-orbitals or qubits) can be represented
the University of Oxford Advanced Research Com-
by a K × K unitary matrix with entries upq :
puting (ARC) facility in carrying out this work
(http://dx.doi.org/10.5281/zenodo.22558), as well as fi- X
nancial support by the EPSRC Hub in Quantum Com- ϕ̃p = ϕq upq , (B1)
q
puting and Simulation (EP/T001062/1). H.G.A.B. ac-
knowledges funding from New College, Oxford (Astor
28

where the sets {ϕp } and {ϕ̃p } are the original and the The cost for a transformation between two spatial or-
rotated basis set. Given a many-body state, this single- bitals is therefore 2 Cpq . The linear scaling with the or-
particle rotation is equivalent to applying a linear trans- bital indices p − q stems from the requirement of imple-
formation on the second-quantized operators: menting the Pauli-Z-strings that arise when mapping the
X fermionic operators onto qubit operators via the Jordan-
ã†p = upq a†pq (B2a) Wigner encoding (see SI in Ref. 26). To avoid this over-
q head, we can geometrically arrange the ordering of qubits
such that the orbitals that contribute to the same orbital
X transformation neighbor each other.
ãp = u∗pq aq (B2b)
q
2. Basis rotations for diatomic bond breaking
Following Thouless’ theorem,[109] the action of the
single-particle rotation on a many-body wavefunction |ψ⟩ For the diatomic systems discussed in Section II and
can be expressed by the following operator: Ref. 52, transforming a delocalized orbital to a localized
X  one requires building a linear combination of two spatial
U (u) = exp [log(u)]pq (a†p aq − a†q ap ) . (B3) orbitals. For N spin-coupled electrons, we have N such
pq transformations.
N,1
Consider the spin-coupled states |O0,0 ⟩ (Appendix A).
While this operator in general acts on the entire Hilbert
K To obtain the localized orbitals from the Hartree–Fock
space (its dimension is D × D, where  D = 2 for
K orbitals, we need rotations of the form ϕ̃L = √12 (ϕ1 +ϕ2 ),
an K-qubit computational basis or N for a particle-
number-conserving basis of N fermions in K single- ϕ̃R = √12 (ϕ1 − ϕ2 ) for each pair of spin-coupled electrons.
particle states), it can be implemented efficiently on a Each rotation involves two spatial orbitals, or four spin-
quantum circuit as follows.[110] orbitals. In total, it requires four one-body operators of
Following the approach outlined in Ref. 110, the uni- the form in Eq. (B6), each with cost Cpq , to localize two
tary operator U (u) can be decomposed into a series of electrons/spatial orbitals. The total cost for localizing or
Givens rotations with the form delocalizing N electrons/orbitals is
Y
exp θpq (a†p aq − a†q ap ) Crot,N = 2N Cpq . (B7)

U (θ) = (B4)
pq
For example, for N2 , we have
without any trotterization error. The corresponding ro-
tational angles θpq can be identified by performing a 1 1
zL = √ (σg + σu ), zR = √ (σg − σu ),
QR decomposition of the orbital transformation matrix 2 2
log(u), which can be solved classically. Each exponenti- 1 1
ated one-body operator can be implemented individually xL = √ (πu,x + πg,x ), xR = √ (πu,x − πg,x ), (B8)
2 2
using efficient quantum circuits.[110, 111] In general, the 1 1
indices run over all elements p, q ∈ {1, 2, .., K} with p > q yL = √ (πu,y + πg,y ), yR = √ (πg,y − πg,y ).
2 2
and thus there are K

2 such one-body operators. How-
ever, for many of the systems that we consider in this as described in Section II B. Using the qubit (spin-
work, the matrix u has many zero entries since only few orbital) ordering {σg , σ̄g , σu , σ̄u , πg,x , π̄g, x , πg,y , π̄g, y },
orbitals in {ϕq } contribute to each ϕ̃p , and therefore
 the where the absence (presence) of an overbar indicates a
cost of the basis transformations is lower than K 2 . α (β) spin-orbital, respectively, we always have p − q = 2
The CNOT cost of implementing the exponential of a and thus Cpq = 2 (Eq. (B6)). Therefore, to localize 2, 4,
one-body operator and 6 spatial orbitals, as required to apply the circuits
for preparation of (|ϕ2x ⟩, |ϕ2y ⟩), |ϕ4 ⟩, and |ϕ6 ⟩ (Section
exp[θpq (a†p aq − a†q ap )] (B5) II), the cost is 4N i.e. 8, 16, and 24 CNOTs, respectively.
that rotates between spin-orbitals p and q on a quantum
circuit is [18, 112] Appendix C: Quantum circuits for spin
( eigenfunctions: explicit decompositions and cost
2(p − q) + 1, p−q >2
Cpq = (B6)
2, p − q ∈ {1, 2}. Below we analyze the exact cost of the CSF prepara-
tion circuit in Section IV. We focus on counting CNOT
In this work, we use restricted orbitals and therefore gates because entangling gates are typically the biggest
the transformations are the same for both spin-up and sources of noise on devices without error correction.[89]
spin-down orbitals that share the same spatial orbital. In Section C 4, we also compute the Toffoli gate counts for
29

longer-term fault-tolerant implementations of these cir- 2-qubit gates 3-qubit gates CSn
cuits, since Toffoli gates dominate the cost when applying S2 1 0 3
error-correcting codes.[77] We separately consider qubit S3 2 1 11
architectures with all-to-all interactions between qubits, S4 3 3 24
as well as devices with restricted nearest-neighbor-only S5 4 6 42
connectivity; specifically, linear, and planar (grid-like) S6 5 10 1 65
connectivity. The latter require decomposing gates be-
tween distant qubits into nearest-neigbor gates, which TABLE III: Number of two-qubit and three-qubit gate
introduces a gate overhead. Noten that1 we • had make • blocks (Figs.
Ryto(✓) n 2b,l 2c)•required
Ry (✓) •
for application of
some circuit design choices to get concrete numbers for symmetric
n l+1 state preparation • unitaries S , and cost CSn
the gate counts, but these are notn necessarily optimal
• n
(number of CNOT
··· gates) in the circuit for Sn after
and could possibly be improved if tailored for a particu- decomposition into elementary gates (Eq. (C1)).
(a)
lar architecture or for the simulated system. Two-qubit gate n •
The circuit for preparing a single spin eigenfunction of
N,1
the form |O0,0 ⟩ contains three parts (see Sections IV E, (b) Three-qubit
decomposed into single-qubit rotationsgate
and two CNOTs
IV F and the circuit diagram in Fig. 3): (Fig. 10) to give CNOT cost of 2(n−1) = N −2; the third
step requires n = N/2 CNOT gates. The total CNOT
1. Preparation of the input FIG.
state 1: Gate blocks used forinput
count for application of Dicke
state preparation state
is therefore
2. Application of the two Snpreparation
unitaries unitaries, as derived in Bartschi.
all 3N
Cin = − 2. (C2)
2
3. Mapping from spin to Fock space.
Below we discuss their cost in terms of the number
of CNOT gates, assuming all-to-all Ry (✓)
i connectivity. While i Ry (✓/2) Ry ( ✓/2)
the unitaries Sn can be implemented using circuits with
j •
only linear (nearest-neighbor) connectivity, this is not the
j • •
j,i and the
case for circuits that implement the(a)input
CRstate
y (✓) gate. (b) Decomposed
FIG. 10: Decomposition of CRyj,igate.
(θ).
mapping from spin to Fock space. We therefore compare
the cost assuming all-to-all connectivity and the cost with The mapping from spin to Fock space costs
connectivity restrictions (linearFIG. 2: Decomposition
and planar topologies). of CRyj,i (✓) into two CNOTs.
all
Cmap =N (C3)

1. All-to-all connectivity for hardware with all-to-all connectivity, giving a total


cost of
Let C2 and C3 denote the number of CNOT gates re- all
Ctot = 2CSn + Cinall all
+ Cmap
quired to implement the two- and three-qubit gate blocks 5 9  3 
required for the Dicke state preparation circuits (Figs. = N 2 − N + 4 + N − 2 +N (C4)
2b, 2c). Each two-qubit gate block consists of a con- 4 2 2
5 2
trolled Ry -gate conjugated by two CNOTS. This can be = N − 2N + 2.
decomposed into 3 CNOT gates: C2 = 3 (Fig. 3 in [112]). 4
The three-qubit gate consists of a Ry -gate controlled by
two qubits, which can be decomposed into 5 CNOT gates: 2. Linear connectivity
C3 = 5 (Fig. 3 in [63]). Since each unitary Ml,l−1 con-
sists of one two-qubit gate block and l − 2 three-qubit
For hardware with linear connectivity, the cir-
gate blocks (Section IV C), the overall CNOT cost for
cuit Sn can be implemented directly without any
implementation of the operator Sn is
modification.[63] However, in the last step of the in-
n
X n 
X  put start preparation circuits, some CNOT gates act
CSn = CMl,l−1 = C2 + C3 (l − 2)) between distant qubits and therefore cannot be imple-
l=2 l=2
(C1) mented as such on a linear topology. The circuit can be
n  adapted for hardware with nearest-neighbor connectivity
X 5
 9
= 3 + 5(l − 2) = n2 − n + 2, via a recompilation of the CNOT accordion into nearest-
2 2
l=2 neighbor CNOTs (see Fig. 11). Evidently, some of the
where CMl,l−1 denotes the CNOT cost of implementing CNOT gates cancel out after the decomposition. The
the unitary Ml,l−1 . Table III shows the cost up to n = 6. CNOT count for the accordion with linear connectivity
The cost of the circuit for input state preparation is is
linear in the number of CNOT gates (Section IV E). The N
−2
first step only contains X gates; the second step re-
2
X N2
Cacc (N ) = N − 1 + 2 (N − 2i − 2) = − 1. (C5)
quires controlled rotation gates CRy (θi ) which can be i=0
2
5
6
7
8
30
1,8
Uin,8 S41,4 ⌦ S45,8 Umap,8
8,1 8
FIG. overall
The 3: A circuit that implements
CNOT the unitary Vof
cost for preparation 8 to prepare
the inputthestate
spin-coupled
on state |O0,0 iOne
a grid. = V8 |00i
can . carefully
It consists of
align the spin-orbitals to
three linear
with parts, delimited by the dashed
connectivity is vertical lines: preparation of the inputavoid
state the
(left),overhead
preparation
ofofrecompiling
the spin the two-qubit (CNOT)
eigenfunction through Dicke state preparation circuits (center), and mapping from spin to spatial orbital occupation
(right). gates required by the spin-to-Fock space mapping Umap .
N2  N2
lin
Cin = (N − 2) + −1 = + N − 3. (C6) Specifically, starting from a basis of localized orbitals
2 2 where {iL } and {iR } are spin-up orbitals localized on
the left and right atom, and {īL } and {īR } are the cor-
responding spin-down orbitals, the circuits for prepara-
• • N,1
• • • • • • tion of a single CSF |O0,0 ⟩ require the following pairs of
• • • • • • • • • qubits/spin-orbitals to be directly connected to avoid any
• • • • • •
• •
gate overhead due to recompiling CNOT gates between
non-neigboring qubits:
CX 1,6 CX 2,5 CX 3,4
1. Spin-up orbitals iL and (i + 1)L for all i ∈
FIG. 4: Recompilation of CNOT accordion for N = 6, namely CX 1,6 CX 2,5 CX 3,4{1,
, into
2, gates
...N restricted
− 1}, mustto linear
be
connected to each other.
FIG. 11: Recompilation of CNOT accordion for N = 6,
connectivity.
namely CX 1,6 CX 2,5 CX 3,4 , into gates restricted to This is required for the circuits for input state
linear connectivity. This corresponds to the last part of preparation Uin,N and the circuits SN/2 (left and
the circuit for preparation of the input state (see Fig. 3 middle section in Fig. 3). The ordering within the
for N = 8 example). left and right subsystem is irrelevant due to the
symmetry of the state within that subspace.
To implement Umap , one could apply the same naive 2. Every spin-up orbital iL/R must be connected with
decomposition of the CNOTs into linear connectivity as their corresponding spin-down orbital īL/R . This is
in Fig. 11, where in this case it is not directly obvious required for the spin-to-Fock-space mapping Umap
if any CNOTs cancel out. This has a large overhead as (right section in Fig. 3).
all
what was previously Cmap = N CNOTs now becomes
3. For the basis rotation circuits (Appendix B 2), each
lin
= N N + 2(N − 1) + (N − 2) = 4N 2 − 4N. (C7)
 
Cmap left orbital iL or īL must be connected to its corre-
sponding right orbital iR or īR .
The total CNOT cost for hardware with linear connec-
tivity would thus be Assuming a rectangular grid, there is no ordering that
satisfies all three conditions. If we only wish to prepare
lin lin lin
Ctot = 2CSn + Cin + Cmap a single CSF without performing orbital rotations, we
5 9   1    can place all spin-up qubits iL/R in one line, and all the
= N 2 − N + 4 + N 2 + N − 3 + 4N 2 − 4N corresponding spin-down qubits īL/R in a line below or
4 2 2
23 2 15 above it. With this ordering, the first two conditions are
= N − N + 1. met and therefore there is no overhead in implementing
4 2
(C8) the input and Dicke state preparation circuits. However,
basis rotations would require applying SWAP gates, with
Due to the decomposition of the CNOT accordion into an overhead scaling as O(N 2 ).
nearest-neigbor gates, the depths of the circuits for in- If we wish to use multiple CSFs (including localized
put state preparation Uin,N and for spin-to-Fock map- and delocalized states) in a quantum algorithm, we can
ping Umap scale as O(N 2 ), which is worse than the O(N ) prepare the CSF in a basis of localized orbitals and then
scaling of the depth for the Dicke circuits Sn . Further- rotate the basis as in Appendix B 2. This is preferably
lin done using the following arrangement to exploit planar
more, Cmap dominates the gate count of CSF preparation
with linear connectivity (Eq. (C8)). It is possible that connectivity without a large gate cost: set all left spin-up
this circuit can be improved through further design opti- orbitals {iL } , i = {1, 2, ...n} in a horizontal line, and all
mization. Nevertheless, since the optimal recompilation right spin-up orbitals {iR }, i = {1, 2, ...n} in parallel, be-
often depends on the details of the particular hardware low the spin-up line. Then, set the left spin-down orbitals
architecture (including the type of native gate set that is īL }, i = {1, 2, ...n} in a line on top, and the right spin-
available), we do not consider these optimisations in this down orbitals īR }, i = {1, 2, ...n} in a line below. This
work. Below, we show that this scaling overhead can be enables applying the circuits for preparation of a single
removed in the case of planar connectivity. CSF with nearly the same cost as if we had full all-to-
all connectivity. The overhead using this grid structure
comes from the need of decomposing the CNOT accor-
3. Planar connectivity dion for preparation of the input state to a linear array
(Appendix C 2). Since the spin-up orbitals iL and iR are
Planar (two-dimensional) structures are the natural already connected ∀i ∈ {1, 2, ..., n}, those rotations can
topology of superconducting hardware.[88] For example, be implemented directly. However, to connect the spin-
Google’s Sycamore processor connects adjacent qubits down orbitals īL and īR , we must swap iL and īL and
31

iR and īR for all i. Then, we can implement the corre- affecting the number of T -gates per rotation gate is
sponding basis rotation circuits. The overhead for going
from all-to-all to planar connectivity is therefore 2n = N NT R = c⌈log(1/ϵr )⌉, (C11)
SWAP gates, or 3N CNOT gates. This linear overhead where c is a constant that depends on the implementa-
is small compared to the total cost for CSF preparation, tion. For deterministic algorithms, c is lower-bounded by
which is quadratic in N (Eq. (C4)) 3,[114, 115] but this can be improved using randomized
The total CNOT count for state preparation of a single methods such as the repeat-until-success (RUS) approach
CSF with N spin-coupled electrons in the localized basis in Ref. 116, which finds an empirical average of c = 1.15.
is: Here, we consider the probabilistic RUS technique, where
pl lin all 7 5 the cost includes an extra additive factor:
Ctot = 2CSn + Cin + Cmap = N 2 − N + 1. (C9)
4 2
NT R ≈ c⌈log(1/ϵr )⌉ + 9.2. (C12)
Preparing that state and then rotating M pairs of spa-
tial orbitals (electrons) to a delocalized basis has an ad- Since we later compare the cost of initial state prepa-
ditional cost of ration with the cost of implementing quantum phase
pl
estimation, we translate the T -gate counts into Tof-
Crot = 3M + 2M Cp,q = 3M + 4M = 7M, (C10) foli counts to compare with the Toffoli counts in state-
of-the-art qubitization-based algorithms for electronic
where M = N if we rotate all orbitals (Eq. (B7)). This
structure.[64, 97] The corresponding number of Toffoli
might not be needed depending on the application (e.g.
gates is roughly TR ≈ 12 NT R in a surface code imple-
when preparing linear combinations of CSFs are needed,
mentation, since the cost of applying a Toffoli gate is ap-
we must rotate a subset of the orbitals for each CSF).
proximately twice the cost of applying a T -gate using the
magic state factories introduced in Ref. 113. Thus, the
4. Fault-tolerant circuits and non-Clifford cost number of T -gates per rotation is 21 (cR⌈log(1/ϵr )⌉ + 4.6).
N,1
The circuits for preparation of |O0,0 ⟩ also require ap-
Universal gate sets for fault-tolerant quantum com- plication of controlled Ry gates CRy and CCRy (Section
putation include Clifford + Toffoli gates or Clifford + IV). Controlling rotation gates does not have any non-
T . Since Clifford gates such as the set of Clifford Clifford overhead if the rotation angle is given by a classi-
group generators {H, S, CX} are efficiently classically cal register, as we can simply use CNOT gates to change
simulable,[51] most fault-tolerant resource estimates typ- the direction of the rotation θ ↔ −θ.[117] This is indeed
ically assume that their cost is negligible compared to the the case in the circuits for preparation of the spin eigen-
N,1
overall cost.[31, 64] On the other hand, expensive magic- function |O0,0 ⟩, since the entire state preparation circuit
state distillation protocols are required to apply noisy (including the angles) can be specified classically before
non-Clifford gates with high fidelity within an error- any quantum computation. To determine the number of
correcting code, resulting in a large space-time overhead Toffoli gates, we must therefore simply count the number
orders of magnitude larger than the cost of fault-tolerant of rotations, irrespective of whether these are controlled
Clifford gates.[113] We therefore focus on minimizing and or not.
counting non-Clifford gates. This allows direct compari- Finally, consider a circuit with R rotations with over-
son with the literature for Hamiltonian simulation of elec- all error ϵ. A naive error analysis based on a triangle
PR
tronic structure (see [31, 64, 97, 101] and Table III in [64] inequality ϵ ≤ r=1 ϵr would lead us to allocate errors
for an overview). as ϵr = ϵ/R, taking the number of bits to represent each
N,1
The circuit VN for preparation of the CSF |O0,0 ⟩, pre- rotation angle as b = log(1/ϵr ). Ref. 118 showed that if
sented in Section IV E, requires converting the continu- one considers the errors as random rather than coherent,
ous rotation gates Ry (θ) to a discrete gate set of Clif- one can analyze them in terms of a random walk. This
ford + Toffoli gates. Due to the finite precision in the gives an √error bound that is tighter by a quadratic fac-
binary representation of continuous numbers, this intro- tor ϵ → ϵ, or equivalently halves the number of bits b.
duces an error ϵr per rotation gate that can be exponen- With this, we obtain a reduced average of
tially suppressed since the gate and qubit counts scale as 1
O(log(1/ϵr )). Since the number of rotations in VN scales TR = cR⌈log(1/ϵr )⌉ + 4.6 (C13)
4
as O(N 2 ), the asymptotic Toffoli complexity is Õ(N 2 ),
1
where the tilde indicates suppression of polylogarithmic Toffoli gates per rotation, where b = 2 ⌈log(1/ϵr )⌉ and
factors. Below, we derive the exact cost. c = 1.15.

a. Rotation gate synthesis b. Toffoli cost for preparation of spin eigenfunctions

A commonly-used approach is to synthesize rotations As discussed in Section IV C, each subcircuit Ml,l−1


in terms of Clifford + T gates. The dominant factor that implements the symmetric state preparation unitary
32

Sn (Eq. (22)) consists of one two-qubit gate gate block bond lengths, it is necessary to prepare linear combi-
(Fig. 2b) and l − 2 three-qubit gate blocks. The total nations of spin eigenfunctions (Section IV I). This can
number of rotations for implementation of Sn is therefore be achieved by controlling the circuits for preparation
N,1
n of a single CSF of the form |O0,0 ⟩. It does not affect
X 1 the asymptotic scaling with N but introduces a constant
RSn = (l − 1) = n(n − 1). (C14)
2 factor overhead. The overhead is small because one only
l=2
needs to control a small part of the circuit for input state
This must be implemented twice to implement VN : preparation, Uin,N , as well as the circuit for Umap .
N,1 N
|O0,0 ⟩ = VN |00⟩ . The number of rotation gates re- To control the circuit for input state preparation and
quired for preparation of the input state is n (Section the Dicke circuits SN/2 , it is sufficient to simply control
IV E). Thus, the total number of rotation gates for VN , the single-qubit Pauli X-gates at the beginning of Uin,N ,
where n = N/2 is as well as the first single-qubit rotation gate Ry (θN/2 ).
If the control qubit is in 0, the remaining part of the cir-
1 2 cuit Uin,N and the entire circuit SN/2 ⊗ SN/2 act like the
R = 2RSN/2 + Rin,N = n(n − 1) + n = n2 = N . (C15)
4 identity, because the CNOT and rotation gates therein
The total (average) Toffoli cost thus becomes have no effect, and thus this effectively controls the entire
  circuit. Therefore the only overhead for controlling this
T = R × TR = R 0.2875⌈log(R/ϵ)⌉ + 4.6 . (C16) is n = N/2 CNOT gates, as well as the cost of convert-
ing the Ry (θn ) to a CRy (θn ) gate. The CRy (θn ) can be
The method requires only a single ancilla qubit to check implemented with 2 CNOTs and two Ry gates (Figure
if each rotation was implemented successfully.[116] 10) and therefore has an overhead of 2 CNOT gates and
Our goal is to use the spin eigenfunctions as initial one Ry gate.
states in fault-tolerant quantum algorithms. Thus, we Finally, consider the circuit Umap , consisting of N
must consider two sources of errors: the error due to i,j
CNOT gates of the form CX = Xi CX i,j Xi . This can
state preparation ϵSP , and the error due to implementa-
be controlled by simply controlling the two X gates and
tion of the quantum algorithm itself, e.g. quantum phase
therefore requires 2N CNOTs gates. The overhead (cost
estimation, ϵQPE . The total error will at worst be
to add on top of state preparation cost without controlled
ϵtot ≤ ϵSP + ϵQPE . (C17) qubits) is 3N − N = 2N . The total CNOT overhead be-
comes:
Since our circuits for initial state preparation are very N  5
efficient, a reasonable strategy would be to allocate most Cctrl = + 2 + 2N = N + 2. (C18)
of the error to the quantum algorithm itself, rather than 2 2
the state preparation task. For example, in the con- In fault-tolerant hardware, the dominant cost comes
text of quantum phase estimation, one could choose from implementing the non-Clifford (Toffoli) gates, not
ϵtot = 0.0016 (in Hartree atomic units) to achieve chem- CNOT gates. As discussed in Appendix C 4, rotations
ical accuracy in the energy estimation. We divide this controlled by arbitrary qubits only have non-Clifford
into ϵSP = 10−7 and ϵQPE = ϵtot − 10−7 . Inserting these overhead compared to uncontrolled rotations.[117]
values and Eq. (C15) into Eq. (C16), we get the Toffoli Therefore, there is no Toffoli overhead for controlling any
N,1
counts for preparation of |O0,0 ⟩ states reported in Ta- part of the state preparation circuit VN .
ble I. Even for the largest systems with N = 34, the cost
is only T = 3989 ≈ 4 × 103 . This is indeed very low
compared to the typical cost of quantum phase estima- 6. Other spin eigenfunctions
tion, which is in the order of 1010 using state-of-the-art
techniques.[64, 97] We briefly discuss the cost for preparing the state
Note that we have also considered the elegant phase
#N/2
gradient technique from Ref. 119 as an alternative to ro-
"
N,2 1
tation gate synthesis. While we found that the Toffoli |O0,0 ⟩ = √ (|αβ⟩ − |βα⟩) . (C19)
cost per rotation can be slightly lower depending on the 2
choice of ϵ, the advantage is washed away due to the
one-off cost of preparing the phase gradient state (which (Section IV H, Eq. 32). Working entirely in the spin
in itself requires rotation gate synthesis), even for the space, where this maps to the qubit state
largest system we consider, where N = 34. " #N/2
1
√ (|1001⟩ − |0110⟩) , (C20)
2
5. Controlling state preparation circuits
state preparation only requires one CNOT per two-
In some quantum algorithms, e.g. when preparing ini- electron singlet, therefore N/2 CNOTs. Two additional
tial states for VQE or phase estimation at intermediate CNOT gates are required to implement the spin-to-Fock
33

space mapping Umap . Therefore the total CNOT cost as- Using Eq. (D2), explicit expressions for the derivatives
all
suming all-to-all connectivity is Ctot = N/2+N = 3/2N . with respect to the linear coefficients can be obtained as
pla
This is the same for planar connectivity, Ctot =
all
Ctot if we carefully align the spin-orbitals on the qubit ∂E h i
= 2 ⟨ΨI (θI )|Ĥ|Ψ(Θ, C)⟩ − ⟨ΨI (θI )|Ψ(Θ, C)⟩ ,
grid using the strategy in Appendix C 3. For lin- ∂CI
ear connectivity, we can choose the qubit ordering X M

{α1 , β1 , α2 , β2 , ..., αN , βN , } to minimize the cost of Umap =2 [HIJ − SIJ ]CJ


to 2 × (N/2). The CNOT between orbitals αi and αi+1 , J=1
which are separated by the βi qubit, can be decomposed (D5)
into 3 CNOTs using the CNOT accordion (Fig. 11). The
total cost for linear connectivity thus becomes where the Hamiltonian and overlap coupling elements are

lin 5
Ctot = N. (C21) HIJ = ⟨ΨI (θI )|Ĥ|ΨJ (θJ )⟩ , (D6a)
2
SIJ = ⟨ΨI (θI )|ΨJ (θJ )⟩ , (D6b)
Controlling these circuits only requires controlling the
Ry (−θ/2) rotation for each two-electron singlet, which and we assume that the linear coefficients C are real
costs two CNOTs (Fig. 10). Adding the cost of control- valued. These derivatives may be computed using the es-
ling Umap , the CNOT overhead for controlling the prepa- tablished circuits for measuring nonorthogonal coupling
N,2
ration of |O0,0 ⟩ is N/2 + N = 3/2N . On a fault-tolerant terms outlined in Ref. 14, which are also used in subspace
device, Ry (−θ/2) can be factorized into HX, therefore diagonalization approaches. Similarly, using Eq. (D1),
the state preparation and its controlled version can be the derivatives with respect to the gate parameters are
implemented entirely with Clifford gates.
∂E Xh
= 2 CI ⟨∂θIi ΨI (θI )|Ĥ|ΨJ (θJ )⟩
∂θIi
J (D7)
Appendix D: Energy and gradients in i
Nonorthogonal VQE − ⟨∂θIi ΨI (θI )|ΨJ (θJ )⟩ CJ ,

The nonorthogonal VQE (NO-VQE) algorithm vari- where the partial derivative of the wavefunction is
ationally optimizes the energy of a wavefunction corre-
sponding to the linear combination defined in Eq. (39).
Y ∂UIi (θIi ) Y
|∂θIi ΨI (θI )⟩ = UIj (θIj ) UIj (θIj ) |ΦI ⟩ .
A simultaneous optimization of the gate parameters Θ = j<i
∂θIi
k>i
(θ1 , . . . , θM ) and linear coefficients C requires the gradi- (D8)
ent of the energy with respect to each variable. In what For the QNP ansatz, the circuits for these partial deriva-
follows, we define the correlated basis states as tives can be constructed using the parameter shift rules
Y detailed in Ref. 18. Therefore, the coupling terms
|ΨI (θI )⟩ = UIi (θIi ) |ΦI ⟩ , (D1) ⟨∂θIi ΨI (θI )|Ĥ|ΨJ (θJ )⟩ and ⟨∂θIi ΨI (θI )|ΨJ (θJ )⟩ can be
i evaluated with the same circuit architecture used to eval-
such that the full wavefunction is given by uate the Hamiltonian and overlap terms in Eq. (D6) with
a constant prefactor.
L
X With these gradient expressions, the NO-VQE algo-
|Ψ(Θ, C)⟩ = CI |ΨI (θI )⟩ . (D2) rithm proceeds using the standard L-BFGS optimiza-
I=1 tion approach. The initial linear coefficients C are ob-
tained by solving the generalized eigenvalue problem with
Since the states {|ΨI (θI )⟩} are not mutually orthogonal, Θ = 0. Although optimizing the expectation value of the
the VQE optimization requires gradients of the energy energy means that the linear expansion [Eq. (D2)] does
expectation value not need to be normalized, we obtain more stable opti-
mization by normalizing C on each iteration.
⟨Ψ(Θ, C)|Ĥ|Ψ(Θ, C)⟩
E(Θ, C) = , (D3)
⟨Ψ(Θ, C)|Ψ(Θ, C)⟩
Appendix E: Matrix elements for QSD based on
which are obtained through the quotient rule as real-time evolution
∂E h i
= 2 ⟨∂Θ Ψ|Ĥ|Ψ⟩ − E ⟨∂Θ Ψ|Ψ⟩ , (D4) For simplicity, we restrict the analysis below to the case
∂Θ
where we only time-evolve a single reference state |Ψ0 ⟩,
∂E
and likewise for ∂C . Here, ∂Θ ≡ ∂Θ ∂
and we have and note that the conclusions remain unchanged for cases
exploited the Hermitian symmetry, e.g. ⟨Ψ|Ĥ|∂Θ Ψ⟩ = with multiple reference states. Choosing a linear time
grid, tj = j∆t with j = 0, 1, ..., NT , we form a subspace
⟨∂Θ Ψ|Ĥ|Ψ⟩.
34

of M = NT + 1 states. The overlap matrix elements where χi is the bond dimension, and each bond dimension
between the time-evolved states forming the subspace are is bounded by the maximal bond dimension, χi ≤ χ.
The preparation of an MPS on a quantum computer
Sj,k = ⟨Ψj |Ψk ⟩ = ⟨Φ0 |e−iH∆t(k−j) |Φ0 ⟩ . (E1) can be achieved using the sequential method by Schön et
al.[107] This scales as O(M ) in the number of gates and
has depth O(M ). Although an improved scaling of depth
Replacing the Hamiltonian operator with the time-
O(log(M )) can be reached through the recent technique
evolution operator U (∆t) := e−iH∆t , we can write Sj,k =
by Malz et al. [120], which is provably optimal,[120] this
⟨Φ0 |[U (∆t)]k−j |Φ0 ⟩. The matrix elements of U (∆t) in
only applies to MPS with short-range correlations. For
the basis of expansion states are:
long-range correlated states such as GHZ states, it has
been proven that state preparation circuits of depth M
Uj,k = ⟨Φj |U (∆t)|Φk ⟩ = ⟨Φ0 |e−iH∆t(k−j+1) |Φ0 ⟩ are optimal based on Lieb-Robinson bounds.[120, 121]
(E2)
= Sj,k+1 = Sj−1,k . Since we do not expect any area-laws to apply to the en-
tanglement of most molecular eigenstates,[106] we must
consider the method for general MPS.[107]
These expressions show that, if we reformulate the The scaling of the deterministic method in Ref. 107 is
eigenvalue problem in Eq. (41) to use the time-evolution also polynomial in the bond dimension. Specifically, for
operator U (∆t) rather than the Hamiltonian, the ma- generic MPS, each circuit contains M gate blocks, one
trix elements U (∆t)j,k become equivalent to the overlap [m ]
for each of the tensors Ai i . Each gate block is a multi-
matrix elements (shifted by one row or column), which
qubit unitary acting on na = log(χi ) ancilla qubits (cor-
has a Toeplitz structure. This equivalence is advanta-
responding to the virtual Hilbert space) and one physical
geous for implementations on quantum hardware, as it
qubit. Decomposing a block into single and two-qubit
only requires measuring the overlap matrix rather than
gates requires a circuit with depth exponential in na , or
separately measuring the Hamiltonian and overlap (intu-
O(χi ).[120, 122, 123] Therefore, the total gate complex-
itively, the matrix elements in Eq. (E2) correspond to an
ity is O(M χ2 ), which is proportional to the scaling of
autocorrelation function).[39, 42] The number of matrix
the dimension of the overall tensor.[8] Although approx-
elements is reduced from 2M 2 to M +1. This result holds
imate methods might ameliorate the scaling,[122–124] it
even when the time-evolution operator is Trotterized, as
is unclear if they can retain sufficient accuracy and their
long as the time grid is linear.[42]
error cannot be theoretically bound.[122]
Formichev et al. provided a detailed derivation of the
cost and a more explicit circuit implementation of the
Appendix F: Quantum circuits for preparation of
MPS state preparation circuits from Ref. 107 using mod-
matrix product states ern quantum linear algebra techniques.[73] This imple-
mentation has Toffoli cost of χi−1 [8χi d + b log(χi d) +
log(χi d)] for each block. Assuming χi = χ for all i, we
Here, we discuss the scaling of known quantum circuits obtain the following approximate cost for the entire MPS
for preparation of matrix product states (MPS). An MPS preparation circuit:
of M sites has the form
M
X
M TMPS = χi−1 [8χi d + b log(χi d) + log(χi d)]
[m1 ] [m2 ] [m ] (F2)
X Y
|Ψ⟩ = A1 A2 · · · AM M |mi ⟩ . (F1) i=2
m i=1 ≈ (M − 1)χ[32χ + (b + 1) log(4χ)].

Here, mi ∈ {0, 1, ..., d − 1} is the physical index that


runs over the possible occupations of the Hilbert space Appendix G: Computational details
of a single site (of local dimension d), and the vector
m = (m1 , m2 , ..., mM ) defines the occupation numbers We obtained the Hamiltonians using PySCF[125, 126]
for each site. In the case of quantum chemistry, each site and used Openfermion[127] to define the operator matri-
can be mapped to a spatial orbital, therefore the site’s ces. NO-VQE and ADAPT-VQE calculations were per-
Hilbert space dimension is d = 4, and M is the number formed using a developmental version of GMIN.[128] We
[m ]
of spatial orbitals.[8]. Each Ai i is a tensor of order χi , developed in-house Python code for all other tasks.

[1] F. A. Evangelista, Perspective: Multireference coupled [2] A. Gilyén, Y. Su, G. H. Low, N. Wiebe, Quantum singu-
cluster theories of dynamical electron correlation, Jour- lar value transformation and beyond: Exponential im-
nal of Chemical Physics 149, 30901 (2018). provements for quantum matrix arithmetics, Proceed-
35

ings of the Annual ACM Symposium on Theory of Com- [23] U. Baek, et al., Say no to optimization: A nonorthog-
puting pp. 193–204 (2019). onal quantum eigensolver, PRX Quantum 4, 030307
[3] J. Kempe, A. Kitaev, O. Regev, The complexity of the (2023).
local hamiltonian problem, SIAM Journal on Comput- [24] R. D’Cunha, T. D. Crawford, M. Motta, J. E. Rice,
ing 35, 1070 (2006). Challenges in the use of quantum computing hardware-
[4] B. O’Gorman, S. Irani, J. Whitfield, B. Fefferman, In- efficient anstze in electronic structure theory, Journal of
tractability of electronic structure in a fixed basis, PRX Physical Chemistry A 127, 3437 (2023).
Quantum 3, 020322 (2022). [25] C. Gustiani, R. Meister, S. C. Benjamin, Exploiting sub-
[5] W. Kohn, Nobel lecture: Electronic structure of mat- space constraints and ab initio variational methods for
terwave functions and density functionals, Reviews of quantum chemistry, New Journal of Physics 25, 073019
Modern Physics 71, 1253 (1999). (2023).
[6] J. H. V. Vleck, Nonorthogonality and ferromagnetism, [26] H. G. Burton, D. Marti-Dafcik, D. P. Tew, D. J. Wales,
Physical Review 49, 232 (1936). Exact electronic states with shallow quantum circuits
[7] T. Helgaker, P. Jörgensen, J. Olsen, Molecular from global optimisation, npj Quantum Information
Electronic-Structure Theory (John Wiley & Sons, 2014). 2023 9:1 9, 1 (2023).
[8] G. K.-L. Chan, S. Sharma, The density matrix renor- [27] K. Dalton, et al., Quantifying the effect of gate errors
malization group in quantum chemistry, Annual Review on variational quantum eigensolvers for quantum chem-
of Physical Chemistry 62, 465 (2011). istry, npj Quantum Information 2024 10:1 10, 1 (2024).
[9] N. M. Tubman, et al., Postponing the orthogo- [28] E. Farhi, J. Goldstone, S. Gutmann, M. Sipser, Quan-
nality catastrophe: efficient state preparation for tum computation by adiabatic evolution, arXiv:quant-
electronic structure simulations on quantum devices, ph/0001106 (2000).
arXiv:1809.05523 (2018). [29] A. Aspuru-Guzik, A. D. Dutoi, P. J. Love, M. Head-
[10] A. Peruzzo, et al., A variational eigenvalue solver on a Gordon, Simulated quantum computation of molecular
photonic quantum processor, Nature Communications energies, Science 309, 1704 (2005).
5, 1 (2014). [30] L. Veis, J. Pittner, Adiabatic state preparation study
[11] J. R. McClean, J. Romero, R. Babbush, A. Aspuru- of methylene, Journal of Chemical Physics 140, 224109
Guzik, The theory of variational hybrid quantum- (2014).
classical algorithms, New Journal of Physics 18 (2016). [31] M. Reiher, N. Wiebe, K. M. Svore, D. Wecker,
[12] H. R. Grimsley, S. E. Economou, E. Barnes, N. J. May- M. Troyer, Elucidating reaction mechanisms on quan-
hall, An adaptive variational algorithm for exact molec- tum computers, Proceedings of the National Academy
ular simulations on a quantum computer, Nature Com- of Sciences of the United States of America 114, 7555
munications 10 (2019). (2017).
[13] O. Higgott, D. Wang, S. Brierley, Variational quantum [32] T. Albash, D. A. Lidar, Adiabatic quantum computa-
computation of excited states, Quantum 3, 156 (2019). tion, Reviews of Modern Physics 90, 015002 (2018).
[14] W. J. Huggins, J. Lee, U. Baek, B. O’Gorman, K. B. [33] V. Kremenetski, C. Mejuto-Zaera, S. J. Cotton, N. M.
Whaley, A non-orthogonal variational quantum eigen- Tubman, Simulation of adiabatic quantum computing
solver, New Journal of Physics 22, 073009 (2020). for molecular ground states, The Journal of Chemical
[15] Z. Cai, Resource estimation for quantum variational Physics 155, 234106 (2021).
simulations of the hubbard model, Physical Review Ap- [34] K. Sugisaki, K. Toyota, K. Sato, D. Shiomi, T. Takui,
plied 14, 14059 (2020). Adiabatic state preparation of correlated wave func-
[16] Y. S. Yordanov, V. Armaos, C. H. Barnes, D. R. tions with nonlinear scheduling functions and broken-
Arvidsson-Shukur, Qubit-excitation-based adaptive symmetry wave functions, Communications Chemistry
variational quantum eigensolver, Communications 2022 5:1 5, 1 (2022).
Physics 4, 1 (2021). [35] S. Lee, et al., Evaluating the evidence for exponential
[17] B. V. Straaten, B. Koczor, Measurement cost of metric- quantum advantage in ground-state quantum chemistry,
aware variational quantum algorithms, PRX Quantum Nature Communications 2023 14:1 14, 1 (2023).
2, 030324 (2021). [36] J. R. McClean, M. E. Kimchi-Schwartz, J. Carter, W. A.
[18] G.-L. Anselmetti, D. Wierichs, C. Gogolin, R. M. Par- de Jong, Hybrid quantum-classical hierarchy for mitiga-
rish, Local, expressive, quantum-number-preserving vqe tion of decoherence and determination of excited states,
ansatze for fermionic systems, New Journal of Physics Physical Review A 95, 042308 (2017).
(2021). [37] J. I. Colless, et al., Computation of molecular spectra on
[19] J. Hu, et al., Benchmarking variational quantum a quantum processor with an error-resilient algorithm,
eigensolvers for quantum chemistry, arXiv:2211.12775 Physical Review X 8, 011021 (2018).
(2022). [38] M. Motta, et al., Determining eigenstates and thermal
[20] E. R. Anschuetz, B. T. Kiani, Quantum variational al- states on a quantum computer using quantum imagi-
gorithms are swamped with traps, Nature Communica- nary time evolution, Nature Physics 16, 205 (2019).
tions 2022 13:1 13, 1 (2022). [39] R. M. Parrish, P. L. McMahon, Quantum filter diag-
[21] J. M. Arrazola, et al., Universal quantum circuits for onalization: Quantum eigendecomposition without full
quantum chemistry, Quantum 6, 742 (2022). quantum phase estimation, arXiv:1909.08925 (2019).
[22] J. F. Gonthier, et al., Measurements as a roadblock to [40] N. H. Stair, R. Huang, F. A. Evangelista, A multirefer-
near-term practical quantum advantage in chemistry: ence quantum krylov algorithm for strongly correlated
Resource analysis, Physical Review Research 4, 033154 electrons, Journal of Chemical Theory and Computation
(2022). 16, 2236 (2020).
36

[41] K. Seki, S. Yunoki, Quantum power method by a su- [61] K. Gratsea, J. S. Kottmann, P. D. Johnson, A. A. Ku-
perposition of time-evolved states, PRX Quantum 2, nitsa, Comparing classical and quantum ground state
010333 (2021). preparation heuristics, arXiv:2401.05306 (2024).
[42] K. Klymko, et al., Real-time evolution for ultracompact [62] R. Pauncz, Spin Eigenfunctions: Construction and Use
hamiltonian eigenstates on quantum hardware, PRX (Springer US (New York), 1979).
Quantum 3, 020323 (2022). [63] A. Bärtschi, S. Eidenbenz, Deterministic preparation of
[43] E. N. Epperly, L. Lin, Y. Nakatsukasa, A theory of dicke states, Lecture Notes in Computer Science (in-
quantum subspace diagonalization, SIAM Journal on cluding subseries Lecture Notes in Artificial Intelligence
Matrix Analysis and Applications 43, 1263 (2022). and Lecture Notes in Bioinformatics) 11651 LNCS,
[44] C. L. Cortes, S. K. Gray, Quantum krylov subspace al- 126 (2019).
gorithms for ground- and excited-state energy estima- [64] J. Lee, et al., Even more efficient quantum compu-
tion, Physical Review A 105, 022417 (2022). tations of chemistry through tensor hypercontraction,
[45] Y. Shen, et al., Real-time krylov theory for quantum PRX Quantum 2, 030305 (2021).
computing algorithms, Quantum 7, 1066 (2023). [65] P. Jordan, E. P. Wigner, Über das paulische
[46] N. H. Stair, C. L. Cortes, R. M. Parrish, J. Cohn, äquivalenzverbot, Zeitschrift fuer Physik (1928).
M. Motta, Stochastic quantum krylov protocol with [66] A. J. W. Thom, Stochastic coupled cluster theory, Phys-
double-factorized hamiltonians, Physical Review A 107, ical Review Letters 105, 263004 (2010).
032414 (2023). [67] J. Lee, W. J. Huggins, M. Head-Gordon, K. B. Whaley,
[47] W. Kirby, M. Motta, A. Mezzacapo, Exact and efficient Generalized unitary coupled cluster wave functions for
lanczos method on a quantum computer, Quantum 7, quantum computation, Journal of Chemical Theory and
1018 (2023). Computation 15, 311 (2019).
[48] W. Kirby, Analysis of quantum krylov algorithms with [68] The state of the frozen core orbitals was explicitly rep-
errors, arXiv:2401.01246 (2024). resented as |C⟩ in Ref. 52, but we drop it in this paper
[49] A. Y. Kitaev, Quantum measurements and the abelian for compactness.
stabilizer problem, arXiv:quant-ph/9511026 (1995). [69] D. Aliverti-Piuri, et al., What can quantum information
[50] D. S. Abrams, S. Lloyd, Quantum algorithm providing theory offer to quantum chemistry?, arXiv:2403.08045
exponential speed increase for finding eigenvalues and (2024).
eigenvectors, Physical Review Letters 83, 5162 (1999). [70] B. Bauer, S. Bravyi, M. Motta, G. K. L. Chan, Quan-
[51] M. A. Nielsen, I. L. Chuang, Quantum Computation and tum algorithms for quantum chemistry and quantum
Quantum Information (Cambridge University Press, materials science, Chemical Reviews (2020).
2010). [71] R. Izsák, A. V. Ivanov, N. S. Blunt, N. Holzmann,
[52] D. Marti-Dafcik, N. Lee, H. G. A. Burton, D. P. Tew, F. Neese, Measuring electron correlation: The impact
Spin-coupled molecular orbitals: chemical intuition of symmetry and orbital transformations, Journal of
meets quantum chemistry, arXiv:2402.08858 (2024). Chemical Theory and Computation 19, 2703 (2023).
[53] D. P. Tew, W. Klopper, T. Helgaker, Electron correla- [72] T. Helgaker, W. Klopper, D. P. Tew, Quantitative quan-
tion: The many-body problem at the heart of chemistry, tum chemistry, Molecular Physics 106, 2107 (2010).
Journal of Computational Chemistry 28, 1307 (2007). [73] S. Fomichev, et al., Initial state preparation
[54] R. J. Bartlett, M. Musia, Coupled-cluster theory in for quantum chemistry on quantum computers,
quantum chemistry, Reviews of Modern Physics 79, 291 arXiv:2310.18410 (2023).
(2007). [74] This assumes a basis of localized orbitals, and L would
[55] G. H. Booth, A. J. Thom, A. Alavi, Fermion monte carlo be significantly higher using delocalized basis functions
without fixed nodes: A game of life, death, and annihi- such as Hartree–Fock orbitals.
lation in slater determinant space, Journal of Chemical [75] V. V. Shende, S. S. Bullock, I. L. Markov, Synthe-
Physics 131, 054106 (2009). sis of quantum-logic circuits, IEEE Transactions on
[56] A. A. Holmes, N. M. Tubman, C. J. Umrigar, Heat-bath Computer-Aided Design of Integrated Circuits and Sys-
configuration interaction: An efficient selected configu- tems 25, 1000 (2006).
ration interaction algorithm inspired by heat-bath sam- [76] A. M. Childs, N. Wiebe, Hamiltonian simulation using
pling, Journal of Chemical Theory and Computation 12, linear combinations of unitary operations, Quantum In-
3674 (2016). formation and Computation 12, 901 (2012).
[57] N. M. Tubman, J. Lee, T. Y. Takeshita, M. Head- [77] R. Babbush, et al., Encoding electronic spectra in quan-
Gordon, K. B. Whaley, A deterministic alternative tum circuits with linear t complexity, Physical Review
to the full configuration interaction quantum monte X 8, 041015 (2018).
carlo method, Journal of Chemical Physics 145, 44112 [78] K. Sugisaki, et al., Quantum chemistry on quantum
(2016). computers: A polynomial-time quantum algorithm for
[58] G. H. Low, V. Kliuchnikov, L. Schaeffer, Trading t-gates constructing the wave functions of open-shell molecules,
for dirty qubits in state preparation and unitary synthe- Journal of Physical Chemistry A 120, 6459 (2016).
sis, arXiv:1812.00954 (2018). [79] K. Sugisaki, et al., Open shell electronic state calcu-
[59] G. Wang, S. Sim, P. D. Johnson, State preparation lations on quantum computers: A quantum circuit for
boosters for early fault-tolerant quantum computation, the preparation of configuration state functions based
Quantum 6, 829 (2022). on serber construction, Chemical Physics Letters: X 1,
[60] K. Gratsea, C. Sun, P. D. Johnson, When to reject a 100002 (2019).
ground state preparation algorithm, arXiv:2212.09492 [80] A. Carbone, D. E. Galli, M. Motta, B. Jones, Quantum
(2022). circuits for the preparation of spin eigenfunctions on
37

quantum computers, Symmetry 2022, Vol. 14, Page 624 nally corrected coupled cluster with quantum inputs,
14, 624 (2022). arXiv:2312.08110 (2023).
[81] K. Sugisaki, et al., Quantum chemistry on quantum [99] A. M. Childs, Y. Su, M. C. Tran, N. Wiebe, S. Zhu,
computers: A method for preparation of multiconfigu- Theory of trotter error with commutator scaling, Phys-
rational wave functions on quantum computers without ical Review X 11, 011020 (2021).
performing post-hartree-fock calculations, ACS Central [100] G. H. Low, I. L. Chuang, Hamiltonian simulation by
Science 5, 167 (2019). qubitization, Quantum 3, 163 (2019).
[82] A. Brtschi, S. Eidenbenz, Short-depth circuits for dicke [101] D. W. Berry, C. Gidney, M. Motta, J. R. McClean,
state preparation, 2022 IEEE International Confer- R. Babbush, Qubitization of arbitrary basis quantum
ence on Quantum Computing and Engineering (QCE) chemistry leveraging sparsity and low rank factoriza-
(2022). tion, Quantum 3, 208 (2019).
[83] T. Bastin, et al., Operational determination of multi- [102] H. R. Larsson, H. Zhai, C. J. Umrigar, G. K. L. Chan,
qubit entanglement classes via tuning of local opera- The chromium dimer: Closing a chapter of quantum
tions, Physical Review Letters 102, 053601 (2009). chemistry, Journal of the American Chemical Society
[84] D. B. Hume, C. W. Chou, T. Rosenband, D. J. 144, 15932 (2022).
Wineland, Preparation of dicke states in an ion chain, [103] P. J. Ollitrault, et al., Enhancing initial state over-
Physical Review A - Atomic, Molecular, and Optical lap through orbital optimization for faster molec-
Physics 80, 052302 (2009). ular electronic ground-state energy estimation,
[85] D. Bacon, I. L. Chuang, A. W. Harrow, Efficient quan- arXiv:2404.08565 (2024).
tum circuits for schur and clebsch-gordan transforms, [104] D. Cleland, G. H. Booth, A. Alavi, Communications:
Physical Review Letters 97, 170502 (2006). Survival of the fittest: Accelerating convergence in full
[86] A. Bärtschi, S. Eidenbenz, Grover mixers for qaoa: configuration-interaction quantum monte carlo, Journal
Shifting complexity from mixer design to state prepa- of Chemical Physics 132, 41103 (2010).
ration, Proceedings - IEEE International Conference on [105] M. Möttönen, J. J. Vartiainen, V. Bergholm, M. M.
Quantum Computing and Engineering, QCE 2020 pp. Salomaa, Transformation of quantum states using uni-
72–82 (2020). formly controlled rotations, Quantum Information and
[87] D. W. Berry, A. M. Childs, R. Cleve, R. Kothari, R. D. Computation 5, 467 (2004).
Somma, Simulating hamiltonian dynamics with a trun- [106] G. K. L. Chan, Low entanglement wavefunctions, Wi-
cated taylor series, Physical Review Letters 114, 090502 ley Interdisciplinary Reviews: Computational Molecular
(2015). Science 2, 907 (2012).
[88] G. A. Quantum, Hartree-fock on a superconducting [107] C. Schn, E. Solano, F. Verstraete, J. I. Cirac, M. M.
qubit quantum computer, Science 369, 1084 (2020). Wolf, Sequential generation of entangled multiqubit
[89] M. Cerezo, et al., Variational quantum algorithms, Na- states, Physical Review Letters 95, 110503 (2005).
ture Reviews Physics 2021 pp. 1–20 (2021). [108] M. B. Hastings, An area law for one-dimensional quan-
[90] S. Sharma, K. Sivalingam, F. Neese, G. K. L. Chan, tum systems, Journal of Statistical Mechanics: Theory
Low-energy spectrum of ironsulfur clusters directly from and Experiment 2007, P08024 (2007).
many-particle quantum mechanics, Nature Chemistry [109] D. J. Thouless, Stability conditions and nuclear rota-
2014 6:10 6, 927 (2014). tions in the hartree-fock theory, Nuclear Physics 21,
[91] Z. Li, S. Guo, Q. Sun, G. K. L. Chan, Electronic land- 225 (1960).
scape of the p-cluster of nitrogenase as revealed through [110] I. D. Kivlichan, et al., Quantum simulation of electronic
many-electron quantum wavefunction simulations, Na- structure with linear depth and connectivity, Physical
ture Chemistry 2019 11:11 11, 1026 (2019). Review Letters 120, 110501 (2018).
[92] Z. Li, J. Li, N. S. Dattani, C. J. Umrigar, G. K. L. [111] S. McArdle, S. Endo, A. Aspuru-Guzik, S. C. Benjamin,
Chan, The electronic complexity of the ground-state of X. Yuan, Quantum computational chemistry, Reviews of
the femo cofactor of nitrogenase as relevant to quan- Modern Physics 92, 15003 (2020).
tum simulations, The Journal of Chemical Physics 150, [112] Y. S. Yordanov, D. R. Arvidsson-Shukur, C. H. Barnes,
024302 (2019). Efficient quantum circuits for quantum computational
[93] L. Lin, Y. Tong, Near-optimal ground state preparation, chemistry, Physical Review A 102, 062612 (2020).
Quantum 4, 372 (2020). [113] C. Gidney, A. G. Fowler, Efficient magic state factories
[94] Y. Dong, L. Lin, Y. Tong, Ground-state preparation and with a catalyzed |ccz⟩ to 2|t⟩ transformation, Quantum
energy estimation on early fault-tolerant quantum com- 3, 135 (2019).
puters via quantum eigenvalue transformation of uni- [114] P. Selinger, Efficient clifford+t approximation of single-
tary matrices, PRX Quantum 3, 040305 (2022). qubit operators, Quantum Information and Computa-
[95] L. Lin, Y. Tong, Heisenberg-limited ground-state energy tion 15, 159 (2012).
estimation for early fault-tolerant quantum computers, [115] N. J. Ross, P. Selinger, Optimal ancilla-free clifford+t
PRX Quantum 3, 010318 (2022). approximation of z-rotations, Quantum Information
[96] H. G. A. Burton, Accurate and gate-efficient quantum and Computation 16, 0901 (2014).
ansätze for electronic states without adaptive optimisa- [116] A. Bocharov, M. Roetteler, K. M. Svore, Efficient syn-
tion, arxiv:2312.09761 (2023). thesis of universal repeat-until-success quantum cir-
[97] V. von Burg, et al., Quantum computing enhanced cuits, Physical Review Letters 114, 080502 (2015).
computational catalysis, Physical Review Research 3, [117] Y. R. Sanders, et al., Compilation of fault-tolerant
033055 (2021). quantum heuristics for combinatorial optimization,
[98] M. Scheurer, G.-L. R. Anselmetti, O. Oumarou, PRX Quantum 1, 020312 (2020).
C. Gogolin, N. C. Rubin, Tailored and exter-
38

[118] G. H. Low, Halving the cost of quantum multiplexed states into shallow quantum circuits, Quantum Science
rotations, arXiv:2110.13439 (2021). and Technology 9, 015012 (2023).
[119] C. Gidney, Halving the cost of quantum addition, Quan- [124] S. J. Ran, Encoding of matrix product states into quan-
tum 2, 74 (2018). tum circuits of one- a nd two-qubit gates, Physical Re-
[120] D. Malz, G. Styliaris, Z.-Y. Wei, J. I. Cirac, Prepara- view A 101, 032310 (2020).
tion of matrix product states with log-depth quantum [125] Q. Sun, et al., Pyscf: the pythonbased simulations of
circuits, Physical Review Letters 132, 040404 (2024). chemistry framework, WIREs Computational Molecular
[121] S. Bravyi, M. B. Hastings, F. Verstraete, Lieb-robinson Science 8 (2018).
bounds and the generation of correlations and topologi- [126] Q. Sun, et al., Recent developments in the pyscf pro-
cal quantum order, Physical Review Letters 97, 050401 gram package, The Journal of Chemical Physics 153,
(2006). 024109 (2020).
[122] M. B. Dov, D. Shnaiderov, A. Makmal, E. G. D. Torre, [127] J. R. McClean, et al., Openfermion: The electronic
Approximate encoding of quantum states using shallow structure package for quantum computers, Quantum
circuits, arXiv:2207.00028 (2022). Science and Technology pp. 1–22 (2020).
[123] M. S. Rudolph, J. Chen, J. Miller, A. Acharya, [128] Gmin: A program for finding global minima and
A. Perdomo-Ortiz, Decomposition of matrix product calculating thermodynamic properties, http://www-
wales.ch.cam.ac.uk/software.html.

You might also like