State preparation and evolution in quantum computing: a perspective from Hamiltonian moments
Abstract
Quantum algorithms on noisy intermediate-scale quantum (NISQ) devices are expected to soon have the ability to simulate quantum systems that are classically intractable, demonstrating a quantum advantage. However, the non-negligible gate error present on NISQ devices impedes the implementation of conventional quantum algorithms. Practical strategies usually exploit hybrid quantum-classical algorithms to demonstrate potentially useful applications of quantum computing in the NISQ era. Among the numerous hybrid quantum-classical algorithms, recent efforts highlight the development of quantum algorithms based upon quantum computed Hamiltonian moments, (), with respect to quantum state . In this tutorial review, we will give a brief review of these quantum algorithms with focuses on the typical ways of computing Hamiltonian moments using quantum hardware and improving the accuracy of the estimated state energies based on the quantum computed moments. Furthermore, we will present an example to show how we can measure and compute the Hamiltonian moments of a four-site Heisenberg model, and compute the energy and magnetization of the model utilizing the imaginary time evolution on the real IBM-Q NISQ hardware. Along this line, we will further discuss some practical issues associated with these algorithms. We will conclude this tutorial review by discussing some possible developments and applications in this direction in the near future.
I Introduction
The original idea of solving many-body quantum mechanical problems on a quantum computer dates back to R. Feynman’s insight almost four decades ago Feynman (1982), when high-performance classical computing was at its early state and quantum computing was only a thought experiment. In the last forty years, developments in classical quantum approaches and high performance computing technology yielded adequate tools to perform highly accurate small- to medium-size quantum simulations and qualitative large-scale quantum simulations. Despite these exciting developments and achievements, the classical quantum approach is gradually approaching the computing limit attributed to the unfavorable scaling along with the expanded system size. For instance, consider the well-known examples of full Configuration Interaction (full CI) approach David Sherrill and Schaefer (1999) and Coupled Cluster (CC) approach Bartlett and Musiał (2007). Classically, in order to achieve the chemical accuracy (1 kcal/mol) for the energy evalulation, the runtime in the full CI and the “golden standard” CCSD(T) scale as and , respectively, with respect to the system size (and number of electrons ), which impedes their routine applications for large systems. On the other hand, quantum computing (simulation performed on a quantum computer) is a next-generation computing technology that bears the hope of outperforming high-performance classical computing techniques to solve a wide class of problems encountered in scientific research and industrial applications. Quantum computation enables the encoding of an exponential amount of information about many-body quantum systems into a polynomial number of qubits, which provides a more efficient way to expand and explore the computational state-space in polynomial time Sugisaki et al. (2019).
Since the first application of quantum computing algorithm for computational chemistry problems in 2005 Aspuru-Guzik et al. (2005), numerous quantum algorithms have been proposed aiming at demonstrating the potential quantum speedup, in comparison to the classical approaches, for solving problems in domain science such as chemistry, nuclear physics, quantum field theory, and high energy physics O’Malley et al. (2016); Linke et al. (2017); Kandala et al. (2017); Dumitrescu et al. (2018); Klco et al. (2018); Colless et al. (2018); McCaskey et al. (2019). However, many proposed quantum algorithms featuring favorable scaling typically have deep circuit demands, and are not suitable for the near-term noisy intermediate-scale quantum (NISQ) devices Preskill (2018). In the search for quantum algorithms suitable running on the NISQ devices, one should mention hybrid quantum-classical Variational Quantum algorithm (VQA) Peruzzo et al. (2014); McClean et al. (2016); Romero et al. (2018); Shen et al. (2017); Kandala et al. (2017, 2019); Colless et al. (2018); Huggins et al. (2020), quantum approximate optimization algorithm (QAOA) Farhi, Goldstone, and Gutmann (2014), quantum annealing Bharti et al. (2021); Albash and Lidar (2018), gaussian boson sampling Aaronson and Arkhipov (2011), analog quantum simulation Trabesinger (2012); Georgescu, Ashhab, and Nori (2014), iterative quantum assisted eigensolver McArdle et al. (2019); Motta et al. (2020); Parrish and McMahon (2019); Kyriienko (2020), and many others, which bear the hope to capitalize the near-term NISQ quantum devices. Nevertheless, these NISQ quantum algorithms have their own limitations. For example, VQAs face challenges such as Barren plateau McClean et al. (2018); Cerezo et al. (2021); Cerezo and Coles (2021), ansatz searching Herasymenko and O’Brien (2019), and measurement overhead Bonet-Monroig, Babbush, and O’Brien (2020) which stimulate the emerging of some many recent improvements. Numerous comprehensive reviews on the NISQ quantum algorithms and techniques have appeared in recent years to pave the foundation of and provide guidance to future studies (see e.g. Refs. 35; 21 for more recent ones). However, due to the rapid development of this emerging field, it becomes challenging for the reviews to comprehensively cover many newly developed quantum algorithms and techniques in a timely manner. Furthermore, it is usually more efficient, in particular for non-specialists, to choose a thread to understand and digest the basic ideas and the interconnection of these algorithms by going through some tutorials. With these concerns in mind, instead of giving another comprehensive review covering some newly developed quantum algorithms, here we present a tutorial review on some recently developed quantum algorithms based on the quantum computation of Hamiltonian moments. The Hamiltonian moment here is referring to the expectation value of Hamiltonian powers with respect to a given state for the systems of interest (), and can be considered as one of the building blocks for performing variational or perturbative calculations solving the energy and property of a many-body quantum system. We will introduce the state preparation and evolution, and hybrid algorithms associated with quantum Hamiltonian moments. With the examples and discussions, we hope to provide a relatively easy-to-read review with a clear demonstration of the connections and differences between some relevant quantum algorithms.
This tutorial review is organized as follows. Section II reviews the ways of quantum computing Hamiltonian moments. Section III reviews some recently proposed hybrid approaches based on quantum Hamiltonian moments (see Fig. 1 for a typical workflow of the hybrid algorithm of this kind). Section IV gives a tutorial of how to target the ground state and magnetization of a generalized four-site Heisenberg model employing quantum Hamilonian moments in the imaginary time evolution approach on IBM-Q quantum hardware. Section IV also includes the discussions of some practical issues associated with performing quantum computing on the NISQ device with a focus on the impact of these issues on the results of quantum simulations. We will conclude this tutorial review by pointing out possible developments and applications in this direction in the near future.

II Quantum computation of Hamiltonian moments
Given a system Hamiltonian of interest and a trial state , the -th order () Hamiltonian moment with respect to is defined as
(1) |
In classical computations, ’s are obtained by power iteration, i.e. a repeated multiplication of and . However, as the dimension of the Hilbert space grows exponentially with the system size, and the number of terms that constitutes would quickly blow up as the power becomes large, the classical computations of ’s would quickly become intractable. On the other hand, quantum computing allows for a more straightforward encoding of the quantum states defined in a Hilbert space of potentially large and classically intractable dimensions. Therefore, we consider (a) how to utilize quantum resources for the direct computation of , and (b) how to exploit the noisy ’s computed on NISQ devices to accurately evaluate ground/excited states and dynamics of quantum many-body systems. In this section, we try to answer the first question by briefly reviewing some of the methods that have been proposed for directly computing on quantum devices. We leave the discussion of the second question to Section III.
To directly compute on quantum devices for a quantum many-body system, a first step is to encode the many-body Hamiltonian in terms of qubits. There have been many ways proposed to encode the system Hamiltonian. For example, in the Jordan-Wigner Jordan and Wigner (1928) and Bravyi-Kitaev methods Bravyi and Kitaev (2002); Seeley, Richard, and Love (2012); Tranter et al. (2015), the -qubit system Hamiltonian is encoded as a linear combination of tensor products of Pauli qubit operators,
(2) |
where is a scalar and is a Pauli string with being either identity matrix or one of the three Pauli matrices
(9) |
for the -th qubit. From (2), a naive way to directly compute is to plug (2) into (1), and evaluate the expectation values of the products of Pauli strings that constitute . Apparently, without any simplification or reduction, the naive way would require evaluating terms and would quickly become intractable as and/or scale up. Therefore, major efforts in this direction focus on mitigating this evaluation overhead as much as possible by introducing some simplifications and reductions based on the properties of the Pauli strings.
II.1 Term-by-term measurement
One straightforward way is to apply Pauli reduction and commutativity to reduce the number of terms, and then do term-by-term measurements. The Pauli reduction simply utilizes the commutation and anti-commutation relations between Pauli matrices
(12) |
where are Pauli matrices, is the structure constant following the Levi-Civita symbol Tyldesley (1975), and is the Kronecker delta. From (12) it is straightforward to see the product of Pauli strings is another Pauli string with an appropriate phase factor. Therefore, in comparison to the evaluation of , the evaluation of does not increase the circuit depth. Furthermore, we notice that the measurement of a single Pauli string could be used to determine classes of contributions to Hamiltonian moments of arbitrary order. For example, since every Pauli string is unitary, the evaluation of can simultaneously contribute all the ’s () Kowalski and Peng (2020). Suchsland et al. systematically analyzed the actual number of Pauli strings corresponding to () for a collection of 11 molecules, including Hn (), LiH, NaH, H2O, NH3, and N2, that require up to qubits Suchsland et al. (2021). It was found that the actual numbers of Pauli strings in these systems after applying (12) exhibit at most , , and scalings in comparison to the predicted , , and scalings for () respectively.

On the top of the above simplification, a more significant term reduction can be achieved by utilizing the commutativity of the Pauli strings. To see this method works mathematically, consider operators and that commute, i.e. , then there exists a complete orthonormal eigenbasis () that simultaneously diagonalizes both operators, and we can write
(13) |
with and being the eigenvalues of and , respectively. Now the expectation values of and with respect to a trial state can be expressed as
(14) |
Therefore, if we measure each ’s, we can deduce and based on (14).
The simplest commutativity among Pauli strings is qubitwise commutativity (QWC) Kandala et al. (2017); McClean et al. (2016), where two Pauli strings qubitwise commute if two single-qubit Pauli matrices at each index commute, i.e.
(15) |
For example, is a QWC class, since any pair of Pauli strings in this class commute (a more pictorial QWC grouping procedure is shown in Fig. 2). For the observables grouped in one QWC class, one can easily find a shared eigenbasis that simultaneously diagonalizes all the observables. Take the previous example, if we want to measure the QWC class in the computational basis, utilizing the Clifford representation of single Pauli matrices and
(16) |
we will only need perform the following measurement
![[Uncaptioned image]](https://cdn.awesomepapers.org/papers/7dcbae80-b802-4c9e-895a-475c743e1025/measurement1.png)
which gives the probabilities of the variational circuit rotated into the shared eigenbasis, and the expectation values of each Pauli string can then be recovered by classically combining these probabilities with the eigenvalues of these Pauli strings.
Now, given a large amount of Pauli strings that constitute , how should we find an optimal grouping of these Pauli strings such that the total number of measurements is minimized? Unfortunately, it turns out that finding the optimal grouping is NP-hard Karp (2010); Gokhale et al. (2020), and can be converted to minimum clique cover or minimum graph coloring problem, which in practice can be approximately solved through heuristic approaches that scale quadratically with the number of the Pauli strings Yen, Verteletskyi, and Izmaylov (2020). In terms of term reduction, for small molecules the QWC grouping can bring roughly 75% saving for evaluating , and even bigger savings for for (). Early studies of applying QWC bases to evaluating for the Heisenberg models represented by 10 to 40 qubits exhibits a sub-linear scaling of the number of measurements in the number of the qubits Vallury et al. (2020). Furthermore, for some small molecular systems, it has been shown that the number of QWC bases grouping the Pauli strings that constitute eventually reach a plateau, regardless of the power Claudino et al. (2021).
Essentially, the QWC grouping has its roots in mutually unbiased bases (MUB) Schwinger (1960); Klappenecker and Rötteler (2004) from quantum information theory associated with maximizing the information learned from a single measurement. In general, MUBs group the potentially -qubit Pauli strings (excluding identity string) into commuting groups of maximal size, of which using QWCs would allow reducing this requirement to Altepeter, James, and Kwiat (2004). Further, it is known that there in principle exists an MUB grouping for -qubit that can further reduce this requirement to even with maximum distinct Pauli strings in each commuting group, however, the entanglement in the MUBs Lawrence, Brukner, and Zeilinger (2002) makes MUB quantum state tomography challenging. In practice, beside QWC, other grouping techniques have alternatively been proposed, including general commutativity Gokhale et al. (2020); Yen, Verteletskyi, and Izmaylov (2020), unitary partitioning Izmaylov et al. (2020), and Fermionic basis rotation grouping Huggins et al. (2021). Generally speaking, the applications of these grouping rules for evaluating have shown that, at a cost of introducing additional one-/multi-qubit unitary transformation before the measurement, the total number of terms can be significantly reduced from to . In particular, for simpler cases where the entire Hamiltonian or its power could even be transformed by a single unitary, the evaluation of the corresponding expectation values can be done in a single set of measurements.
It is worth mentioning that the above discussion is limited to the number of terms that can be efficiently reduced through the groupings that feature the commutativity of Pauli strings, while the total number of measurements required from the number of groups also critically depends on the covariance between Pauli strings, , and the desired precision . Given that we can always write the Hamiltonian powers as linear combination of Pauli strings similar to Eq. (2), the total number of measurements for evaluating the moments can then be expressed as Gonthier et al. (2020); Rubin, Babbush, and J. (2018); Wecker, Hastings, and Troyer (2015)
(17) |
with indexing the groups. Therefore, it is likely that the number of groups decreases at the cost of introducing larger covariances that could essentially increase the total number of measurements required to achieve a desired precision.
There are also many other advanced measurement schemes proposed recently. One example is to simultaneously obtain expectation values of multiple observables by randomly measuring and projecting the quantum state into classical shadows Huang, Kueng, and Preskill (2020, 2021); Aaronson (2018); Struchalin et al. (2021); Chen et al. (2021); Zhao, Rubin, and Miyake (2021); Acharya, Saha, and Sengupta (2021); Hadfield (2021); Hillmich et al. (2021); Zhang et al. (2021). The algorithm in principle enables the measurements of low-weight observables using only samples. The practical performance of the algorithm for model and molecular Hamiltonians on NISQ device, in terms of accuracy and efficiency, is still under intense study.
II.2 Expansion by Chebyshev polynomials via quantum walk
Another way to exactly implement is by Chebyshev polynomial expansion. We can exactly express a monomial on as a finite sum of Chebyshev polynomials of the first kind, i.e.
(18) |
where
(19) |
and is the degree Chebyshev polynomial of the first kind Subramanian, Brierley, and Jozsa (2019). This expansion is useful because a “quantum walk” can exactly produce the effect of Chebyshev polynomials in , where is the sparsity of the matrix. A quantum walk is a process in which the Hilbert space is enlarged, and then a walk operator that works in the enlarged Hilbert space is repeatedly implemented. For details on quantum walks and their implementations, see e.g. Refs. 68; 69. With amplitude amplification to obtain a constant success probability of preparing a quantum state and a flag qubit to indicate if the preparation was successful, the query complexity of this method is Subramanian, Brierley, and Jozsa (2019). Using this method without amplitude amplification, the query complexity of this implementation is . Here, denotes the smallest (magnitude) eigenvalue of , with the assumption that does not have 0 as an eigenvalue. However, this method has its own drawbacks. The implementation of the quantum walk can be expensive, requiring a number of ancillary qubits that grows linearly with the system size Subramanian, Brierley, and Jozsa (2019). The Chebyshev polynomial expansion’s potential application for computing the imaginary time evolution of an unnormalized quantum state, and in general for evaluating partition function has been discussed in Ref. 70, where they use the recursion relation for Chebyshev polynomials.
II.3 Linear combination of unitary time propagator
Recently, Seki and Yunoki proposed a quantum power method to evaluate Seki and Yunoki (2021), where the Hamiltonian power is first expressed as the -th time derivative of at , then the time derivative is approximated through the central finite difference (CFD) which can be alternatively expressed as a linear combination of ’s at different time variables ’s in close proximity to , i.e.
(20) |
To implement the approach to quantum circuit, each is further decomposed using symmetric Suzuki-Trotter decomposition (SSTD) Trotter (1959); Suzuki (1990). There are two sources of error in this approach, the CFD error and SSTD error. The systematic CFD error scales quadratically with , i.e. , while the -th order SSTD brings the systematic error of . However, there are several advantages in the proposed approach. First, both aforementioned errors can be systematically suppressed through Richardson extrapolation Richardson and Gaunt (1927); Temme, Bravyi, and Gambetta (2017) at the cost of involving more terms in the linear combination in Eq. (20), which on the other hand offers the space for using low order SSTD. Second, the number of gates required for approximate for an -qubit Hamiltonian scales as . In principle, the quantum power method can be combined with many classical approaches, which will be elaborated in Section III for targeting the ground/excited states and properties of a quantum many-body system. As shown in Ref. 71, numerical noiseless simulations employing the quantum power method in Krylov-subspace diagonalization for targeting ground state of model systems demonstrate systematically improved accuracy over the conventional VQE.
Inspired by the classical inverse power iteration approach for finding the dominant eigenstate of a given hermitian matrix with a more favorable logarithmic complexity in the iteration depth Sachdeva and Vishnoi (2014), Kyriienko recently proposed a quantum inverse iteration algorithm Kyriienko (2020) for approximately computing the inverse of -th power of the Hamiltonian, . The approach extends the discretized Fourier approximation of the Hamiltonian inverse Childs, Kothari, and Somma (2017) to the -th power,
(21) |
where is a normalization factor, and with the integers and and intervals defining a discretization grid space, and unitary is given by . Then is evaluated through the SWAP test Ekert et al. (2002); Higgott, Wang, and Brierley (2019), overlap measurement Mitarai and Fujii (2019), or Bell-type-like measurement Kyriienko (2020). However, this method requires a number of calls to a time-evolution oracle that is dependent on the condition number of the system, as well as a requirement that the Hamiltonian has been shifted such that all of the eigenvalues are positive. This latter requirement requires the user to have some knowledge of the ground state energy of the system a priori. A suitable guess can be found, for example, using the minimum label finding algorithm of Ref. 81.
Alternatively, if considering a fault-tolerant implementation with favorable resource scaling or can be done through amplitude amplification approach Brassard et al. (2002), Hamiltonian simulation Berry et al. (2015), qubitization Low and Chuang (2017), or the direct block-encoding Gilyén et al. (2019) methods at the cost of introducing deeper circuit and implementing controlled- operations which, however, are still challenging to be implemented in the NISQ devices. Some of these methods have deep connection to linear combination of unitaries (LCU) technique Childs and Wiebe (2012). The basic idea of LCU can be briefly demonstrated from the following toy example. Suppose we want to implement on . We can start by introducing an ancilla qubit and preparing with . Then by performing the controlled unitary on we are able to obtain the state . Now, if we measure the ancilla qubit in the basis and obtain the outcome associated with , then we obtain a state proportional to .
III Classical computation of target states from quantum computed moments
Now supposing we already have necessary Hamiltonian moments computed from quantum simulation, how can we proceed to accurately evaluate the energy and/or properties of the Hamiltonian? In this section, we will try to give our answers from some classical approaches.
III.1 Lanczos approach
One of the most common approaches that can be exploited is the Lanczos approach Lanczos (1950). In the Lanczos approach, an orthonormal set of quantum states is generated through a three-term recursion
(22) |
where , . The three-term recursion essentially transforms the Hamiltonian into a tridiagonal form
(27) |
where the matrix elements and can essentially be represented recursively in terms of Hamiltonian moments. Explicitly, it can be shown
(28) | ||||
Now the strict upper bound of the ground state energy of the Hamiltonian can be obtained by directly diagonalizing . It has been shown that the classical Lanczos scheme can directly be applied even employing the noisy Suchsland et al. (2021), which effectively provides a way to improve the energy estimate from real quantum hardware without being subject to specific hardware and an explicit description of the underlying noise. Preliminary tests on H2, H3, and four-site tetrahedral Heisenberg model demonstrate that combining the Lanczos scheme with VQE algorithm helps correct the noise and enhance the quality of the ansatz.
Alternatively, the matrix elements and can be expressed in terms of the so-called connected moments (or cumulant) Horn and Weinstein (1984). An -th order connected moments is defined as
(31) |
from which (28) can be generalized as -expansions of and Hollenberg (1993); Witte and Hollenberg (1994); Hollenberg and Witte (1996)
(32) |
where with the recursion index and the volume of the system. From the bound analysis of orthogonal polynomials van Doorn (1987); Ismail and Li (1992), it has been shown that the true ground state energy of the Hamiltonian can be approximated through the greatest lower bound (i.e. infimum) of the expansion (32) Hollenberg and Witte (1996),
(33) |
For example, the first order in gives Hollenberg and Witte (1994)
(34) |
Approach (34) has been employed to demonstrate the energy estimate of 2D quantum magnetism model on lattices up to 25 qubits on IBM-Q quantum hardware Vallury et al. (2020), where the results show a consistent improvement of the estimated energies in comparison to the VQE results for the same ansatz. Another application of the Lanczos method is the continued fractions expansion of the Green’s function. The Green’s function can be calculated in the Lehmann representation using the and parameters from the Lanczos method. For full details, see e.g. Chapter 8 of Ref. 95.
It is worth mentioning that standard Lanczos is a variational approach while the infimum approach (33) does not offer a strict upper bound of the true ground state energy, which, as exemplified by (34), is essentially an alternative size-extensive connected moment expansion of the ground state energy. Other polynomial expansion approaches, in particular connected moment expansions, will be discussed in details in Section III.3.
III.2 Real time evolution
Alternatively, one can resort to time evolution approaches to target the energy of the Hamiltonian. Here the time evolution approaches include both real time evolution (RTE) approach and its imaginary analogue. In RTE, the dynamics of a quantum state is governed by the time-dependent Schrödinger equation
(35) |
In the Schrödinger picture the solution of Eq. (35), , can be expressed by acting a time-evolution operator (or its truncated Taylor expansion in practice) on an initial state ,
(36) |
The introduction of in (36) effectively builds a non-orthogonal Krylov subspace. Now consider an order-() Krylov subspace , the ground state can then be approximated as a linear combination of the Krylov basis
(37) |
under the constraint with ’s being the expansion coefficients to be determined. Plugging the expanded form (37) to Eq. (35) and projecting the approximate evolution on to the subspace , we end up with a set of time-dependent coupled equations in terms of the first Hamiltonian moments, which in matrix form can be expressed as
(38) |
where the vector and is its time-derivative. Hamiltonian moments matrices and are defined as follows,
(43) | ||||
(48) |
Eq. (38) is a first order, linear ordinary differential equation (ODE), whose solution depends on the eigenvalues of the matrix that essentially approximate some of the eigenvalues of . In a recent study Guzman and Lacroix (2021), this approach is demonstrate to be able to to target both the ground and excited states for pairing Hamiltonian. Alternatively, as shown in Ref. 71, employing Rayleigh-Ritz technique, and constitute general eigenvalue problem
(49) |
with the eigenpair () approximate the energy and state vector of the true ground state. To solve Eq. (49) for (), employ canonical orthogonalization to do the following transformation
(50) |
where . Since is a hermitian that can be diagonalized by a unitary matrix with eigenvalues constituting a diagonal matrix , i.e. , can be constructed through . Similar subspace diagonalization schemes have been adopted in many other hybrid quantum-classical algorithms such as quantum subspace expansion McClean et al. (2017); Colless et al. (2018), and some variants of VQE Parrish et al. (2019); Nakanishi, Mitarai, and Fujii (2019); Huggins et al. (2020).
III.3 Imaginary time evolution
The imaginary time evolution (ITE) approach has a long history of being a robust computational approach to solve the ground state of a many-body quantum system. To see how it works, let’s first assume the quantum state at time , , is now expanded in the eigen-space of , , with ’s being the corresponding eigenvalues, then as we show in the previous section can be expressed as
(51) |
with ’s being the expansion coefficients at the initial time. Now, if we consider a replacement ( is often called imaginary time) and confine , Eq. (51) becomes
(52) |
Comparing Eqs. (51) and (52), we see that the wavefunction is driven from an “oscillating” superposition of the Hamiltonian eigenstates to an “exponential decaying” superposition of the eigenstates with the decay rate proportional to . More importantly, in the limit of large
(53) |
the ground state is “screened out” because its exponential decay rate is the smallest in the eigen spectrum of . Briefly speaking, given a trial wavefunction that has non-zero overlap with the true ground state wavefunction , , the ITE approach, in comparison to the RTE of the trial state, guarantees a monotonically decreasing energy functional in that converges to the ground state energy at , i.e.
(54) |
Here the ITE of the trial wavefunction, , is defined as
(55) |
and the first derivative of with respect to is non-positive
(56) |
where the equality sign holds if and only if is an exact eigenstate of (e.g. ).
The development and application of ITE approach targeting the ground state wave function and energy dates back to 1970s, when the similar random-walk imaginary-time technique were developed for diffusion Monte-Carlo methods Davies et al. (1980); Anderson (1975, 1979, 1980).
Later on, in the pursuit of non-perturbative analytical tool for a wide variety of Hamiltonian systems that can be systematically improved, various polynomial expansions of Eq. (54) have been studied extensively. For example, a straightforward way is to utilize a finite number of Hamiltonian moments, , to constitute a truncated Taylor expansion of which we will have a detailed examination in Section IV.
t-expansion Beyond the conventional Taylor expansion, to reproduce asymptotic behavior of over a longer imaginary time, Horn and Weinstein introduced a power series expansion in imaginary time for in the early 1980s Horn and Weinstein (1984),
(57) |
where is the connected moments defined in Eq. (31).
The behavior of was then reconstructed through Padé approximants to Eq. (57). Furthermore, it was found that it might be more efficient and error-resilient to first construct ()-Padé approximants () to rather than to , and then integrate the Padé approximant from 0 to the critical -value (where Eq. (56) becomes positive) to obtain a larger set of approximations to . Early applications of the Padé approximants of Eq. (57) on Heisenberg and Ising models demonstrated remarkable improvement upon mean-field results.
Connected moment expansion Based on the Horn-Weinstein theorem, Cioslowski further derived a more practical re-summation technique, the so-called connected moment expansion (CMX), to Eq. (57) to address the algebraic -independent form of Cioslowski (1987),
(58) |
where the ’s are defined from the recursion
It is worth mentioning that further analysis of the CMX recursion had suggested that Eq. (58) might be recast in a compact matrix form Knowles (1987); Stubbins (1988)
(59) |
where vector is the solution of the following linear system
(66) |
The analytical properties of CMX and its comparison with other methods (e.g. Lanczos approach) have been extensively discussed in the literature Knowles (1987); Prie et al. (1994); Mancini, Zhou, and Meier (1994); Ullah (1995); Mancini et al. (1995); Fessatidis et al. (2006, 2010). In particular, similar to perturbational theories, CMX is conceptually simple and size-extensive, and the accuracy can be easily tuned through the rank of the connected moments included in the approximation and/or the quality of the trial wave function. For the latter, there have been discussions on using multi-configurational Cioslowski (1987) or even correlated wavefunction (e.g. truncated CI or CC wave functions Noga, Szabados, and Surján (2002)) in the CMX framework for accelerating the CMX convergence rate. Nevertheless, there were two major problems associated with the CMX calculations. First, the algebraic structure of the series expansion in the CMX could cause singularity Mancini, Zhou, and Meier (1994). Early efforts tried to address this issue with limited success by employing alternative moments expansion Mancini, Zhou, and Meier (1994); Ullah (1995); Mancini et al. (1995), generalized moments expansion Fessatidis et al. (2006), or generalized Padé expansion Knowles (1987). Second, CMX results are not variational. In comparison with the variational Lanczos methods Mancini, Prie, and Massano (1991); Prie et al. (1994), it was found that CMX might be considered as a limiting case of the strict Lanczos scheme, and their exact equivalence only holds in certain regions of parameter space. A systematic analysis of low order Lanczos and CMX ground state energy for model Hamiltonians such as harmonic oscilaltor, anharmonic oscillator, and Kondo model concluded that the accuracy of both approaches was dependent on the region of parameter space being studied Prie et al. (1994).
Peeters-Devreese-Soldatov approach Regarding the variational expansion, there had been another interesting yet less known moment approach proposed by Peeters and Devreese Peeters and Devreese (1984), and further analyzed by Soldatov Soldatov (1995) in the 1990s, which we will refer to as the PDS approach. This approach originates in generalizing the Bogolubov inequality Bogolubov (1947) and the Feynman inequality Feynman (1955) through the analysis of their Laplace transforms to obtain the upper bounds for the free energy. In particular, in the operator formalism, given a system Hamiltonian operator with ground state , and a complex scalar with , we can write the following inequality Peeters and Devreese (1984); Devreese, Evrard, and Kartheuser (1975)
(67) |
with and being the Laplace transform and inverse Laplace transform operators, and the expectation value with respect to a given trial wavefunction . The PDS formalism is based on expanding and analyzing using a simple identity (with introducing a parameter , and )
(68) |
Remarkably, the expectation value of the third term on the right hand side of identity (68), defined as a residual term , is non-negative,
(69) |
By analyzing its first and second derivatives with respect to the induced parameter , it is easy to show that when , reaches its local minima.
Now if we recursively expand in (68) one more time, and introducing another parameter parameter , and , we have
(70) |
with a refined residual term also being non-negative for
(71) |
Examining the first and second derivatives of with respect to , shows when
(74) |
reaches its local minima.
Note that there is an important feature about (), a little transform of (74) shows
(75) |
from which it is straightforward to show
(76) |
therefore all the terms in (70) are non-negative, and
(77) |
(76) and (77) tell us that by keeping recursively refining using identity (68), we are able to get tighter upper bounds to . This then constitutes the basic idea of PDS approach. For example, if we recursively expand (68) times () with each time introducing a new parameter , ), we will get a more refined non-negative residual term
(78) |
and
(79) |
The -th order PDS formalism, PDS(), is then associated with determining the introduced real parameters that minimize the value of through
(80) |
which can be alternatively translated to
(81) |
with each () being an strict upper bound to , and can be expressed as
(82) |
To numerically solve ’s in a PDS() approach, take Eq. (74) as an exmaple, we can first express Eq. (74) in a matrix form
(83) |
with
(88) | ||||
(94) |
Then, if we assume , then and exists, therefore we are able to act on both sides of (83) to get
(95) |
with
(100) | ||||
(103) |
It can be easily seen that given (), (or equivalently ) can be solved from (95), and () are then the solution of the polynomial
(104) |
Now generalizing (83)-(104) we get the working equation for the PDS() approach, where one need to first solve (95) for an auxiliary vector with matrix elements defined as
(107) |
and then solve the following polynomial
(108) |
for that provide upper bounds to the exact ground and excited state energies of the Hamiltonian characterized by either discrete or continuous spectral resolutions, or both.
III.4 Variational simulation of real and imaginary time dynamics
In the RTE and ITE, the target quantum state is approximated through the evolutions of the trial state as shown in Eqs. (36) and (55). In practice, the trial state can usually be prepared by applying a sequence of parametrized gates to the initial state . For example, we can write
(109) |
where with and the -th one-/two-qubit gate parametrized by rotation . In the conventional VQE, based on the measurement outcome of the trial state, classical optimization routines are then employed to find optimal that minimizes the cost function defined as the measurement outcome of the quantum state ,
(110) |
which gives a lowest upper bound to the true ground state state energy of the many-body Hamiltonian in the parameter space. Similar variational procedure can also be applied following the RTE and ITE approaches by replacing the in the VQE with the time-evolved state approximated as a linear combination of Krylov bases as shown in Eqs. (36) and (55). There are at least two advantages of combining variational optimization with the time evolution of the quantum state. First of all, since the variational optimization also contributes to bring down the cost function, the combination would help reduce the number of Krylov bases employed to evolve the trial state, which in turn helps reduce the measurement requirement. Second, the Krylov bases employed to improve the fidelity of the trial state helps restructure the potential energy surface in the same parameter space, which in turn offers an effective approach to navigate the dynamics to be free from getting trapped in the local minima of the old potential energy surface.
These advantages have been demonstrated by Peng and Kowalski in a recent study Peng and Kowalski (2021), where Eq. (104) is effectively treated as an alternative energy functional for the variational quantum solver whose analytical energy derivative can be exploted to drive the optimization
(111) |
Here is the learning rate for updating at step , and is the Riemannian metric matrix at that is flexible to characterize the singular point of the parameter space and is essentially related to the indistinguishability of Yamamoto (2019). Note that similar formulation is quite often used in, for example, gradient descent approach Lemaéchal (2012) with the metric being often replaced by the Hessian matrix. Different from the conventional gradient descent approach, Eq. (111) has its origin in the general nonlinear optimization framework featuring natural gradient for targeting machine learning problems Amari (1998). Here, the natural gradient accounts for the geometric structure of the parameter space, and is expressed as the product of the Fisher information matrix and the gradient of the cost function often with the attempt of circumventing the plateaus in the parameter space McArdle et al. (2019); Yamamoto (2019); Stokes et al. (2020). Analogously, in the quantum computing, the Fisher information matrix is replaced by the quantum Fubini-Study metric to describe the curvature of the ansatz class.
There have been discussions about deriving and applying quantum Fubini-Study metric in the quantum simulations, and comparing its performance with classical Fisher metric in some recent reports McArdle et al. (2019); Yamamoto (2019); Stokes et al. (2020). For example, following Ref. 26 one can show that the RTE or ITE approach can be translated to a variational approach for a given ansatz , for targeting the target state . Take ITE approach as an example, following McLachlan’s variational principle McLachlan (1964)
(112) |
where
(113) |
and
(114) |
with , we have
(115) |
Note that since is normalized, , we also have
(116) |
and Eq. (115) can be simplified as a linear system in matrix form
(117) |
which is essentially the dynamics (111) with the matrix elements given by
(118) |
Same dynamics and Riemannian metric might be derived from other variational principles. As shown in Ref. 126, classically there are three variational principles, namely the Dirac and Frenkel variational principle Dirac (1930); Frenkel (1934), the McLachlan variational principle McLachlan (1964), and the time-dependent variational principle Kramer and Saraceno (1981); Broeckhove et al. (1988), that lead to the same evolution equation. However, when parameters are confined to be real, the Dirac and Frenkel and the McLachlan variational principles are equivalent, while the time-dependent variational principle cannot lead to a nontrivial evolution of the parameters.
It is also worth mentioning that a caveat in the above derivation is that we implicitly assume . If there is time-dependent global phase difference between and , their time-derivatives can be very different and the dynamics (117) would be incorrect McArdle et al. (2019). In essence, the defined in Eq. (118) is not gauge invariant thus not qualified for measuring the quantum distance Cheng (2013). The problem can be fixed by either explicitly introducing a time-dependent phase gate to the trial state, or defining a gauge invariant quantum geometric tensor
(119) |
which is essentially associated with quantum natural gradient Stokes et al. (2020).
IV Quantum simulation of a model Hamiltonian
In the preceding sections we have given a brief review of hybrid quantum-classical algorithms featuring quantum computed Hamiltonian moments and their classical post-processing approaches that are in principle capable of providing systematically improvable estimates for the target energy and quantum state. In this section, we will go over a simple tutor to demonstrate how one can use the algorithms of these kinds in practice for solving some practical problems.
In this tutorial, we will study a model Hamiltonian
(120) |
which, as suggested in Ref. 132, when combining with different choices of scalars , , and , abstracts a large range of physical topics such as the integrability Essler and Fagotti (2016), quantum magnetism Vasiliev et al. (2018) and many-body localization Abanin and Papić (2017); Nandkishore and Huse (2015). In the following we will show how to employ a Hamiltonian moment based hybrid quantum-classical approach to compute the energy and magnetism of this model Hamiltonian (120) governing four sites mapped to four qubits.
IV.1 Computation preparation
To launch the quantum computation, we first need to prepare our trial state. The following hardware efficient ansatz is used for the preparation of our trial state
(121) |
where is a controlled-Z operation with control qubit and target qubit , is a rotation of on qubit around -axis, and two rotations . We employ the hybrid ITE approach for our demonstration where the Hamiltonian moments are evaluated from IBM-Q NISQ quantum hardware, and the energy and magnetism computations are performed using classical ITE approach on quantum computed Hamiltonian moments. Specifically, the ITE propagator is approximated using a -order Taylor expansion,
(122) |
and therefore and can be approximated as linear combinations of .
IV.2 Quantum measurement
As mentioned in preceding sections, since can be generally expressed as a linear combination of Pauli strings, ’s, is then evaluated based on the measurements of ’s. In quantum computation, there are generally two ways of measuring observables, direct and indirect measurement. In the former, the measured state collapses to the measurement basis that one can freely choose. In the latter, the measured state is not completely destroyed. One simplest and important example is the Hadamard test Aharonov, Jones, and Landau (2009). In the Hadamard test of in the computational basis for a trial state governed by a unitary , one is able to reuse the state after the measurement. This property can be exploited for designing more efficient and less resource-demanding version of some quantum algorithms (for example, iterative version of the quantum phase estimation, see Refs. 138; 139). Here for the demonstration we first employ the Hadamard test to measure each that contributes , and then to show how to group ’s to enable simultaneous measurements.
As discussed in Section II.1, the Pauli cyclic rules (12) allow for some reduction of the number of Pauli strings that need to be measured. Here, for the four-site Heisenberg model, we found that 72 Pauli strings form a complete basis for reformulating arbitrary . This finding suggests that once ’s are measured one can quickly (classically) compute approximate and using arbitrary order expansion for any . Moreover, the relatively few number of required ’s allows the feasibility of measuring ’s on physical quantum computers. Here, four IBM-Q’s publicly available devices, ibmq_quito, ibmq_santiago, ibmq_bogota, and ibmq_manila are employed, whose topologies and noise metrics are shown in Fig. 3.

IV.3 Estimated ground state energy and magnetism
We estimated the ground state energies of the four-site Heisenberg model Hamiltonian for varied and values with employing approximate ITE approach, where is approximated by a 15th order Taylor expansion. In all the calculations, we set . A detailed discussion with respect to different truncation of the Taylor expansion and the choice of will be given in Section IV.4.3.
Fig. 4a exhibits the computed ground state energies of the model Hamiltonian with varied , which are in an excellent agreement with the exact solutions shown in Fig. 4b, and the average mean squared error between the two is only 0.0025 a.u. Based on the measured ’s we are also able to evaluate the magnetization of the model Hamiltonian. The magnetization operator is given by . The contour of the expectation values of magnetization with respect to the estimated ground states under different and values is shown in Fig. 4c, which are again in great agreement with the exact solutions shown in Fig. 4d.

IV.4 Discussions
IV.4.1 In comparison with VQE
Remarkably, the ground state energy and magnetization estimations in the present study exhibit great improvement in comparison with the VQE solutions of the same model on the real hardware reported in previous work Kandala et al. (2017). It is also worth mentioning that (i) the ansatz we used here is much simpler and shallower, and thus less robust, than the one used in the previous study, (ii) the measurement requirement for the model Hamiltonian is trivial in the present study, and does not heuristically depends on the quality of the ansatz and the number of iterations as it does in the conventional VQE practice, and (iii) as shown in Fig. 4e the deviation between measured and analytical ’s could be as large as a.u. given . Nevertheless, despite of less robust ansatz employed and the noisy ’s, the ITE approach is able to give very accurate estimation based on trivial measurements, thus features a robust error mitigation ability from the algorithmic perspective. Similar feature has also been reported in the real quantum application of other Krylov subspace methods, for example the hybrid and quantum Lanczos approaches Suchsland et al. (2021); Motta et al. (2020).
IV.4.2 Grouping and readout error mitigation
As discussed in the preceding section, we can further reduce the number of terms to be measured by grouping the Pauli strings that commute and performing simultaneous measurements Yen, Verteletskyi, and Izmaylov (2020). Here, by applying QWC grouping, we found that the number of terms to be measured can be further reduced from 72 to 25 for the model Hamiltonian.
Another advantage of performing simultaneous measurement is that the readout error of the simultaneous measurement could be largely mitigated through calibration. To see how it works, suppose we perform -qubit simultaneous measurements for totally computational bases, the normalized resulting counts from the measurement of each basis, when collected, would constitute a matrix of dimension. Under the ideal noise-free condition, we know that is an identity matrix of the same dimension. In the real NISQ quantum computing, however, with the effect of noise deviates from the identity matrix, and the extent of deviation implies how noisy the real measurement outcomes are. Based on this relation, if is known, we can algebraically build a connection between any ideal counts vector and its noisy analogue from the real measurement through
(123) |

Apparently, the construction of matrix eats up the resource quickly due to the exponential growth of the dimension. Nevertheless, for small size problem like the four-site model Hamiltonian studied in the present work, is a matrix, and a direction construction of from simultaneously measuring 16 computation bases is still feasible. Here, after grouping the commuting Pauli strings into 25 QWC bases, we performed the simultaneous measurement in the computational basis for each QWC basis five times on four IBM-Q machines, and then mitigated the measurement outcomes employing (123). The raw and calibrated measurement outcomes are exhibited in Fig. 5a-d. As can be seen, after calibration the measurement outcomes are greatly improved showing great consistency with the analytical ones and significantly suppressing the deviation. Based on the measurement outcomes, the estimated ground state energies of the model Hamiltonian are in excellent agreement with the exact solutions (see Fig. 5e), and the MSE between the estimated and exact energies over all the ()’s is only 0.0045 a.u.
IV.4.3 Approximation in ITE
In the case of the model Hamiltonian studied here, as stated at the beginning of this section we approximate the imaginary time operator employing truncated Taylor expansion. Generally speaking, for , the truncated expansion of an exponential function is -approximation when using terms in the expansion Sachdeva and Vishnoi (2014). Regarding the ploynomial approximations to , before figuring out the number of terms in the expansion, the variable needs to be determined first as it rescales the spectrum of and indirectly affects the level of accuracy of the employed expansion. To do so, as exemplified in Fig. 6, we fix the number of terms in the expansion and evolve the energy as a function of to find its optimal value that gives the lowest energy. This is essentially a one dimensional optimization problem over , and can be solved classically by Golden-Section Search Kiefer (1953), or similar methods. In the present study, we found providing well converged ground state energies estimates for the given and ranges. It worth mentioning that to improve the estimate of using polynomial expansions in the larger regime, one can alternatively resort to other polynomial expansions, such as CMX as discussed in Section III.3, or Padé expansion as originally proposed in Ref. 88 and recently applied in Ref. 96.

After fixing , we can test the accuracy of the truncated Taylor expansion using different numbers of terms. As shown in Fig. 7 the level of accuracy improves as more terms used in the truncated Taylor expansion, and it verifies that at least a 15th order expansion is needed in order to give accurate energy estimates for the entire and ranges.

V Conclusion and outlook
In this tutorial review, we have gone over some recent developments of hybrid quantum-classical algorithms based upon quantum
computation of Hamiltonian moments. In particular, we have given an overview about how to rely on quantum resources to compute Hamiltonian
moments, and what classical methods can be used following the quantum computation of the Hamiltonian moments to give rise to energy
estimation for the target state. The levels of accuracy of these hybrid quantum-classical approaches are usually systematically improvable,
even under the condition of working with noisy Hamiltonian moments, and thus provide error mitigation at the algorithmic level regardless of
detailed description of the noise. Essentially, the algorithmic error mitigation can be attributed to the fact that these
hybrid approaches are exploring the Krylov subspace of the entire Hilbert space, and if the ansatz is not totally orthogonal
the true state, the power iteration would in principle guarantee a convergence given sufficiently high order Hamiltonian moments.
On the other hand, what comes with the systematically improvable accuracy that is based on the cumulative construction of Krylov subspace is
the price paid for obtaining higher order Hamiltonian moments, which will significantly increase the measurement requirement and bring
hurdles for targeting larger quantum systems. The ongoing efforts are focused on either introducing some more efficient grouping rules
to reduce the number of terms to be measured, or polishing the quality of the ansatz based on some physical and chemical inspiration
such that lower order Hamiltonian moments will be sufficient for approaching certain level of accuracy. The latter can even be
combined with variational quantum solver for the same purpose, which also helps the variational solver circumvent the barren plateaus.
Another active direction is to combine these techniques with complete active space and/or Hamiltonian downfolding techniques
(for recent studies, see e.g. Refs. 141; 142) that can target larger systems.
All these combinations are worth exploring in the NISQ era. Ultimately, from the accuracy and efficiency improvements brought
by these algorithms and/or their combinations with some other techniques, the critical question becomes which problems are most
benefited from quantum computed Hamiltonian moments and how much quantum advantage can be eventually gained in these contexts.
VI Acknowledgement
J. C. A., T. K., and B. P. acknowledge Dr. Karol Kowalski and Dr. Niri Govind for their helpful discussions on some topics during the preparation of this paper.
VII Funding Information
J. C. A. was supported in part by the U.S. Department of Energy, Office of Science, Office of Workforce Development for Teachers and Scientists (WDTS) under the Science Undergraduate Laboratory Internships Program (SULI). T. K. is supported by the U. S. Department of Energy, Office of Science, Office of Workforce Development for Teachers and Scientists, Office of Science Graduate Student Research (SCGSR) program. The SCGSR program is administered by the Oak Ridge Institute for Science and Education (ORISE) for the DOE. ORISE is managed by ORAU under contract number DE-SC0014664. B. P. was supported by the “Embedding QC into Many-body Frameworks for Strongly Correlated Molecular and Materials Systems” project, which is funded by the U.S. Department of Energy, Office of Science, Office of Basic Energy Sciences (BES), the Division of Chemical Sciences, Geosciences, and Biosciences. This material is based upon work supported by the U.S. Department of Energy, Office of Science, National Quantum Information Science Research Centers.
VIII Data Availability Statement
The data and code that support the findings of this study are available at https://github.com/jaulicino/HamiltonianMoment/
References
- Feynman (1982) R. P. Feynman, “Simulating physics with computers,” Int. J. Theor. Phys. 21, 467–488 (1982).
- David Sherrill and Schaefer (1999) C. David Sherrill and H. F. Schaefer, “The configuration interaction method: Advances in highly correlated approaches,” (Academic Press, 1999) pp. 143–269.
- Bartlett and Musiał (2007) R. J. Bartlett and M. Musiał, “Coupled-cluster theory in quantum chemistry,” Rev. Mod. Phys. 79, 291–352 (2007).
- Sugisaki et al. (2019) K. Sugisaki, S. Nakazawa, K. Toyota, K. Sato, D. Shiomi, and T. Takui, “Quantum chemistry on quantum computers: A method for preparation of multiconfigurational wave functions on quantum computers without performing post-hartree–fock calculations,” ACS Central Sci. 5, 167–175 (2019).
- Aspuru-Guzik et al. (2005) A. Aspuru-Guzik, A. D. Dutoi, P. J. Love, and M. Head-Gordon, “Simulated quantum computation of molecular energies,” Science 309, 1704–1707 (2005).
- O’Malley et al. (2016) P. J. J. O’Malley, R. Babbush, I. D. Kivlichan, J. Romero, J. R. McClean, R. Barends, J. Kelly, P. Roushan, A. Tranter, N. Ding, B. Campbell, Y. Chen, Z. Chen, B. Chiaro, A. Dunsworth, A. G. Fowler, E. Jeffrey, E. Lucero, A. Megrant, J. Y. Mutus, M. Neeley, C. Neill, C. Quintana, D. Sank, A. Vainsencher, J. Wenner, T. C. White, P. V. Coveney, P. J. Love, H. Neven, A. Aspuru-Guzik, and J. M. Martinis, “Scalable quantum simulation of molecular energies,” Phys. Rev. X 6, 031007 (2016).
- Linke et al. (2017) N. M. Linke, D. Maslov, M. Roetteler, S. Debnath, C. Figgatt, K. A. Landsman, K. Wright, and C. Monroe, “Experimental comparison of two quantum computing architectures,” Proc. Natl. Acad. Sci. U. S. A. 114, 3305–3310 (2017).
- Kandala et al. (2017) A. Kandala, A. Mezzacapo, K. Temme, M. Takita, M. Brink, J. M. Chow, and J. M. Gambetta, “Hardware-efficient variational quantum eigensolver for small molecules and quantum magnets,” Nature 549, 242–246 (2017).
- Dumitrescu et al. (2018) E. F. Dumitrescu, A. J. McCaskey, G. Hagen, G. R. Jansen, T. D. Morris, T. Papenbrock, R. C. Pooser, D. J. Dean, and P. Lougovski, “Cloud quantum computing of an atomic nucleus,” Phys. Rev. Lett. 120, 210501 (2018).
- Klco et al. (2018) N. Klco, E. F. Dumitrescu, A. J. McCaskey, T. D. Morris, R. C. Pooser, M. Sanz, E. Solano, P. Lougovski, and M. J. Savage, “Quantum-classical computation of schwinger model dynamics using quantum computers,” Phys. Rev. A 98, 032331 (2018).
- Colless et al. (2018) J. I. Colless, V. V. Ramasesh, D. Dahlen, M. S. Blok, M. E. Kimchi-Schwartz, J. R. McClean, J. Carter, W. A. de Jong, and I. Siddiqi, “Computation of molecular spectra on a quantum processor with an error-resilient algorithm,” Phys. Rev. X 8, 011021 (2018).
- McCaskey et al. (2019) A. J. McCaskey, Z. P. Parks, J. Jakowski, S. V. Moore, T. D. Morris, T. S. Humble, and R. C. Pooser, “Quantum chemistry as a benchmark for near-term quantum computers,” npj Quantum Inf. 5, 98 (2019).
- Preskill (2018) J. Preskill, “Quantum Computing in the NISQ era and beyond,” Quantum 2, 79 (2018).
- Peruzzo et al. (2014) A. Peruzzo, J. R. McClean, P. Shadbolt, M.-H. Yung, X.-Q. Zhou, P. J. Love, A. Aspuru-Guzik, and J. L. O’brien, “A variational eigenvalue solver on a photonic quantum processor,” Nat. Commun. 5, 4213 (2014).
- McClean et al. (2016) J. R. McClean, J. Romero, R. Babbush, and A. Aspuru-Guzik, “The theory of variational hybrid quantum-classical algorithms,” New J. Phys. 18, 023023 (2016).
- Romero et al. (2018) J. Romero, R. Babbush, J. R. McClean, C. Hempel, P. J. Love, and A. Aspuru-Guzik, “Strategies for quantum computing molecular energies using the unitary coupled cluster ansatz,” Quantum Sci. Technol. 4, 014008 (2018).
- Shen et al. (2017) Y. Shen, X. Zhang, S. Zhang, J.-N. Zhang, M.-H. Yung, and K. Kim, “Quantum implementation of the unitary coupled cluster for simulating molecular electronic structure,” Phys. Rev. A 95, 020501 (2017).
- Kandala et al. (2019) A. Kandala, K. Temme, A. D. Corcoles, A. Mezzacapo, J. M. Chow, and J. M. Gambetta, “Error mitigation extends the computational reach of a noisy quantum processor,” Nature 567, 491–495 (2019).
- Huggins et al. (2020) W. J. Huggins, J. Lee, U. Baek, B. O’Gorman, and K. B. Whaley, “A non-orthogonal variational quantum eigensolver,” New J. Phys. 22, 073009 (2020).
- Farhi, Goldstone, and Gutmann (2014) E. Farhi, J. Goldstone, and S. Gutmann, “A quantum approximate optimization algorithm,” arXiv preprint arXiv:1411.4028 (2014).
- Bharti et al. (2021) K. Bharti, A. Cervera-Lierta, T. H. Kyaw, T. Haug, S.Alperin-Lea, A. Anand, M. Degroote, H. Heimonen, J. S. Kottmann, T. Menke, W.-K. Mok, S. Sim, L.-C. Kwek, and A. Aspuru-Guzik, “Noisy intermediate-scale quantum (nisq) algorithms,” arXiv preprint arXiv:2101.08448 (2021).
- Albash and Lidar (2018) T. Albash and D. A. Lidar, “Adiabatic quantum computation,” Rev. Mod. Phys. 90, 015002 (2018).
- Aaronson and Arkhipov (2011) S. Aaronson and A. Arkhipov, “Proceedings of the forty-third annual acm symposium on theory of computing,” (2011) pp. 333–342.
- Trabesinger (2012) A. Trabesinger, “Quantum simulation,” Nat. Phys. 8, 263 (2012).
- Georgescu, Ashhab, and Nori (2014) I. M. Georgescu, S. Ashhab, and F. Nori, “Quantum simulation,” Rev. Mod. Phys. 86, 153–185 (2014).
- McArdle et al. (2019) S. McArdle, T. Jones, S. Endo, Y. Li, S. C. Benjamin, and X. Yuan, “Variational ansatz-based quantum simulation of imaginary time evolution,” npj Quantum Inf. 5, 1–6 (2019).
- Motta et al. (2020) M. Motta, C. Sun, A. T. K. Tan, M. J. O’Rourke, E. Ye, A. J. Minnich, F. G. S. L. Brandão, and G. K.-L. Chan, “Determining eigenstates and thermal states on a quantum computer using quantum imaginary time evolution,” Nat. Phys. 16, 205–210 (2020).
- Parrish and McMahon (2019) R. M. Parrish and P. L. McMahon, “Quantum filter diagonalization: Quantum eigendecomposition without full quantum phase estimation,” arXiv preprint arXiv:1909.08925 (2019).
- Kyriienko (2020) O. Kyriienko, “Quantum inverse iteration algorithm for programmable quantum simulators,” npj Quantum Inf. 6, 1–8 (2020).
- McClean et al. (2018) J. R. McClean, S. Boixo, V. N. Smelyanskiy, R. Babbush, and H. Neven, “Barren plateaus in quantum neural network training landscapes,” Nat. Commun. 9, 4812 (2018).
- Cerezo et al. (2021) M. Cerezo, A. Sone, T. Volkoff, L. Cincio, and P. J. Coles, “Cost function dependent barren plateaus in shallow parametrized quantum circuits,” Nat. Commun. 12, 1791 (2021).
- Cerezo and Coles (2021) M. Cerezo and P. J. Coles, “Higher order derivatives of quantum neural networks with barren plateaus,” Quantum Sci. Technol. 6, 035006 (2021).
- Herasymenko and O’Brien (2019) Y. Herasymenko and T. E. O’Brien, “A diagrammatic approach to variational quantum ansatz construction,” arXiv preprint arXiv:1907.08157 (2019).
- Bonet-Monroig, Babbush, and O’Brien (2020) X. Bonet-Monroig, R. Babbush, and T. E. O’Brien, “Nearly optimal measurement scheduling for partial tomography of quantum states,” Phys. Rev. X 10, 031064 (2020).
- Endo et al. (2021) S. Endo, Z. Cai, S. C. Benjamin, and X. Yuan, “Hybrid quantum-classical algorithms and quantum error mitigation,” J. Phys. Soc. Jpn. 90, 032001 (2021).
- Jordan and Wigner (1928) P. Jordan and E. Wigner, “Über das paulische Äquivalenzverbot,” Z. Physik 47, 631–651 (1928).
- Bravyi and Kitaev (2002) S. B. Bravyi and A. Y. Kitaev, “Fermionic quantum computation,” Ann. Phys. 298, 210–226 (2002).
- Seeley, Richard, and Love (2012) J. T. Seeley, M. J. Richard, and P. J. Love, “The bravyi-kitaev transformation for quantum computation of electronic structure,” J. Chem. Phys. 137, 224109 (2012).
- Tranter et al. (2015) A. Tranter, S. Sofia, J. Seeley, M. Kaicher, J. McClean, R. Babbush, P. V. Coveney, F. Mintert, F. Wilhelm, and P. J. Love, “The bravyi–kitaev transformation: Properties and applications,” Int. J. Quantum Chem. 115, 1431–1441 (2015).
- Tyldesley (1975) J. R. Tyldesley, An Introduction to Tensor Analysis for Engineers and Applied Scientists (Longman, 1975).
- Kowalski and Peng (2020) K. Kowalski and B. Peng, “Quantum simulations employing connected moments expansions,” J. Chem. Phys. 153, 201102 (2020).
- Suchsland et al. (2021) P. Suchsland, F. Tacchino, M. H. Fischer, T. Neupert, P. K. Barkoutsos, and I. Tavernelli, “Algorithmic Error Mitigation Scheme for Current Quantum Processors,” Quantum 5, 492 (2021).
- Karp (2010) R. M. Karp, “Reducibility among combinatorial problems,” in 50 Years of Integer Programming 1958-2008: From the Early Years to the State-of-the-Art, edited by M. Jünger, T. M. Liebling, D. Naddef, G. L. Nemhauser, W. R. Pulleyblank, G. Reinelt, G. Rinaldi, and L. A. Wolsey (Springer Berlin Heidelberg, Berlin, Heidelberg, 2010) pp. 219–241.
- Gokhale et al. (2020) P. Gokhale, O. Angiuli, Y. Ding, K. Gui, T. Tomesh, M. Suchara, M. Martonosi, and F. T. Chong, “ measurement cost for variational quantum eigensolver on molecular hamiltonians,” IEEE Trans. Qunatum Eng. 1, 1–24 (2020).
- Yen, Verteletskyi, and Izmaylov (2020) T.-C. Yen, V. Verteletskyi, and A. F. Izmaylov, “Measuring all compatible operators in one series of single-qubit measurements using unitary transformations,” J. Chem. Theory Comput. 16, 2400–2409 (2020).
- Vallury et al. (2020) H. J. Vallury, M. A. Jones, C. D. Hill, and L. C. L. Hollenberg, “Quantum computed moments correction to variational estimates,” Quantum 4, 373 (2020).
- Claudino et al. (2021) D. Claudino, B. Peng, N. P. Bauman, K. Kowalski, and T. S. Humble, “Improving the accuracy and efficiency of quantum connected moments expansions,” Quantum Sci. Technol. 6, 034012 (2021).
- Schwinger (1960) J. Schwinger, “Unitary operator bases,” Proc. Nat. Acad. Sci. USA 46, 570–579 (1960).
- Klappenecker and Rötteler (2004) A. Klappenecker and M. Rötteler, “Constructions of mutually unbiased bases,” in Finite Fields and Applications, edited by G. L. Mullen, A. Poli, and H. Stichtenoth (Springer Berlin Heidelberg, Berlin, Heidelberg, 2004) pp. 137–144.
- Altepeter, James, and Kwiat (2004) J. B. Altepeter, D. F. V. James, and P. G. Kwiat, “4 qubit quantum state tomography,” in Quantum State Estimation, edited by M. Paris and J. Řeháček (Springer Berlin Heidelberg, Berlin, Heidelberg, 2004) pp. 113–145.
- Lawrence, Brukner, and Zeilinger (2002) J. Lawrence, i. c. v. Brukner, and A. Zeilinger, “Mutually unbiased binary observable sets on n qubits,” Phys. Rev. A 65, 032320 (2002).
- Izmaylov et al. (2020) A. F. Izmaylov, T.-C. Yen, R. A. Lang, and V. Verteletskyi, “Unitary partitioning approach to the measurement problem in the variational quantum eigensolver method,” J. Chem. Theory Comput. 16, 190–195 (2020).
- Huggins et al. (2021) W. J. Huggins, J. R. McClean, N. C. Rubin, Z. Jiang, N. Wiebe, K. B. Whaley, and R. Babbush, “Efficient and noise resilient measurements for quantum chemistry on near-term quantum computers,” npj Quantum Inf. 7 (2021), 10.1038/s41534-020-00341-7.
- Gonthier et al. (2020) J. F. Gonthier, M. D. Radin, C. Buda, E. J. Doskocil, C. M. Abuan, and J. Romero, “Identifying challenges towards practical quantum advantage through resource estimation: the measurement roadblock in the variational quantum eigensolver,” (2020), arXiv:2012.04001 [quant-ph] .
- Rubin, Babbush, and J. (2018) N. C. Rubin, R. Babbush, and M. J., “Application of fermionic marginal constraints to hybrid quantum algorithms,” New J. Phys. 20, 053020 (2018).
- Wecker, Hastings, and Troyer (2015) D. Wecker, M. B. Hastings, and M. Troyer, “Progress towards practical quantum variational algorithms,” Phys. Rev. A 92, 042303 (2015).
- Huang, Kueng, and Preskill (2020) H. Y. Huang, R. Kueng, and J. Preskill, “Predicting many properties of a quantum system from very few measurements,” Nat. Phys. 16, 1050–1057 (2020).
- Huang, Kueng, and Preskill (2021) H.-Y. Huang, R. Kueng, and J. Preskill, “Efficient estimation of pauli observables by derandomization,” Phys. Rev. Lett. 127, 030503 (2021).
- Aaronson (2018) S. Aaronson, “Shadow tomography of quantum states,” in Proceedings of the 50th Annual ACM SIGACT Symposium on Theory of Computing, STOC 2018 (Association for Computing Machinery, New York, NY, USA, 2018) pp. 325–338.
- Struchalin et al. (2021) G. I. Struchalin, Y. A. Zagorovskii, E. V. Kovlakov, S. S. Straupe, and S. P. Kulik, “Experimental estimation of quantum state properties from classical shadows,” PRX Quantum 2, 010307 (2021).
- Chen et al. (2021) S. Chen, W. Yu, P. Zeng, and S. T. Flammia, “Robust shadow estimation,” arXiv preprint arXiv:2011.09636 (2021).
- Zhao, Rubin, and Miyake (2021) A. Zhao, N. C. Rubin, and A. Miyake, “Fermionic partial tomography via classical shadows,” Phys. Rev. Lett. 127, 110504 (2021).
- Acharya, Saha, and Sengupta (2021) A. Acharya, S. Saha, and A. M. Sengupta, “Informationally complete povm-based shadow tomography,” arXiv preprint arXiv:2105.05992 (2021).
- Hadfield (2021) C. Hadfield, “Adaptive pauli shadows for energy estimation,” arXiv preprint arXiv:2105.12207 (2021).
- Hillmich et al. (2021) S. Hillmich, C. Hadfield, R. Raymond, A. Mezzacapo, and R. Wille, “Decision diagrams for quantum measurements with shallow circuits,” arXiv preprint arXiv:2105.06932 (2021).
- Zhang et al. (2021) T. Zhang, J. Sun, X.-X. Fang, X. Zhang, X. Yuan, and H. Lu, “Experimental quantum state measurement with classical shadows,” arXiv preprint arXiv:2106.10190 (2021).
- Subramanian, Brierley, and Jozsa (2019) S. Subramanian, S. Brierley, and R. Jozsa, “Implementing smooth functions of a hermitian matrix on a quantum computer,” J. Phys. Commun. 3, 065002 (2019).
- Venegas-Andraca (2012) S. Venegas-Andraca, “Quantum walks: a comprehensive review,” Quantum Inf. Process. 11, 1015–1106 (2012).
- Kempe (2003) J. Kempe, “Quantum random walks: An introductory overview,” Contemp. Phys. 44, 307–327 (2003).
- Patel and Priyadarsini (2018) A. Patel and A. Priyadarsini, “Efficient quantum algorithms for state measurement and linear algebra applications,” Int. J. Quantum Inf. 16, 1850048 (2018).
- Seki and Yunoki (2021) K. Seki and S. Yunoki, “Quantum power method by a superposition of time-evolved states,” PRX Quantum 2, 010333 (2021).
- Trotter (1959) H. F. Trotter, “On the product of semi-groups of operators,” Proc. Amer. Math. Soc. 10, 545–551 (1959).
- Suzuki (1990) M. Suzuki, “Fractal decomposition of exponential operators with applications to many-body theories and monte carlo simulations,” Phys. Lett. A 146, 319–323 (1990).
- Richardson and Gaunt (1927) L. F. Richardson and J. A. Gaunt, “Viii. the deferred approach to the limit,” Philosophical Transactions of the Royal Society of London. Series A, Containing Papers of a Mathematical or Physical Character 226, 299–361 (1927).
- Temme, Bravyi, and Gambetta (2017) K. Temme, S. Bravyi, and J. M. Gambetta, “Error mitigation for short-depth quantum circuits,” Phys. Rev. Lett. 119, 180509 (2017).
- Sachdeva and Vishnoi (2014) S. Sachdeva and N. K. Vishnoi, Faster Algorithms via Approximation Theory (Now Foundations and Trends, 2014).
- Childs, Kothari, and Somma (2017) A. M. Childs, R. Kothari, and R. D. Somma, “Quantum algorithm for systems of linear equations with exponentially improved dependence on precision,” SIAM Journal on Computing 46, 1920–1950 (2017).
- Ekert et al. (2002) A. K. Ekert, C. M. Alves, D. K. L. Oi, M. Horodecki, P. Horodecki, and L. C. Kwek, “Direct estimations of linear and nonlinear functionals of a quantum state,” Phys. Rev. Lett. 88, 217901 (2002).
- Higgott, Wang, and Brierley (2019) O. Higgott, D. Wang, and S. Brierley, “Variational Quantum Computation of Excited States,” Quantum 3, 156 (2019).
- Mitarai and Fujii (2019) K. Mitarai and K. Fujii, “Methodology for replacing indirect measurements with direct measurements,” Phys. Rev. Research 1, 013006 (2019).
- Ge, Tura, and Cirac (2019) Y. Ge, J. Tura, and J. I. Cirac, “Faster ground state preparation and high-precision ground energy estimation with fewer qubits,” J. Math. Phys. 60, 1–25 (2019).
- Brassard et al. (2002) G. Brassard, P. Hoyer, M. Mosca, and A. Tapp, “Quantum amplitude amplification and estimation,” in Quantum Computation and Quantum Information, Vol. 305, edited by J. S. J. Lomonaco (AMS Contemporary Mathematics, 2002) pp. 53–74.
- Berry et al. (2015) D. W. Berry, A. M. Childs, R. Cleve, R. Kothari, and R. D. Somma, “Simulating hamiltonian dynamics with a truncated taylor series,” Phys. Rev. Lett. 114, 090502 (2015).
- Low and Chuang (2017) G. H. Low and I. L. Chuang, “Optimal hamiltonian simulation by quantum signal processing,” Phys. Rev. Lett. 118, 010501 (2017).
- Gilyén et al. (2019) A. Gilyén, Y. Su, G. H. Low, and N. Wiebe, “Quantum singular value transformation and beyond: Exponential improvements for quantum matrix arithmetics,” in Proceedings of the 51st Annual ACM SIGACT Symposium on Theory of Computing, STOC 2019 (Association for Computing Machinery, New York, NY, USA, 2019) p. 193–204.
- Childs and Wiebe (2012) A. M. Childs and N. Wiebe, “Hamiltonian simulation using linear combinations of unitary operations,” Quantum Inf. Comput. 12 (2012), 10.26421/qic12.11-12.
- Lanczos (1950) C. Lanczos, “An iteration method for the solution of the eigenvalue problem of linear differential and integral operators,” J. Res. Natl. Bur. Stand. B 45, 255–283 (1950).
- Horn and Weinstein (1984) D. Horn and M. Weinstein, “The expansion: A nonperturbative analytic tool for hamiltonian systems,” Phys. Rev. D 30, 1256–1270 (1984).
- Hollenberg (1993) L. C. L. Hollenberg, “Plaquette expansion in lattice hamiltonian models,” Phys. Rev. D 47, 1640–1644 (1993).
- Witte and Hollenberg (1994) N. S. Witte and L. C. L. Hollenberg, “Plaquette expansion proof and interpretation,” Z. Phys. B 95 (1994).
- Hollenberg and Witte (1996) L. C. L. Hollenberg and N. S. Witte, “Analytic solution for the ground-state energy of the extensive many-body problem,” Phys. Rev. B 54, 16309–16312 (1996).
- van Doorn (1987) E. A. van Doorn, “Representations and bounds for zeros of orthogonal polynomials and eigenvalues of sign-symmetric tri-diagonal matrices,” J. Approx. Theory 51, 254–266 (1987).
- Ismail and Li (1992) M. E. H. Ismail and X. Li, “Bound on the extreme zeros of orthogonal polynomials,” Proc. Amer. Math. Soc. 115, 131–140 (1992).
- Hollenberg and Witte (1994) L. C. L. Hollenberg and N. S. Witte, “General nonperturbative estimate of the energy density of lattice hamiltonians,” Phys. Rev. D 50, 3382–3386 (1994).
- Pavarini et al. (2011) E. Pavarini, E. Koch, A. Lichtenstein, and D. Vollhardt, eds., The LDA+DMFT approach to strongly correlated materials, Schriften des Forschungszentrums Jülich. Reihe modeling and simulation, Vol. 1 (Forschungszenrum Jülich GmbH Zentralbibliothek, Verlag, Jülich, 2011).
- Guzman and Lacroix (2021) E. A. R. Guzman and D. Lacroix, “Predicting ground state, excited states and long-time evolution of many-body systems from short-time evolution on a quantum computer,” arXiv preprint arXiv:2104.08181 (2021).
- McClean et al. (2017) J. R. McClean, M. E. Kimchi-Schwartz, J. Carter, and W. A. de Jong, “Hybrid quantum-classical hierarchy for mitigation of decoherence and determination of excited states,” Phys. Rev. A 95, 042308 (2017).
- Parrish et al. (2019) R. M. Parrish, E. G. Hohenstein, P. L. McMahon, and T. J. Martínez, “Quantum computation of electronic transitions using a variational quantum eigensolver,” Phys. Rev. Lett. 122, 230401 (2019).
- Nakanishi, Mitarai, and Fujii (2019) K. M. Nakanishi, K. Mitarai, and K. Fujii, “Subspace-search variational quantum eigensolver for excited states,” Phys. Rev. Research 1, 033062 (2019).
- Davies et al. (1980) K. T. R. Davies, H. Flocard, S. Krieger, and M. S. Weiss, “Application of the imaginary time step method to the solution of the static hartree-fock problem,” Nuc. Phys. A 342, 111–123 (1980).
- Anderson (1975) J. B. Anderson, “A random‐walk simulation of the schrödinger equation: H+3,” J. Chem. Phys. 63, 1499–1503 (1975).
- Anderson (1979) J. B. Anderson, “Quantum chemistry by random walk: H4 square,” Int. J. Quantum Chem. 15, 109–120 (1979).
- Anderson (1980) J. B. Anderson, “Quantum chemistry by random walk: Higher accuracy,” J. Chem. Phys. 73, 3897–3899 (1980).
- Cioslowski (1987) J. Cioslowski, “Connected moments expansion: A new tool for quantum many-body theory,” Phys. Rev. Lett. 58, 83–85 (1987).
- Knowles (1987) P. J. Knowles, “On the validity and applicability of the connected moments expansion,” Chem. Phys. Lett. 134, 512–518 (1987).
- Stubbins (1988) C. Stubbins, “Methods of extrapolating the t-expansion series,” Phys. Rev. D 38, 1942 (1988).
- Prie et al. (1994) J. D. Prie, D. Schwall, J. D. Mancini, D. Kraus, and W. J. Massano, “On the relation between the connected-moments expansion and the lanczos variational scheme,” Nuov. Cim. D 16, 433–448 (1994).
- Mancini, Zhou, and Meier (1994) J. D. Mancini, Y. Zhou, and P. F. Meier, “Analytic properties of connected moments expansions,” Int. J. Quantum Chem. 50, 101–107 (1994).
- Ullah (1995) N. Ullah, “Removal of the singularity in the moment-expansion formalism,” Phys. Rev. A 51, 1808–1810 (1995).
- Mancini et al. (1995) J. D. Mancini, W. J. Massano, J. D. Prie, and Y. Zhuo, “Avoidance of singularities in moments expansions: a numerical study,” Phys. Lett. A 209, 107–112 (1995).
- Fessatidis et al. (2006) V. Fessatidis, J. D. Mancini, R. Murawski, and S. P. Bowen, “A generalized moments expansion,” Phys. Lett. A 349, 320–323 (2006).
- Fessatidis et al. (2010) V. Fessatidis, F. A. Corvino, J. D. Mancini, R. K. Murawski, and J. Mikalopas, “Analytic properties of moments matrices,” Phys. Lett. A 374, 2890–2893 (2010).
- Noga, Szabados, and Surján (2002) J. Noga, A. Szabados, and P. Surján, “On the use of connected moments expansion with coupled cluster reference,” Int. J. Mol. Sci. 3, 508–521 (2002).
- Mancini, Prie, and Massano (1991) J. D. Mancini, J. D. Prie, and W. J. Massano, “Approximations to the ground-state energy of the single-impurity kondo model using variational and connected-moments expansions,” Phys. Rev. A 43, 1777–1782 (1991).
- Peeters and Devreese (1984) F. M. Peeters and J. T. Devreese, “Upper bounds for the free energy. a generalisation of the bogolubov inequality and the feynman inequality,” Journal of Physics A: Mathematical and General 17, 625 (1984).
- Soldatov (1995) A. V. Soldatov, “Generalized variational principle in quantum mechanics,” Int. J. Mod. Phys. B 9, 2899–2936 (1995).
- Bogolubov (1947) N. Bogolubov, “On the theory of superfluidity,” J. Phys. 11, 23–32 (1947).
- Feynman (1955) R. P. Feynman, “Slow electrons in a polar crystal,” Phys. Rev. 97, 660–665 (1955).
- Devreese, Evrard, and Kartheuser (1975) J. T. Devreese, R. Evrard, and E. Kartheuser, “Self-consistent equation-of-motion approach for polarons,” Phys. Rev. B 12, 3353–3367 (1975).
- Peng and Kowalski (2021) B. Peng and K. Kowalski, “Variational quantum solver employing the pds energy functional,” Quantum 5, 473 (2021).
- Yamamoto (2019) N. Yamamoto, “On the natural gradient for variational quantum eigensolver,” arXiv preprint arXiv:1909.05074 (2019).
- Lemaéchal (2012) C. Lemaéchal, “Cauchy and the gradient method,” Doc. Math. Extra Optimization Stories, 251–254 (2012).
- Amari (1998) S. Amari, “Natural gradient works efficiently in learning,” Neural Comput. 10, 251–276 (1998).
- Stokes et al. (2020) J. Stokes, J. Izaac, N. Killoran, and G. Carleo, “Quantum Natural Gradient,” Quantum 4, 269 (2020).
- McLachlan (1964) A. D. McLachlan, “A variational solution of the time-dependent schrodinger equation,” Mol. Phys. 8, 39–44 (1964).
- Yuan et al. (2019) X. Yuan, S. Endo, Q. Zhao, Y. Li, and S. C. Benjamin, “Theory of variational quantum simulation,” Quantum 3, 191 (2019).
- Dirac (1930) P. A. M. Dirac, “Note on exchange phenomena in the thomas atom,” Math. Proc. Camb. Philos. Soc. 26, 376–385 (1930).
- Frenkel (1934) J. Frenkel, Wave Mechanics: Advanced General Theory (Clarendon Press Oxford, 1934).
- Kramer and Saraceno (1981) P. Kramer and M. Saraceno, eds., “The time-dependent variational principle (tdvp),” in Geometry of the Time-Dependent Variational Principle in Quantum Mechanics (Springer Berlin Heidelberg, Berlin, Heidelberg, 1981) pp. 3–14.
- Broeckhove et al. (1988) J. Broeckhove, L. Lathouwers, E. Kesteloot, and P. Van Leuven, “On the equivalence of time-dependent variational principles,” Chem. Phys. Lett. 149, 547–550 (1988).
- Cheng (2013) R. Cheng, “Quantum geometric tensor (fubini-study metric) in simple quantum system: A pedagogical introduction,” arXiv preprint arXiv:1012.1337 (2013).
- Smith et al. (2019) A. Smith, M. Kim, F. Pollmann, and J. Knolle, “Simulating quantum many-body dynamics on a current digital quantum computer,” npj Quantum Inf. 5 (2019), 10.1038/s41534-019-0217-0.
- Essler and Fagotti (2016) F. H. L. Essler and M. Fagotti, “Quench dynamics and relaxation in isolated integrable quantum spin chains,” J. Stat. Mech.: Theory Exp. 2016, 064002 (2016).
- Vasiliev et al. (2018) A. Vasiliev, O. Volkova, E. Zvereva, and M. Markina, “Milestones of low-d quantum magnetism,” npj Quant. Mater. 3, 18 (2018).
- Abanin and Papić (2017) D. A. Abanin and Z. Papić, “Recent progress in many-body localization,” Ann. Phys. (Berl.) 529, 1700169 (2017).
- Nandkishore and Huse (2015) R. Nandkishore and D. A. Huse, “Many-body localization and thermalization in quantum statistical mechanics,” Ann. Rev. Condens. Matter Phys. 6, 15–38 (2015).
- Aharonov, Jones, and Landau (2009) D. Aharonov, V. Jones, and Z. Landau, “A polynomial quantum algorithm for approximating the jones polynomial,” Algorithmica 55, 395–421 (2009).
- Knill, Ortiz, and Somma (2007) E. Knill, G. Ortiz, and R. D. Somma, “Optimal quantum measurements of expectation values of observables,” Phys. Rev. A 75, 012328 (2007).
- Dobšíček et al. (2007) M. Dobšíček, G. Johansson, V. Shumeiko, and G. Wendin, “Arbitrary accuracy iterative quantum phase estimation algorithm using a single ancillary qubit: A two-qubit benchmark,” Phys. Rev. A 76, 030306 (2007).
- Kiefer (1953) J. Kiefer, “Sequential minimax search for a maximum,” Proc. Am. Math. Soc. 4, 502–506 (1953).
- Bauman et al. (2019) N. P. Bauman, E. J. Bylaska, S. Krishnamoorthy, G. H. Low, N. Wiebe, C. E. Granade, M. Roetteler, M. Troyer, and K. Kowalski, “Downfolding of many-body hamiltonians using active-space models: Extension of the sub-system embedding sub-algebras approach to unitary coupled cluster formalisms,” J. Chem. Phys. 151, 014107 (2019).
- Mniszewski et al. (2021) S. M. Mniszewski, P. A. Dub, S. Tretiak, P. M. Anisimov, Y. Zhang, and C. F. A. Negre, “Reduction of the molecular hamiltonian matrix using quantum community detection,” Sci. Rep. 11, 4099 (2021).