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

thanks: These two authors contributed equally.thanks: These two authors contributed equally.

Measurement-efficient quantum Krylov subspace diagonalisation

Zongkang Zhang Graduate School of China Academy of Engineering Physics, Beijing 100193, China    Anbang Wang Graduate School of China Academy of Engineering Physics, Beijing 100193, China    Xiaosi Xu Graduate School of China Academy of Engineering Physics, Beijing 100193, China    Ying Li [email protected] Graduate School of China Academy of Engineering Physics, Beijing 100193, China
Abstract

The Krylov subspace methods, being one category of the most important classical numerical methods for linear algebra problems, can be much more powerful when generalised to quantum computing. However, quantum Krylov subspace algorithms are prone to errors due to inevitable statistical fluctuations in quantum measurements. To address this problem, we develop a general theoretical framework to analyse the statistical error and measurement cost. Based on the framework, we propose a quantum algorithm to construct the Hamiltonian-power Krylov subspace that can minimise the measurement cost. In our algorithm, the product of power and Gaussian functions of the Hamiltonian is expressed as an integral of the real-time evolution, such that it can be evaluated on a quantum computer. We compare our algorithm with other established quantum Krylov subspace algorithms in solving two prominent examples. To achieve an error comparable to that of the classical Lanczos algorithm at the same subspace dimension, our algorithm typically requires orders of magnitude fewer measurements than others. Such an improvement can be attributed to the reduced cost of composing projectors onto the ground state. These results show that our algorithm is exceptionally robust to statistical fluctuations and promising for practical applications.

1 Introduction

Finding the ground-state energy of a quantum system is of vital importance in many fields of physics [1, 2, 3]. The Lanczos algorithm [4, 5] is one of the most widely used algorithms to solve this problem. It belongs to the Krylov subspace methods [6], in which the solution usually converges to the true answer with an increasing subspace dimension. However, such methods are unscalable for many-body systems because of the exponentially-growing Hilbert space dimension [7]. In quantum computing, there is a family of hybrid quantum-classical algorithms that can be regarded as a quantum generalisation of the classical Lanczos algorithm [8, 9, 10, 11, 12, 13, 14, 15, 16, 17, 18, 19, 20]. Following Refs. [11, 15, 16], we call them quantum Krylov subspace diagonalisation (KSD). These algorithms are scalable with the system size by carrying out the classically-intractable vector and matrix arithmetic on the quantum computer. They possess potential advantages as the conventional quantum algorithm to solve the ground-state problem, quantum phase estimation, requires considerable quantum resources [21, 22], while the variational quantum eigensolver is limited by the ansatz and classical optimisation bottlenecks [23, 24].

However, Krylov subspace methods are often confronted with the obstacle that small errors can cause large deviations in the ground-state energy. This issue is rooted in the Krylov subspace that is spanned by an almost linearly dependent basis [25]. Contrary to classical computing, in which one can exponentially suppress rounding errors by increasing the number of bits, quantum computing is inherently subject to statistical error. Since statistical error decreases slowly with the measurement number MM as 1/M\propto 1/\sqrt{M}, an extremely large MM can be required to reach an acceptable error. In this case, although quantum KSD algorithms perform well in principle, the measurement cost has to be assessed and optimised for realistic implementations [16, 20].

In this work, we present a general and rigorous analysis of the measurement cost in quantum KSD algorithms. Specifically, we obtain an upper bound formula of the measurement number that is applicable to all quantum KSD algorithms. Then, we propose an algorithm to construct the Hamiltonian-power Krylov subspace [6]. In our algorithm, we express the product of Hamiltonian power and a Gaussian function of Hamiltonian as an integral of real-time evolution. In this way, the statistical error decreases exponentially with the power, which makes our algorithm measurement-efficient. We benchmark quantum KSD algorithms by estimating their measurement costs in solving the anti-ferromagnetic Heisenberg model and the Hubbard model. Various lattices of each model are taken in the benchmarking. It is shown that the measurement cost in our algorithm is typically orders of magnitude fewer than other algorithms to reach an error comparable to the classical Lanczos algorithm. We also demonstrate the measurement efficiency of our algorithm by composing a ground-state projector.

2 Krylov subspace diagonalisation

First, we introduce the KSD algorithm and some relevant notations. The algorithm starts with a reference state |φ|{\varphi}\rangle (or a set of reference states [10, 11]; we focus on the single-reference case in this work). Then we generate a set of basis states

|ϕk=fk(H)|φ,\displaystyle|{\phi_{k}}\rangle=f_{k}(H)|{\varphi}\rangle, (1)

where HH is the Hamiltonian, and f1,f2,,fdf_{1},f_{2},\ldots,f_{d} are linearly-independent (d1)(d-1)-degree polynomials (to generate a dd-dimensional Krylov subspace). For example, it is conventional to take the power (P) function fk(H)=Hk1f_{k}(H)=H^{k-1} in the Lanczos algorithm. These states span a subspace called the Krylov subspace. We compute the ground-state energy by solving the generalised eigenvalue problem

𝐇𝐚=E𝐒𝐚,\displaystyle\mathbf{H}\mathbf{a}=E\mathbf{S}\mathbf{a}, (2)

where 𝐇k,q=ϕk|H|ϕq\mathbf{H}_{k,q}=\langle{\phi_{k}}|H|{\phi_{q}}\rangle and 𝐒k,q=ϕk|ϕq\mathbf{S}_{k,q}=\langle\phi_{k}|\phi_{q}\rangle. Let EminE_{min} be the minimum eigenvalue of the generalised eigenvalue problem. The error in the ground-state energy is

ϵK=EminEg,\displaystyle\epsilon_{K}=E_{min}-E_{g}, (3)

where EgE_{g} is the true ground-state energy. We call ϵK\epsilon_{K} the subspace error. One can also construct generalised Krylov subspaces in which fkf_{k} are functions of HH other than polynomials; see Table 1. Appendix A provides a more detailed introduction to the KSD algorithm.

In the literature, subspaces generated by variational quantum circuits [26, 27, 28] and stochastic time evolution [29] are also proposed. Quantum KSD techniques can also be used to mitigate errors caused by imperfect gates [30, 31, 32]. In this work, we focus on Hamiltonian functions because of their similarities to conventional Krylov subspace methods.

Abbr. fk(x)f_{k}(x) Refs.
P xk1x^{k-1} [19, 12]
CP Tk1(x/htot)T_{k-1}(x/h_{tot}) [20]
GP xk1e12x2τ2x^{k-1}e^{-\frac{1}{2}x^{2}\tau^{2}} This work
IP x(k1)x^{-(k-1)} [18]
ITE eτ(k1)xe^{-\tau(k-1)x} [8, 9]
RTE eixΔt(kd+12)e^{-ix\Delta t\left(k-\frac{d+1}{2}\right)} [10, 11, 12, 13, 14, 15, 16, 17]
F L1l=1Lei[xΔE(k1)]Δt(lL+12)L^{-1}\sum_{l=1}^{L}e^{-i\left[x-\Delta E\left(k-1\right)\right]\Delta t\left(l-\frac{L+1}{2}\right)} [15]
Table 1: Operators fk(x)f_{k}(x) generating the basis of a (generalised) Krylov subspace. Here x=HE0x=H-E_{0}, where E0E_{0} is a constant. In different algorithms, fkf_{k} can be power (P), Chebyshev polynomial (CP), Gaussian-power (GP), inverse power (IP) or exponential [i.e. imaginary-time evolution (ITE), real-time evolution (RTE) and filter (F)] functions of the Hamiltonian. TnT_{n} is the nn-th Chebyshev polynomial of the first kind, and E0=0E_{0}=0 in the CP basis. For a Hamiltonian expressed in the form H=jhjσjH=\sum_{j}h_{j}\sigma_{j}, where σj\sigma_{j} are Pauli operators, htot=j|hj|h_{tot}=\sum_{j}|h_{j}| is the 1-norm of coefficients. τ\tau, Δt\Delta t and ΔE\Delta E are some real parameters.

3 Statistical error and regularisation

In addition to ϵK\epsilon_{K}, the other error source is the statistical error. In quantum KSD algorithms, matrices 𝐇\mathbf{H} and 𝐒\mathbf{S} are obtained by measuring qubits at the end of certain quantum circuits. Measurements yield estimators 𝐇^\hat{\mathbf{H}} and 𝐒^\hat{\mathbf{S}} of the two matrices, respectively. Errors in 𝐇^\hat{\mathbf{H}} and 𝐒^\hat{\mathbf{S}} depend on the measurement number. Suppose the budget for each complex matrix entry is 2M2M measurements (the budget for each part is MM). Variances of matrix entries have upper bounds in the form

Var(𝐇^k,q)2C𝐇2/M\displaystyle\mathrm{Var}(\hat{\mathbf{H}}_{k,q})\leqslant 2C_{\mathbf{H}}^{2}/M (4)

and

Var(𝐒^k,q)2C𝐒2/M,\displaystyle\mathrm{Var}(\hat{\mathbf{S}}_{k,q})\leqslant 2C_{\mathbf{S}}^{2}/M, (5)

where C𝐇C_{\mathbf{H}} and C𝐒C_{\mathbf{S}} are some factors depending on the measurement protocol. For example, suppose each entry of 𝐇\mathbf{H} and 𝐒\mathbf{S} can be expressed in the form sqsφ|Us|φ\sum_{s}q_{s}\langle{\varphi}|U_{s}|{\varphi}\rangle, where UsU_{s} are unitary operators, and we measure each term using the Hadamard test [33], then the variance of the entry has the upper bound 2C2/M2C^{2}/M, where C=s|qs|C=\sum_{s}|q_{s}|.

We quantify the error in a matrix with the spectral norm. The distributions of 𝐇^𝐇2\|\hat{\mathbf{H}}-\mathbf{H}\|_{2} and 𝐒^𝐒2\|\hat{\mathbf{S}}-\mathbf{S}\|_{2} depend on not only the measurement number but also the correlations between matrix entries. We can measure matrix entries independently, then their distributions are uncorrelated. However, certain matrix entries take the same value, and we can measure them collectively. For example, for the P basis, all entries 𝐇kq,q\mathbf{H}_{k-q,q} take the same value as 𝐇k1,1\mathbf{H}_{k-1,1}. We only need to measure one of them on the quantum computer, and then these entries become correlated. In Appendix B, we analyse the distributions of spectral norms for each basis in Table 1.

Errors in matrices cause an error in the ground-state energy in addition to ϵK\epsilon_{K}. It is common that the subspace basis is nearly linearly dependent, which makes the overlap matrix 𝐒\mathbf{S} ill-conditioned. Then, even small errors in matrices may cause a serious error in the ground-state energy. Thresholding is a method to overcome this problem [8, 16]. In this work, we consider the regularisation method; see Algorithm 1. An advantage of the regularisation method is that by taking a proper regularisation parameter η\eta, the resulting energy E^min\hat{E}_{min} is variational, i.e. E^minEg\hat{E}_{min}\geqslant E_{g} up to a controllable failure probability.

Protocol η\eta α\alpha β\beta Basis
IM (Chebyshev) 2d2Mκ\frac{2d^{2}}{\sqrt{M\kappa}} 256κ\frac{256}{\kappa} d6d^{6} All
IM (Hoeffding) 2d2Mln8d2κ\sqrt{\frac{2d^{2}}{M}\ln{\frac{8d^{2}}{\kappa}}} 128ln1κ128\ln\frac{1}{\kappa} d4d^{4}
CM (Real Hankel) 2dMln4dκ\sqrt{\frac{2d}{M}\ln{\frac{4d}{\kappa}}} 64ln1κ64\ln\frac{1}{\kappa} d(2d1)d(2d-1) P, GP, IP, ITE
CM (Real symmetric) 32ln1κ32\ln\frac{1}{\kappa} d2(d+1)d^{2}(d+1) CP, F
CM (Complex Toeplitz) 2(2d1)Mln4dκ\sqrt{\frac{2(2d-1)}{M}\ln{\frac{4d}{\kappa}}} 64ln1κ64\ln\frac{1}{\kappa} (2d1)2(2d-1)^{2} RTE
Table 2: The regularisation parameter η\eta, and measurement cost factors α(κ)\alpha(\kappa) and β(d)\beta(d). These quantities depend on the protocol for measuring matrix entries. Two types of protocols are considered: independent measurement (IM) and collective measurement (CM). Using the Chebyshev and Hoeffding inequalities, we obtain two different results (upper bounds) for the IM protocol. These factors are based on the protocols introduced in Appendices B and E. We remark that for CP, leveraging the characteristic of Chebyshev polynomials, the scaling with dd can be reduced to O(d2)O(d^{2}) [20].

We propose to choose the regularisation parameter η\eta according to the distributions of 𝐇^𝐇2\|\hat{\mathbf{H}}-\mathbf{H}\|_{2} and 𝐒^𝐒2\|\hat{\mathbf{S}}-\mathbf{S}\|_{2}. Let κ\kappa be the permissible failure probability. We take η\eta such that

Pr(𝐇^𝐇2C𝐇ηand𝐒^𝐒2C𝐒η)\displaystyle\Pr\left(\|\hat{\mathbf{H}}-\mathbf{H}\|_{2}\leqslant C_{\mathbf{H}}\eta\ \rm{and}\ \|\hat{\mathbf{S}}-\mathbf{S}\|_{2}\leqslant C_{\mathbf{S}}\eta\right) (6)
\displaystyle\geqslant 1κ.\displaystyle 1-\kappa.

In Table 2, we list η\eta satisfying the above condition for each basis. The value of η\eta depends on the failure probability κ\kappa and the measurement number MM. Taking η\eta in this way, an upper bound of the energy error is given in the following lemma.

Lemma 1.

Let

E(η,𝐚)=𝐚(𝐇+2C𝐇η)𝐚𝐚(𝐒+2C𝐒η)𝐚.\displaystyle E^{\prime}(\eta,\mathbf{a})=\frac{\mathbf{a}^{\dagger}(\mathbf{H}+2C_{\mathbf{H}}\eta)\mathbf{a}}{\mathbf{a}^{\dagger}(\mathbf{S}+2C_{\mathbf{S}}\eta)\mathbf{a}}. (7)

Under conditions Eg<0E_{g}<0 and min𝐚𝟎E(η,𝐚)<0\min_{\mathbf{a}\neq\bf{0}}E^{\prime}(\eta,\mathbf{a})<0, the following inequality holds,

Pr(EgE^minmin𝐚𝟎E(η,𝐚))1κ.\displaystyle\Pr\left(E_{g}\leqslant\hat{E}_{min}\leqslant\min_{\mathbf{a}\neq\bf{0}}E^{\prime}(\eta,\mathbf{a})\right)\geqslant 1-\kappa. (8)

See Appendix C for the proof. Notice that we can always subtract a sufficiently large positive constant from the Hamiltonian to satisfy the conditions. The upper bound of the energy error in the regularisation method is min𝐚𝟎E(η,𝐚)Eg\min_{\mathbf{a}\neq\bf{0}}E^{\prime}(\eta,\mathbf{a})-E_{g}, up to the failure probability.

1:
2:Input 𝐇^\hat{\mathbf{H}}, 𝐒^\hat{\mathbf{S}}, C𝐇C_{\mathbf{H}}, C𝐒C_{\mathbf{S}} and η>0\eta>0.
3:Solve the generalised eigenvalue problem
(𝐇^+C𝐇η)𝐚=E(𝐒^+C𝐒η)𝐚(\hat{\mathbf{H}}+C_{\mathbf{H}}\eta)\mathbf{a}=E(\hat{\mathbf{S}}+C_{\mathbf{S}}\eta)\mathbf{a}
4:Output the minimum eigenvalue E^min\hat{E}_{min}.
Algorithm 1 Regularised quantum Krylov-subspace diagonalisation algorithm.

4 Analysis of the measurement cost

In this work, we rigorously analyse the number of measurements. Let ϵ\epsilon be the target error in the ground-state energy. We will give a measurement number sufficient (i.e. an upper bound of the measurement number necessary) for achieving the target. For the regularisation method, we already have an upper bound of the error min𝐚𝟎E(η,𝐚)Eg\min_{\mathbf{a}\neq\bf{0}}E^{\prime}(\eta,\mathbf{a})-E_{g}. Therefore, we can derive such a measurement number by solving the equation

min𝐚𝟎E(η,𝐚)Eg=ϵ,\displaystyle\min_{\mathbf{a}\neq\bf{0}}E^{\prime}(\eta,\mathbf{a})-E_{g}=\epsilon, (9)

in which η\eta is the unknown. With the solution, we can work out the corresponding measurement number MM according to Table 2. If matrix entries are measured independently, the total measurement number is Mtot=4d2MM_{tot}=4d^{2}M; we have the following theorem.

Theorem 1.

Suppose ϵ>ϵK\epsilon>\epsilon_{K} and Eg+ϵ<0E_{g}+\epsilon<0. The total measurement number

Mtot=α(κ)β(d)16η2\displaystyle M_{tot}=\frac{\alpha(\kappa)\beta(d)}{16\eta^{2}} (10)

is sufficient for achieving the permissible error ϵ\epsilon and failure probability κ\kappa, i.e. E^min[Eg,Eg+ϵ]\hat{E}_{min}\in[E_{g},E_{g}+\epsilon] with a probability of at least 1κ1-\kappa. Here, α(κ)=256/κ\alpha(\kappa)=256/\kappa, β(d)=d6\beta(d)=d^{6}, and η\eta is the solution to Eq. (9).

See Appendix C for the proof. Similar to Lemma 1, we can always satisfy the condition Eg+ϵ<0E_{g}+\epsilon<0 by subtracting a sufficiently large positive constant from the Hamiltonian. The above theorem also holds (up to some approximations) for collective measurements if we replace functions α(κ)\alpha(\kappa) and β(d)\beta(d) according to Table 2.

4.1 Factoring the measurement cost

To compare different bases, we would like to divide the measurement number into three factors. If matrix entries are measured independently, the first two factors are the same for different bases of Krylov subspaces, and the third factor depends on the basis. We find that the third factor is related to the ill-conditioned problem, and we can drastically reduce it with our measurement-efficient algorithm.

We can always write the total measurement number Eq. (10) as a product of three factors,

Mtot=α(κ)H22pg2ϵ2×β(d)×γ.\displaystyle M_{tot}=\frac{\alpha(\kappa)\|H\|_{2}^{2}}{p_{g}^{2}\epsilon^{2}}\times\beta(d)\times\gamma. (11)

The first factor is the cost of measuring the energy. Roughly speaking, it is the cost in the ideal case that one can prepare the ground state by an ideal projection; see Appendix D. The spectral norm H2\|H\|_{2} characterises the range of the energy. Assume that the statistical error and ϵ\epsilon are comparable, MtotH22/ϵ2M_{tot}\propto\|H\|_{2}^{2}/\epsilon^{2} according to the standard scaling of the statistical error. α(κ)\alpha(\kappa) is the overhead for achieving the permissible failure probability κ\kappa. The success of KSD algorithms depends on a finite overlap between the reference state and the true ground state |ψg|{\psi_{g}}\rangle [6, 16, 20]. This requirement is reflected in Mtot1/pg2M_{tot}\propto 1/p_{g}^{2}, where pg=|ψg|φ|2p_{g}=|\langle\psi_{g}|\varphi\rangle|^{2}. The second factor β(d)\beta(d) is the overhead due to measuring d×dd\times d matrices. There are more sources of statistical errors (matrix entries) influencing the eventual result when dd is larger. The remaining factor

γ=pg2ϵ216H22η2\displaystyle\gamma=\frac{p_{g}^{2}\epsilon^{2}}{16\|H\|_{2}^{2}\eta^{2}} (12)

is related to the spectrum of 𝐒\mathbf{S} (i.e. the basis), as we will show in Section 7. Notice that this expression of γ\gamma holds for all the bases. In the numerical result, we find that the typical value of γ\gamma for certain bases can be as large as 101210^{12} to achieve certain ϵ\epsilon. We propose the Gaussian-power (GP) basis (see Table 1), and it can reduce γ\gamma to about 44.

5 Measurement-efficient algorithm

As we have shown, the measurement overhead γ\gamma depends on the overlap matrix 𝐒\mathbf{S}, i.e. how we choose the basis of Krylov subspace. The main advantage of our algorithm is that we utilise a basis resulting in a small γ\gamma.

To generate a standard Krylov subspace, we choose operators

fk=(HE0)k1e12(HE0)2τ2,\displaystyle f_{k}=(H-E_{0})^{k-1}e^{-\frac{1}{2}(H-E_{0})^{2}\tau^{2}}, (13)

where E0E_{0} is a constant up to choice. We call it the GP basis. The corresponding subspace is a conventional Hamiltonian-power Krylov subspace, but the reference state |φ|{\varphi}\rangle has been effectively replaced by e12(HE0)2τ2|φe^{-\frac{1}{2}(H-E_{0})^{2}\tau^{2}}|{\varphi}\rangle.

The spectral norm of fkf_{k} operators for the GP basis [defined in Eq. (13)] has the upper bound

fk2(k1eτ2)k12,\displaystyle\|f_{k}\|_{2}\leqslant(\frac{k-1}{e\tau^{2}})^{\frac{k-1}{2}}, (14)

where ee is Euler’s constant. This upper bound is proven in Lemma 9 in Appendix F. Under the condition eτ2>d1e\tau^{2}>d-1, the norm upper bound (k1eτ2)k12(\frac{k-1}{e\tau^{2}})^{\frac{k-1}{2}} decreasing exponentially with kk. This property is essential for the measurement efficiency of our algorithm.

We propose to realise the Gaussian-power basis by expressing the operator of interest as a linear combination of unitaries (LCU). The same approach has been taken to realise other bases like power [19, 12], inverse power [18], imaginary-time evolution [34], filter [15] and Gaussian function of the Hamiltonian [35].

Suppose we want to measure the quantity φ|A|φ\langle{\varphi}|A|{\varphi}\rangle. Here A=fkHfqA=f_{k}^{\dagger}Hf_{q} and A=fkfqA=f_{k}^{\dagger}f_{q} for 𝐇k,q\mathbf{H}_{k,q} and 𝐒k,q\mathbf{S}_{k,q}, respectively. First, we work out an expression in the from A=sqsUsA=\sum_{s}q_{s}U_{s}, where UsU_{s} are unitary operators, and qsq_{s} are complex coefficients. Then, there are two ways to measure φ|A|φ\langle{\varphi}|A|{\varphi}\rangle. When the expression has finite terms, we can apply AA on the state |φ|{\varphi}\rangle by using a set of ancilla qubits; In the literature, LCU usually refers to this approach [36]. Alternatively, we can evaluate the summation φ|A|φ\langle{\varphi}|A|{\varphi}\rangle using the Monte Carlo method [37] by averaging Hadamard-test [33] estimates of randomly selected terms φ|Us|φ\langle{\varphi}|U_{s}|{\varphi}\rangle. The second way works for an expression with even infinite terms and requires only one or zero ancilla qubits [38, 39]. We focus on the Monte Carlo method in this work.

Now we give our LCU expression. Suppose we already express HH as a linear combination of Pauli operators, it is straightforward to work out the LCU expression of AA given the expression of fkf_{k}. Therefore, we only give the expression of fkf_{k}. Utilising the Fourier transformation of a modified Hermite function (which is proved in Appendix F), we can express fkf_{k} in Eq. (13) as

fk\displaystyle f_{k} =\displaystyle= ik12k12τk1+𝑑tHk1(t2τ)gτ(t)\displaystyle\frac{i^{k-1}}{2^{\frac{k-1}{2}}\tau^{k-1}}\int_{-\infty}^{+\infty}dtH_{k-1}\left(\frac{t}{\sqrt{2}\tau}\right)g_{\tau}(t) (15)
×eixt,\displaystyle\times e^{-ixt},

where Hn(u)H_{n}(u) denotes Hermite polynomials [40], gτ(t)=1τ2πet22τ2g_{\tau}(t)=\frac{1}{\tau\sqrt{2\pi}}e^{-\frac{t^{2}}{2\tau^{2}}} is the normalised Gaussian function, x=HE0x=H-E_{0}, and eixte^{-ixt} is the real-time evolution (RTE) operator. Notice that in Eq. (15) fkf_{k} is expressed as a linear combination of RTE operators eiHte^{-iHt}, and the summation in conventional LCU is replaced with an integral over tt. See Appendix G for how to evaluate this integral using the Monte Carlo method.

5.1 Real-time evolution

In order to implement the LCU expression Eq. (15), we need to implement the RTE eiHt|φe^{-iHt}|{\varphi}\rangle. There are various protocols for implementing RTE, including Trotterisation [41, 42, 43], LCU [36, 44, 45, 37] and qDRIFT [46], etc. In most of the protocols, RTE is inexact but approaches the exact one when the circuit depth increases. Therefore, all these protocols work for our algorithm as long as the circuit is sufficiently deep. In this work, we specifically consider an LCU-type protocol, the zeroth-order leading-order-rotation formula [47].

In the leading-order-rotation protocol, RTE is exact even when the circuit depth is finite. In this protocol, we express RTE in the LCU form

eiHt=[rvr(t/N)Vr(t/N)]N,\displaystyle e^{-iHt}=\left[\sum_{r}v_{r}(t/N)V_{r}(t/N)\right]^{N}, (16)

where Vr(t)=eiϕ(t)σ,σV_{r}(t)=e^{-i\phi(t)\sigma},\sigma is a rotation or Pauli operator, vr(t)v_{r}(t) are complex coefficients, and NN is the time step number. This exact expression of RTE is worked out in the spirit of Taylor expansion and contains infinite terms. Notice that NN determines the circuit depth. For details of the zeroth-order leading-order-rotation formula, see Ref. [47] or Appendix G in this paper, where Eq. (80) gives the specific form of Eq. (16). Substituting Eq. (16) into Eq. (15), we obtain the eventual LCU expression of fkf_{k}, a concatenation of two LCU expressions, i.e.

fk\displaystyle f_{k} =\displaystyle= ik12k12τk1+𝑑tHk1(t2τ)gτ(t)eiE0t\displaystyle\frac{i^{k-1}}{2^{\frac{k-1}{2}}\tau^{k-1}}\int_{-\infty}^{+\infty}dtH_{k-1}\left(\frac{t}{\sqrt{2}\tau}\right)g_{\tau}(t)e^{iE_{0}t} (17)
×[rvr(t/N)Vr(t/N)]N.\displaystyle\times\left[\sum_{r}v_{r}(t/N)V_{r}(t/N)\right]^{N}.

With the above expression, we have written fkf_{k} as the linear combination of Vr(t/N)V_{r}(t/N). Therefore, we can use the Monte Carlo method to realise fkf_{k}: We sample tt and rr according to the coefficients, then we implement the corresponding Vr(t/N)V_{r}(t/N) on the quantum computer.

5.2 Bias and variance

Let A^\hat{A} be the estimator of φ|A|φ\langle{\varphi}|A|{\varphi}\rangle. There are two types of errors in A^\hat{A}, bias and variance. Because RTE is exact in the leading-order-rotation protocol, the estimator A^\hat{A} is unbiased. Therefore, we can focus on the variance from now on.

If we use the Monte Carlo method and the one-ancilla Hadamard test to evaluate the LCU expression of AA, the variance has the upper bound

Var(A^)2CA2M,\displaystyle\mathrm{Var}(\hat{A})\leqslant\frac{2C_{A}^{2}}{M}, (18)

where 2M2M is the measurement number, and CA=s|qs|C_{A}=\sum_{s}|q_{s}| is the 1-norm of coefficients in A=sqsUsA=\sum_{s}q_{s}U_{s}. Here, AA is a summation of unitary operators. The generalisation to integral is straightforward. We call CAC_{A} the cost of the LCU expression. We remark that the variance in this form is universal and valid for all algorithms with some proper factor CAC_{A}.

The cost of an LCU expression is related to the spectral norm. Notice Us2=1\|U_{s}\|_{2}=1, we immediately conclude that CAA2C_{A}\geqslant\|A\|_{2}. Therefore, the best we can achieve is a carefully designed LCU expression that CAC_{A} is as close as possible to A2\|A\|_{2}. Only when fk2\|f_{k}\|_{2} decreases exponentially with kk, it is possible to measure φ|A|φ\langle{\varphi}|A|{\varphi}\rangle with an error decreasing exponentially with kk.

In our algorithm, we find that the cost has the form

CA=htotckcq and ckcq\displaystyle C_{A}=h_{tot}c_{k}c_{q}\text{ and }c_{k}c_{q} (19)

for 𝐇k,q\mathbf{H}_{k,q} and 𝐒k,q\mathbf{S}_{k,q}, respectively. ckc_{k} is the cost due to fkf_{k}, and

ck\displaystyle c_{k} =\displaystyle= 12k12τk1+𝑑t|Hk1(t2τ)|gτ(t)\displaystyle\frac{1}{2^{\frac{k-1}{2}}\tau^{k-1}}\int_{-\infty}^{+\infty}dt\left|H_{k-1}\left(\frac{t}{\sqrt{2}\tau}\right)\right|g_{\tau}(t) (20)
×[r|vr(tN)|]N2(k1eτ2)k12.\displaystyle\times\left[\sum_{r}\left|v_{r}\left(\frac{t}{N}\right)\right|\right]^{N}\leqslant 2\left(\frac{k-1}{e\tau^{2}}\right)^{\frac{k-1}{2}}.

This upper bound holds when N4ehtot2τ2N\geqslant 4eh_{tot}^{2}\tau^{2}, which is a requirement on the circuit depth NN. The proof is in Appendix G. With the upper bound, we can find that we already approach the lower bound of the variance given by the spectral norm (The difference is a factor of 22). As a result, the variance of A^\hat{A} decreases exponentially with kk and qq.

Detailed pseudocodes and analysis are given in Appendix G. There are two ways to apply Theorem 1: First, we can take C𝐇=htotC𝐒C_{\mathbf{H}}=h_{tot}C_{\mathbf{S}} and C𝐒=max{ckcq}C_{\mathbf{S}}=\max\{c_{k}c_{q}\}; Second, we can rescale the basis by taking fk=fk/ckf^{\prime}_{k}=f_{k}/c_{k}. Matrix entries 𝐇^k,q\hat{\mathbf{H}}^{\prime}_{k,q} and 𝐒^k,q\hat{\mathbf{S}}^{\prime}_{k,q} of the rescaled basis {fk|φ}\{f^{\prime}_{k}|{\varphi}\rangle\} have variance upper bounds 2htot2/M2h_{tot}^{2}/M and 2/M2/M, respectively. Therefore, C𝐇=htotC_{\mathbf{H}}=h_{tot} and C𝐒=1C_{\mathbf{S}}=1 for the rescaled basis. We take the rescaled basis in the numerical study.

6 Measurement overhead benchmarking

With Theorem 1, we can benchmark the measurement cost in quantum KSD algorithms listed in Table 1. We focus on the overhead factor γ\gamma, while ignoring the slight differences in the other two factors in different algorithms (See Table 2). Given a target error ϵ\epsilon, first we work out the solution η\eta to Eq. (9), then γ\gamma is calculated according to Eq. (12).

Two models of strongly correlated systems, namely, the anti-ferromagnetic Heisenberg model and Hubbard model [7, 48, 49, 50], are used in benchmarking. For each model, the lattice is taken from two categories: regular lattices, including chain and ladder, and randomly generated graphs. We fix the lattice size such that the Hamiltonian can be encoded into ten qubits. For each Hamiltonian specified by the model and lattice, we evaluate all quantum KSD algorithms with the same subspace dimension dd taken from 2,3,,302,3,\ldots,30. In total, 233233 instances of (model,lattice,d)(model,lattice,d) are generated to benchmark quantum KSD algorithms. The details of the numerical calculations including parameters taken in different algorithms are given in Appendix H.

Fig. 1 illustrates the result of the instance (Heisenberg,chain,d=5)(\mathrm{Heisenberg},\mathrm{chain},d=5), for example. We can find that the error decreases with the cost overhead. When the target error is relatively high, algorithms have similar performance, e.g. when ϵ0.01\epsilon\approx 0.01, the cost γ10\gamma\approx 10 is sufficient in most algorithms. However, in comparison with others, the error in the GP algorithm decreases much faster. When dd is sufficiently large, MM required to achieve the error ϵ\epsilon always follows a scaling of 1/ϵ21/\epsilon^{2}. However, the size of dd at which different algorithms converge to this behavior varies; see Fig. 9. The GP algorithm converges with a smaller dd compared to other algorithms. In this instance, the GP algorithm has already converged when d=5d=5, whereas the other algorithms have not yet converged, resulting in poorer performance.

Refer to caption
Figure 1: Comparison between quantum KSD algorithms listed in Table 1. Here, we take the instance (Heisenberg,chain,d=5)(\mathrm{Heisenberg},\mathrm{chain},d=5) as an example. ϵK\epsilon_{K} is the subspace error of the P algorithm. The grey area illustrates the range of γ\gamma when we take E0[Eg0.1,Eg+0.1]E_{0}\in[E_{g}-0.1,E_{g}+0.1] in the GP algorithm. Notice that H2=1\|H\|_{2}=1. The blue curve represents the result of the GP algorithm with E0=EgE_{0}=E_{g}. Similar results are obtained when other measures are used for benchmarking, including the measurement number MM computed with the optimal α\alpha and β\beta factors, the necessary measurement cost (instead of its upper bound) estimated numerically and the necessary cost when using the thresholding method; see Appendix L.

To confirm this advantage of the GP algorithm, we implement the simulation for a number of instances. We choose the power algorithm as the standard for comparison. We remark that the power algorithm and the classical Lanczos algorithm yield the same result when the implementation is error-free (i.e. without statistical error and rounding error), and the eventual error in this ideal case is the subspace error ϵK\epsilon_{K}. Given the model, lattice and subspace dimension, we compute ϵK\epsilon_{K} of the power algorithm. Then, we take the permissible error ϵ=2ϵK\epsilon=2\epsilon_{K} and compute the overhead factor γ\gamma for each algorithm. Then, the empirical distribution of γ\gamma is shown in Fig. 2.

Refer to caption
Figure 2: Empirical distribution of the measurement overhead γ\gamma for algorithms listed in Table 1. In the Gaussian-power algorithm, we take a random E0E_{0} in the interval [Eg0.1H2,Eg+0.1H2][E_{g}-0.1\|H\|_{2},E_{g}+0.1\|H\|_{2}], i.e. we assume that we have a preliminary estimation of the ground-state energy with an uncertainty as large as 10%10\% of the entire spectrum.

From the distribution, we can find that our Gaussian-power algorithm has the smallest measurement overhead, and γ\gamma is smaller than 10210^{2} for almost all instances. The filter (F) algorithm is the most well-performed one among others. Part of the reason is that we have taken the optimal parameter found by grid search, of which the details are discussed in Appendix H. The median value of γ\gamma is about 44 for the GP algorithm, 1313 for the F algorithm and as large as 101210^{12} for some other algorithms. Comparing the GP and F algorithms, although the median values are similar, the cost overhead is larger than 10410^{4} in more than 20%20\% instances in the F algorithm, which never happens in the GP algorithm.

The performance of the GP algorithm depends on the choice of parameters: τ\tau, E0E_{0} and η\eta. The regularisation parameter η\eta is taken according to Table 2. To choose the value of E0E_{0}, we need prior knowledge about the true ground-state energy EgE_{g}, which can be obtained from certain classical algorithms computing the ground state or the quantum KSD algorithm taking a smaller subspace dimension. Although we prefer an E0E_{0} close to EgE_{g}, the numerical result suggests that a relatively large difference |E0Eg||E_{0}-E_{g}| (10%10\% of the entire spectrum) is tolerable. The Gaussian factor in the GP basis acts as a filter on the spectrum. Therefore, if |E0Eg||E_{0}-E_{g}| is finite and we take an over large τ\tau, we may have a vanishing overlap between the effective reference state e12(HE0)2τ2|φe^{-\frac{1}{2}(H-E_{0})^{2}\tau^{2}}|{\varphi}\rangle and the ground state. To avoid this problem, we need to assume an upper bound |E0Eg|ϵ0|E_{0}-E_{g}|\leqslant\epsilon_{0}, then we suggest taking τ\tau in the range d1e<τd1ϵ0\sqrt{\frac{d-1}{e}}<\tau\leqslant\frac{\sqrt{d-1}}{\epsilon_{0}}, see Appendix K. Although we take the same τ\tau for all fkf_{k} in each instance in the benchmarking, one can flexibly choose different τ\tau in practice.

7 Composing a projector

We show that γ\gamma is related to the spectrum of 𝐒\mathbf{S}. Let’s consider a small η\eta and the Taylor expansion of EE^{\prime}. The minimum value of the zeroth-order term 𝐚𝐇𝐚𝐚𝐒𝐚\frac{\mathbf{a}^{\dagger}\mathbf{H}\mathbf{a}}{\mathbf{a}^{\dagger}\mathbf{S}\mathbf{a}} is EminE_{min} according to the Rayleigh quotient theorem [51]. Take this minimum value, we have EEmin+sηE^{\prime}\simeq E_{min}+s\eta, where s𝐚𝐚𝐚𝐒𝐚s\propto\frac{\mathbf{a}^{\dagger}\mathbf{a}}{\mathbf{a}^{\dagger}\mathbf{S}\mathbf{a}} is the first derivative. The solution to E=Eg+ϵE^{\prime}=E_{g}+\epsilon is η(Eg+ϵEmin)/s\eta\simeq(E_{g}+\epsilon-E_{min})/s, then

γu2(pg𝐚𝐚𝐚𝐒𝐚)2,\displaystyle\gamma\simeq u^{2}\left(\frac{p_{g}\mathbf{a}^{\dagger}\mathbf{a}}{\mathbf{a}^{\dagger}\mathbf{S}\mathbf{a}}\right)^{2}, (21)

where u(C𝐇+H2C𝐒)ϵ2H2(ϵϵK)C𝐒u\leqslant\frac{(C_{\mathbf{H}}+\|H\|_{2}C_{\mathbf{S}})\epsilon}{2\|H\|_{2}(\epsilon-\epsilon_{K})}\approx C_{\mathbf{S}} under assumptions C𝐇H2C𝐒C_{\mathbf{H}}\approx\|H\|_{2}C_{\mathbf{S}} and ϵϵKϵ\epsilon-\epsilon_{K}\approx\epsilon. Eq. (21) is derived in Appendix I. Therefore, when 𝐒\mathbf{S} is close to singular, the overhead can be large.

We can explain the measurement efficiency of our algorithm by composing a projector. In our algorithm with the rescaled basis, the solution is a state in the form |Ψ(𝐚)=k=1dakck1fk|φ|{\Psi(\mathbf{a})}\rangle=\sum_{k=1}^{d}a_{k}c_{k}^{-1}f_{k}|{\varphi}\rangle. Ideally, the linear combination of fkf_{k} realises a projector onto the ground state, i.e. k=1dakck1fk=|ψgψg|\sum_{k=1}^{d}a_{k}c_{k}^{-1}f_{k}=|\psi_{g}\rangle\langle\psi_{g}|. Then, |Ψ(𝐚)=pg|ψg|{\Psi(\mathbf{a})}\rangle=\sqrt{p_{g}}|{\psi_{g}}\rangle up to a phase and 𝐚𝐒𝐚=Ψ(𝐚)|Ψ(𝐚)=pg\mathbf{a}^{\dagger}\mathbf{S}\mathbf{a}=\langle\Psi(\mathbf{a})|\Psi(\mathbf{a})\rangle=p_{g}. Then γ(pg𝐚𝐚𝐚𝐒𝐚)2=(𝐚𝐚)2\gamma\approx\left(\frac{p_{g}\mathbf{a}^{\dagger}\mathbf{a}}{\mathbf{a}^{\dagger}\mathbf{S}\mathbf{a}}\right)^{2}=\left(\mathbf{a}^{\dagger}\mathbf{a}\right)^{2} (notice that C𝐒=1C_{\mathbf{S}}=1), we can find that 𝐚𝐚\mathbf{a}^{\dagger}\mathbf{a} determines γ\gamma. When ckc_{k} is smaller, |ak||a_{k}| required for composing the projector is smaller. In our algorithm, ckc_{k} decreases exponentially with kk, and the speed of decreasing is controllable via the parameter τ\tau. In this way, our algorithm results in a small γ\gamma.

To be concrete, let’s consider a classic way of composing a projector from Hamiltonian powers using Chebyshev polynomials (CPs) TnT_{n}, which has been used for proving the convergence of the Lanczos algorithm [52], RTE and CP quantum KSD algorithms [16, 20]. Without loss of generality, we suppose H2=1\|H\|_{2}=1. An approximate projector in the form of Chebyshev polynomials is

Tn(Z)Tn(z1)=|ψgψg|+Ω,\displaystyle\frac{T_{n}(Z)}{T_{n}(z_{1})}=|\psi_{g}\rangle\langle\psi_{g}|+\Omega, (22)

where Z=1(HEgΔ)Z=1-(H-E_{g}-\Delta), Δ\Delta is the energy gap between the ground state and the first excited state, z1=1+Δz_{1}=1+\Delta, and Ω\Omega is an operator with the upper bound Ω22/(z1n+z1n)\|\Omega\|_{2}\leqslant 2/(z_{1}^{n}+z_{1}^{-n}). Notice that the error Ω\Omega depends on nn, and its upper bound decreases exponentially with nn if Δ\Delta is finite. For simplicity, we focus on the case that E0=EgE_{0}=E_{g} in the Hamiltonian power. Then, in the expansion Tn(Z)=l=0nbl(HEg)lT_{n}(Z)=\sum_{l=0}^{n}b_{l}(H-E_{g})^{l}, coefficients have upper bounds |bl|nlTn(z1)|b_{l}|\leqslant n^{l}T_{n}(z_{1}) increasing exponentially with ll. See Appendix J. In the LCU and Monte Carlo approach, large coefficients lead to large variance. Therefore, it is difficult to realise this projector because of the exponentially increasing coefficients.

In our algorithm, the exponentially decreasing ckc_{k} can cancel out the large blb_{l}. We can compose the CP projector with an additional Gaussian factor. Taking d=n+1d=n+1 and ak=ckbk1/Tn(z1)a_{k}=c_{k}b_{k-1}/T_{n}(z_{1}), we have k=1dakck1fk=Tn(Z)Tn(z1)e12(HEg)2τ2\sum_{k=1}^{d}a_{k}c_{k}^{-1}f_{k}=\frac{T_{n}(Z)}{T_{n}(z_{1})}e^{-\frac{1}{2}(H-E_{g})^{2}\tau^{2}}. Because the Gaussian factor is also a projector onto the ground state, the overall operator is a projector with a smaller error than the CP projector. The corresponding overhead factor is

γ411n3/(eτ2).\displaystyle\gamma\lesssim 4\frac{1}{1-n^{3}/(e\tau^{2})}. (23)

When τ\tau is sufficiently large, γ4\gamma\lesssim 4. This example shows that the Gaussian-power basis can compose a projector with a small measurement cost.

By composing a projector with an explicit expression, we can obtain the worst-case performance of the algorithm. Similar analyses are reported in Refs. [16] and [20] for RTE and CP algorithms, respectively. Because the projector is approximate, it results in a finite error in the ground-state energy ϵP\epsilon_{P}. By solving the generalised eigenvalue problem, we can usually work out a better solution, i.e. the corresponding energy error is lower than ϵP\epsilon_{P}. Our result about the measurement cost is applicable to a target error that is not limited by ϵP\epsilon_{P}. Theorem 1 holds even when ϵ<ϵP\epsilon<\epsilon_{P}, which is beyond the explicit projector analysis.

The method of constructing polynomial projection operators mentioned above can be applied not only to the GP algorithm but also to the P algorithm. Since the variance of the P algorithm does not decrease exponentially with kk, it can be seen that our algorithm has an advantage over the P algorithm in this regard. We remark that for this particular projector, the decomposition into Hamiltonian powers is costly, and the existence of a better decomposition is possible. Compared to other algorithms, we have provided a numerical benchmark test in the previous section. Additionally, for the CP and RTE algorithms, it is possible to construct approximate projection operators with constant bounded expansion coefficients [16, 20]. In this case, further methods are needed to theoretically explain the improvement in measurement cost of our algorithm compared to these algorithms.

8 Discussions and Conclusions

We can apply Theorem 1 to estimate the measurement cost of various quantum KSD algorithms for arbitrary target error ϵ\epsilon and subspace dimension dd. In this paper, we focus on a target error comparable to the Lanczos algorithm at the same dd, i.e. ϵϵK\epsilon\sim\epsilon_{K}. The Lanczos algorithm is a well-known classical algorithm. It is natural to demand quantum KSD algorithms, as quantum counterparts of Lanczos algorithm, to reach a comparable target error. In this case, the measurement cost of the GP algorithm is observed much smaller than the others in the benchmarking.

In certain cases of practical applications, there is definite target error epsilon, e.g. the chemical accuracy. In these cases, these target errors are no longer depending on ϵK\epsilon_{K}. According to the Kaniel–Paige convergence theory [52], by increasing the subspace dimension dd and measurement number MM, we can always achieve arbitrary finite target error. Our results focus on relatively small d, in which case the GP algorithm requires a much smaller MM than the others, as illustrated in Fig. 2. We show in Fig. 9 that MM converges to O(1/ϵ2)O(1/\epsilon^{2}) when dd is sufficiently large, as demonstrated in Refs. [20, 53]. Increasing dd can weaken the advantage of the GP algorithm in the measurement cost. However, increasing dd not only brings more measurement entries (i.e. a larger factor β\beta), but also implies deeper quantum circuits and more gate noises.

In conclusion, we have proposed a regularised estimator of the ground-state energy (see Algorithm 1). Even in the presence of statistical errors, the regularised estimator is variational (i.e. equal to or above the true ground-state energy) and has a rigorous upper bound. Because of these properties, it provides a universal way of analysing the measurement cost in quantum KSD algorithms, with the result summarised in Theorem 1. This approach can be generalised to analyse the effect of other errors, e.g. imperfect quantum gates. The robustness to statistical errors in our algorithm implies the robustness to other errors.

To minimise the measurement cost, we have proposed a protocol for constructing Hamiltonian power with an additional Gaussian factor. This Gaussian factor is important because it leads to the statistical error decreasing exponentially with the power. Consequently, we can construct a Chebyshev-type projector onto the ground state with a small measurement cost. We benchmark quantum KSD algorithms with two models of quantum many-body systems. We find that our algorithm requires the smallest measurement number, and a measurement overhead γ\gamma of a hundred is almost always sufficient for approaching the theoretical-limit accuracy ϵK\epsilon_{K} of the Lanczos algorithm. In addition to the advantage over the classical Lanczos algorithm in scalability, our result suggests that the quantum algorithm is also competitive in accuracy at a reasonable measurement cost.

This paper primarily focuses on the ground-state problem. Notably, our quantum KSD algorithm holds promise for application to low-lying excited states as well. We leave it for future research.

YL thanks Xiaoting Wang and Zhenyu Cai for the helpful discussions. This work is supported by the National Natural Science Foundation of China (Grants No. 12225507 and No. 12088101), Innovation Program for Quantum Science and Technology (Grant No. 2023ZD0300200), and NSAF (Grants No. U2330201 and No. U2330401). The source codes for the numerical simulation are available at GitHub [54].

Appendix A General formalism of KSD algorithm

Given a Hamiltonian HH, a reference state |φ|{\varphi}\rangle and an integer dd, the standard Krylov subspace is 𝒦=Span(|φ,H|φ,H2|φ,,Hd1|φ)\mathcal{K}=\mathrm{Span}\left(|{\varphi}\rangle,H|{\varphi}\rangle,H^{2}|{\varphi}\rangle,\ldots,H^{d-1}|{\varphi}\rangle\right). This can be seen as taking polynomials fk(x)=xk1f_{k}(x)=x^{k-1}. The subspace is the same when {f1,f2,,fd}\{f_{1},f_{2},\ldots,f_{d}\} is an arbitrary set of linearly-independent (d1)(d-1)-degree polynomials. Any state in the Krylov subspace can be expressed as

|ψK\displaystyle|{\psi_{K}}\rangle =\displaystyle= k=1dakfk(H)|φ.\displaystyle\sum_{k=1}^{d}a_{k}f_{k}(H)|{\varphi}\rangle. (24)

The state that minimises the Rayleigh quotient ψK|H|ψK/ψK|ψK\langle\psi_{K}|H|\psi_{K}\rangle/\langle\psi_{K}|\psi_{K}\rangle is regarded as the approximate ground state of the Hamiltonian HH, and the corresponding energy is the approximate ground-state energy, i.e.

|ψg\displaystyle|{\psi_{g}}\rangle \displaystyle\sim |ψmin=argminψK𝒦ψK|H|ψKψK|ψK,\displaystyle|{\psi_{min}}\rangle=\operatorname*{arg\,min}_{\psi_{K}\in\mathcal{K}}\frac{\langle\psi_{K}|H|\psi_{K}\rangle}{\langle\psi_{K}|\psi_{K}\rangle}, (25)
Eg\displaystyle E_{g} \displaystyle\sim Emin=minψK𝒦ψK|H|ψKψK|ψK.\displaystyle E_{min}=\min_{\psi_{K}\in\mathcal{K}}\frac{\langle\psi_{K}|H|\psi_{K}\rangle}{\langle\psi_{K}|\psi_{K}\rangle}. (26)
Definition 1 (Rayleigh quotient).

For any matrix 𝐀d×d\mathbf{A}\in\mathbb{C}^{d\times d}, and any non-zero vector 𝐚d\mathbf{a}\in\mathbb{C}^{d}, the Rayleigh quotient is defined as

𝐀(𝐚)\displaystyle\langle\mathbf{A}\rangle(\mathbf{a}) =\displaystyle= 𝐚𝐀𝐚𝐚𝐚.\displaystyle\frac{\mathbf{a}^{\dagger}\mathbf{A}\mathbf{a}}{\mathbf{a}^{\dagger}\mathbf{a}}. (27)

The approximate ground state energy can be rewritten as

Emin\displaystyle E_{min} =\displaystyle= min𝐚0𝐚𝐇𝐚𝐚𝐒𝐚=min𝐚0𝐇(𝐚)𝐒(𝐚),\displaystyle\min_{\mathbf{a}\neq 0}\frac{\mathbf{a}^{\dagger}\mathbf{H}\mathbf{a}}{\mathbf{a}^{\dagger}\mathbf{S}\mathbf{a}}=\min_{\mathbf{a}\neq 0}\frac{\langle\mathbf{H}\rangle(\mathbf{a})}{\langle\mathbf{S}\rangle(\mathbf{a})}, (28)

where 𝐇\mathbf{H} and 𝐒\mathbf{S} are d×dd\times d Hermitian matrices as defined in the main text. Notice that EminE_{min} is the minimum eigenvalue of Eq. (2).

When kk increases, Hk|φH^{k}|{\varphi}\rangle converges to the eigenstate (having non-zero overlap with |φ|{\varphi}\rangle) with the largest absolute eigenvalue. The eigenstate is usually the ground state (If not, take HHE0H\leftarrow H-E_{0} where E0E_{0} is a positive constant). Therefore, it is justified to express the ground state as a linear combination of {Hk|φ}\{H^{k}|{\varphi}\rangle\} as long as |φ|{\varphi}\rangle has a non-zero overlap with the true ground state |ψg|{\psi_{g}}\rangle. However, the convergence with kk causes inherent trouble that basis vectors of the Krylov subspace are nearly linearly dependent. Then, the overlap matrix 𝐒\mathbf{S}, which is a Gram matrix, becomes nearly singular and has a large condition number. As a result, the denominator of the Rayleigh quotient can be very small, and a tiny error in computing may cause a large deviation of the approximate ground-state energy EminE_{min}.

According to the Rayleigh-Ritz theorem (Rayleigh quotient theorem) [51], the minimisation problem is equivalent to the generalised eigenvalue problem Eq. (2). The generalised eigenvalue problem can be reduced to an eigenvalue problem while the singularity issue remains. For example, one can use the Cholesky decomposition to express 𝐒\mathbf{S} as the product of an upper-triangular matrix and its conjugate transpose, i.e. 𝐒=𝐑𝐑\mathbf{S}=\mathbf{R}^{\dagger}\mathbf{R}, and the eigenvalue problem becomes (𝐑)1𝐇𝐑1(𝐑𝐚)=E𝐑𝐚\left(\mathbf{R}^{\dagger}\right)^{-1}\mathbf{H}\mathbf{R}^{-1}(\mathbf{R}\mathbf{a})=E\mathbf{R}\mathbf{a}. The Cholesky decomposition, however, also requires the matrix 𝐒\mathbf{S} to be positive-definite. When 𝐒\mathbf{S} is nearly singular, the Cholesky decomposition is unstable. The QZ algorithm is a more numerically stable method, and it is built into most standard programming languages. One way to address the singularity issue is by adding a positive diagonal matrix to 𝐒\mathbf{S}. In this work, we also add a diagonal matrix to 𝐇\mathbf{H} to ensure that the Rayleigh quotient is variational, i.e. the energy is not lower than the ground-state energy (with a controllable probability). An alternative way to address the singularity issue is the so-called thresholding method, see Refs. [8, 16].

Appendix B Norm distributions of noise matrices

In this section, we analyse the distributions of 𝐇^𝐇2\|\hat{\mathbf{H}}-\mathbf{H}\|_{2} and 𝐒^𝐒2\|\hat{\mathbf{S}}-\mathbf{S}\|_{2}. For the two matrices, there are a total of 2d22d^{2} matrix entries. We can measure each of them independently on the quantum computer. Under this independent measurement protocol, we obtain the most general results that apply to all the bases in Table 1, which are summarised in Lemma 2 and Lemma 3. In the 2d22d^{2} matrix entries, some of them are equivalent. To reduce the measurement cost, we can only measure one entry in each set of equivalent matrix entries, and we call it the collective measurement protocol. Because the equivalent-entry sets depend on the basis, the corresponding distributions are basis-dependent. The results are summarised in Lemma 4 and Lemma 5; we remark that these results are obtained under the assumption that statistical noise is Gaussian.

Lemma 2 (Based on Chebyshev’s inequality).

Let κ\kappa be the failure probability. The inequality (6) holds when ηM2d2κ\eta\sqrt{M}\geqslant\frac{2d^{2}}{\sqrt{\kappa}}.

Proof.

Recall the variance upper bounds of 𝐇^k,q\hat{\mathbf{H}}_{k,q} and 𝐒^k,q\hat{\mathbf{S}}_{k,q} given in Eq. (4) and (5), respectively. When M4d4κη2M\geqslant\frac{4d^{4}}{\kappa\eta^{2}},

Var(𝐇^k,q)\displaystyle\mathrm{Var}(\hat{\mathbf{H}}_{k,q}) \displaystyle\leqslant κC𝐇2η22d4,\displaystyle\frac{\kappa C_{\mathbf{H}}^{2}\eta^{2}}{2d^{4}}, (29)
Var(𝐒^k,q)\displaystyle\mathrm{Var}(\hat{\mathbf{S}}_{k,q}) \displaystyle\leqslant κC𝐒2η22d4.\displaystyle\frac{\kappa C_{\mathbf{S}}^{2}\eta^{2}}{2d^{4}}. (30)

According to Chebyshev’s inequality, we have

Pr(|𝐇^k,q𝐇k,q|C𝐇ηd)\displaystyle\Pr\left(|\hat{\mathbf{H}}_{k,q}-\mathbf{H}_{k,q}|\geqslant\frac{C_{\mathbf{H}}\eta}{d}\right) \displaystyle\leqslant κ2d2,\displaystyle\frac{\kappa}{2d^{2}}, (31)
Pr(|𝐒^k,q𝐒k,q|C𝐒ηd)\displaystyle\Pr\left(|\hat{\mathbf{S}}_{k,q}-\mathbf{S}_{k,q}|\geqslant\frac{C_{\mathbf{S}}\eta}{d}\right) \displaystyle\leqslant κ2d2.\displaystyle\frac{\kappa}{2d^{2}}. (32)

Since matrix entries are measured independently, estimators 𝐇^k,q\hat{\mathbf{H}}_{k,q} or 𝐒^k,q\hat{\mathbf{S}}_{k,q} are independent. We have

Pr(k,q|𝐇^k,q𝐇k,q|C𝐇ηdand|𝐒^k,q𝐒k,q|C𝐒ηd)\displaystyle\Pr\left(\begin{array}[]{c}\forall k,q\ |\hat{\mathbf{H}}_{k,q}-\mathbf{H}_{k,q}|\leqslant\frac{C_{\mathbf{H}}\eta}{d}\\ {\rm and}\ |\hat{\mathbf{S}}_{k,q}-\mathbf{S}_{k,q}|\leqslant\frac{C_{\mathbf{S}}\eta}{d}\end{array}\right) (35)
\displaystyle\geqslant [(1κ2d2)d2]21κ,\displaystyle\left[\left(1-\frac{\kappa}{2d^{2}}\right)^{d^{2}}\right]^{2}\geqslant 1-\kappa, (36)

where Bernoulli’s inequality is used.

Let 𝐀\mathbf{A} be an m×nm\times n matrix with entries Ai,jA_{i,j}, its spectral norm satisfies

A2\displaystyle\|A\|_{2} \displaystyle\leqslant mnmaxi,j|Ai,j|.\displaystyle\sqrt{mn}\max_{i,j}|A_{i,j}|. (37)

This is because of the well known relation A2AF\|A\|_{2}\leqslant\|A\|_{F}, where the Frobenius norm is AF=(i=1mj=1n|Ai,j|2)1/2\|A\|_{F}=\left(\sum_{i=1}^{m}\sum_{j=1}^{n}|A_{i,j}|^{2}\right)^{1/2}. Therefore, when |𝐇^k,q𝐇k,q|C𝐇ηd|\hat{\mathbf{H}}_{k,q}-\mathbf{H}_{k,q}|\leqslant\frac{C_{\mathbf{H}}\eta}{d} and |𝐒^k,q𝐒k,q|C𝐒ηd|\hat{\mathbf{S}}_{k,q}-\mathbf{S}_{k,q}|\leqslant\frac{C_{\mathbf{S}}\eta}{d} for all kk and qq, we have 𝐇^𝐇2C𝐇η\|\hat{\mathbf{H}}-\mathbf{H}\|_{2}\leqslant C_{\mathbf{H}}\eta and 𝐒^𝐒2C𝐒η\|\hat{\mathbf{S}}-\mathbf{S}\|_{2}\leqslant C_{\mathbf{S}}\eta. ∎

The above proof applies to all bases and measurement protocols and holds without any assumption on the statistical noise. As the most general result, Lemma 2 is used to prove Theorem 1 in Appendix C.

To measure matrix entries on a quantum computer, a practical method is using Hadamard-test circuits. Then, measurement outcomes are independent random variables ±1\pm 1, i.e. the random variables are bounded even taking into account other factors (We will give an example later, see Algorithm 2, where the phase of qsq_{s} can be absorbed into UsU_{s}). In this case, we can use Hoeffding’s inequality [55] to obtain an improved result.

Lemma 3 (Based on Hoeffding’s inequality).

Suppose matrix entries 𝐇^k,q\hat{\mathbf{H}}_{k,q} and 𝐒^k,q\hat{\mathbf{S}}_{k,q} are empirical means of random variables in ranges [C𝐇,C𝐇][-C_{\mathbf{H}},C_{\mathbf{H}}] and [C𝐒,C𝐒][-C_{\mathbf{S}},C_{\mathbf{S}}], respectively. Let κ\kappa be the failure probability. If the matrix elements of 𝐇\mathbf{H} and 𝐒\mathbf{S} are estimated using Algorithm 2, then the inequality (6) holds when ηM2d2ln8d2κ\eta\sqrt{M}\geqslant\sqrt{2d^{2}\ln{\frac{8d^{2}}{\kappa}}}.

Proof.

For each matrix entry, its real (imaginary) part is computed by taking the mean of MM random variables. According to Hoeffding’s inequality, we have,

Pr(|𝐇^k,q𝐇k,q|C𝐇ηd)\displaystyle\Pr\left(|\hat{\mathbf{H}}_{k,q}-\mathbf{H}_{k,q}|\geqslant\frac{C_{\mathbf{H}}\eta}{d}\right) (38)
\displaystyle\leqslant Pr(|Re(𝐇^k,q)Re(𝐇k,q)|C𝐇η2d)\displaystyle\Pr\left(|\mathrm{Re}(\hat{\mathbf{H}}_{k,q})-\mathrm{Re}(\mathbf{H}_{k,q})|\geqslant\frac{C_{\mathbf{H}}\eta}{\sqrt{2}d}\right)
×\displaystyle\times Pr(|Im(𝐇^k,q)Im(𝐇k,q)|C𝐇η2d)\displaystyle\Pr\left(|\mathrm{Im}(\hat{\mathbf{H}}_{k,q})-\mathrm{Im}(\mathbf{H}_{k,q})|\geqslant\frac{C_{\mathbf{H}}\eta}{\sqrt{2}d}\right)
\displaystyle\leqslant 4eMη22d2.\displaystyle 4e^{-\frac{M\eta^{2}}{2d^{2}}}.

Similarly,

Pr(|𝐒^k,q𝐒k,q|C𝐒ηd)4eMη22d2.\displaystyle\Pr\left(|\hat{\mathbf{S}}_{k,q}-\mathbf{S}_{k,q}|\geqslant\frac{C_{\mathbf{S}}\eta}{d}\right)\leqslant 4e^{-\frac{M\eta^{2}}{2d^{2}}}. (39)

When M2d2η2ln8d2κM\geqslant\frac{2d^{2}}{\eta^{2}}\ln{\frac{8d^{2}}{\kappa}}, Eq. (31) and Eq. (32) hold. The rest of the proof is the same as in Lemma 2. ∎

Considering the equivalence between matrix entries, we can find that 𝐇\mathbf{H} and 𝐒\mathbf{S} are real Hankel matrices for the GP and P, IP and ITE bases in Table 1, real symmetric matrices for the CP and F bases and complex Hermitian-Toeplitz matrices for the RTE basis. Under the collective measurement protocol, random matrices 𝐇^𝐇\hat{\mathbf{H}}-\mathbf{H} and 𝐒^𝐒\hat{\mathbf{S}}-\mathbf{S} are matrices of the same type as 𝐇\mathbf{H} and 𝐒\mathbf{S}. To analyse the distribution of these random matrices, we assume that the distributions of 𝐇^k,q𝐇k,q\hat{\mathbf{H}}_{k,q}-\mathbf{H}_{k,q} and 𝐒^k,q𝐒k,q\hat{\mathbf{S}}_{k,q}-\mathbf{S}_{k,q} are Gaussian with the zero mean. Notice that inequivalent entries are independent random variables. Then, we obtain the following results.

Lemma 4 (Real symmetric/Hankel matrices).

Let κ\kappa be the failure probability. If 𝐇^𝐇\hat{\mathbf{H}}-\mathbf{H} and 𝐒^𝐒\hat{\mathbf{S}}-\mathbf{S} are random real symmetric matrices following the Gaussian distribution, then the inequality (6) holds when ηM2dln4dκ\eta\sqrt{M}\geqslant\sqrt{2d\ln{\frac{4d}{\kappa}}}.

Proof.

Applying the matrix Gaussian series theorem [56] to the real symmetric matrix with Gaussian noise, we have

Pr(𝐇^𝐇2C𝐇η)2deMη22d,\displaystyle\Pr\left(\|\hat{\mathbf{H}}-\mathbf{H}\|_{2}\geqslant C_{\mathbf{H}}\eta\right)\leqslant 2de^{-\frac{M\eta^{2}}{2d}}, (40)

where Var(𝐇^k,q)C𝐇2M{\rm Var}(\hat{\mathbf{H}}_{k,q})\leqslant\frac{C_{\mathbf{H}}^{2}}{M} is used. Similarly,

Pr(𝐒^𝐒2C𝐒η)2deMη22d.\displaystyle\Pr\left(\|\hat{\mathbf{S}}-\mathbf{S}\|_{2}\geqslant C_{\mathbf{S}}\eta\right)\leqslant 2de^{-\frac{M\eta^{2}}{2d}}. (41)

Note that the real Hankel matrices with Gaussian noise yield the same inequalities [57].

To make sure 𝐇^𝐇2C𝐇η\|\hat{\mathbf{H}}-\mathbf{H}\|_{2}\leqslant C_{\mathbf{H}}\eta and 𝐒^𝐒2C𝐒η\|\hat{\mathbf{S}}-\mathbf{S}\|_{2}\leqslant C_{\mathbf{S}}\eta with a probability of at least 1κ1-\kappa, we take MM such that

2deMη22d=κ2.\displaystyle 2de^{-\frac{M\eta^{2}}{2d}}=\frac{\kappa}{2}. (42)

Then,

M=2dη2ln4dκ.\displaystyle M=\frac{2d}{\eta^{2}}\ln{\frac{4d}{\kappa}}. (43)

Lemma 5 (Complex Hermitian-Toeplitz matrices).

Let κ\kappa be the failure probability. If 𝐇^𝐇\hat{\mathbf{H}}-\mathbf{H} and 𝐒^𝐒\hat{\mathbf{S}}-\mathbf{S} are random complex Hermitian-Toeplitz matrices following the Gaussian distribution, then the inequality (6) holds when ηM2(2d1)ln4dκ\eta\sqrt{M}\geqslant\sqrt{2(2d-1)\ln{\frac{4d}{\kappa}}}.

Proof.

For complex Hermitian-Toeplitz matrices, the matrix Gaussian series theorem [56] gives

Pr(𝐇^𝐇2C𝐇η)2deMη22(2d1).\displaystyle\Pr\left(\|\hat{\mathbf{H}}-\mathbf{H}\|_{2}\geqslant C_{\mathbf{H}}\eta\right)\leqslant 2de^{-\frac{M\eta^{2}}{2(2d-1)}}. (44)

Similarly,

Pr(𝐒^𝐒2C𝐒η)2deMη22(2d1).\displaystyle\Pr\left(\|\hat{\mathbf{S}}-\mathbf{S}\|_{2}\geqslant C_{\mathbf{S}}\eta\right)\leqslant 2de^{-\frac{M\eta^{2}}{2(2d-1)}}. (45)

To make sure 𝐇^𝐇2C𝐇η\|\hat{\mathbf{H}}-\mathbf{H}\|_{2}\leqslant C_{\mathbf{H}}\eta and 𝐒^𝐒2C𝐒η\|\hat{\mathbf{S}}-\mathbf{S}\|_{2}\leqslant C_{\mathbf{S}}\eta with a probability of at least 1κ1-\kappa, we take

M=2(2d1)η2ln4dκ.\displaystyle M=\frac{2(2d-1)}{\eta^{2}}\ln{\frac{4d}{\kappa}}. (46)

Appendix C Details on the analysis of measurement cost

In this section, we develop theoretical tools for analysing the measurement cost in quantum KSD algorithms. At the end of this section, we give the proofs of Lemma 1 and Theorem 1.

Definition 2.

For simplification, we define the following functions:

E(𝐚)\displaystyle E(\mathbf{a}) =\displaystyle= 𝐇(𝐚)𝐒(𝐚),\displaystyle\frac{\langle\mathbf{H}\rangle(\mathbf{a})}{\langle\mathbf{S}\rangle(\mathbf{a})}, (47)
E^(η,𝐚)\displaystyle\hat{E}(\eta,\mathbf{a}) =\displaystyle= 𝐇^(𝐚)+C𝐇η𝐒^(𝐚)+C𝐒η,\displaystyle\frac{\langle\hat{\mathbf{H}}\rangle(\mathbf{a})+C_{\mathbf{H}}\eta}{\langle\hat{\mathbf{S}}\rangle(\mathbf{a})+C_{\mathbf{S}}\eta}, (48)
E(η,𝐚)\displaystyle E^{\prime}(\eta,\mathbf{a}) =\displaystyle= 𝐇(𝐚)+2C𝐇η𝐒(𝐚)+2C𝐒η.\displaystyle\frac{\langle\mathbf{H}\rangle(\mathbf{a})+2C_{\mathbf{H}}\eta}{\langle\mathbf{S}\rangle(\mathbf{a})+2C_{\mathbf{S}}\eta}. (49)

Here, functions 𝐀(𝐚)\langle\mathbf{A}\rangle(\mathbf{a}) are defined in Eq. (27). min𝐚0E(𝐚)=Emin\min_{\mathbf{a}\neq 0}E(\mathbf{a})=E_{min} is the minimum eigenvalue of the generalised eigenvalue problem in Eq. (2); min𝐚0E^(η,𝐚)=E^min\min_{\mathbf{a}\neq 0}\hat{E}(\eta,\mathbf{a})=\hat{E}_{min} is the minimum eigenvalue of the generalised eigenvalue problem solved in Algorithm 1; and Eq. (49) is equivalent to Eq. (7).

Lemma 6.

If 𝐇^𝐇2C𝐇η\|\hat{\mathbf{H}}-\mathbf{H}\|_{2}\leqslant C_{\mathbf{H}}\eta and 𝐒^𝐒2C𝐒η\|\hat{\mathbf{S}}-\mathbf{S}\|_{2}\leqslant C_{\mathbf{S}}\eta, the following statements hold:

  1. (1)

    E^(η,𝐚)min{E(𝐚),0}\hat{E}(\eta,\mathbf{a})\geqslant\min\{E(\mathbf{a}),0\};

  2. (2)

    If E^(η,𝐚)<0\hat{E}(\eta,\mathbf{a})<0, then E^(η,𝐚)E(η,𝐚)\hat{E}(\eta,\mathbf{a})\leqslant E^{\prime}(\eta,\mathbf{a});

  3. (3)

    If E(η,𝐚)<0E^{\prime}(\eta,\mathbf{a})<0, then E^(η,𝐚)<0\hat{E}(\eta,\mathbf{a})<0.

Proof.

Consider the error in the denominator,

|𝐒^(𝐚)𝐒(𝐚)|=|𝐚𝐒^𝐚𝐚𝐒𝐚𝐚𝐚|\displaystyle|\langle\hat{\mathbf{S}}\rangle(\mathbf{a})-\langle\mathbf{S}\rangle(\mathbf{a})|=\left|\frac{\mathbf{a}^{\dagger}\hat{\mathbf{S}}\mathbf{a}-\mathbf{a}^{\dagger}\mathbf{S}\mathbf{a}}{\mathbf{a}^{\dagger}\mathbf{a}}\right|
=|𝐚(𝐒^𝐒)𝐚𝐚𝐚||𝐚|𝐒^𝐒2|𝐚|𝐚𝐚\displaystyle=\left|\frac{\mathbf{a}^{\dagger}(\hat{\mathbf{S}}-\mathbf{S})\mathbf{a}}{\mathbf{a}^{\dagger}\mathbf{a}}\right|\leqslant\frac{|\mathbf{a}^{\dagger}|\|\hat{\mathbf{S}}-\mathbf{S}\|_{2}|\mathbf{a}|}{\mathbf{a}^{\dagger}\mathbf{a}}
C𝐒η,\displaystyle\leqslant C_{\mathbf{S}}\eta, (50)

from which we obtain

0<𝐒(𝐚)\displaystyle 0<\langle\mathbf{S}\rangle(\mathbf{a}) \displaystyle\leqslant 𝐒^(𝐚)+C𝐒η\displaystyle\langle\hat{\mathbf{S}}\rangle(\mathbf{a})+C_{\mathbf{S}}\eta (51)
\displaystyle\leqslant 𝐒(𝐚)+2C𝐒η.\displaystyle\langle\mathbf{S}\rangle(\mathbf{a})+2C_{\mathbf{S}}\eta.

Similarly, we have

𝐇(𝐚)\displaystyle\langle\mathbf{H}\rangle(\mathbf{a}) \displaystyle\leqslant 𝐇^(𝐚)+C𝐇η\displaystyle\langle\hat{\mathbf{H}}\rangle(\mathbf{a})+C_{\mathbf{H}}\eta (52)
\displaystyle\leqslant 𝐇(𝐚)+2C𝐇η.\displaystyle\langle\mathbf{H}\rangle(\mathbf{a})+2C_{\mathbf{H}}\eta.

Let aa, bb, cc and dd be positive numbers, it can be shown that

ab\displaystyle\frac{-a}{b} <\displaystyle< a+cb+d.\displaystyle\frac{-a+c}{b+d}. (53)

If E^(η,𝐚)<0\hat{E}(\eta,\mathbf{a})<0, then 𝐇^(𝐚)+C𝐇η<0\langle\hat{\mathbf{H}}\rangle(\mathbf{a})+C_{\mathbf{H}}\eta<0. Under this condition, by using Eqs. (51)-(53), we get

𝐇(𝐚)𝐒(𝐚)\displaystyle\frac{\langle\mathbf{H}\rangle(\mathbf{a})}{\langle\mathbf{S}\rangle(\mathbf{a})} \displaystyle\leqslant 𝐇^(𝐚)+C𝐇η𝐒^(𝐚)+C𝐒η\displaystyle\frac{\langle\hat{\mathbf{H}}\rangle(\mathbf{a})+C_{\mathbf{H}}\eta}{\langle\hat{\mathbf{S}}\rangle(\mathbf{a})+C_{\mathbf{S}}\eta} (54)
\displaystyle\leqslant 𝐇(𝐚)+2C𝐇η𝐒(𝐚)+2C𝐒η,\displaystyle\frac{\langle\mathbf{H}\rangle(\mathbf{a})+2C_{\mathbf{H}}\eta}{\langle\mathbf{S}\rangle(\mathbf{a})+2C_{\mathbf{S}}\eta},

i.e. E(𝐚)E^(η,𝐚)E(η,𝐚)E(\mathbf{a})\leqslant\hat{E}(\eta,\mathbf{a})\leqslant E^{\prime}(\eta,\mathbf{a}). Both the first and second statements are proven.

If E(η,𝐚)<0E^{\prime}(\eta,\mathbf{a})<0, then 𝐇^(𝐚)+C𝐇η𝐇(𝐚)+2C𝐇η<0\langle\hat{\mathbf{H}}\rangle(\mathbf{a})+C_{\mathbf{H}}\eta\leqslant\langle\mathbf{H}\rangle(\mathbf{a})+2C_{\mathbf{H}}\eta<0. Therefore, E^(η,𝐚)<0\hat{E}(\eta,\mathbf{a})<0. The third statement is proven. ∎

Under the condition E(η,𝐚)<0E^{\prime}(\eta,\mathbf{a})<0, Lemma 6 states E(𝐚)E^(η,𝐚)E(η,𝐚)E(\mathbf{a})\leqslant\hat{E}(\eta,\mathbf{a})\leqslant E^{\prime}(\eta,\mathbf{a}), where the first inequality shows that the algorithm is variational, and the second inequality gives the error bound. In the following, we will show that this relation still holds after minimisation.

Lemma 7.

Under conditions

  1. (1)

    𝐇^𝐇2C𝐇η\|\hat{\mathbf{H}}-\mathbf{H}\|_{2}\leqslant C_{\mathbf{H}}\eta and 𝐒^𝐒2C𝐒η\|\hat{\mathbf{S}}-\mathbf{S}\|_{2}\leqslant C_{\mathbf{S}}\eta, and

  2. (2)

    Eg<0E_{g}<0 and min𝐚𝟎E(η,𝐚)<0\min_{\mathbf{a}\neq\bf{0}}E^{\prime}(\eta,\mathbf{a})<0,

the following statement holds,

Egmin𝐚𝟎E^(η,𝐚)min𝐚𝟎E(η,𝐚).\displaystyle E_{g}\leqslant\min_{\mathbf{a}\neq\bf{0}}\hat{E}(\eta,\mathbf{a})\leqslant\min_{\mathbf{a}\neq\bf{0}}E^{\prime}(\eta,\mathbf{a}). (55)
Proof.

According to the first statement in Lemma 6, 𝐚𝟎\forall\mathbf{a}\neq\bf{0}, E^(η,𝐚)min{E(𝐚),0}\hat{E}(\eta,\mathbf{a})\geqslant\min\{E(\mathbf{a}),0\}. Since E(𝐚)EgE(\mathbf{a})\geqslant E_{g} and 0>Eg0>E_{g}, we have min𝐚𝟎E^(η,𝐚)Eg\min_{\mathbf{a}\neq\bf{0}}\hat{E}(\eta,\mathbf{a})\geqslant E_{g}.

Let 𝐚=argmin𝐚𝟎E(η,𝐚)\mathbf{a}^{*}=\operatorname*{arg\,min}_{\mathbf{a}\neq\bf{0}}E^{\prime}(\eta,\mathbf{a}). According to the third statement in Lemma 6, the condition E(η,𝐚)<0E^{\prime}(\eta,\mathbf{a}^{*})<0 implies that E^(η,𝐚)<0\hat{E}(\eta,\mathbf{a}^{*})<0. Under this condition, according to the second statement in Lemma 6, E^(η,𝐚)E(η,𝐚)\hat{E}(\eta,\mathbf{a}^{*})\leqslant E^{\prime}(\eta,\mathbf{a}^{*}). Therefore, we have

min𝐚𝟎E^(η,𝐚)E^(η,𝐚)\displaystyle\min_{\mathbf{a}\neq\bf{0}}\hat{E}(\eta,\mathbf{a})\leqslant\hat{E}(\eta,\mathbf{a}^{*}) (56)
\displaystyle\leqslant E(η,𝐚)=min𝐚𝟎E(η,𝐚).\displaystyle E^{\prime}(\eta,\mathbf{a}^{*})=\min_{\mathbf{a}\neq\bf{0}}E^{\prime}(\eta,\mathbf{a}).

Lemma 8.

Suppose 𝐒\mathbf{S} is invertible, ϵ>ϵK\epsilon>\epsilon_{K} and Eg+ϵ<0E_{g}+\epsilon<0. Eq. (9) has a positive solution in η\eta, and the solution is unique.

Proof.

Let 𝐚η=argmin𝐚𝟎E(η,𝐚)\mathbf{a}^{*}_{\eta}=\operatorname*{arg\,min}_{\mathbf{a}\neq\bf{0}}E^{\prime}(\eta,\mathbf{a}).

First, we prove that min𝐚𝟎E(η,𝐚)\min_{\mathbf{a}\neq\bf{0}}E^{\prime}(\eta,\mathbf{a}) is a strictly monotonically increasing function of η\eta when η\eta is in the range {η0|min𝐚𝟎E(η,𝐚)<0}\{\eta\geqslant 0|\min_{\mathbf{a}\neq\bf{0}}E^{\prime}(\eta,\mathbf{a})<0\}. Let η2\eta_{2} and η1\eta_{1} be in the range and η2>η1\eta_{2}>\eta_{1}. By Eq. (53), we have E(η1,𝐚η2)<E(η2,𝐚η2)E^{\prime}(\eta_{1},\mathbf{a}^{*}_{\eta_{2}})<E^{\prime}(\eta_{2},\mathbf{a}^{*}_{\eta_{2}}). Since min𝐚𝟎E(η1,𝐚)E(η1,𝐚η2)\min_{\mathbf{a}\neq\bf{0}}E^{\prime}(\eta_{1},\mathbf{a})\leqslant E^{\prime}(\eta_{1},\mathbf{a}^{*}_{\eta_{2}}),

min𝐚𝟎E(η1,𝐚)<min𝐚𝟎E(η2,𝐚).\displaystyle\min_{\mathbf{a}\neq\bf{0}}E^{\prime}(\eta_{1},\mathbf{a})<\min_{\mathbf{a}\neq\bf{0}}E^{\prime}(\eta_{2},\mathbf{a}). (57)

Second, we prove that min𝐚𝟎E(η,𝐚)\min_{\mathbf{a}\neq\bf{0}}E^{\prime}(\eta,\mathbf{a}) is a continuous function of η\eta when η0\eta\geqslant 0. When η0\eta\geqslant 0 and 𝐚𝟎\mathbf{a}\neq\bf{0}, E(η,𝐚)E^{\prime}(\eta,\mathbf{a}) is a continuous function of η\eta. Consequently, δE>0\forall\delta_{E}>0, δη>0\exists\delta_{\eta}>0 such that E(η+δη,𝐚η)E(η,𝐚η)<δEE^{\prime}(\eta+\delta_{\eta},\mathbf{a}^{*}_{\eta})-E^{\prime}(\eta,\mathbf{a}^{*}_{\eta})<\delta_{E}. Then,

0\displaystyle 0 <\displaystyle< E(η+δη,𝐚η+δη)E(η,𝐚η)\displaystyle E^{\prime}(\eta+\delta_{\eta},\mathbf{a}^{*}_{\eta+\delta_{\eta}})-E^{\prime}(\eta,\mathbf{a}^{*}_{\eta}) (58)
<\displaystyle< E(η+δη,𝐚η)E(η,𝐚η)δE.\displaystyle E^{\prime}(\eta+\delta_{\eta},\mathbf{a}^{*}_{\eta})-E^{\prime}(\eta,\mathbf{a}^{*}_{\eta})\leqslant\delta_{E}.

Therefore, δE>0\forall\delta_{E}>0, δη>0\exists\delta_{\eta}>0 such that

0\displaystyle 0 <\displaystyle< min𝐚𝟎E(η+δη,𝐚)min𝐚𝟎E(η,𝐚)\displaystyle\min_{\mathbf{a}\neq\bf{0}}E^{\prime}(\eta+\delta_{\eta},\mathbf{a})-\min_{\mathbf{a}\neq\bf{0}}E^{\prime}(\eta,\mathbf{a}) (59)
<\displaystyle< δE.\displaystyle\delta_{E}.

Combining the two, we conclude that min𝐚𝟎E(η,𝐚)\min_{\mathbf{a}\neq\bf{0}}E^{\prime}(\eta,\mathbf{a}) is a strictly monotonically increasing continuous function of η\eta when η0\eta\geqslant 0 and min𝐚𝟎E(η,𝐚)<0\min_{\mathbf{a}\neq\bf{0}}E^{\prime}(\eta,\mathbf{a})<0. Taking η=0\eta=0, we have min𝐚𝟎E(0,𝐚)=Eg+ϵK\min_{\mathbf{a}\neq\bf{0}}E^{\prime}(0,\mathbf{a})=E_{g}+\epsilon_{K}. Therefore, when ϵ>ϵK\epsilon>\epsilon_{K} and Eg+ϵ<0E_{g}+\epsilon<0, the equation min𝐚𝟎E(η,𝐚)=Eg+ϵ\min_{\mathbf{a}\neq\bf{0}}E^{\prime}(\eta,\mathbf{a})=E_{g}+\epsilon has a unique solution. ∎

C.1 Proof of Lemma 1

Proof.

For a regularisation parameter η\eta satisfying Eq. (6), the first condition of Lemma 7 holds up to the failure probability κ\kappa. Under conditions Eg<0E_{g}<0 and min𝐚𝟎E(η,𝐚)<0\min_{\mathbf{a}\neq\bf{0}}E^{\prime}(\eta,\mathbf{a})<0, according to Lemma 7, EgE^minmin𝐚𝟎E(η,𝐚)E_{g}\leqslant\hat{E}_{min}\leqslant\min_{\mathbf{a}\neq\bf{0}}E^{\prime}(\eta,\mathbf{a}) holds up to the failure probability κ\kappa. Eq. (8) is proven. ∎

C.2 Proof of Theorem 1

Proof.

Suppose ϵ>ϵK\epsilon>\epsilon_{K} and Eg+ϵ<0E_{g}+\epsilon<0. The existence of the solution η\eta is proved in Lemma 8. With the solution η\eta, Eq. (9) is satisfied.

If we take the measurement number

M=4d4κη2,\displaystyle M=\frac{4d^{4}}{\kappa\eta^{2}}, (60)

𝐇^𝐇2C𝐇η\|\hat{\mathbf{H}}-\mathbf{H}\|_{2}\leqslant C_{\mathbf{H}}\eta and 𝐒^𝐒2C𝐒η\|\hat{\mathbf{S}}-\mathbf{S}\|_{2}\leqslant C_{\mathbf{S}}\eta hold up to the failure probability κ\kappa as proved in Lemma 2. Then, according to Lemma 7, Egmin𝐚𝟎E^(η,𝐚)min𝐚𝟎E(η,𝐚)=Eg+ϵE_{g}\leqslant\min_{\mathbf{a}\neq\bf{0}}\hat{E}(\eta,\mathbf{a})\leqslant\min_{\mathbf{a}\neq\bf{0}}E^{\prime}(\eta,\mathbf{a})=E_{g}+\epsilon holds up to the failure probability.

In the independent measurement protocol, the total number of measurements is

Mtot\displaystyle M_{tot} =\displaystyle= M×2×d2×2=16d6κη2.\displaystyle M\times 2\times d^{2}\times 2=\frac{16d^{6}}{\kappa\eta^{2}}. (61)

Notice that for each matrix entry 𝐇^k,q\hat{\mathbf{H}}_{k,q} or 𝐒^k,q\hat{\mathbf{S}}_{k,q}, we need MM measurements for its real part and MM measurements for its imaginary part. There are two matrices, and each matrix has d2d^{2} entries. To match Eq. (61), we can take α(κ)=256/κ\alpha(\kappa)=256/\kappa and β(d)=d6\beta(d)=d^{6} in Eq. (10). ∎

In the above proof, we have used Lemma 2 to give the most general and conservative estimation of the measurement cost. In fact, according to Appendix B, when considering details of the quantum KSD algorithms, MM (and MtotM_{tot}) can be optimised, so as α(κ)\alpha(\kappa) and β(d)\beta(d); see Appendix E.

C.3 Remarks on the measurement number

First, in our algorithm (i.e. GP) and some other KSD algorithms (i.e. P, CP, IP, ITE and F), matrices 𝐇\mathbf{H} and 𝐒\mathbf{S} are real. In this case, we only need to measure the real part. The cost is directly reduced by a factor of two. Only measuring the real part also reduces variances by a factor of two, i.e. from Var(𝐇^k,q)2C𝐇2/M\mathrm{Var}(\hat{\mathbf{H}}_{k,q})\leqslant 2C_{\mathbf{H}}^{2}/M and Var(𝐒^k,q)2C𝐒2/M\mathrm{Var}(\hat{\mathbf{S}}_{k,q})\leqslant 2C_{\mathbf{S}}^{2}/M to Var(𝐇^k,q)C𝐇2/M\mathrm{Var}(\hat{\mathbf{H}}_{k,q})\leqslant C_{\mathbf{H}}^{2}/M and Var(𝐒^k,q)C𝐒2/M\mathrm{Var}(\hat{\mathbf{S}}_{k,q})\leqslant C_{\mathbf{S}}^{2}/M. Because of the reduced variance, the cost is reduced by another factor of two. Overall, the measurement cost is reduced by a factor of four in algorithms using real matrices.

Second, when the failure probability κ\kappa is low, typically 𝐇^𝐇2C𝐇η\|\hat{\mathbf{H}}-\mathbf{H}\|_{2}\ll C_{\mathbf{H}}\eta and 𝐒^𝐒2C𝐒η\|\hat{\mathbf{S}}-\mathbf{S}\|_{2}\ll C_{\mathbf{S}}\eta. In this case, the error is overestimated in EE^{\prime}, and the typical error is approximately given by

E′′(η,𝐚)=𝐚(𝐇+C𝐇η)𝐚𝐚(𝐒+C𝐒η)𝐚.\displaystyle E^{\prime\prime}(\eta,\mathbf{a})=\frac{\mathbf{a}^{\dagger}(\mathbf{H}+C_{\mathbf{H}}\eta)\mathbf{a}}{\mathbf{a}^{\dagger}(\mathbf{S}+C_{\mathbf{S}}\eta)\mathbf{a}}. (62)

Notice that a factor of two has been removed from the denominator and numerator compared with EE^{\prime}. If we consider the typical error ϵ=E′′Eg\epsilon=E^{\prime\prime}-E_{g} instead of the error upper bound ϵ=EEg\epsilon=E^{\prime}-E_{g}, the required sampling cost is reduced by a factor of four.

Appendix D Cost for measuring the energy with an ideal projection

We consider the ideal case that we can realise an ideal projection onto the true ground state. To use the above results, we assume that f1=|ψgψg|f_{1}=|\psi_{g}\rangle\langle\psi_{g}| is the projection operator and take d=1d=1 in the KSD algorithm. Then, we have

𝐇1,1\displaystyle\mathbf{H}_{1,1} =\displaystyle= φ|f1Hf1|φ=pgEg,\displaystyle\langle{\varphi}|f_{1}^{\dagger}Hf_{1}|{\varphi}\rangle=p_{g}E_{g}, (63)
𝐒1,1\displaystyle\mathbf{S}_{1,1} =\displaystyle= φ|f1f1|φ=pg,\displaystyle\langle{\varphi}|f_{1}^{\dagger}f_{1}|{\varphi}\rangle=p_{g}, (64)

and Emin=𝐇1,1/𝐒1,1=EgE_{min}=\mathbf{H}_{1,1}/\mathbf{S}_{1,1}=E_{g}, i.e. ϵK=0\epsilon_{K}=0.

We suppose C𝐇=H2C_{\mathbf{H}}=\|H\|_{2} and C𝐒=1C_{\mathbf{S}}=1. When we take η=pgϵ4H2\eta=\frac{p_{g}\epsilon}{4\|H\|_{2}}, min𝐚𝟎E(η,𝐚)Eg+ϵ\min_{\mathbf{a}\neq\bf{0}}E^{\prime}(\eta,\mathbf{a})\leqslant E_{g}+\epsilon. Notice that min𝐚𝟎E(η,𝐚)=(𝐇1,1+2H2η)/(𝐒1,1+2η)Eg+4H2η/pg\min_{\mathbf{a}\neq\bf{0}}E^{\prime}(\eta,\mathbf{a})=(\mathbf{H}_{1,1}+2\|H\|_{2}\eta)/(\mathbf{S}_{1,1}+2\eta)\leqslant E_{g}+4\|H\|_{2}\eta/p_{g}. Accordingly, the total measurement number in the ideal case has an upper bound

Mtot=α(κ)H22pg2ϵ2,\displaystyle M_{tot}=\frac{\alpha(\kappa)\|H\|_{2}^{2}}{p_{g}^{2}\epsilon^{2}}, (65)

which is the first factor in Eq. (11).

Appendix E Optimised α\alpha and β\beta factors

Theorem 1 is proved taking the bound given by Lemma 2. However, Lemma 2 aims to give a general result, leading to a conservative estimation of α\alpha and β\beta. In this section, based on other results in Appendix B, we will show that these two factors can be further reduced. We summarise the reduced α\alpha and β\beta in Table 2 as well.

For the independent measurement protocol, Lemma 3 guarantees that the lower bound of the total measurement number is

Mtot=4d2M=8d4η2ln8d2κ.\displaystyle M_{tot}=4d^{2}M=\frac{8d^{4}}{\eta^{2}}\ln{\frac{8d^{2}}{\kappa}}. (66)

Assuming κ18d2\kappa\ll\frac{1}{8d^{2}} and neglecting 8d28d^{2} in logarithm yields α128ln1κ\alpha\approx 128\ln\frac{1}{\kappa} and βd4\beta\approx d^{4}.

Definition 3.

Let 𝐠2d1\mathbf{g}\in\mathbb{R}^{2d-1} and 𝐆d×d\mathbf{G}\in\mathbb{R}^{d\times d}. If 𝐆i,j=𝐠i+j1\mathbf{G}_{i,j}=\mathbf{g}_{i+j-1}, then 𝐆\mathbf{G} is a d×dd\times d real Hankel matrix.

In our algorithm (i.e. GP) and some other KSD algorithms (i.e. P, IP and ITE), 𝐇\mathbf{H} and 𝐒\mathbf{S} are real Hankel matrices. The collective measurement protocol only requires measuring 2d12d-1 matrix entries for each matrix to construct 𝐇^\hat{\mathbf{H}} and 𝐒^\hat{\mathbf{S}}. Specifically, we measure 𝐇l2,l2\mathbf{H}_{\lceil\frac{l}{2}\rceil,\lfloor\frac{l}{2}\rfloor} and 𝐒l2,l2\mathbf{S}_{\lceil\frac{l}{2}\rceil,\lfloor\frac{l}{2}\rfloor} with l=2,3,,2dl=2,3,\ldots,2d. Then we take 𝐇i,j=𝐇i+j2,i+j2\mathbf{H}_{i,j}=\mathbf{H}_{\lceil\frac{i+j}{2}\rceil,\lfloor\frac{i+j}{2}\rfloor} and 𝐒i,j=𝐒i+j2,i+j2\mathbf{S}_{i,j}=\mathbf{S}_{\lceil\frac{i+j}{2}\rceil,\lfloor\frac{i+j}{2}\rfloor} for all ii and jj. According to Lemma 4, the total measurement number is

Mtot\displaystyle M_{tot} =\displaystyle= M×(2d1)×2\displaystyle M\times(2d-1)\times 2 (67)
=\displaystyle= 4d(2d1)η2ln4dκ.\displaystyle\frac{4d(2d-1)}{\eta^{2}}\ln{\frac{4d}{\kappa}}.

Assuming κ14d\kappa\ll\frac{1}{4d} and neglecting 4d4d in logarithm yields α=64ln1κ\alpha=64\ln\frac{1}{\kappa} and β=d(2d1)\beta=d(2d-1).

For the CP and F algorithms, 𝐇\mathbf{H} and 𝐒\mathbf{S} are real symmetric matrices. We need to measure d(d+1)/2d(d+1)/2 matrix entries to construct 𝐇^\hat{\mathbf{H}} and 𝐒^\hat{\mathbf{S}}, respectively. According to Lemma 4, the total measurement number is

Mtot\displaystyle M_{tot} =\displaystyle= M×d(d+1)/2×2\displaystyle M\times d(d+1)/2\times 2 (68)
=\displaystyle= 2d2(d+1)η2ln4dκ.\displaystyle\frac{2d^{2}(d+1)}{\eta^{2}}\ln{\frac{4d}{\kappa}}.

Assuming κ14d\kappa\ll\frac{1}{4d} and neglecting 4d4d in logarithm yields α=32ln1κ\alpha=32\ln\frac{1}{\kappa} and β=d2(d+1)\beta=d^{2}(d+1).

The RTE algorithm is the only one having complex matrices 𝐇\mathbf{H} and 𝐒\mathbf{S}. Considering its Hermitian–Toeplitz structure, we need to measure dd real and d1d-1 imaginary matrix entries to estimate 𝐇\mathbf{H} and 𝐒\mathbf{S}, respectively. According to Lemma 5, the total measurement number is

Mtot\displaystyle M_{tot} =\displaystyle= M×(2d1)×2\displaystyle M\times(2d-1)\times 2 (69)
=\displaystyle= 4(2d1)2η2ln4dκ.\displaystyle\frac{4(2d-1)^{2}}{\eta^{2}}\ln{\frac{4d}{\kappa}}.

Assuming κ14d\kappa\ll\frac{1}{4d} and neglecting 4d4d in logarithm yields α=64ln1κ\alpha=64\ln\frac{1}{\kappa} and β=(2d1)2\beta=(2d-1)^{2}.

Finally, the optimised α\alpha and β\beta are listed in Table 2. Compared with cases, Hankel matrices have smaller α\alpha and β\beta. For the GP algorithm, taking α=16ln1κ\alpha=16\ln\frac{1}{\kappa} and β=d(2d1)\beta=d(2d-1) is sufficient.

Appendix F Gaussian-power basis

In this section, we prove Eq. (14) and Eq. (15), respectively.

F.1 Norm upper bound

Lemma 9.

The Gaussian-power basis operators fkf_{k} defined in Eq. (13) satisfy Eq. (14).

Proof.

By the spectral decomposition of the operator (HE0)τ(H-E_{0})\tau, we have

τk1fk2maxy|yk1e12y2|,\displaystyle\tau^{k-1}\|f_{k}\|_{2}\leqslant\max_{y\in\mathbb{R}}|y^{k-1}e^{-\frac{1}{2}y^{2}}|, (70)

where yy corresponds to eigenvalues of (HE0)τ(H-E_{0})\tau. It can be found that

maxy|yk1e12y2|=(k1e)k12.\displaystyle\max_{y\in\mathbb{R}}|y^{k-1}e^{-\frac{1}{2}y^{2}}|=\left(\frac{k-1}{e}\right)^{\frac{k-1}{2}}. (71)

Combining Eqs. (70) and (71), Eq. (14) is proved. ∎

F.2 The integral

The generating function for the Hermite polynomials is [40]

n=0Hn(u)tnn!=e2utt2.\displaystyle\sum_{n=0}^{\infty}H_{n}(u)\frac{t^{n}}{n!}=e^{2ut-t^{2}}. (72)
Lemma 10.
+𝑑uHn(u)eu2iyu=π(iy)ne14y2\displaystyle\int_{-\infty}^{+\infty}duH_{n}(u)e^{-u^{2}-iyu}=\sqrt{\pi}(-iy)^{n}e^{-\frac{1}{4}y^{2}} (73)

holds for arbitrary yy\in\mathbb{R}.

Proof.

Denote In=+𝑑uHn(u)eu2iyuI_{n}=\int_{-\infty}^{+\infty}duH_{n}(u)e^{-u^{2}-iyu}. Taking Eq. (72) into account, we have [58]

n=0Intnn!\displaystyle\sum_{n=0}^{\infty}I_{n}\frac{t^{n}}{n!} =\displaystyle= +𝑑ueu2+(2tiy)ut2\displaystyle\int_{-\infty}^{+\infty}due^{-u^{2}+(2t-iy)u-t^{2}} (74)
=\displaystyle= πe14y2iyt\displaystyle\sqrt{\pi}e^{-\frac{1}{4}y^{2}-iyt}
=\displaystyle= πe14y2n=0(iy)ntnn!,\displaystyle\sqrt{\pi}e^{-\frac{1}{4}y^{2}}\sum_{n=0}^{\infty}(-iy)^{n}\frac{t^{n}}{n!},

where the Gaussian integral is used. From Eq. (74), we conclude that In=π(iy)ne14y2I_{n}=\sqrt{\pi}(-iy)^{n}e^{-\frac{1}{4}y^{2}}. ∎

Lemma 11.

Eq. (15) is true.

Proof.

Taking y=2xτy=\sqrt{2}x\tau and u=t2τu=\frac{t}{\sqrt{2}\tau} in Eq. (73), Eq. (15) is proved directly. ∎

Appendix G Algorithm and variance

In this section, first, we review the zeroth-order leading-order-rotation formula [47], then we analyse the variance and work out the upper bounds of the cost, finally, we give the pseudocode of our measurement-efficient algorithm.

G.1 Zeroth-order leading-order-rotation formula

Assume that the Hamiltonian is expressed as a linear combination of Pauli operators, i.e.

H=jhjσj.\displaystyle H=\sum_{j}h_{j}\sigma_{j}. (75)

The Taylor expansion of the time evolution operator is

eiHΔt=𝟙iHΔt+T0(Δt),\displaystyle e^{-iH\Delta t}=\mathbbm{1}-iH\Delta t+T_{0}(\Delta t), (76)

where the summation of high-order terms is

T0(Δt)\displaystyle T_{0}(\Delta t) =\displaystyle= k=2j1,,jka=1k(ihjaΔt)k!\displaystyle\sum_{k=2}^{\infty}\sum_{j_{1},\ldots,j_{k}}\frac{\prod_{a=1}^{k}(-ih_{j_{a}}\Delta t)}{k!} (77)
×σjkσj1.\displaystyle\times\sigma_{j_{k}}\cdots\sigma_{j_{1}}.

The leading-order terms can be expressed as a linear combination of rotation operators,

𝟙iHΔt=jβj(Δt)eisgn(hj)ϕ(Δt)σj,\displaystyle\mathbbm{1}-iH\Delta t=\sum_{j}\beta_{j}(\Delta t)e^{-i\mathrm{sgn}(h_{j})\phi(\Delta t)\sigma_{j}}, (78)

where ϕ(Δt)=arctan(htotΔt)\phi(\Delta t)=\arctan(h_{tot}\Delta t), βj(Δt)=|hj|Δt/sinϕ(Δt)\beta_{j}(\Delta t)=|h_{j}|\Delta t/\sin\phi(\Delta t) and htot=j|hj|h_{tot}=\sum_{j}|h_{j}|. The zeroth-order leading-order-rotation formula is

eiHΔt\displaystyle e^{-iH\Delta t} =\displaystyle= jβj(Δt)eisgn(hj)ϕ(Δt)σj\displaystyle\sum_{j}\beta_{j}(\Delta t)e^{-i\mathrm{sgn}(h_{j})\phi(\Delta t)\sigma_{j}} (79)
+T0(Δt),\displaystyle+T_{0}(\Delta t),

which is a linear combination of rotation and Pauli operators. Accordingly, for the evolution time tt and time step number NN, the LCU expression of the time evolution operator is

eiHt\displaystyle e^{-iHt} =\displaystyle= [jβj(t/N)eisgn(hj)ϕ(t/N)σj\displaystyle\left[\sum_{j}\beta_{j}(t/N)e^{-i\mathrm{sgn}(h_{j})\phi(t/N)\sigma_{j}}\right. (80)
+T0(t/N)]N.\displaystyle\left.+T_{0}(t/N)\right]^{N}.

The above equation is the explicit form of the expression eiHt=[rvr(t/N)Vr(t/N)]Ne^{-iHt}=\left[\sum_{r}v_{r}(t/N)V_{r}(t/N)\right]^{N} referred in the main text.

Now, we consider the cost factor, i.e. the 1-norm of coefficients in an LCU expression. For Eq. (79), the corresponding cost factor is

c(Δt)=j|βj(Δt)|+k=2j1,,jka=1k|hjaΔt|k!\displaystyle c(\Delta t)=\sum_{j}|\beta_{j}(\Delta t)|+\sum_{k=2}^{\infty}\sum_{j_{1},\ldots,j_{k}}\frac{\prod_{a=1}^{k}|h_{j_{a}}\Delta t|}{k!}
=1+htot2Δt2+ehtot|Δt|(1+htot|Δt|).\displaystyle=\sqrt{1+h_{tot}^{2}\Delta t^{2}}+e^{h_{tot}|\Delta t|}-(1+h_{tot}|\Delta t|). (81)

Accordingly, the cost factor of Eq. (80) is [c(t/N)]N[c(t/N)]^{N}.

Lemma 12.
c(Δt)ee2htot2Δt2.\displaystyle c(\Delta t)\leqslant e^{\frac{e}{2}h_{tot}^{2}\Delta t^{2}}. (82)
Proof.

Let x=htot|Δt|x=h_{tot}|\Delta t|, thus x0x\geqslant 0. Then

c(Δt)\displaystyle c(\Delta t) =\displaystyle= 1+x2+ex(1+x)\displaystyle\sqrt{1+x^{2}}+e^{x}-(1+x) (83)
\displaystyle\leqslant ex22+ex(1+x).\displaystyle e^{\frac{x^{2}}{2}}+e^{x}-(1+x).

Define the function

y(x)=ee2x2+(1+x)(ex+ex22).\displaystyle y(x)=e^{\frac{e}{2}x^{2}}+(1+x)-(e^{x}+e^{\frac{x^{2}}{2}}). (84)

It follows that y(0)=0y(0)=0 and

y(x)\displaystyle y^{\prime}(x) =\displaystyle= exee2x2+1(ex+xex22)\displaystyle exe^{\frac{e}{2}x^{2}}+1-(e^{x}+xe^{\frac{x^{2}}{2}}) (85)
\displaystyle\geqslant (e1)xee2x2+1ex\displaystyle(e-1)xe^{\frac{e}{2}x^{2}}+1-e^{x}
\displaystyle\geqslant (e1)x+1ex.\displaystyle(e-1)x+1-e^{x}. (86)

Let z(x)=(e1)x+1exz(x)=(e-1)x+1-e^{x}. Since z(0)=z(1)=0z(0)=z(1)=0, it is easy to see that z(x)0z(x)\geqslant 0 when 0x10\leqslant x\leqslant 1, which indicates that y(x)0y^{\prime}(x)\geqslant 0 when 0x10\leqslant x\leqslant 1. Based on Eq. (85) and the fact that e1>1e-1>1, we have y(x)1y^{\prime}(x)\geqslant 1 when x1x\geqslant 1. As a result, y(x)0y^{\prime}(x)\geqslant 0 when x0x\geqslant 0, which means y(x)0y(x)\geqslant 0. Therefore,

c(Δt)ee2x2y(x)ee2x2.\displaystyle c(\Delta t)\leqslant e^{\frac{e}{2}x^{2}}-y(x)\leqslant e^{\frac{e}{2}x^{2}}. (87)

According to Lemma 12, the cost factor eiHte^{-iHt} has the upper bound

[c(t/N)]Neehtot2t22N.\displaystyle[c(t/N)]^{N}\leqslant e^{\frac{eh_{tot}^{2}t^{2}}{2N}}. (88)

Therefore, the cost factor approaches one when the time step number NN increases.

G.2 Variance analysis

In this section, first, we prove the general upper bound of the variance, then we work out the upper bound of the cost ckc_{k}.

G.2.1 Variance upper bound

Refer to caption
Figure 3: The Hadamard-test circuit 𝒞s\mathcal{C}_{s}. The qubit on the top is the ancilla qubit. The unitary operator UφU_{\varphi} prepares the state |φ|{\varphi}\rangle, i.e. Uφ|0n=|φU_{\varphi}|{0}\rangle^{\otimes n}=|{\varphi}\rangle. When the ancilla qubit is measured in the XX (or YY) basis, the measurement outcome is an eigenvalue of the Pauli operator XX (or YY), i.e. μX=±1\mu^{X}=\pm 1 (or μY=±1\mu^{Y}=\pm 1).

Given the LCU expression A=sqsUsA=\sum_{s}q_{s}U_{s}, we can compute φ|A|φ\langle{\varphi}|A|{\varphi}\rangle using the Monte Carlo method and the one-ancilla Hadamard-test circuit shown in Fig. 3. In the Monte Carlo calculation, we sample the unitary operator UsU_{s} with a probability of |qs|/CA|q_{s}|/C_{A} in the spirit of importance sampling. The corresponding algorithm is given in Algorithm 2.

1:Input |φ|{\varphi}\rangle, {(qs,Us)}\{(q_{s},U_{s})\} and MM.
2:CAs|qs|C_{A}\leftarrow\sum_{s}|q_{s}|
3:A^0\hat{A}\leftarrow 0
4:for l=1l=1 to MM do
5:     Choose ss with a probability of |qs|/CA|q_{s}|/C_{A}.
6:     Implement the circuit 𝒞s\mathcal{C}_{s} for one shot, measure the ancilla qubit in the XX basis and record the measurement outcome μX\mu^{X}.
7:     Implement the circuit 𝒞s\mathcal{C}_{s} for one shot, measure the ancilla qubit in the YY basis and record the measurement outcome μY\mu^{Y}.
8:     A^A^+eiarg(qs)(μX+iμY)\hat{A}\leftarrow\hat{A}+e^{i\arg(q_{s})}(\mu^{X}+i\mu^{Y})
9:Output A^CAMA^\hat{A}\leftarrow\frac{C_{A}}{M}\hat{A}.
Algorithm 2 Monte Carlo evaluation of an LCU expression of AA.
Lemma 13.

According to Algorithm 2, the estimator A^\hat{A} is unbiased, and the variance upper bound in Eq. (18) is true.

Proof.

First, we rewrite the LCU expression as

A=CAs|qs|CAeiarg(qs)Us.\displaystyle A=C_{A}\sum_{s}\frac{|q_{s}|}{C_{A}}e^{i\arg(q_{s})}U_{s}. (89)

Then,

φ|A|φ=CAs|qs|CAeiarg(qs)φ|Us|φ.\displaystyle\langle{\varphi}|A|{\varphi}\rangle=C_{A}\sum_{s}\frac{|q_{s}|}{C_{A}}e^{i\arg(q_{s})}\langle{\varphi}|U_{s}|{\varphi}\rangle. (90)

Each unitary operator UsU_{s} has a corresponding Hadamard-test circuit denoted by 𝒞s\mathcal{C}_{s}, which is shown in Fig. 3. When the ancilla qubit is measured in the W=X,YW=X,Y basis, the measurement has a random outcome μW=±1\mu^{W}=\pm 1. Let PμWWP^{W}_{\mu^{W}} be the probability of the measurement outcome μW\mu^{W} in 𝒞s\mathcal{C}_{s}. According to Ref. [33], we have

φ|Us|φ\displaystyle\langle{\varphi}|U_{s}|{\varphi}\rangle =\displaystyle= μX=±1PμXXμX\displaystyle\sum_{\mu^{X}=\pm 1}P^{X}_{\mu^{X}}\mu^{X} (91)
+iμY=±1PμYYμY.\displaystyle+i\sum_{\mu^{Y}=\pm 1}P^{Y}_{\mu^{Y}}\mu^{Y}.

The probability of (s,μX,μY)(s,\mu^{X},\mu^{Y}) is (|qs|/CA)PμXXPμYY(|q_{s}|/C_{A})P^{X}_{\mu^{X}}P^{Y}_{\mu^{Y}}. Using Eqs. (90) and (91), we have

φ|A|φ\displaystyle\langle{\varphi}|A|{\varphi}\rangle =\displaystyle= CAs,μX,μY|qs|CAPμXXPμYYeiarg(qs)\displaystyle C_{A}\sum_{s,\mu^{X},\mu^{Y}}\frac{|q_{s}|}{C_{A}}P^{X}_{\mu^{X}}P^{Y}_{\mu^{Y}}e^{i\arg(q_{s})} (92)
×(μX+iμY).\displaystyle\times(\mu^{X}+i\mu^{Y}).

The corresponding Monte Carlo algorithm is given in Algorithm 2.

The estimator A^\hat{A} is unbiased. According to Eq. (92), the expected value of CAeiarg(qs)(μX+iμY)C_{A}e^{i\arg(q_{s})}(\mu^{X}+i\mu^{Y}) is φ|A|φ\langle{\varphi}|A|{\varphi}\rangle. Therefore, the expected value of A^\hat{A} is also φ|A|φ\langle{\varphi}|A|{\varphi}\rangle. Notice that A^\hat{A} is the average of CAeiarg(qs)(μX+iμY)C_{A}e^{i\arg(q_{s})}(\mu^{X}+i\mu^{Y}) over MM samples.

The variance of A^\hat{A} is

Var(A^)=1M[s,μX,μY|qs|CAPμXXPμYY\displaystyle\mathrm{Var}(\hat{A})=\frac{1}{M}\left[\sum_{s,\mu^{X},\mu^{Y}}\frac{|q_{s}|}{C_{A}}P^{X}_{\mu^{X}}P^{Y}_{\mu^{Y}}\right.
×|CAeiarg(qs)(μX+iμY)|2|φ|A|φ|2]\displaystyle\left.\times\left|C_{A}e^{i\arg(q_{s})}(\mu^{X}+i\mu^{Y})\right|^{2}-\left|\langle{\varphi}|A|{\varphi}\rangle\right|^{2}\right]
2CA2M.\displaystyle\leqslant\frac{2C_{A}^{2}}{M}. (93)

Here, we remark that the summation-form LCU expressions can be generalised to the integral-form LCU expressions, which inherit the same properties of the former.

G.2.2 Cost upper bound

We use a two-level LCU to construct fkf_{k}: First, fkf_{k} is an integral of RTE operators eiHte^{-iHt}; Second, RTE is realised using the leading-order-rotation method. Substituting Eq. (80) into Eq. (15), we can obtain the eventual LCU expression of fkf_{k}, i.e.

fk=ik12k12τk1+𝑑tHk1(t2τ)gτ(t)eiE0t\displaystyle f_{k}=\frac{i^{k-1}}{2^{\frac{k-1}{2}}\tau^{k-1}}\int_{-\infty}^{+\infty}dtH_{k-1}\left(\frac{t}{\sqrt{2}\tau}\right)g_{\tau}(t)e^{iE_{0}t}
×[jβj(t/N)eisgn(hj)ϕ(t/N)σj+T0(t/N)]N.\displaystyle\times\left[\sum_{j}\beta_{j}(t/N)e^{-i\mathrm{sgn}(h_{j})\phi(t/N)\sigma_{j}}+T_{0}(t/N)\right]^{N}. (94)

Substituting LCU expressions of fkf_{k} and fqf_{q} as well as the expression of HH into A=fkHfq,fkfqA=f_{k}^{\dagger}Hf_{q},f_{k}^{\dagger}f_{q}, we can obtain the LCU expression of AA. Then, CA=htotckcq,ckcqC_{A}=h_{tot}c_{k}c_{q},c_{k}c_{q}, respectively, where

ck\displaystyle c_{k} =\displaystyle= 12k12τk1+𝑑t|Hk1(t2τ)|gτ(t)\displaystyle\frac{1}{2^{\frac{k-1}{2}}\tau^{k-1}}\int_{-\infty}^{+\infty}dt\left|H_{k-1}\left(\frac{t}{\sqrt{2}\tau}\right)\right|g_{\tau}(t) (95)
×[c(t/N)]N.\displaystyle\times[c(t/N)]^{N}.

Here, we have replaced [r|vr(tN)|]N\left[\sum_{r}\left|v_{r}\left(\frac{t}{N}\right)\right|\right]^{N} with [c(t/N)]N[c(t/N)]^{N} in Eq. (20).

To work out the cost of AA, we have used that the cost factor is additive and multiplicative. Suppose LCU expressions of A1A_{1} and A2A_{2} are A1=q1U1+q1U1A_{1}=q_{1}U_{1}+q_{1}^{\prime}U_{1}^{\prime} and A2=q2U2+q2U2A_{2}=q_{2}U_{2}+q_{2}^{\prime}U_{2}^{\prime}, respectively. The cost factors of A1A_{1} and A2A_{2} are C1=|q1|+|q1|C_{1}=|q_{1}|+|q_{1}^{\prime}| and C2=|q2|+|q2|C_{2}=|q_{2}|+|q_{2}^{\prime}|, respectively. Substituting LCU expressions of A1A_{1} and A2A_{2} into A=A1+A2A=A_{1}+A_{2}, the expression of AA reads A=q1U1+q1U1+q2U2+q2U2A=q_{1}U_{1}+q_{1}^{\prime}U_{1}^{\prime}+q_{2}U_{2}+q_{2}^{\prime}U_{2}^{\prime}. Then, the cost factor of AA is CA=|q1|+|q1|+|q2|+|q2|=C1+C2C_{A}=|q_{1}|+|q_{1}^{\prime}|+|q_{2}|+|q_{2}^{\prime}|=C_{1}+C_{2}. Substituting LCU expressions of A1A_{1} and A2A_{2} into A=A1A2A^{\prime}=A_{1}A_{2}, the expression of AA^{\prime} reads A=(q1U1+q1U1)(q2U2+q2U2)=q1q2U1U2+q1q2U1U2+q1q2U1U2+q1q2U1U2A^{\prime}=(q_{1}U_{1}+q_{1}^{\prime}U_{1}^{\prime})(q_{2}U_{2}+q_{2}^{\prime}U_{2}^{\prime})=q_{1}q_{2}U_{1}U_{2}+q_{1}q_{2}^{\prime}U_{1}U_{2}^{\prime}+q_{1}^{\prime}q_{2}U_{1}^{\prime}U_{2}+q_{1}^{\prime}q_{2}^{\prime}U_{1}^{\prime}U_{2}^{\prime}. Then, the cost factor of AA^{\prime} is CA=|q1q2|+|q1q2|+|q1q2|+|q1q2|=C1C2C_{A}^{\prime}=|q_{1}q_{2}|+|q_{1}q_{2}^{\prime}|+|q_{1}^{\prime}q_{2}|+|q_{1}^{\prime}q_{2}^{\prime}|=C_{1}C_{2}.

The upper bound of ckc_{k} is

ckckub\displaystyle c_{k}\leqslant c_{k}^{ub} =\displaystyle= 12k12τk1+𝑑t|Hk1(t2τ)|gτ(t)\displaystyle\frac{1}{2^{\frac{k-1}{2}}\tau^{k-1}}\int_{-\infty}^{+\infty}dt\left|H_{k-1}\left(\frac{t}{\sqrt{2}\tau}\right)\right|g_{\tau}(t) (96)
×eχt2τ2,\displaystyle\times e^{\chi\frac{t^{2}}{\tau^{2}}},

where χ=ehtot2τ22N\chi=\frac{eh_{tot}^{2}\tau^{2}}{2N}. Here, we have used Lemma 12. Notice that ckubc_{k}^{ub} increases monotonically with χ\chi (i.e. decreases monotonically with NN). Taking χ=18\chi=\frac{1}{8} (i.e. N=4ehtot2τ2N=4eh_{tot}^{2}\tau^{2}), we numerically evaluate the upper bound ckubc_{k}^{ub} and plot it in Fig. 4. We can find that

ckub2(k1eτ2)k12.\displaystyle c_{k}^{ub}\leqslant 2\left(\frac{k-1}{e\tau^{2}}\right)^{\frac{k-1}{2}}. (97)
Refer to caption
Figure 4: The ratio between cost upper bound and norm upper bound, ckub/[(k1)/(eτ2)](k1)/2c_{k}^{ub}/[(k-1)/(e\tau^{2})]^{(k-1)/2}.

G.3 Pseudocode

Given the LCU expression of AA, we can compute φ|A|φ\langle{\varphi}|A|{\varphi}\rangle according to Algorithm 2. We would like to repeat how we compose the LCU expression of AA: Substituting Eq. (80) into Eq. (15), we can obtain the eventual LCU expression of fkf_{k}; Substituting LCU expressions of fkf_{k} and fqf_{q} as well as the expression of HH into A=fkHfqA=f_{k}^{\dagger}Hf_{q} and A=fkfqA=f_{k}^{\dagger}f_{q} (corresponding to 𝐇k,q\mathbf{H}_{k,q} and 𝐒k,q\mathbf{S}_{k,q}, respectively), we can obtain the LCU expression of AA. Algorithm 2 does not involve details of the LCU expression. In this section, we give pseudocodes involving details.

We remark that matrix entries of the rescaled basis fk=fk/ckf^{\prime}_{k}=f_{k}/c_{k} are 𝐇^k,q=𝐇^k,q/(ckcq)\hat{\mathbf{H}}^{\prime}_{k,q}=\hat{\mathbf{H}}_{k,q}/(c_{k}c_{q}) and 𝐒^k,q=𝐒^k,q/(ckcq)\hat{\mathbf{S}}^{\prime}_{k,q}=\hat{\mathbf{S}}_{k,q}/(c_{k}c_{q}), where 𝐇^k,q\hat{\mathbf{H}}_{k,q} and 𝐒^k,q\hat{\mathbf{S}}_{k,q} are computed according to the pseudocodes.

In pseudocodes, we use notations 𝐡=(h1,h2,)\mathbf{h}=(h_{1},h_{2},\ldots) and 𝝈=(σ1,σ2,)\bm{\sigma}=(\sigma_{1},\sigma_{2},\ldots) to represent the Hamiltonian H=jhjσjH=\sum_{j}h_{j}\sigma_{j}. We define probability density functions

pτ,k(t)\displaystyle p_{\tau,k}(t) =\displaystyle= 1ck2k12τk1|Hk1(t2τ)|gτ(t)\displaystyle\frac{1}{c_{k}2^{\frac{k-1}{2}}\tau^{k-1}}\left|H_{k-1}\left(\frac{t}{\sqrt{2}\tau}\right)\right|g_{\tau}(t) (98)
×[c(t/N)]N,\displaystyle\times[c(t/N)]^{N},

and probabilities

PL(Δt)\displaystyle P_{\mathrm{L}}(\Delta t) =\displaystyle= 1+htot2Δt2c(Δt),\displaystyle\frac{\sqrt{1+h_{tot}^{2}\Delta t^{2}}}{c(\Delta t)}, (99)
PT(Δt)\displaystyle P_{\mathrm{T}}(\Delta t) =\displaystyle= ehtot|Δt|(1+htot|Δt|)c(Δt).\displaystyle\frac{e^{h_{tot}|\Delta t|}-(1+h_{tot}|\Delta t|)}{c(\Delta t)}. (100)

We use POISSON(λ)\text{P{\scriptsize OISSON}}(\lambda) to denote a subroutine that returns kk with a probability of λkeλ/k!\lambda^{k}e^{-\lambda}/k!. The LCU expression of fkf_{k} is sampled according to Algorithms 5, 6 and 7. In Algorithm 5, we generate random time tt according to the distribution pτ,k(t)p_{\tau,k}(t). In Algorithm 6, we realise the RTE eiHte^{-iHt} by NN steps leading-order-rotation. In Algorithm 7, we randomly sample rotations or Pauli operators according to Eq. (80).

1:Input |φ|{\varphi}\rangle, (𝐡,𝝈)(\mathbf{h},\bm{\sigma}), E0E_{0}, τ\tau, NN and MM.
2:htotj|hj|h_{tot}\leftarrow\sum_{j}|h_{j}|
3:Compute ckc_{k} and cqc_{q}.
4:𝐇^k,q0\hat{\mathbf{H}}_{k,q}\leftarrow 0
5:for l=1l=1 to MM do
6:     Choose jj with a probability of |hj|/htot|h_{j}|/h_{tot}.
7:     (v¯k,V¯k)(\bar{v}_{k},\bar{V}_{k}) \leftarrow BasisGen(𝐡\mathbf{h},𝝈\bm{\sigma},E0E_{0},τ\tau,NN,kk)
8:     (v¯q,V¯q)(\bar{v}_{q},\bar{V}_{q}) \leftarrow BasisGen(𝐡\mathbf{h},𝝈\bm{\sigma},E0E_{0},τ\tau,NN,qq)
9:     Implement the circuit 𝒞s\mathcal{C}_{s} with Us=V¯kσjV¯qU_{s}=\bar{V}_{k}^{\dagger}\sigma_{j}\bar{V}_{q} for one shot, measure the ancilla qubit in the XX basis and record the measurement outcome μX\mu^{X}.
10:     Implement the circuit 𝒞s\mathcal{C}_{s} with Us=V¯kσjV¯qU_{s}=\bar{V}_{k}^{\dagger}\sigma_{j}\bar{V}_{q} for one shot, measure the ancilla qubit in the YY basis and record the measurement outcome μY\mu^{Y}.
11:     𝐇^k,q𝐇^k,q+eiarg(v¯khjv¯q)(μX+iμY)\hat{\mathbf{H}}_{k,q}\leftarrow\hat{\mathbf{H}}_{k,q}+e^{i\arg(\bar{v}_{k}^{*}h_{j}\bar{v}_{q})}(\mu^{X}+i\mu^{Y})
12:Output 𝐇^k,qhtotckcqM𝐇^k,q\hat{\mathbf{H}}_{k,q}\leftarrow\frac{h_{tot}c_{k}c_{q}}{M}\hat{\mathbf{H}}_{k,q}.
Algorithm 3 Measurement of matrix entry 𝐇k,q\mathbf{H}_{k,q}.
1:Input |φ|{\varphi}\rangle, (𝐡,𝝈)(\mathbf{h},\bm{\sigma}), E0E_{0}, τ\tau, NN and MM.
2:Compute ckc_{k} and cqc_{q}.
3:𝐒^k,q0\hat{\mathbf{S}}_{k,q}\leftarrow 0
4:for l=1l=1 to MM do
5:     (v¯k,V¯k)(\bar{v}_{k},\bar{V}_{k}) \leftarrow BasisGen(𝐡\mathbf{h},𝝈\bm{\sigma},E0E_{0},τ\tau,NN,kk)
6:     (v¯q,V¯q)(\bar{v}_{q},\bar{V}_{q}) \leftarrow BasisGen(𝐡\mathbf{h},𝝈\bm{\sigma},E0E_{0},τ\tau,NN,qq)
7:     Implement the circuit 𝒞s\mathcal{C}_{s} with Us=V¯kV¯qU_{s}=\bar{V}_{k}^{\dagger}\bar{V}_{q} for one shot, measure the ancilla qubit in the XX basis and record the measurement outcome μX\mu^{X}.
8:     Implement the circuit 𝒞s\mathcal{C}_{s} with Us=V¯kV¯qU_{s}=\bar{V}_{k}^{\dagger}\bar{V}_{q} for one shot, measure the ancilla qubit in the YY basis and record the measurement outcome μY\mu^{Y}.
9:     𝐒^k,q𝐒^k,q+eiarg(v¯kv¯q)(μX+iμY)\hat{\mathbf{S}}_{k,q}\leftarrow\hat{\mathbf{S}}_{k,q}+e^{i\arg(\bar{v}_{k}^{*}\bar{v}_{q})}(\mu^{X}+i\mu^{Y})
10:Output 𝐒^k,qckcqM𝐒^k,q\hat{\mathbf{S}}_{k,q}\leftarrow\frac{c_{k}c_{q}}{M}\hat{\mathbf{S}}_{k,q}.
Algorithm 4 Measurement of matrix entry 𝐒k,q\mathbf{S}_{k,q}.
1:function BasisGen(𝐡\mathbf{h},𝝈\bm{\sigma},E0E_{0},τ\tau,NN,kk)
2:     Generate random time tt according to the distribution pτ,k(t)p_{\tau,k}(t).
3:     (v¯,V¯)(\bar{v},\bar{V}) \leftarrow RTE(𝐡\mathbf{h},𝝈\bm{\sigma},NN,tt)
4:     v¯v¯×ik12k12τk1Hk1(t2τ)gτ(t)eiE0t\bar{v}\leftarrow\bar{v}\times\frac{i^{k-1}}{2^{\frac{k-1}{2}}\tau^{k-1}}H_{k-1}\left(\frac{t}{\sqrt{2}\tau}\right)g_{\tau}(t)e^{iE_{0}t}
5:     Output (v¯,V¯)(\bar{v},\bar{V}).
Algorithm 5 Basis generator fkf_{k}.
1:function RTE(𝐡\mathbf{h},𝝈\bm{\sigma},NN,tt)
2:     v¯1\bar{v}\leftarrow 1
3:     V¯𝟙\bar{V}\leftarrow\mathbbm{1}
4:     for i=1i=1 to NN do
5:         (v,V)(v,V) \leftarrow LORF(𝐡\mathbf{h},𝝈\bm{\sigma},t/Nt/N)
6:         v¯v¯×v\bar{v}\leftarrow\bar{v}\times v
7:         V¯V¯×V\bar{V}\leftarrow\bar{V}\times V      
8:     Output (v¯,V¯)(\bar{v},\bar{V}).
Algorithm 6 Real-time evolution.
1:function LORF(𝐡\mathbf{h},𝝈\bm{\sigma},Δt\Delta t)
2:     Choose O\mathrm{O} from L\mathrm{L} and T\mathrm{T} with probabilities of PL(Δt)P_{\mathrm{L}}(\Delta t) and PT(Δt)P_{\mathrm{T}}(\Delta t), respectively.
3:     if O=L\mathrm{O}=\mathrm{L} then
4:         Choose jj with a probability of |hj|/htot|h_{j}|/h_{tot}.
5:         vβj(Δt)v\leftarrow\beta_{j}(\Delta t)
6:         Veisgn(hj)ϕ(Δt)σjV\leftarrow e^{-i\mathrm{sgn}(h_{j})\phi(\Delta t)\sigma_{j}}
7:     else if O=T\mathrm{O}=\mathrm{T} then
8:         while k<2k<2 do
9:              kPoisson(htot|Δt|)k\leftarrow\textsc{Poisson}(h_{tot}|\Delta t|)          
10:         v1/k!v\leftarrow 1/k!
11:         V𝟙V\leftarrow\mathbbm{1}
12:         for a=1a=1 to kk do
13:              Choose jj with a probability of |hj|/htot|h_{j}|/h_{tot}.
14:              vv×(ihjΔt)v\leftarrow v\times(-ih_{j}\Delta t)
15:              VV×σjV\leftarrow V\times\sigma_{j}               
16:     Output (v,V)(v,V).
Algorithm 7 Leading-order-rotation formula.

Appendix H Details in numerical calculation

In this section, first, we describe models and reference states taken in the benchmarking, and then we give details about instances and parameters.

H.1 Models and reference states

Two models are used in the benchmarking: the anti-ferromagnetic Heisenberg model

H=Ji,j(σiXσjX+σiYσjY+σiZσjZ),\displaystyle H=J\sum_{\langle i,j\rangle}(\sigma^{X}_{i}\sigma^{X}_{j}+\sigma^{Y}_{i}\sigma^{Y}_{j}+\sigma^{Z}_{i}\sigma^{Z}_{j}), (101)

where σiX\sigma^{X}_{i}, σiY\sigma^{Y}_{i} and σiZ\sigma^{Z}_{i} are Pauli operators of the iith qubit; and the Hubbard model

H\displaystyle H =\displaystyle= Ji,jσ=,(ai,σaj,σ+aj,σai,σ)\displaystyle-J\sum_{\langle i,j\rangle}\sum_{\sigma=\uparrow,\downarrow}(a_{i,\sigma}^{\dagger}a_{j,\sigma}+a_{j,\sigma}^{\dagger}a_{i,\sigma}) (102)
+\displaystyle+ Ui(ai,ai,𝟙2)(ai,ai,𝟙2),\displaystyle U\sum_{i}\left(a_{i,\uparrow}^{\dagger}a_{i,\uparrow}-\frac{\mathbbm{1}}{2}\right)\left(a_{i,\downarrow}^{\dagger}a_{i,\downarrow}-\frac{\mathbbm{1}}{2}\right),

where ai,σa_{i,\sigma} is the annihilation operator of the iith orbital and spin σ\sigma. For the Heisenberg model, we take the spin number as ten, and for the Hubbard model, we take the orbital number as five. Then both models can be encoded into ten qubits. For the Hubbard model, we take U=JU=J. Without loss of generality, we normalise the Hamiltonian for simplicity: We take JJ such that H2=1\|H\|_{2}=1, i.e. eigenvalues of HH are in the interval [1,1][-1,1].

For each model, we consider three types of lattices: chain, ladder and randomly generated graphs, see Fig. 5. We generate a random graph in the following way: i) For a vertex vv, randomly choose a vertex uvu\neq v and add the edge (v,u)(v,u); ii) Repeat step-i two times for the vertex vv; iii) Implement steps i and ii for each vertex vv. On a graph generated in this way, each vertex is connected to four vertices on average. Notice that some pairs of vertices are connected by multiple edges, and then the interaction between vertices in the pair is amplified by the number of edges.

Refer to caption
Figure 5: Lattices of the Heisenberg model and Hubbard model. One of the randomly generated graphs for each model is illustrated in the figure as an example.

It is non-trivial to choose and prepare a good reference state that has a sufficiently large overlap with the true ground state. Finding the ground-state energy of a local Hamiltonian is QMA-hard [59, 60]. If one can prepare a good reference state, then quantum computing is capable of solving the ground-state energy using quantum phase estimation [61]. So far, there does not exist a universal method for reference state preparation; see Ref. [62] for relevant discussions. Despite this, there are some practical ways of choosing and preparing the reference state. For fermion systems, we can take the mean-field approximate solution, i.e. a Hartree-Fock state, as the reference state. For general systems, one may try adiabatic state preparation, etc. We stress that preparing good reference states is beyond the scope of this work. Here, we choose two reference states which are likely to overlap with true ground states as examples. For the Heisenberg model, we take the pairwise singlet state

|φ\displaystyle|{\varphi}\rangle =\displaystyle= |Φ1,2|Φ3,4,\displaystyle|{\Phi}\rangle_{1,2}\otimes|{\Phi}\rangle_{3,4}\otimes\cdots, (103)

where

|Φi,j=12(|0i|1j|1i|0j)\displaystyle|{\Phi}\rangle_{i,j}=\frac{1}{\sqrt{2}}\left(|{0}\rangle_{i}\otimes|{1}\rangle_{j}-|{1}\rangle_{i}\otimes|{0}\rangle_{j}\right) (104)

is the singlet state of spins ii and jj. For the Hubbard model, we take a Hartree-Fock state as the reference state: We compute the ground state of the one-particle Hamiltonian (which is equivalent to taking U=0U=0) and take the ground state of the one-particle Hamiltonian (a Slater determinant) as the reference state.

H.2 Instances

We test the algorithms listed in Table 1 with many instances. Each instance is a triple (model,lattice,d)(model,lattice,d), in which modelmodel takes one of the two models (Heisenberg model and Hubbard model), latticelattice takes a chain, ladder or random graph, and dd is the dimension of the subspace.

Instances for computing empirical distributions in Fig. 2 consist of six groups: the Heisenberg model on the chain, ladder and random graphs, and the Hubbard model on the chain, ladder and random graphs. For chain and ladder, we evaluate each subspace dimension d=2,3,,30d=2,3,\ldots,30; For a random graph, we only evaluate one subspace dimension randomly taken in the interval. Some instances are discarded. To avoid any potential issue caused by the precision of floating-point numbers, we discard instances that the subspace error ϵK\epsilon_{K} of the P algorithm is lower than 10910^{-9} due to a large dd. We also discard instances that ϵK\epsilon_{K} of the P algorithm is higher than 10210^{-2} due to a small dd, because the dimension may be too small to achieve an interesting accuracy in this case. For randomly generated graphs, the two reference states may have a small overlap with the true ground state. Because a good reference state is necessary for all quantum KSD algorithms, we discard graphs with pg<103p_{g}<10^{-3}. Eventually, we have eight instances of the Heisenberg chain, seven instances of the Heisenberg Ladder, twenty-two instances of the Hubbard chain and twenty instances of the Hubbard ladder. For each model, we generate a hundred random-graph instances. The total number of instances is 233233.

H.3 Parameters

Abbr. E0E_{0} τ\tau Δt\Delta t ΔE\Delta E C𝐇C_{\mathbf{H}} C𝐒C_{\mathbf{S}}
P Eg+1E_{g}+1 N/A N/A N/A 11 11
CP 0 N/A N/A N/A 11 11
GP [Eg0.1,Eg+0.1][E_{g}-0.1,E_{g}+0.1] Solving Eq. (106) N/A N/A htot1h_{tot}\geqslant 1 11
IP Eg1E_{g}-1 N/A N/A N/A 11 11
ITE EgE_{g} Solving Eq. (107) N/A N/A 11 11
RTE EgE_{g} N/A Minimising ϵK\epsilon_{K} N/A 11 11
F EgE_{g} Solving Eq. (106) 2τ/(L1)2\tau/(L-1) Minimising ϵK\epsilon_{K} 11 11
Table 3: Parameters taken in the numerical calculation.

In this section, we give parameters taken in each algorithm listed in Table 1, namely, E0E_{0}, τ\tau, Δt\Delta t, ΔE\Delta E, C𝐇C_{\mathbf{H}} and C𝐒C_{\mathbf{S}}. We summarise these parameters in Table 3.

For E0E_{0}, we expect that taking E0E_{0} close to EgE_{g} is optimal for our GP algorithm. As the exact value of EgE_{g} is unknown, we assume that we have a preliminary estimation of the ground-state energy with uncertainty as large as 10%10\% of the entire spectrum, i.e. we take E0[Eg0.1,Eg+0.1]E_{0}\in[E_{g}-0.1,E_{g}+0.1]. For other algorithms, we take E0E_{0} by assuming that the exact value of EgE_{g} is known. In the P algorithm, we take E0=Eg+1E_{0}=E_{g}+1, such that the ground state is the eigenstate of HE0H-E_{0} with the largest absolute eigenvalue and HE02=1\|H-E_{0}\|_{2}=1. Similarly, in the IP algorithm, we take E0=Eg1E_{0}=E_{g}-1, such that the ground state is the eigenstate of (HE0)1(H-E_{0})^{-1} with the largest absolute eigenvalue and (HE0)12=1\|(H-E_{0})^{-1}\|_{2}=1. In the ITE algorithm, E0E_{0} causes a constant factor eτ(k1)E0e^{\tau(k-1)E_{0}} in each operator fkf_{k}, i.e. determines the spectral norm of fkf_{k}. Because the variance is related to the norm, a large E0E_{0} is problematic. Therefore, in the ITE algorithm, we take E0=EgE_{0}=E_{g}, such that fk2=1\|f_{k}\|_{2}=1 for all kk. In the F algorithm, we expect that E0=EgE_{0}=E_{g} is optimal because f1f_{1} is an energy filter centred at the ground-state energy in this case. Therefore, we take E0=EgE_{0}=E_{g} in the F algorithm. The RTE algorithm is closely related to the F algorithm. Therefore, we also take E0=EgE_{0}=E_{g} in the RTE algorithm.

We remark that in the GP algorithm, we take random E0E_{0} uniformly distributed in the interval [Eg0.1,Eg+0.1][E_{g}-0.1,E_{g}+0.1] to generate data in Fig. 2, and we take E0=Eg+iδEE_{0}=-E_{g}+i\delta E, where i=50,9,,1,0,1,,9,50i=-50,-9,\ldots,-1,0,1,\ldots,9,50 and δE=0.002\delta E=0.002, to generate data in Fig. 1. We remark that we have normalised the Hamiltonian such that H2=1\|H\|_{2}=1.

For Δt\Delta t in the F algorithm, we take Δt=2τ/(L1)\Delta t=2\tau/(L-1). For simplicity, we take the limit L+L\rightarrow+\infty, i.e.

fk\displaystyle f_{k} =\displaystyle= 12τττ𝑑tei[HE0ΔE(k1)]t\displaystyle\frac{1}{2\tau}\int_{-\tau}^{\tau}dte^{-i\left[H-E_{0}-\Delta E\left(k-1\right)\right]t} (105)
=\displaystyle= sin[HE0ΔE(k1)]τ[HE0ΔE(k1)]τ.\displaystyle\frac{\sin\left[H-E_{0}-\Delta E\left(k-1\right)\right]\tau}{\left[H-E_{0}-\Delta E\left(k-1\right)\right]\tau}.

Notice that this fkf_{k} is an energy filter centred at E0+ΔE(k1)E_{0}+\Delta E\left(k-1\right), and the filter is narrower when τ\tau is larger.

Now, we have three algorithms that have the parameter τ\tau: GP, ITE and F. Similar to the F algorithm, f1f_{1} in the GP algorithm is an energy filter centred at E0E_{0}. For these two algorithms, if E0=EgE_{0}=E_{g}, f1|φf_{1}|{\varphi}\rangle converges to the true ground state in the limit τ+\tau\rightarrow+\infty. It is also similar in the ITE algorithm, in which fdf_{d} is a projector onto the ground state, and fd|φf_{d}|{\varphi}\rangle converges to the true ground state in the limit τ+\tau\rightarrow+\infty. Realising a large τ\tau is at a certain cost, specifically, the circuit depth increases with τ\tau [42, 8]. Therefore, it is reasonable to take a finite τ\tau. Next, we give the protocol for determining the value of τ\tau in each algorithm.

Without getting into details about realising the filters and projectors, we choose τ\tau under the assumption that if filters and projectors have the same power in computing the ground-state energy, they probably require similar circuit depths. In GP and F algorithms, if E0=EgE_{0}=E_{g}, the energy error achieved by filters reads 𝐇1,1/𝐒1,1Eg\mathbf{H}_{1,1}/\mathbf{S}_{1,1}-E_{g}; and in the ITE algorithm, the energy error achieved by the projector reads 𝐇d,d/𝐒d,dEg\mathbf{H}_{d,d}/\mathbf{S}_{d,d}-E_{g}. We take τ\tau such that errors achieved by filters and projectors take the same value ϵB\epsilon_{B}. Specifically, for the GP and F algorithms, we determine the value of τ\tau by solving the equation (taking E0=EgE_{0}=E_{g})

𝐇1,1(τ)𝐒1,1(τ)Eg=ϵB;\displaystyle\frac{\mathbf{H}_{1,1}(\tau)}{\mathbf{S}_{1,1}(\tau)}-E_{g}=\epsilon_{B}; (106)

and for the ITE algorithm, we determine the value of τ\tau by solving the equation

𝐇d,d(τ)𝐒d,d(τ)Eg=ϵB.\displaystyle\frac{\mathbf{H}_{d,d}(\tau)}{\mathbf{S}_{d,d}(\tau)}-E_{g}=\epsilon_{B}. (107)

To choose the value of ϵB\epsilon_{B}, we take the P algorithm as the standard, in which fdf_{d} is a projector onto the ground state, i.e. fd|φf_{d}|{\varphi}\rangle converges to the true ground state in the limit d+d\rightarrow+\infty. Overall, we determine the value of τ\tau in the following way: Given an instance (model,lattice,d)(model,lattice,d), i) first, compute the energy error achieved by the projector in the P algorithm, take ϵB=𝐇d,d/𝐒d,dEg\epsilon_{B}=\mathbf{H}_{d,d}/\mathbf{S}_{d,d}-E_{g}; then, ii) solve equations of τ\tau. In this way, filters and projectors in P, GP, ITE and F algorithms have the same power.

For Δt\Delta t in the RTE algorithm and ΔE\Delta E in the F algorithm, there are works on how to choose them in the literature [14, 17, 15]. In this work, we simply determine their values by a grid search. For the RTE algorithm, we take Δt=iδt\Delta t=i\delta t, where i=1,2,,100i=1,2,\ldots,100 and δt=2π100\delta t=\frac{2\pi}{100}; we compute the subspace error ϵK\epsilon_{K} for all Δt\Delta t; and we choose Δt\Delta t of the minimum ϵK\epsilon_{K}. For the F algorithm, we take ΔE=iδE\Delta E=i\delta E, where i=1,2,,100i=1,2,\ldots,100 and δE=2100d\delta E=\frac{2}{100d} (In this way, when we take the largest ΔE\Delta E, filters span the entire spectrum); we compute the subspace error ϵK\epsilon_{K} for all ΔE\Delta E; and we choose ΔE\Delta E of the minimum ϵK\epsilon_{K}.

For C𝐇C_{\mathbf{H}} and C𝐒C_{\mathbf{S}}, we take C𝐇=htotC_{\mathbf{H}}=h_{tot} and C𝐒=1C_{\mathbf{S}}=1 in our GP algorithm following the analysis in Appendix G. For other algorithms, we take these two parameters in the following way. For P, IP, ITE and RTE algorithms, we take C𝐇C_{\mathbf{H}} and C𝐒C_{\mathbf{S}} according to spectral norms (the lower bound of cost) without getting into details about measuring matrix entries. In these four algorithms, fkHfq2=fkfq2=1\|f_{k}^{\dagger}Hf_{q}\|_{2}=\|f_{k}^{\dagger}f_{q}\|_{2}=1 for all kk and qq, therefore, we take C𝐇=C𝐒=1C_{\mathbf{H}}=C_{\mathbf{S}}=1. In the CP algorithm, we take C𝐇C_{\mathbf{H}} and C𝐒C_{\mathbf{S}} in a similar way. Because H/htot21\|H/h_{tot}\|_{2}\leqslant 1, the spectrum of H/htotH/h_{tot} is in the interval [1,1][-1,1]. In this interval, the Chebyshev polynomial of the first kind takes values in the interval [1,1][-1,1]. Therefore, Tk(H/htot)21\|T_{k}(H/h_{tot})\|_{2}\leqslant 1, and consequently fkHfq2,fkfq21\|f_{k}^{\dagger}Hf_{q}\|_{2},\|f_{k}^{\dagger}f_{q}\|_{2}\leqslant 1. These norms depend on the spectrum of HH. For simplicity, we take the upper bound of norms, i.e. C𝐇=C𝐒=1C_{\mathbf{H}}=C_{\mathbf{S}}=1, in the CP algorithm. In the F algorithm, each fkf_{k} is a linear combination of eiHte^{-iHt} operators, i.e. fkf_{k} is expressed in the LCU form, and the cost factor of the LCU expression is one. Therefore, we take C𝐇=C𝐒=1C_{\mathbf{H}}=C_{\mathbf{S}}=1 in the F algorithm, assuming that HH is expressed in the LCU form with a cost factor of one (Notice that one is the lower bound).

Appendix I Derivation of Eq. (21)

Define the function f(η)=E(η,𝐚)=𝐚𝐇𝐚+2C𝐇η𝐚𝐚𝐚𝐒𝐚+2C𝐒η𝐚𝐚f(\eta)=E^{\prime}(\eta,\mathbf{a})=\frac{\mathbf{a}^{\dagger}\mathbf{H}\mathbf{a}+2C_{\mathbf{H}}\eta\mathbf{a}^{\dagger}\mathbf{a}}{\mathbf{a}^{\dagger}\mathbf{S}\mathbf{a}+2C_{\mathbf{S}}\eta\mathbf{a}^{\dagger}\mathbf{a}}. Consider a small η\eta and the Taylor expansion of f(η)f(\eta). We have f(0)=𝐚𝐇𝐚𝐚𝐒𝐚f(0)=\frac{\mathbf{a}^{\dagger}\mathbf{H}\mathbf{a}}{\mathbf{a}^{\dagger}\mathbf{S}\mathbf{a}} and

f(η)\displaystyle f^{\prime}(\eta) =\displaystyle= 2𝐚𝐚(C𝐇𝐚𝐒𝐚C𝐒𝐚𝐇𝐚)(𝐚𝐒𝐚+2C𝐒η𝐚𝐚)2,\displaystyle\frac{2\mathbf{a}^{\dagger}\mathbf{a}(C_{\mathbf{H}}\mathbf{a}^{\dagger}\mathbf{S}\mathbf{a}-C_{\mathbf{S}}\mathbf{a}^{\dagger}\mathbf{H}\mathbf{a})}{(\mathbf{a}^{\dagger}\mathbf{S}\mathbf{a}+2C_{\mathbf{S}}\eta\mathbf{a}^{\dagger}\mathbf{a})^{2}},
f(0)\displaystyle f^{\prime}(0) =\displaystyle= 2𝐚𝐚(C𝐇𝐚𝐒𝐚C𝐒𝐚𝐇𝐚)(𝐚𝐒𝐚)2\displaystyle\frac{2\mathbf{a}^{\dagger}\mathbf{a}(C_{\mathbf{H}}\mathbf{a}^{\dagger}\mathbf{S}\mathbf{a}-C_{\mathbf{S}}\mathbf{a}^{\dagger}\mathbf{H}\mathbf{a})}{(\mathbf{a}^{\dagger}\mathbf{S}\mathbf{a})^{2}} (108)
=\displaystyle= 𝐚𝐚𝐚𝐒𝐚2(C𝐇C𝐒𝐚𝐇𝐚𝐚𝐒𝐚).\displaystyle\frac{\mathbf{a}^{\dagger}\mathbf{a}}{\mathbf{a}^{\dagger}\mathbf{S}\mathbf{a}}2\left(C_{\mathbf{H}}-C_{\mathbf{S}}\frac{\mathbf{a}^{\dagger}\mathbf{H}\mathbf{a}}{\mathbf{a}^{\dagger}\mathbf{S}\mathbf{a}}\right).

The minimum value of the zeroth-order term 𝐚𝐇𝐚𝐚𝐒𝐚\frac{\mathbf{a}^{\dagger}\mathbf{H}\mathbf{a}}{\mathbf{a}^{\dagger}\mathbf{S}\mathbf{a}} is EminE_{min} cccording to the Rayleigh quotient theorem [51]. Take this minimum value, we have EEmin+sηE^{\prime}\simeq E_{min}+s\eta, where s=f(0)𝐚𝐚𝐚𝐒𝐚s=f^{\prime}(0)\propto\frac{\mathbf{a}^{\dagger}\mathbf{a}}{\mathbf{a}^{\dagger}\mathbf{S}\mathbf{a}} is the first derivative. The solution to E=Eg+ϵE^{\prime}=E_{g}+\epsilon is η(Eg+ϵEmin)/s\eta\simeq(E_{g}+\epsilon-E_{min})/s, then

γ\displaystyle\gamma \displaystyle\simeq ϵ216H224(C𝐇C𝐒𝐚𝐇𝐚𝐚𝐒𝐚)2(Eg+ϵEmin)2(pg𝐚𝐚𝐚𝐒𝐚)2\displaystyle\frac{\epsilon^{2}}{16\|H\|_{2}^{2}}\frac{4\left(C_{\mathbf{H}}-C_{\mathbf{S}}\frac{\mathbf{a}^{\dagger}\mathbf{H}\mathbf{a}}{\mathbf{a}^{\dagger}\mathbf{S}\mathbf{a}}\right)^{2}}{(E_{g}+\epsilon-E_{min})^{2}}\left(\frac{p_{g}\mathbf{a}^{\dagger}\mathbf{a}}{\mathbf{a}^{\dagger}\mathbf{S}\mathbf{a}}\right)^{2} (109)
=\displaystyle= u2(pg𝐚𝐚𝐚𝐒𝐚)2,\displaystyle u^{2}\left(\frac{p_{g}\mathbf{a}^{\dagger}\mathbf{a}}{\mathbf{a}^{\dagger}\mathbf{S}\mathbf{a}}\right)^{2},

where

u\displaystyle u =\displaystyle= ϵ|C𝐇C𝐒𝐚𝐇𝐚𝐚𝐒𝐚|2H2|Eg+ϵEmin|\displaystyle\frac{\epsilon\left|C_{\mathbf{H}}-C_{\mathbf{S}}\frac{\mathbf{a}^{\dagger}\mathbf{H}\mathbf{a}}{\mathbf{a}^{\dagger}\mathbf{S}\mathbf{a}}\right|}{2\|H\|_{2}\left|E_{g}+\epsilon-E_{min}\right|} (110)
\displaystyle\leqslant (C𝐇+H2C𝐒)ϵ2H2(ϵϵK)C𝐒\displaystyle\frac{(C_{\mathbf{H}}+\|H\|_{2}C_{\mathbf{S}})\epsilon}{2\|H\|_{2}(\epsilon-\epsilon_{K})}\approx C_{\mathbf{S}}

under assumptions C𝐇H2C𝐒C_{\mathbf{H}}\approx\|H\|_{2}C_{\mathbf{S}} and ϵϵKϵ\epsilon-\epsilon_{K}\approx\epsilon.

Appendix J Composing Chebyshev polynomials

J.1 Chebyshev-polynomial projector

In this section, we explain how to use Chebyshev polynomials as projectors onto the ground state.

The nnth Chebyshev polynomial of the first kind is

Tn(z)={cos(narccosz),if |z|1,sgn(z)ncosh(narccosh|z|),if |z|>1.\displaystyle T_{n}(z)=\left\{\begin{array}[]{cc}\cos(n\arccos z),&\text{if }|z|\leqslant 1,\\ \mathrm{sgn}(z)^{n}\cosh(n\,\mathrm{arccosh}|z|),&\text{if }|z|>1.\end{array}\right. (113)

Chebyshev polynomials have the following properties: i) When |z|1|z|\leqslant 1, |Tn(z)|1|T_{n}(z)|\leqslant 1; and ii) when |z|>1|z|>1, |Tn(z)|>(|z|n+|z|n)/2|T_{n}(z)|>(|z|^{n}+|z|^{-n})/2 increases exponentially with nn.

Let’s consider the spectral decomposition of the Hamiltonian, H=m=1dEm|ψmψm|H=\sum_{m=1}^{d_{\mathcal{H}}}E_{m}|\psi_{m}\rangle\langle\psi_{m}|. Here, dd_{\mathcal{H}} is the dimension of the Hilbert space, EmE_{m} are eigenenergies of the Hamiltonian, and |ψm|{\psi_{m}}\rangle are eigenstates. We suppose that eigenenergies are sorted in ascending order, i.e. EiEjE_{i}\leqslant E_{j} if i<ji<j. Then, E1=EgE_{1}=E_{g} is the ground-state energy (accordingly, |ψ1=|ψg|{\psi_{1}}\rangle=|{\psi_{g}}\rangle is the ground state) and E2E_{2} is the energy of the first excited state. The energy gap between the ground state and the first excited state is Δ=E2E1\Delta=E_{2}-E_{1}.

To compose a projector, we take

Z=1HE2H2.\displaystyle Z=1-\frac{H-E_{2}}{\|H\|_{2}}. (114)

Notice that ZZ and HH have the same eigenstates. The spectral decomposition of ZZ is Z=m=1dzm|ψmψm|Z=\sum_{m=1}^{d_{\mathcal{H}}}z_{m}|\psi_{m}\rangle\langle\psi_{m}|, where

zm=1EmE2H2.\displaystyle z_{m}=1-\frac{E_{m}-E_{2}}{\|H\|_{2}}. (115)

Then, the ground state corresponds to z1=1+ΔH2>1z_{1}=1+\frac{\Delta}{\|H\|_{2}}>1 (suppose the gap is finite), and the first excited state corresponds to z2=1z_{2}=1. Because |EmE2|2H2|E_{m}-E_{2}|\leqslant 2\|H\|_{2}, zm1z_{m}\geqslant-1 for all mm. Therefore, except the ground state, zmz_{m} of all excited states (i.e. m2m\geqslant 2) is in the interval [1,1][-1,1].

The projector reads

Tn(Z)Tn(z1)\displaystyle\frac{T_{n}(Z)}{T_{n}(z_{1})} =\displaystyle= m=1dTn(zm)Tn(z1)|ψmψm|\displaystyle\sum_{m=1}^{d_{\mathcal{H}}}\frac{T_{n}(z_{m})}{T_{n}(z_{1})}|\psi_{m}\rangle\langle\psi_{m}| (116)
=\displaystyle= |ψgψg|+Ω,\displaystyle|\psi_{g}\rangle\langle\psi_{g}|+\Omega,

where

Ω\displaystyle\Omega =\displaystyle= m=2dTn(zm)Tn(z1)|ψmψm|.\displaystyle\sum_{m=2}^{d_{\mathcal{H}}}\frac{T_{n}(z_{m})}{T_{n}(z_{1})}|\psi_{m}\rangle\langle\psi_{m}|. (117)

Ω\Omega and HH have the same eigenstates. Because |Tn(zm)|1|T_{n}(z_{m})|\leqslant 1 when m2m\geqslant 2,

Ω21Tn(z1).\displaystyle\|\Omega\|_{2}\leqslant\frac{1}{T_{n}(z_{1})}. (118)

The key in using Chebyshev polynomials as projectors is laying all excited states in the interval z[1,1]z\in[-1,1] and leaving the ground state out of the interval. For the CP basis, all eigenstates are in the interval z[1,1]z\in[-1,1]. Therefore, fkf_{k} operators of the CP basis are not projectors in general.

J.2 Expanding the projector as Hamiltonian powers

In this section, we expand the Chebyshev-polynomial projector as a linear combination of Hamiltonian powers (HE0)k1(H-E_{0})^{k-1}. We focus on the case that E0=EgE_{0}=E_{g}.

The explicit expression of Tn(z)T_{n}(z) (n>0n>0) is

Tn(z)\displaystyle T_{n}(z) =\displaystyle= nm=0n(2)m(n+m1)!(nm)!(2m)!\displaystyle n\sum_{m=0}^{n}(-2)^{m}\frac{(n+m-1)!}{(n-m)!(2m)!} (119)
×(1z)m.\displaystyle\times(1-z)^{m}.

Because

(1Z)m\displaystyle(1-Z)^{m} =\displaystyle= l=0mm!(ml)!l!\displaystyle\sum_{l=0}^{m}\frac{m!}{(m-l)!l!} (120)
×(E0E2)ml(HE0)lH2m,\displaystyle\times\frac{(E_{0}-E_{2})^{m-l}(H-E_{0})^{l}}{\|H\|_{2}^{m}},

we have

Tn(Z)=l=0nbl(HE0)lH2l,\displaystyle T_{n}(Z)=\sum_{l=0}^{n}b_{l}\frac{(H-E_{0})^{l}}{\|H\|_{2}^{l}}, (121)

where

bl\displaystyle b_{l} =\displaystyle= nm=ln(2)m(n+m1)!(nm)!(2m)!m!(ml)!l!\displaystyle n\sum_{m=l}^{n}(-2)^{m}\frac{(n+m-1)!}{(n-m)!(2m)!}\frac{m!}{(m-l)!l!} (122)
×(E0E2)mlH2ml.\displaystyle\times\frac{(E_{0}-E_{2})^{m-l}}{\|H\|_{2}^{m-l}}.

Now, we give an upper bound of |bl||b_{l}|. First, we consider l=0l=0. In this case,

|b0|\displaystyle|b_{0}| \displaystyle\leqslant nm=0n2m(n+m1)!(nm)!(2m)!|E0E2|mH2m\displaystyle n\sum_{m=0}^{n}2^{m}\frac{(n+m-1)!}{(n-m)!(2m)!}\frac{|E_{0}-E_{2}|^{m}}{\|H\|_{2}^{m}} (123)
=\displaystyle= Tn(z1).\displaystyle T_{n}(z_{1}).

Then, for an arbitrary ll,

|bl|\displaystyle|b_{l}| \displaystyle\leqslant nlTn(z1),\displaystyle n^{l}T_{n}(z_{1}), (124)

where the inequality

m!(ml)!l!\displaystyle\frac{m!}{(m-l)!l!} \displaystyle\leqslant mlnl\displaystyle m^{l}\leqslant n^{l} (125)

has been used.

Finally, taking d=n+1d=n+1 and

ak=ckbk1Tn(z1)H2k1,\displaystyle a_{k}=\frac{c_{k}b_{k-1}}{T_{n}(z_{1})\|H\|_{2}^{k-1}}, (126)

we have

k=1dakck1fk=Tn(Z)Tn(z1)e12(HE0)2τ2,\displaystyle\sum_{k=1}^{d}a_{k}c_{k}^{-1}f_{k}=\frac{T_{n}(Z)}{T_{n}(z_{1})}e^{-\frac{1}{2}(H-E_{0})^{2}\tau^{2}}, (127)

where fkf_{k} are operators defined in Eq. (13). Because of the upper bound

|ak|\displaystyle|a_{k}| \displaystyle\leqslant 2(k1eτ2)k12(nH2)k1\displaystyle 2\left(\frac{k-1}{e\tau^{2}}\right)^{\frac{k-1}{2}}\left(\frac{n}{\|H\|_{2}}\right)^{k-1} (128)
\displaystyle\leqslant 2(n3eH22τ2)k12,\displaystyle 2\left(\frac{n^{3}}{e\|H\|_{2}^{2}\tau^{2}}\right)^{\frac{k-1}{2}},

the overhead factor is

γ\displaystyle\gamma \displaystyle\approx (𝐚𝐚)24k=1d(n3eH22τ2)k1\displaystyle\left(\mathbf{a}^{\dagger}\mathbf{a}\right)^{2}\leqslant 4\sum_{k=1}^{d}\left(\frac{n^{3}}{e\|H\|_{2}^{2}\tau^{2}}\right)^{k-1} (129)
\displaystyle\leqslant 411n3/(eH22τ2).\displaystyle 4\frac{1}{1-n^{3}/(e\|H\|_{2}^{2}\tau^{2})}.

Appendix K The choice of τ\tau

There are two bounds determining the range of τ\tau. First, we take τ>d1e\tau>\sqrt{\frac{d-1}{e}}, such that the upper bound of fk2\|f_{k}\|_{2} decrease exponentially with kk. Second, under the assumption |E0Eg|ϵ0|E_{0}-E_{g}|\leqslant\epsilon_{0}, we take τd1ϵ0\tau\leqslant\frac{\sqrt{d-1}}{\epsilon_{0}}. To justify the second bound, we plot the absolute value of the rescaled Gaussian-power function |fk(x)|=|fk(x)|/ck|f^{\prime}_{k}(x)|=|f_{k}(x)|/c_{k} with different kk in Fig. 6, where the instance (Heisenberg,chain,d=5)(\mathrm{Heisenberg},\mathrm{chain},d=5) is taken as an example. This function has maximum points at (±k1τ,(k1eτ2)k12/ck)\left(\pm\frac{\sqrt{k-1}}{\tau},(\frac{k-1}{e\tau^{2}})^{\frac{k-1}{2}}/c_{k}\right). When kk increases from 11 to dd, the peaks of fk(x)f^{\prime}_{k}(x) span in the range [d1τ,d1τ]\left[-\frac{\sqrt{d-1}}{\tau},\frac{\sqrt{d-1}}{\tau}\right], making the basis a filter [2]. It is preferred that the ground-state energy is in the range, which leads to the second bound. Notice that we can always normalise the Hamiltonian by taking HH/htotH\leftarrow H/h_{tot}. Then, H21\|H\|_{2}\leqslant 1, |E0Eg|1|E_{0}-E_{g}|\leqslant 1 for all E0[1,0]E_{0}\in[-1,0], and the two bounds are consistent.

Refer to caption
Figure 6: The absolute value of the rescaled Gaussian-power function |fk(x)|=|xk1ex2τ2/2|/ck|f^{\prime}_{k}(x)|=|x^{k-1}e^{-x^{2}\tau^{2}/2}|/c_{k} with k=1k=1 to dd and τ=5\tau=5. Here, we take the instance (Heisenberg,chain,d=5)(\mathrm{Heisenberg},\mathrm{chain},d=5) as an example.

Appendix L Numerical results of the measurement number

In Theorem 1, we give a theoretical result about the sufficient measurement number, which is used in later numerical study. In this section, we numerically calculate the necessary measurement number and compare the regularisation method to the thresholding method.

L.1 Necessary measurement costs

Due to the statistical error, the output energy E^min\hat{E}_{min} follows a certain distribution in the vicinity of the true ground-state energy EgE_{g}. Given the permissible energy error ϵ\epsilon and failure probability κ\kappa, a measurement number is said to be sufficient if E^min\hat{E}_{min} is in the range [Egϵ,Eg+ϵ][E_{g}-\epsilon,E_{g}+\epsilon] with a probability not lower than 1κ1-\kappa; and the necessary measurement number is the minimum sufficient measurement number.

Taking the instance (Heisenberg,chain,d=5)(\mathrm{Heisenberg},\mathrm{chain},d=5) as an example, Fig. 7 shows the errors in the ground-state energy and the corresponding necessary/sufficient measurement numbers. We can find that the sufficient measurement costs are close to the necessary measurement numbers.

Refer to caption
Figure 7: The variational ground-state energy errors ϵ\epsilon and the corresponding measurement numbers MM for the quantum KSD algorithms listed in Table 1. The instance is (Heisenberg,chain,d=5)(\mathrm{Heisenberg},\mathrm{chain},d=5) and Eg=1E_{g}=-1. The solid curves indicate the sufficient measurement numbers computed through Theorem 1. The dots denote errors at various measurement numbers computed following Algorithm 1. At each measurement number MM, Algorithm 1 is implemented for 100100 times (simulated on a classical computer), and (M,|E^minEg|)(M,|\hat{E}_{min}-E_{g}|) of each implementation is plotted in the figure. The dashed lines are taken such that 10%10\% dots are above them, i.e. the dashed lines corresponding to the necessary measurement number for the failure probability κ=0.1\kappa=0.1. For the GP algorithm, the solid and dashed lines represent the result of the GP algorithm with E0=EgE_{0}=E_{g}. The light blue area illustrates the range of necessary measurement cost when we take E0[Eg0.1,Eg+0.1]E_{0}\in[E_{g}-0.1,E_{g}+0.1] in the GP algorithm.

When estimating the necessary measurement numbers, we need to numerically simulate the estimators 𝐇^\hat{\mathbf{H}} and 𝐒^\hat{\mathbf{S}}, which can be regarded as the summations of the matrices 𝐇\mathbf{H} and 𝐒\mathbf{S} and random noise matrices 𝐇^𝐇\hat{\mathbf{H}}-\mathbf{H} and 𝐒^𝐒\hat{\mathbf{S}}-\mathbf{S}, respectively. Considering Gaussian statistical error and the collective measurement protocol, the random matrices 𝐇^𝐇\hat{\mathbf{H}}-\mathbf{H} and 𝐒^𝐒\hat{\mathbf{S}}-\mathbf{S} can be numerically generated. The real and imaginary parts of the random matrix entries obey the Gaussian distribution with the mean value 0 (assume the estimation is unbiased for all algorithms) and the standard deviations are C𝐇/MC_{\mathbf{H}}/\sqrt{M} and C𝐒/MC_{\mathbf{S}}/\sqrt{M}, respectively. Under collective measurement protocols, the random matrices have the same structure as 𝐇\mathbf{H} and 𝐒\mathbf{S}.

L.2 Comparison using the thresholding method

Ref. [16] provides a detailed analysis of the thresholding procedure. The optimal thresholds are roughly close to the noise rate. Here, we set them to 10C𝐒/M10C_{\mathbf{S}}/\sqrt{M} to get relatively stable results. With the same instance (Heisenberg,chain,d=5)(\mathrm{Heisenberg},\mathrm{chain},d=5), the results shown in Fig. 8 are similar to Fig. 7. This indicates that regardless of which method is used to deal with the ill-conditioned problem, GP algorithm outperforms other algorithms in terms of measurement costs.

Refer to caption
Figure 8: The absolute energy error ϵ\epsilon and the corresponding necessary measurement numbers MM when using the thresholding method. The instance is (Heisenberg,chain,d=5)(\mathrm{Heisenberg},\mathrm{chain},d=5) and Eg=1E_{g}=-1. Similar to Fig. 7, the dashed lines indicate the necessary measurement costs with κ=0.1\kappa=0.1. The blue dashed line represents the result of the GP algorithm with E0=EgE_{0}=E_{g}. The light blue area illustrates the range of necessary measurement cost when we take E0[Eg0.1,Eg+0.1]E_{0}\in[E_{g}-0.1,E_{g}+0.1] in the GP algorithm.
Refer to caption
Figure 9: The estimated measurement number MM and the corresponding energy error ϵ\epsilon at different dd for quantum Krylov subspace diagonalisation algorithms. We take the 1010-qubit Heisenberg chain model as an example. The dashed lines indicate the 1/M1/\sqrt{M} scaling. The failure rate κ=0.1\kappa=0.1. For the GP basis, E0E_{0} is randomly chosen in [Eg0.1,Eg+0.1][E_{g}-0.1,E_{g}+0.1]. For other bases, we use optimal parameters and minimum cost factors.

References

  • [1] Elbio Dagotto. “Correlated electrons in high-temperature superconductors”. Reviews of Modern Physics 66, 763 (1994).
  • [2] Michael R Wall and Daniel Neuhauser. “Extraction, through filter-diagonalization, of general quantum eigenvalues or classical normal mode frequencies from a small number of residues or a short-time segment of a signal. i. theory and application to a quantum-dynamics model”. The Journal of chemical physics 102, 8011–8022 (1995).
  • [3] Etienne Caurier, Gabriel Martínez-Pinedo, Fréderic Nowacki, Alfredo Poves, and AP Zuker. “The shell model as a unified view of nuclear structure”. Reviews of modern Physics 77, 427 (2005).
  • [4] Cornelius Lanczos. “An iteration method for the solution of the eigenvalue problem of linear differential and integral operators”. Journal of research of the National Bureau of Standards 45, 255–282 (1950).
  • [5] Yousef Saad. “Numerical methods for large eigenvalue problems”. Society for Industrial and Applied Mathematics.  (2011).
  • [6] Åke Björck. “Numerical methods in matrix computations”. Volume 59 of Texts in Applied Mathematics. Springer.  (2015).
  • [7] Adolfo Avella and Ferdinando Mancini. “Strongly correlated systems”. Volume 176 of Springer Series in Solid-State Sciences. Springer.  (2013).
  • [8] Mario Motta, Chong Sun, Adrian T. K. Tan, Matthew J. O’Rourke, Erika Ye, Austin J. Minnich, Fernando G. S. L. Brandão, and Garnet Kin-Lic Chan. “Determining eigenstates and thermal states on a quantum computer using quantum imaginary time evolution”. Nature Physics 16, 205–210 (2019).
  • [9] Kübra Yeter-Aydeniz, Raphael C Pooser, and George Siopsis. “Practical quantum computation of chemical and nuclear energy levels using quantum imaginary time evolution and lanczos algorithms”. npj Quantum Information 6, 1–8 (2020).
  • [10] Robert M. Parrish and Peter L. McMahon. “Quantum filter diagonalization: Quantum eigendecomposition without full quantum phase estimation” (2019). arXiv:1909.08925.
  • [11] Nicholas H. Stair, Renke Huang, and Francesco A. Evangelista. “A multireference quantum krylov algorithm for strongly correlated electrons”. Journal of Chemical Theory and Computation 16, 2236–2245 (2020).
  • [12] Tatiana A Bespalova and Oleksandr Kyriienko. “Hamiltonian operator approximation for energy measurement and ground-state preparation”. PRX Quantum 2, 030318 (2021).
  • [13] Jeffrey Cohn, Mario Motta, and Robert M Parrish. “Quantum filter diagonalization with compressed double-factorized hamiltonians”. PRX Quantum 2, 040352 (2021).
  • [14] Katherine Klymko, Carlos Mejuto-Zaera, Stephen J Cotton, Filip Wudarski, Miroslav Urbanek, Diptarka Hait, Martin Head-Gordon, K Birgitta Whaley, Jonathan Moussa, Nathan Wiebe, et al. “Real-time evolution for ultracompact hamiltonian eigenstates on quantum hardware”. PRX Quantum 3, 020323 (2022).
  • [15] Cristian L Cortes and Stephen K Gray. “Quantum krylov subspace algorithms for ground-and excited-state energy estimation”. Physical Review A 105, 022417 (2022).
  • [16] Ethan N Epperly, Lin Lin, and Yuji Nakatsukasa. “A theory of quantum subspace diagonalization”. SIAM Journal on Matrix Analysis and Applications 43, 1263–1290 (2022).
  • [17] Yizhi Shen, Katherine Klymko, James Sud, David B. Williams-Young, Wibe A. de Jong, and Norm M. Tubman. “Real-Time Krylov Theory for Quantum Computing Algorithms”. Quantum 7, 1066 (2023).
  • [18] Oleksandr Kyriienko. “Quantum inverse iteration algorithm for programmable quantum simulators”. npj Quantum Information 6, 1–8 (2020).
  • [19] Kazuhiro Seki and Seiji Yunoki. “Quantum power method by a superposition of time-evolved states”. PRX Quantum 2, 010333 (2021).
  • [20] William Kirby, Mario Motta, and Antonio Mezzacapo. “Exact and efficient Lanczos method on a quantum computer”. Quantum 7, 1018 (2023).
  • [21] Ryan Babbush, Craig Gidney, Dominic W. Berry, Nathan Wiebe, Jarrod McClean, Alexandru Paler, Austin Fowler, and Hartmut Neven. “Encoding electronic spectra in quantum circuits with linear t complexity”. Phys. Rev. X 8, 041015 (2018).
  • [22] Joonho Lee, Dominic W. Berry, Craig Gidney, William J. Huggins, Jarrod R. McClean, Nathan Wiebe, and Ryan Babbush. “Even more efficient quantum computations of chemistry through tensor hypercontraction”. PRX Quantum 2, 030305 (2021).
  • [23] Alberto Peruzzo, Jarrod McClean, Peter Shadbolt, Man-Hong Yung, Xiao-Qi Zhou, Peter J Love, Alán Aspuru-Guzik, and Jeremy L O’brien. “A variational eigenvalue solver on a photonic quantum processor”. Nature communications 5, 4213 (2014).
  • [24] Jarrod R McClean, Sergio Boixo, Vadim N Smelyanskiy, Ryan Babbush, and Hartmut Neven. “Barren plateaus in quantum neural network training landscapes”. Nature communications 9, 4812 (2018).
  • [25] Beresford N. Parlett. “The symmetric eigenvalue problem”. Society for Industrial and Applied Mathematics.  (1998).
  • [26] Robert M Parrish, Edward G Hohenstein, Peter L McMahon, and Todd J Martínez. “Quantum computation of electronic transitions using a variational quantum eigensolver”. Physical review letters 122, 230401 (2019).
  • [27] Ken M Nakanishi, Kosuke Mitarai, and Keisuke Fujii. “Subspace-search variational quantum eigensolver for excited states”. Physical Review Research 1, 033062 (2019).
  • [28] William J Huggins, Joonho Lee, Unpil Baek, Bryan O’Gorman, and K Birgitta Whaley. “A non-orthogonal variational quantum eigensolver”. New Journal of Physics 22, 073009 (2020).
  • [29] Nicholas H. Stair, Cristian L. Cortes, Robert M. Parrish, Jeffrey Cohn, and Mario Motta. “Stochastic quantum krylov protocol with double-factorized hamiltonians”. Phys. Rev. A 107, 032414 (2023).
  • [30] Jarrod R. McClean, Mollie E. Kimchi-Schwartz, Jonathan Carter, and Wibe A. de Jong. “Hybrid quantum-classical hierarchy for mitigation of decoherence and determination of excited states”. Phys. Rev. A 95, 042308 (2017).
  • [31] Jarrod R McClean, Zhang Jiang, Nicholas C Rubin, Ryan Babbush, and Hartmut Neven. “Decoding quantum errors with subspace expansions”. Nature communications 11, 636 (2020).
  • [32] Nobuyuki Yoshioka, Hideaki Hakoshima, Yuichiro Matsuzaki, Yuuki Tokunaga, Yasunari Suzuki, and Suguru Endo. “Generalized quantum subspace expansion”. Phys. Rev. Lett. 129, 020502 (2022).
  • [33] Artur K Ekert, Carolina Moura Alves, Daniel KL Oi, Michał Horodecki, Paweł Horodecki, and Leong Chuan Kwek. “Direct estimations of linear and nonlinear functionals of a quantum state”. Physical review letters 88, 217901 (2002).
  • [34] Mingxia Huo and Ying Li. “Error-resilient monte carlo quantum simulation of imaginary time”. Quantum 7, 916 (2023).
  • [35] Pei Zeng, Jinzhao Sun, and Xiao Yuan. “Universal quantum algorithmic cooling on a quantum computer” (2021). arXiv:2109.15304.
  • [36] Andrew M Childs and Nathan Wiebe. “Hamiltonian simulation using linear combinations of unitary operations”. Quantum Information & Computation 12, 901–924 (2012).
  • [37] Paul K. Faehrmann, Mark Steudtner, Richard Kueng, Maria Kieferova, and Jens Eisert. “Randomizing multi-product formulas for Hamiltonian simulation”. Quantum 6, 806 (2022).
  • [38] Thomas E O’Brien, Stefano Polla, Nicholas C Rubin, William J Huggins, Sam McArdle, Sergio Boixo, Jarrod R McClean, and Ryan Babbush. “Error mitigation via verified phase estimation”. PRX Quantum 2, 020317 (2021).
  • [39] Sirui Lu, Mari Carmen Bañuls, and J Ignacio Cirac. “Algorithms for quantum simulation at finite energies”. PRX Quantum 2, 020321 (2021).
  • [40] G.B. Arfken, H.J. Weber, and F.E. Harris. “Mathematical methods for physicists: A comprehensive guide”. Elsevier.  (2012).
  • [41] Seth Lloyd. “Universal quantum simulators”. Science 273, 1073–1078 (1996).
  • [42] Dominic W Berry, Graeme Ahokas, Richard Cleve, and Barry C Sanders. “Efficient quantum algorithms for simulating sparse hamiltonians”. Communications in Mathematical Physics 270, 359–371 (2007).
  • [43] Nathan Wiebe, Dominic Berry, Peter Høyer, and Barry C Sanders. “Higher order decompositions of ordered operator exponentials”. Journal of Physics A: Mathematical and Theoretical 43, 065203 (2010).
  • [44] Dominic W Berry, Andrew M Childs, Richard Cleve, Robin Kothari, and Rolando D Somma. “Simulating hamiltonian dynamics with a truncated taylor series”. Physical review letters 114, 090502 (2015).
  • [45] Richard Meister, Simon C. Benjamin, and Earl T. Campbell. “Tailoring Term Truncations for Electronic Structure Calculations Using a Linear Combination of Unitaries”. Quantum 6, 637 (2022).
  • [46] Earl Campbell. “Random compiler for fast hamiltonian simulation”. Physical review letters 123, 070503 (2019).
  • [47] Yongdan Yang, Bing-Nan Lu, and Ying Li. “Accelerated quantum monte carlo with mitigated error on noisy quantum computer”. PRX Quantum 2, 040361 (2021).
  • [48] Philip W Anderson. “The resonating valence bond state in la2cuo4 and superconductivity”. science 235, 1196–1198 (1987).
  • [49] Patrick A Lee, Naoto Nagaosa, and Xiao-Gang Wen. “Doping a mott insulator: Physics of high-temperature superconductivity”. Reviews of modern physics 78, 17 (2006).
  • [50] Kazuhiro Seki, Tomonori Shirakawa, and Seiji Yunoki. “Symmetry-adapted variational quantum eigensolver”. Phys. Rev. A 101, 052340 (2020).
  • [51] Roger A. Horn and Charles R. Johnson. “Matrix analysis”. Cambridge University Press.  (2012). 2 edition.
  • [52] Germund Dahlquist and Åke Björck. “Numerical methods in scientific computing, volume i”. Society for Industrial and Applied Mathematics.  (2008).
  • [53] William Kirby. “Analysis of quantum krylov algorithms with errors” (2024). arXiv:2401.01246.
  • [54] https://github.com/ZongkangZhang/QKSD.
  • [55] Wassily Hoeffding. “Probability inequalities for sums of bounded random variables”. Journal of the American Statistical Association 58, 13–30 (1963).
  • [56] Joel A. Tropp. “An introduction to matrix concentration inequalities”. Foundations and Trends® in Machine Learning 8, 1–230 (2015).
  • [57] Raymundo Albert and Cecilia G. Galarza. “Model order selection for sum of complex exponentials”. In 2021 IEEE URUCON. Pages 561–565.  (2021).
  • [58] D. Babusci, G. Dattoli, and M. Quattromini. “On integrals involving hermite polynomials”. Applied Mathematics Letters 25, 1157–1160 (2012).
  • [59] Alexei Yu Kitaev, Alexander Shen, and Mikhail N Vyalyi. “Classical and quantum computation”. Graduate studies in mathematics. American Mathematical Society.  (2002).
  • [60] Julia Kempe, Alexei Kitaev, and Oded Regev. “The complexity of the local hamiltonian problem”. Siam journal on computing 35, 1070–1097 (2006).
  • [61] Chris Cade, Marten Folkertsma, Sevag Gharibian, Ryu Hayakawa, François Le Gall, Tomoyuki Morimae, and Jordi Weggemans. “Improved Hardness Results for the Guided Local Hamiltonian Problem”. In Kousha Etessami, Uriel Feige, and Gabriele Puppis, editors, 50th International Colloquium on Automata, Languages, and Programming (ICALP 2023). Volume 261 of Leibniz International Proceedings in Informatics (LIPIcs), pages 32:1–32:19. Dagstuhl, Germany (2023). Schloss Dagstuhl – Leibniz-Zentrum für Informatik.
  • [62] Seunghoon Lee, Joonho Lee, Huanchen Zhai, Yu Tong, Alexander M Dalzell, Ashutosh Kumar, Phillip Helms, Johnnie Gray, Zhi-Hao Cui, Wenyuan Liu, et al. “Evaluating the evidence for exponential quantum advantage in ground-state quantum chemistry”. Nature communications 14, 1952 (2023).