\LetLtxMacro\ORIGselectlanguage\ORIGselectlanguage

english

Learning topological states from randomized measurements using
variational tensor network tomography

Yanting Teng Department of Physics, Harvard University, Cambridge, MA 02138, USA    Rhine Samajdar Department of Physics, Princeton University, Princeton, NJ 08544, USA Princeton Center for Theoretical Science, Princeton University, Princeton, NJ 08544, USA    Katherine Van Kirk Department of Physics, Harvard University, Cambridge, MA 02138, USA    Frederik Wilde Dahlem Center for Complex Quantum Systems, Freie Universität Berlin, 14195 Berlin, Germany    Subir Sachdev Department of Physics, Harvard University, Cambridge, MA 02138, USA    Jens Eisert Dahlem Center for Complex Quantum Systems, Freie Universität Berlin, 14195 Berlin, Germany    Ryan Sweke IBM Quantum, Almaden Research Center, San Jose, CA 95120, USA    Khadijeh Najafi IBM Quantum, IBM T.J. Watson Research Center, Yorktown Heights, NY 10598 USA MIT-IBM Watson AI Lab, Cambridge, MA 02142, USA
(June 28, 2024)
Abstract

Learning faithful representations of quantum states is crucial to fully characterizing the variety of many-body states created on quantum processors. While various tomographic methods such as classical shadow and MPS tomography have shown promise in characterizing a wide class of quantum states, they face unique limitations in detecting topologically ordered two-dimensional states. To address this problem, we implement and study a heuristic tomographic method that combines variational optimization on tensor networks with randomized measurement techniques. Using this approach, we demonstrate its ability to learn the ground state of the surface code Hamiltonian as well as an experimentally realizable quantum spin liquid state. In particular, we perform numerical experiments using MPS ansätze and systematically investigate the sample complexity required to achieve high fidelities for systems of sizes up to 48484848 qubits. In addition, we provide theoretical insights into the scaling of our learning algorithm by analyzing the statistical properties of maximum likelihood estimation. Notably, our method is sample-efficient and experimentally friendly, only requiring snapshots of the quantum state measured randomly in the X𝑋Xitalic_X or Z𝑍Zitalic_Z bases. Using this subset of measurements, our approach can effectively learn any real pure states represented by tensor networks, and we rigorously prove that random-XZ𝑋𝑍XZitalic_X italic_Z measurements are tomographically complete for such states.

I Introduction

In recent years, quantum simulators have proven instrumental to studying a diverse range of correlated quantum many-body systems, facilitating the investigation of phenomena that are often difficult to probe in conventional solid-state settings [1]. Today, there exist a host of candidate quantum simulation architectures, including ultracold atoms in optical lattices [2], neutral atom arrays [3], trapped ions [4], photonic devices [5], and superconducting quantum processors [6]; additionally, many of these platforms can be operated in either analog or digital modes. A common use case for such simulators is preparing the ground (or low-energy) states of some desired Hamiltonian. Importantly, going beyond simple symmetry-breaking ground states, quantum simulation has also enabled the experimental realization of highly entangled states with topological order [7, 8], characterized by fractionalized excitations and emergent gauge fields [9, 10].

Tomographic and other recovery methods for extracting information from these experiments are crucial, in particular, when they provide detailed diagnostic information about the prepared quantum state [11]. Indeed, recent approaches using randomized measurements have shown promise for measuring specific properties, such as the Renyi entropy [8, 12, 13, 14]. However, to extract Von Neumann entanglement entropy across arbitrary subsystem sizes using these methods might require multiple copies, substantially more experimental control, or more samples than what is currently available. Quantum state tomography (QST) is designed to reconstruct a quantum state from informationally complete measurements and is useful for systems where the desired observables are not known a priori [15] or cannot be measured due to the constraints of the experimental setup. Although QST is generally not scalable [16, 17], it becomes significantly more effective when applied with priors for particular classes of states. This makes QST a valuable alternative for extracting many properties from such systems, particularly when traditional methods are limited. Established state tomographic methods like matrix product state (MPS) tomography [18, 19, 20] provide a natural description with robust guarantees for one-dimensional systems but fall short in accurately capturing two-dimensional quantum states, especially when they exhibit non-injective properties [21] and complex entanglement patterns [22]. While neural network quantum state tomography (NNQST) [23, 24] offers yet another option, its efficacy for topological states also remains unclear due to their intrinsic long-range entanglement and intricate phase structures across different components of the wavefunction [25].

Refer to caption
Figure 1: Schematic illustration of variational tensor network tomography. Data collection: A target state vector |ϕketitalic-ϕ|\phi\rangle| italic_ϕ ⟩ is measured in multiple bases for a collection of N𝑁Nitalic_N samples defined in  Eq. (1). For instance, the first panel shows measurements in a random X𝑋Xitalic_X or Z𝑍Zitalic_Z basis for each qubit. Optimization: A tensor network model is variationally optimized from N𝑁Nitalic_N samples using gradient descent, which outputs tensors A^(N)superscript^𝐴𝑁\hat{A}^{(N)}over^ start_ARG italic_A end_ARG start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT representing the state vector |ψ(A^(N))ket𝜓superscript^𝐴𝑁|\psi(\hat{A}^{(N)})\rangle| italic_ψ ( over^ start_ARG italic_A end_ARG start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT ) ⟩. The middle panel illustrates an MPS model prior, as numerically investigated in this work. The loss function given by Eq. (5) includes the negative log-likelihood (NLL), typically used in maximum likelihood estimation (MLE). It is optionally regularized by physical observables estimated from the dataset (e.g., using classical shadow tomography represented by the pink dashed box). Evaluation: The algorithm outputs a tensor network or, in the current work, an MPS representation A^(N)superscript^𝐴𝑁\hat{A}^{(N)}over^ start_ARG italic_A end_ARG start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT as the learned state. We sometimes omit the number of samples and abbreviate the output tensor as A^^𝐴\hat{A}over^ start_ARG italic_A end_ARG. In the right panel, we evaluate the performance of our learning protocol by computing the fidelity F=|ϕ|ψ(A^)|𝐹inner-productitalic-ϕ𝜓^𝐴F=\lvert\mbox{$\langle\phi|\psi(\hat{A})\rangle$}\rvertitalic_F = | ⟨ italic_ϕ | italic_ψ ( over^ start_ARG italic_A end_ARG ) ⟩ |. From the reconstructed state, we can extract arbitrary physical observables, including the von Neumann entanglement entropy and highly nonlocal Pauli strings (the inset shows an example of the nonlocal Z𝑍Zitalic_Z string operators considered in Fig. 4).

To address the need for viable tomographic techniques for exotic two-dimensional systems [7], here, we demonstrate a heuristic state tomography method combining a tensor-network prior with maximum likelihood estimation (MLE) [26, 27, 28, 29] and randomized measurements [25, 30]. Previous works have considered learning one-dimensional quantum states from uniformly random single-qubit measurements [26] or two-dimensional quantum channels with random Paulli measurements [29]. In this work, we focus on two-dimensional real pure states and show that they are completely characterized by only randomized XZ𝑋𝑍XZitalic_X italic_Z measurements, which randomly measure each qubit in either the X𝑋Xitalic_X or the Z𝑍Zitalic_Z basis. The states considered here are not only potentially more complex than in previous studies but also easy to implement with high fidelity in experimental systems [31]. Moreover, we prove the tomographic completeness of randomized XZ𝑋𝑍XZitalic_X italic_Z measurements on the space of pure and real quantum states, allowing the method to be extendable to nonstoquastic real Hamiltonians’ eigenstates [32]. Thus, our method provides a pragmatic pathway for characterizing two-dimensional states on near-term quantum devices, making use of only easy-to-implement measurements. We provide a detailed discussion of related works in Appendix A.

In our numerical experiments, we use an MPS ansatz as our model prior. While MPS-based representations of two-dimensional states are inherently limited, our approach sets the stage for higher-dimensional tensor networks, such as projected entangled pair states (PEPS), for future investigations. In addition, our method provides a scalable approach that aligns with the existing capabilities of modern simulators, thereby providing a benchmark for more advanced tomographic techniques. To supplement our numerical results, we leverage the statistical properties of MLE to provide insights into the sample complexity of our approach. More specifically, we provide a pedagogical overview of standard techniques for analyzing MLE [33, 34, 35], and discuss where they fail in this context, due to gauge freedom in MPS representations of quantum states. We then sketch how these difficulties can be circumvented by considering MLE on a manifold [36], and apply these results to the variational manifold given by our MPS prior [37].

The rest of this paper is structured as follows. In Sec. II, we present our protocol in conjunction with an in-depth discussion of our state-tomographic measurement schemes. Then, in Sec. III, we demonstrate the effectiveness of our protocol when applied to two topological states: a ground state of the surface code Hamiltonian and a 2subscript2\mathbb{Z}_{2}blackboard_Z start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT quantum spin liquid state of Rydberg atoms arrayed on the ruby lattice. Our results are immediately relevant to current experiments since we only require feasible XZ𝑋𝑍XZitalic_X italic_Z measurements and consider system sizes achievable with state-of-the-art simulators. Finally, we conclude by discussing the prospects and potential directions for improvement of our methodology in Sec. IV.

II Variational tensor network tomography

In this section, we introduce our variational tensor network (TN) tomography protocol. To begin, we briefly review the framework of randomized measurements, wherein randomly chosen unitaries are applied to a target state before measuring it in the computational basis. Then, we utilize this data to variationally find the optimal tensor network approximation of the target state. We do so via maximum likelihood estimation, a standard technique for statistical inference; we also have the ability to regularize the MLE solution by estimations of physical quantities via classical shadows. This is done by adding a term to the MLE cost function. These steps are described below and represented by the schematic in Fig. 1.

Dataset and measurements.

We follow the notation commonly used in randomized measurements [38, 25]. Let |ϕketitalic-ϕ|\phi\rangle| italic_ϕ ⟩ be the unknown n𝑛nitalic_n-qubit pure state prepared by the quantum simulator, and let 𝒰𝒰\mathcal{U}caligraphic_U be the unitary ensemble with a uniform distribution over a set {U}𝑈\{U\}{ italic_U }. We apply a unitary U𝒰similar-to𝑈𝒰U\sim\mathcal{U}italic_U ∼ caligraphic_U before measuring in the computational basis. Performing this process repeatedly, we collect N𝑁Nitalic_N unitary-bit-string tuples as our dataset

{(Ui,bi)}i=1N,superscriptsubscriptsubscript𝑈𝑖subscript𝑏𝑖𝑖1𝑁\{(U_{i},\,b_{i})\}_{i=1}^{N},{ ( italic_U start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) } start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT , (1)

where each (U,b)𝑈𝑏(U,b)( italic_U , italic_b ) is effectively sampled with probability

1|{U}|×b|U|ϕϕ|U|b.1𝑈delimited-⟨⟩𝑏𝑈italic-ϕdelimited-⟨⟩italic-ϕsuperscript𝑈𝑏\frac{1}{\lvert\{U\}\rvert}\times\langle b\lvert U\rvert\phi\rangle\!\langle% \phi\lvert U^{\dagger}\rvert b\rangle.divide start_ARG 1 end_ARG start_ARG | { italic_U } | end_ARG × ⟨ italic_b | italic_U | italic_ϕ ⟩ ⟨ italic_ϕ | italic_U start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT | italic_b ⟩ . (2)

In this work, motivated by the available prior knowledge of the target state (which we discuss in the next subsection) as well as what measurements are experimentally feasible with high precision [31, 12], we consider two measurement schemes: (1) 𝒰Globalsubscript𝒰Global\mathcal{U}_{\textnormal{Global}}caligraphic_U start_POSTSUBSCRIPT Global end_POSTSUBSCRIPT: measuring every qubit in the X𝑋Xitalic_X basis or every qubit in the Z𝑍Zitalic_Z basis, and (2) 𝒰XZsubscript𝒰𝑋𝑍\mathcal{U}_{XZ}caligraphic_U start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT: randomly measuring each qubit in either the X𝑋Xitalic_X or the Z𝑍Zitalic_Z basis. For the ensemble 𝒰Globalsubscript𝒰Global\mathcal{U}_{\textnormal{Global}}caligraphic_U start_POSTSUBSCRIPT Global end_POSTSUBSCRIPT, the unitaries would be uniformly sampled from

{iUiX,iUiZ}.subscriptproduct𝑖superscriptsubscript𝑈𝑖𝑋subscriptproduct𝑖superscriptsubscript𝑈𝑖𝑍\displaystyle\left\{\prod_{i}U_{i}^{X},\,\prod_{i}U_{i}^{Z}\right\}.{ ∏ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_U start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_X end_POSTSUPERSCRIPT , ∏ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_U start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_Z end_POSTSUPERSCRIPT } . (3)

The unitary Uiαsuperscriptsubscript𝑈𝑖𝛼U_{i}^{\alpha}italic_U start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT rotates the measurement basis of the i𝑖iitalic_i-th qubit to the α=X,Z𝛼𝑋𝑍\alpha=X,Zitalic_α = italic_X , italic_Z basis. Note that the most general global controls have been considered in earlier work [15], in which the same SU(2)2(2)( 2 ) unitary is applied to each qubit. Here, we only focus on a discrete subset, of Pauli Xnsuperscript𝑋tensor-productabsent𝑛X^{\otimes n}italic_X start_POSTSUPERSCRIPT ⊗ italic_n end_POSTSUPERSCRIPT and Znsuperscript𝑍tensor-productabsent𝑛Z^{\otimes n}italic_Z start_POSTSUPERSCRIPT ⊗ italic_n end_POSTSUPERSCRIPT measurements, also considered in previous work [27]. For our second ensemble 𝒰XZsubscript𝒰𝑋𝑍\mathcal{U}_{XZ}caligraphic_U start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT, the unitaries are uniformly sampled from

{iUiα,α=X,Z}.formulae-sequencesubscriptproduct𝑖superscriptsubscript𝑈𝑖𝛼𝛼𝑋𝑍\displaystyle\left\{\prod_{i}U_{i}^{\alpha},\,\alpha=X,Z\right\}.{ ∏ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_U start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT , italic_α = italic_X , italic_Z } . (4)

In general, one can exploit priors on the target state to reduce the set of random unitaries required [39, 15]. For instance, consider learning the eigenstates of a real Hamiltonian. These states can be chosen to be real, and random-XZ𝑋𝑍XZitalic_X italic_Z measurements can learn any property of a state within such a space of real and pure states. This can be formally stated as follows.

Theorem 1 (Tomographic completeness).

Random-XZ𝑋𝑍XZitalic_X italic_Z measurements 𝒰XZsubscript𝒰𝑋𝑍\mathcal{U}_{XZ}caligraphic_U start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT are tomographically complete on the space of real and pure states.

The intuition here is that when |ϕketitalic-ϕ|\phi\rangle| italic_ϕ ⟩ is decomposed in terms of the computational basis states, all amplitudes can be learned by measuring every qubit in the computational (Z𝑍Zitalic_Z) basis. The (relative) signs can be learned by rotating some qubits to the X𝑋Xitalic_X basis. We refer interested readers to Appendices C and D for details of the proof. In Appendix C, we formalize our random-XZ𝑋𝑍XZitalic_X italic_Z measurement protocol and rigorously derive the associated measurement channel in the framework of classical shadow tomography. Furthermore, in Appendix D, we motivate our restricted random-XZ𝑋𝑍XZitalic_X italic_Z measurement scheme, proving that it is tomographically complete in the space of real and pure states considered in this work. We note that this scenario is more general than that of a fully positive eigenstate of a stoquastic Hamiltonian, for which measurements in only the computational basis suffice [32].

Loss function.

To estimate the target state from the dataset {(Ui,bi)}i=1Nsuperscriptsubscriptsubscript𝑈𝑖subscript𝑏𝑖𝑖1𝑁\{(U_{i},\,b_{i})\}_{i=1}^{N}{ ( italic_U start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) } start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT, we use a tensor network ansatz |ψ(A)ket𝜓𝐴|\psi({A})\rangle| italic_ψ ( italic_A ) ⟩, where a sequence of local tensors is denoted by A=(A[1],,A[n])𝐴superscript𝐴delimited-[]1superscript𝐴delimited-[]𝑛{A}=\left({A}^{[1]},\cdots,{A}^{[n]}\right)italic_A = ( italic_A start_POSTSUPERSCRIPT [ 1 ] end_POSTSUPERSCRIPT , ⋯ , italic_A start_POSTSUPERSCRIPT [ italic_n ] end_POSTSUPERSCRIPT ). The advantage of this formulation is that one can efficiently compute the probability of a given measurement outcome in any basis by applying the corresponding unitary to the tensor network model as shown in the middle panel of Fig. 1. This allows us to evaluate the loss function, whose minimum leads to our estimate of the target state. The loss function is

L(A)𝐿𝐴\displaystyle L({A})italic_L ( italic_A ) =1Ni=1Nlog|bi|Ui|ψ(A)|2+βR(A),\displaystyle=-\frac{1}{N}\sum_{i=1}^{N}\log\lvert\langle b_{i}\lvert U_{i}% \rvert\psi({A})\rangle\rvert^{2}+\beta R({A}),= - divide start_ARG 1 end_ARG start_ARG italic_N end_ARG ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT roman_log | ⟨ italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | italic_U start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | italic_ψ ( italic_A ) ⟩ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_β italic_R ( italic_A ) , (5)

where R(A)𝑅𝐴R(A)italic_R ( italic_A ) denotes the regularizer and its coefficient β𝛽\betaitalic_β describes how much we prioritize the regularization over MLE. The first term in Eq. (5) above, the negative log-likelihood (NLL), arises from the Kullback–Leibler divergence, which measures the difference between two probability distributions. In our case, these two probability distributions are over measurement outcomes in randomly selected bases, defined by the target and the model state, respectively. Minimizing the N𝑁Nitalic_N-sample NLL leads to an MLE solution A^(N)superscript^𝐴𝑁\hat{A}^{(N)}over^ start_ARG italic_A end_ARG start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT. To improve our training, we add the regularization term R(A)𝑅𝐴R(A)italic_R ( italic_A ): it penalizes the differences between our model’s prediction of a physical observable ψ(A)|O|ψ(A)delimited-⟨⟩𝜓𝐴𝑂𝜓𝐴\langle\psi({A})\lvert O\rvert\psi({A})\rangle⟨ italic_ψ ( italic_A ) | italic_O | italic_ψ ( italic_A ) ⟩ and the estimation ONsubscriptdelimited-⟨⟩𝑂𝑁\langle O\rangle_{N}⟨ italic_O ⟩ start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT inferred directly from the dataset. For instance, we can choose R(A)=ψ(A)|O|ψ(A)ON2𝑅𝐴subscriptnormdelimited-⟨⟩𝜓𝐴𝑂𝜓𝐴subscriptdelimited-⟨⟩𝑂𝑁2R(A)=\|\langle\psi({A})\lvert O\rvert\psi({A})\rangle-\langle O\rangle_{N}\|_{2}italic_R ( italic_A ) = ∥ ⟨ italic_ψ ( italic_A ) | italic_O | italic_ψ ( italic_A ) ⟩ - ⟨ italic_O ⟩ start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT ∥ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT with a k𝑘kitalic_k-body Pauli string O=Xk𝑂superscript𝑋tensor-productabsent𝑘O=X^{\otimes k}italic_O = italic_X start_POSTSUPERSCRIPT ⊗ italic_k end_POSTSUPERSCRIPT, where the estimation XkNsubscriptdelimited-⟨⟩superscript𝑋tensor-productabsent𝑘𝑁\langle X^{\otimes k}\rangle_{N}⟨ italic_X start_POSTSUPERSCRIPT ⊗ italic_k end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT can be accurately determined from our global-XZ𝑋𝑍XZitalic_X italic_Z measurements. Broadly speaking, if we are confident about certain properties of the target state, our model should reflect those properties. The choices of regularization observables, therefore, depend on the systems of interest, which we discuss in Sec. III and Appendix E, where we describe the loss function and physical observables used for our numerical results.

Refer to caption
Figure 2: Numerical results for the surface code. (a) Schematic representation of the surface code of dimension (Lx,Ly)subscript𝐿𝑥subscript𝐿𝑦(L_{x},\,L_{y})( italic_L start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT , italic_L start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT ) with Z𝑍Zitalic_Z stabilizers (red patches) and X𝑋Xitalic_X stabilizers (blue patches). The MPS “snakes” through the system along the path shown (black arrow). (b) Scaling of the infidelity with the number of samples N𝑁Nitalic_N for a 3×3333\times 33 × 3 surface code with global-XZ𝑋𝑍XZitalic_X italic_Z (cross) or random-XZ𝑋𝑍XZitalic_X italic_Z (circle) measurements. Shades of blue indicates the regularization strength β𝛽\betaitalic_β. (c) To investigate larger system sizes, we focus on the surface code in a strip geometry (Ly=3subscript𝐿𝑦3L_{y}=3italic_L start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT = 3) with a perturbation (hz=0.1subscript𝑧0.1h_{z}=0.1italic_h start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = 0.1). The three lines indicate the infidelity for different numbers of qubits, n𝑛nitalic_n, with random-XZ𝑋𝑍XZitalic_X italic_Z measurements. Dark blue indicates a regularization β=5𝛽5\beta=5italic_β = 5. The dashed line marks the fidelity threshold for n=9𝑛9n=9italic_n = 9. (d) Sample complexity to reach a fixed local fidelity threshold Flocal=(0.99)nsubscript𝐹localsuperscript0.99𝑛F_{\text{local}}=(0.99)^{n}italic_F start_POSTSUBSCRIPT local end_POSTSUBSCRIPT = ( 0.99 ) start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT extracted from (c). Colors and markers are the same as in (b).

MPS prior.

Matrix product states, which are one of the most extensively studied families of tensor networks, have enjoyed remarkable success in accurately representing many-body quantum states. Thanks to an effective variational algorithm known as the density-matrix renormalization group (DMRG) [40], MPS ansätze can be efficiently employed to describe the ground states of various many-body Hamiltonians. Strictly speaking, two-dimensional systems require the generalization of MPS, as the quasi-1D MPS description is limited to capturing short-range correlations and thus necessarily fails for sufficiently large and entangled systems. However, in practice, MPS ansätze are often used to represent ground states even in two dimensions, due to their relative advantage of efficient contractions. For the remainder of this paper, we utilize an MPS as our model prior, and we variationally find a state that is most consistent with the observed measurements, as shown in Fig. 1. The optimization of the MPS is carried out by taking gradients simultaneously with respect to all the tensor components; we provide the details in Appendix G.

III Results

In this section, we demonstrate the effectiveness of our variational tomography protocol for two topologically ordered states. While our eventual goal is to perform tomography on a state in the laboratory, for our results here, we perform numerical “experiments” using simulated target states obtained via DMRG, as detailed in Appendix F. Our protocol assumes that the laboratory state is close to a pure state that could be approximated as an MPS (albeit with a possibly large bond dimension). After obtaining our simulated measurements by sampling the target state, we initiate our protocol with random initial parameters. These parameters are iteratively adjusted until they converge to the trained state. Then, we evaluate the efficacy of this training process by computing the fidelity between the trained and target states. Here, we analyze 10101010 such random initializations, which are shown in our numerical results below. Of course, when using realistic experimental data, one would lack access to the “ground truth”, i.e., a classical description of the target state, which is needed to compute the fidelity exactly. In this case, we perform cross validation by comparing converged models on a test dataset [Appendix G.2], or alternatively, one could consider efficient certifications of high fidelity [41, 42].

III.1 Learning the surface code

In our first demonstration, we consider learning the ground state of the perturbed surface code Hamiltonian [43], Hsc=lSZ,llSX,lhziZisubscript𝐻scsubscript𝑙subscript𝑆𝑍𝑙subscript𝑙subscript𝑆𝑋𝑙subscript𝑧subscript𝑖subscript𝑍𝑖H_{\text{sc}}=-\sum_{l}S_{Z,\,l}-\sum_{l}S_{X,\,l}-h_{z}\sum_{i}Z_{i}italic_H start_POSTSUBSCRIPT sc end_POSTSUBSCRIPT = - ∑ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT italic_S start_POSTSUBSCRIPT italic_Z , italic_l end_POSTSUBSCRIPT - ∑ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT italic_S start_POSTSUBSCRIPT italic_X , italic_l end_POSTSUBSCRIPT - italic_h start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_Z start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, defined on the square lattice. Here, SP,l=ilPisubscript𝑆𝑃𝑙subscriptproduct𝑖subscript𝑙subscript𝑃𝑖S_{P,\,l}=\prod_{i\in\square_{l}}P_{i}italic_S start_POSTSUBSCRIPT italic_P , italic_l end_POSTSUBSCRIPT = ∏ start_POSTSUBSCRIPT italic_i ∈ □ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT denotes the l𝑙litalic_l-th stabilizer, which is a product of Pauli P𝑃Pitalic_P operators (P=X𝑃𝑋P=Xitalic_P = italic_X or P=Z𝑃𝑍P=Zitalic_P = italic_Z) on a plaquette or boundary [Fig. 2(a)], and hzsubscript𝑧h_{z}italic_h start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT is a perturbative field. Such a perturbation drives the ground state away from the exactly solvable limit, but it can still be well approximated by a tensor network. Let |ϕketitalic-ϕ|\phi\rangle| italic_ϕ ⟩ be the target MPS representation of the ground state along a snake path. Although our target state is possibly perturbed away from the exact stabilizer-state limit, the expectation values of the stabilizer operators should be smoothly connected to their fixed-point values (1absent1\approx 1≈ 1) as a function of the perturbation strength. This knowledge informs our choice of the set of stabilizers as observables for regularization. We estimate these stabilizers from the finite dataset either via direct measurement in the global-XZ𝑋𝑍XZitalic_X italic_Z scheme or from classical shadows for random-XZ𝑋𝑍XZitalic_X italic_Z measurements.

Our results for the surface code are presented in Fig. 2. To begin, we consider the surface code without any external perturbation (hz=0subscript𝑧0h_{z}=0italic_h start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = 0): for n=9𝑛9n=9italic_n = 9 qubits arrayed on a square geometry, our protocol outputs the state vector |ψ(A^)ket𝜓^𝐴|\psi(\hat{A})\rangle| italic_ψ ( over^ start_ARG italic_A end_ARG ) ⟩, and we evaluate the error of our learning protocol using the infidelity, 1F=1|ϕ|ψ(A^)|1𝐹1inner-productitalic-ϕ𝜓^𝐴1-F=1-\lvert\mbox{$\langle\phi|\psi(\hat{A})\rangle$}\rvert1 - italic_F = 1 - | ⟨ italic_ϕ | italic_ψ ( over^ start_ARG italic_A end_ARG ) ⟩ |. Figure 2(b) shows that the infidelity improves with the number of samples N𝑁Nitalic_N, for both the global-XZ𝑋𝑍XZitalic_X italic_Z and random-XZ𝑋𝑍XZitalic_X italic_Z measurements discussed in Sec. II. For the latter, we find that regularization over the stabilizers, by increasing β𝛽\betaitalic_β, further reduces the infidelity in the sample-limited regime (N<1000𝑁1000N<1000italic_N < 1000). We observe that both the global and random measurement schemes succeed in learning these states, and furthermore, the former outperforms the latter in terms of requiring fewer samples to achieve the same infidelity. However, while global-XZ measurements 𝒰Globalsubscript𝒰Global\mathcal{U}_{\textnormal{Global}}caligraphic_U start_POSTSUBSCRIPT Global end_POSTSUBSCRIPT are sufficient for certain highly structured states such as a surface code (or GHZ) state, they may not necessarily be suited for learning generic states, wherefore random-XZ𝑋𝑍XZitalic_X italic_Z measurements 𝒰XZsubscript𝒰𝑋𝑍\mathcal{U}_{XZ}caligraphic_U start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT could be beneficial. Finally, we note that the infidelity scales polynomially with the number of samples N𝑁Nitalic_N, similarly to observations in earlier work on one-dimensional states [26]. In particular, focusing on the exact surface code ground state (for hz=0subscript𝑧0h_{z}=0italic_h start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = 0) trained without regularization (β=0𝛽0\beta=0italic_β = 0), we find an excellent fit of 1F1/Nαproportional-to1𝐹1superscript𝑁𝛼1-F\propto 1/N^{\alpha}1 - italic_F ∝ 1 / italic_N start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT, as shown in Fig. 2(b), with α=1.16𝛼1.16\alpha=1.16italic_α = 1.16 (α=1.05𝛼1.05\alpha=1.05italic_α = 1.05) for random-XZ𝑋𝑍XZitalic_X italic_Z (global-XZ𝑋𝑍XZitalic_X italic_Z) measurements. To substantiate these numerical observations regarding the sample complexity needed to achieve a certain fidelity, we rigorously establish in Theorem 2 that the infidelity can be probabilistically upper bounded such that

[1|ϕ|ψ(A^(N))|ϵ(N)]1δ,delimited-[]1inner-productitalic-ϕ𝜓superscript^𝐴𝑁italic-ϵ𝑁1𝛿\mathbb{P}\left[1-\lvert\mbox{$\langle\phi|\psi(\hat{A}^{(N)})\rangle$}\rvert% \leq\epsilon(N)\right]\geq 1-\delta,blackboard_P [ 1 - | ⟨ italic_ϕ | italic_ψ ( over^ start_ARG italic_A end_ARG start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT ) ⟩ | ≤ italic_ϵ ( italic_N ) ] ≥ 1 - italic_δ , (6)

where ϵ(N)=O([nχmax2/(Nδ)]1/2)italic-ϵ𝑁𝑂superscriptdelimited-[]𝑛superscriptsubscript𝜒max2𝑁𝛿12\epsilon(N)=O\big{(}\big{[}\displaystyle n\chi_{\mathrm{max}}^{2}/(N\delta)% \big{]}^{1/2}\big{)}italic_ϵ ( italic_N ) = italic_O ( [ italic_n italic_χ start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / ( italic_N italic_δ ) ] start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT ) and χmaxsubscript𝜒\chi_{\max}italic_χ start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT represents the maximum bond dimension. Interestingly, our numerical results, which exhibit α>0.5𝛼0.5\alpha>0.5italic_α > 0.5, suggest better performance than our theoretical bound for arbitrary target states. Further details and substantial analytical and rigorous insights into the sample complexity, drawing from the statistical properties of MLE, are discussed in Appendix B.

Next, we add a perturbation to the Hamiltonian, hz=0.1subscript𝑧0.1h_{z}=0.1italic_h start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = 0.1, and analyze the performance of our protocol for various system sizes with up to n=45𝑛45n=45italic_n = 45 qubits. Given the limitations of an MPS ansatz, which is inherently quasi-one-dimensional, here, we only increase the length of the system Lxsubscript𝐿𝑥L_{x}italic_L start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT along the x𝑥xitalic_x-axis. In Fig. 2(c), we illustrate how the infidelity scales for different system sizes (indicated by different lines) using random-XZ𝑋𝑍XZitalic_X italic_Z measurements and find that the fidelity improves with N𝑁Nitalic_N. Finally, we analyze how the sample complexity scales with the system size n𝑛nitalic_n. The requirement of achieving a constant fidelity is stringent as the overlap between two many-body states can vanish exponentially with n𝑛nitalic_n, a phenomenon known as the orthogonality catastrophe [44]. For a finite system size, however, the overlap still serves as a useful diagnostic indicator [45]. Therefore, we relax the requirement of a fixed global fidelity by considering instead the per-site local fidelity FlocalF1/nsubscript𝐹localsuperscript𝐹1𝑛F_{\text{local}}\equiv F^{1/n}italic_F start_POSTSUBSCRIPT local end_POSTSUBSCRIPT ≡ italic_F start_POSTSUPERSCRIPT 1 / italic_n end_POSTSUPERSCRIPT. Extrapolating from the data in (c), Fig. 2(d) plots the samples required to achieve a local fidelity threshold, e.g., as indicated by the dashed line in (c) for n=9𝑛9n=9italic_n = 9 qubits. For the largest system size, n=45𝑛45n=45italic_n = 45, we find that only a few thousand measurements are necessary to attain a fidelity of Flocal=0.99subscript𝐹local0.99F_{\text{local}}=0.99italic_F start_POSTSUBSCRIPT local end_POSTSUBSCRIPT = 0.99.

Refer to caption
Figure 3: Numerical results for learning quantum spin liquids. (a) The tuning parameters of the Rydberg Hamiltonian are the Rabi frequency ΩΩ\Omegaroman_Ω, the detuning ΔΔ\Deltaroman_Δ, and the blockade radius Rbsubscript𝑅𝑏R_{b}italic_R start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT. The gray arrow shows the map** between a particular measurement bit-string on the ruby lattice and a dimer configuration on the kagome lattice: Light (dark) blue circles correspond to an atom being in the excited |rket𝑟|r\rangle| italic_r ⟩ (ground |gket𝑔|g\rangle| italic_g ⟩) state and maps to the presence (absence) of a dimer. A careful choice of the Rydberg blockade radius Rb=2asubscript𝑅𝑏2𝑎R_{b}=2aitalic_R start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT = 2 italic_a, where a𝑎aitalic_a is the lattice spacing, ensures that the dimers do not overlap. Bottom panel: The two phases as a function of the detuning ratio Δ/ΩΔΩ\Delta/\Omegaroman_Δ / roman_Ω. (b) The ground state belongs to the trivial disordered phase at Δ/Ω=0.5ΔΩ0.5\Delta/\Omega=0.5roman_Δ / roman_Ω = 0.5. The infidelity decreases with an increasing number of samples N𝑁Nitalic_N. For random-XZ𝑋𝑍XZitalic_X italic_Z measurements with N40000𝑁40000N\leq 40000italic_N ≤ 40000, a regularization of β=1𝛽1\beta=1italic_β = 1 (red circles) further improves the fidelity compared to the case without regularization β=0𝛽0\beta=0italic_β = 0 (red squares). Global-XZ𝑋𝑍XZitalic_X italic_Z measurements (gray crosses) perform worse than their random-XZ𝑋𝑍XZitalic_X italic_Z counterparts (red). (c) Same as in (b) but for the spin-liquid state at Δ/Ω=1.7ΔΩ1.7\Delta/\Omega=1.7roman_Δ / roman_Ω = 1.7. Relative to the case with no regularization β=0𝛽0\beta=0italic_β = 0 (light blue), regularization β=1𝛽1\beta=1italic_β = 1 (dark blue) does not improve the fidelity but reduces the number of nonconvergent outliers (above the black dashed line). Global-XZ𝑋𝑍XZitalic_X italic_Z (gray cross) measurements perform similarly to random-XZ𝑋𝑍XZitalic_X italic_Z ones (blue).

III.2 Learning a Rydberg quantum spin liquid

For a second application of our tomographic framework, we turn to learning ground states of strongly interacting arrays of neutral atoms. Over the last decade, these Rydberg atom arrays have evolved into mature platforms for quantum simulation and have shown great potential for probing a variety of correlated quantum phases of matter [46, 47, 3, 48, 7]. The effective Hamiltonian of this system can be written as

HRydsubscript𝐻Ryd\displaystyle H_{\rm Ryd}italic_H start_POSTSUBSCRIPT roman_Ryd end_POSTSUBSCRIPT =Ω2XΔ12(1+Z)absentΩ2subscriptsubscript𝑋Δsubscript121subscript𝑍\displaystyle=\frac{\Omega}{2}\sum_{\ell}X_{\ell}-\Delta\sum_{\ell}\frac{1}{2}% (1+Z_{\ell})= divide start_ARG roman_Ω end_ARG start_ARG 2 end_ARG ∑ start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT italic_X start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT - roman_Δ ∑ start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( 1 + italic_Z start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT )
+12,V,4(1+Z)(1+Z),12subscriptsuperscriptsubscript𝑉superscript41subscript𝑍1subscript𝑍superscript\displaystyle+\frac{1}{2}\sum_{\ell,\,\ell^{\prime}}\frac{V_{\ell,\,\ell^{% \prime}}}{4}(1+Z_{\ell})(1+Z_{\ell^{\prime}}),+ divide start_ARG 1 end_ARG start_ARG 2 end_ARG ∑ start_POSTSUBSCRIPT roman_ℓ , roman_ℓ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT divide start_ARG italic_V start_POSTSUBSCRIPT roman_ℓ , roman_ℓ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT end_ARG start_ARG 4 end_ARG ( 1 + italic_Z start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT ) ( 1 + italic_Z start_POSTSUBSCRIPT roman_ℓ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ) , (7)

where ΩΩ\Omegaroman_Ω represents the Rabi frequency, ΔΔ\Deltaroman_Δ is the detuning of the laser drive, and V,subscript𝑉superscriptV_{\ell,\,\ell^{\prime}}italic_V start_POSTSUBSCRIPT roman_ℓ , roman_ℓ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT is the van der Waals interaction potential. The strong Rydberg-Rydberg interactions prevent neighboring atoms lying within a “blockade radius”, Rbsubscript𝑅𝑏R_{b}italic_R start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT, from being simultaneously excited to the Rydberg state, thereby engendering strong quantum correlations. For an appropriate choice of the blockade radius, configurations of Rydberg atoms on the ruby lattice can be mapped to a set of “dimers” on the kagome lattice [Fig. 3(a)]; such quantum dimer models have long been known to host topological quantum spin liquids [49, 50]. In the Rydberg system, depending on the detuning parameter Δ/ΩΔΩ\Delta/\Omegaroman_Δ / roman_Ω, the ground state (for this specifically chosen Rbsubscript𝑅𝑏R_{b}italic_R start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT) can be either a trivial disordered phase or a topologically ordered 2subscript2\mathbb{Z}_{2}blackboard_Z start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT quantum spin liquid phase [51, 52, 7, 53]; see Fig. 3(a). In the computational basis, these two phases are difficult to distinguish as they both lack any symmetry-breaking order. Here, we select two parameters in the phase diagram belonging to the spin-liquid (Δ/Ω=1.7ΔΩ1.7\Delta/\Omega=1.7roman_Δ / roman_Ω = 1.7) or the disordered (Δ/Ω=0.5ΔΩ0.5\Delta/\Omega=0.5roman_Δ / roman_Ω = 0.5) phase and perform DMRG simulations to find these ground states (see Appendix F for more details).

Refer to caption
Figure 4: Prediction of spin-liquid properties. (a) A quantum spin liquid (QSL) state can be viewed as a superpostions of dimers in the computational basis. (b) Definition of measurable diagonostic quantities such as the Z𝑍Zitalic_Z string operators. Top panel: A dashed line across a bond is equivalent to applying the Pauli Z𝑍Zitalic_Z operator to the qubit on the same bond. The Z𝑍Zitalic_Z logical string PLogicalsubscript𝑃LogicalP_{\mathrm{Logical}}italic_P start_POSTSUBSCRIPT roman_Logical end_POSTSUBSCRIPT applies the Z𝑍Zitalic_Z string operator across the system. Bottom panel: An open string diagonal FM operator PFMsubscript𝑃FMP_{\mathrm{FM}}italic_P start_POSTSUBSCRIPT roman_FM end_POSTSUBSCRIPT, renormalized by the expectation value of a closed-loop operator; e.g., the semicircle corresponds to the product of 6666 Pauli Z𝑍Zitalic_Z operators. (c) The expectation values of the PFMsubscript𝑃FMP_{\mathrm{FM}}italic_P start_POSTSUBSCRIPT roman_FM end_POSTSUBSCRIPT (red) and PLogicalsubscript𝑃LogicalP_{\mathrm{Logical}}italic_P start_POSTSUBSCRIPT roman_Logical end_POSTSUBSCRIPT (blue) operators for different numbers of samples, N𝑁Nitalic_N, converge to their exact values (lines). (d) Predictions (crosses) of the entanglement entropy for different sizes of the bipartition and (e) the Schmidt values (cut at site i=21𝑖21i=21italic_i = 21) from MPS ansätze trained with N𝑁Nitalic_N samples converge to the exact values (open circles).

Importantly, for this system, we do not have prior knowledge of which observables are important to regularize, unlike for the surface code studied above. Hence, analogously to MPS tomography, we consider estimations of all estimable Pauli strings that are supported on a local subsystem (taken to be 6 qubits here) and use them to regularize our MLE loss function. This subsystem should be chosen such that it captures the important correlations of the state and that the Paulis supported on this subsystem can be estimated accurately using randomized measurements 111Note that we could also consider regularization from global measurements, but then we would only be able to estimate Xksuperscript𝑋tensor-productabsent𝑘X^{\otimes k}italic_X start_POSTSUPERSCRIPT ⊗ italic_k end_POSTSUPERSCRIPT or Zksuperscript𝑍tensor-productabsent𝑘Z^{\otimes k}italic_Z start_POSTSUPERSCRIPT ⊗ italic_k end_POSTSUPERSCRIPT.. With regard to the task of learning the states of n=48𝑛48n=48italic_n = 48 qubits, we now show that our protocol (see Appendix E) is capable of learning both regimes in the phase diagram—i.e., the spin liquid as well as the disordered state—which cannot be distinguished by a local order parameter due to the topological nature of the former.

Focusing first on the disordered state shown in Fig. 3(b), we see that the fidelity improves with the number of measurements N𝑁Nitalic_N for all measurement schemes. Moreover, we observe a slight improvement with a regularization strength of β=1𝛽1\beta=1italic_β = 1 in the sample-limited regime (N40000𝑁40000N\leq 40000italic_N ≤ 40000). Next, we turn to the spin liquid illustrated in 4(a). For this state as well, the infidelity generally reduces with N𝑁Nitalic_N, as shown in Fig. 3(c), albeit with outliers that did not converge (indicated by squares above the dashed line). However, in contrast to the disordered phase, we observe that regularization, with strength β=1𝛽1\beta=1italic_β = 1 (indicated by the dark blue circles), does not improve the fidelity beyond that obtained with no regularization β=0𝛽0\beta=0italic_β = 0 (light blue squares), but the number of outliers (above the dashed line) is reduced. This lack of substantial improvement upon regularization suggests that the important operators capturing the correlations of the spin-liquid state are supported on a larger subsystem than used in our current numerical experiments.

Finally, we show that, using a successfully trained MPS, we can evaluate physical properties that may be difficult to measure directly given limited experimental controls. Certain diagnostic physical observables, such as the diagonal Fredenhagen-Marcu (FM) order parameter PFMsubscript𝑃FMP_{\mathrm{FM}}italic_P start_POSTSUBSCRIPT roman_FM end_POSTSUBSCRIPT as well as the Z𝑍Zitalic_Z logical loop operator Plogicalsubscript𝑃logicalP_{\mathrm{logical}}italic_P start_POSTSUBSCRIPT roman_logical end_POSTSUBSCRIPT and its X𝑋Xitalic_X counterpart, have been previously suggested for the spin-liquid state [7, 52, 54], as illustrated in Figs. 4(a,b). These measurable observables can be used to verify the consistency of our learned states, and we see that the predictions from our trained MPS converge well to the target values [Fig. 4(c)]. Moreover, our protocol allows for the extraction of additional quantities such as the von Neumann entanglement entropy, which is essential to determining the topological entanglement entropy—a key quantity that characterizes a gapped spin liquid [55, 56]. Figure 4(d) demonstrates that the bipartite entanglement entropy converges to the target’s as the bipartition varies in subsystem size for N90000𝑁90000N\approx 90000italic_N ≈ 90000. For a cut across roughly half of the system, the Schmidt values in Fig. 4(e) also closely match the exact values of the target state.

IV Discussion and outlook

In this work, we combine randomized measurements with tensor networks to perform quantum state tomography via MLE, with optional regularization using classical shadows. While the randomized measurements toolbox is useful for estimating many physical properties, tensor networks are known to efficiently represent physically relevant quantum states. Here, we have considered a more restricted measurement setting, random-XZ𝑋𝑍XZitalic_X italic_Z shadows, that is well motivated by both its tomographic completeness on real and pure states and its feasibility in near-term experiments. Using these XZ𝑋𝑍XZitalic_X italic_Z measurements, we can regularize our MLE loss function with observables estimated via the classical shadow framework. We showcase the performance of our protocol by learning the topological ground states of the surface code Hamiltonian as well as the Rydberg Hamiltonian on the ruby lattice. For both of these states, our protocol with random-XZ𝑋𝑍XZitalic_X italic_Z measurements can accurately reconstruct the target state up to a fidelity of over 0.950.950.950.95 with under 105superscript10510^{5}10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT samples. In addition, we test our shadow regularization technique on these two applications: we find that requiring the stabilizer expectations values to agree with estimations from classical shadows further improves the fidelity for the surface code, whereas for the ruby-lattice Rydberg spin liquid, this only reduces the number of outliers observed during the optimization process.

Using higher-dimensional TNs combined with randomized measurements would, in principle, allow us to learn any states of interest. In the present work, we have numerically benchmarked our protocol using MPS as our model prior, given its advantages of exact contractions and sampling. An immediate extension of our work would be to generalize this model’s architecture to represent 2222D area-law entangled states, such as string bond states [57], projected entangled pair states (PEPS) [58]. We leave the numerical demonstrations of our tomographic method for such tensor-network states to future work.

It constitutes an interesting perspective to extend these methods to the study of fermionic quantum states, which are the focus of attention in the study of correlated electron systems. Our analysis could be extended to quantum gas microscope measurements of optical lattice Fermi-Hubbard systems [59] which yield snapshots with the charge and spin on each site. While in this setting, quench dynamics generated by noninteracting Hamiltonians are presumably most conceivable [60], it is possible that the spin degrees of freedom can be measured using a similar unitary ensemble such as our random-XZ𝑋𝑍XZitalic_X italic_Z measurements. The variational optimization can be carried out using fermionic neural networks [61] or auxiliary wavefunctions [62, 63], replacing the MPS we have employed in the present work. To enhance the optimization process using neural-network ansätze, it would be especially interesting to extend recent results [42] on locally-randomized single-qubit measurements to such fermionic systems.

While our method can be generalized to quantum states realized in a wide class of quantum processors, more specifically, for Rydberg quantum simulators, it can be easily extended to learning logically encoded quantum states. In particular, our random-XZ𝑋𝑍XZitalic_X italic_Z measurements are well suited for near-term encodings, such as the color codes recently implemented in neutral atom arrays [31]. In such a Rydberg array, every logical qubit, encoded, for example, in a Steane code, can be measured in either the X𝑋Xitalic_X or the Z𝑍Zitalic_Z basis, allowing for both error detection via post-selection and the ability to perform high-fidelity random-XZ𝑋𝑍XZitalic_X italic_Z measurements. Even if current experiments do not aim to perform full logical quantum state tomography, our characterization and derivation of the tools required for random-XZ𝑋𝑍XZitalic_X italic_Z classical shadows can be immediately applied to these logical encodings.

In settings where qubits are not logically encoded, our techniques are also experimentally amenable for physical qubits. Even as the degree of control over large quantum systems improves, single-qubit rotations will likely always remain preferable compared to many-qubit gates due to their orders-of-magnitude higher fidelity and short implementation times [64]. Therefore, the ability to accurately estimate a laboratory state from random measurements will remain attractive beyond the NISQ era. Our protocol can also be straightforwardly extended to random-Pauli measurements [25]. Moreover, this technique could be made even more attractive by incorporating adaptivity into our measurement protocol. Judiciously altering the measurement ensemble as more data is obtained allows for more efficient learning and thus reduces the sample complexity. To implement an adaptive learning scheme, one may incorporate ideas from Ref. [65] into our learning technique in order to use fewer measurements to achieve the same infidelity.

Finally, our work also lends itself to learning about experimental imperfections in state preparation. In this sense, our scheme goes beyond state verification: it learns not only the state prepared but also the unwanted deviations from the desired state, which constitute feedback to the experiment on how to better achieve a higher-fidelity preparation over many iterations. This variational state preparation is practical from a sample-complexity point of view when the underlying state has some efficient representation; in other words, an exponential number of degrees of freedom need not be learned. This is precisely the case in many condensed-matter contexts, where typical states exhibit entanglement structures that are well captured by tensor networks [22].

V Code and data availability

The numerical code used for producing results and generating the figures in this paper is available at: https://github.com/teng10/tn-shadow-qst.

Acknowledgements.
We thank Jacob Barandes, Christian Bertoni, Jonathan Conrad, Zoe Holmes, Robert Huang, Dmitrii Kochkov, Nathan Leitao, Alexander Nietner, Jakob Unfried, Guifre Vidal, and Manuel Rudolph for useful discussions. We thank Hong-Ye Hu and Leandro Aolita for feedback on the manuscript. Y.T. is grateful to Dmitrii Kochkov for his inputs on the initial prototype of the numerical code base and systematic code reviews. Y.T. also thanks the groups of Zoe Holmes and Jens Eisert for their hospitality during her visits, when part of this work was completed. R.S. is supported by the Princeton Quantum Initiative Fellowship. K.V.K. acknowledges support from the Fannie and John Hertz Foundation and the National Defense Science and Engineering Graduate (NDSEG) Fellowship. The Berlin team acknowledges support from the BMBF (MuniQCAtoms), the DFG (CRC 183, FOR 2724), the Munich Quantum Valley, and the ERC (DebuQC). Y.T. and S.S. have been supported by the U.S. Department of Energy under Grant DE-SC0019030.

References

  • Cirac and Zoller [2012] J. I. Cirac and P. Zoller, Goals and opportunities in quantum simulation, Nat. Phys. 8, 264 (2012).
  • Bloch et al. [2012] I. Bloch, J. Dalibard, and S. Nascimbene, Quantum simulations with ultracold quantum gases, Nat. Phys. 8, 267 (2012).
  • Ebadi et al. [2021] S. Ebadi, T. T. Wang, H. Levine, A. Keesling, G. Semeghini, A. Omran, D. Bluvstein, R. Samajdar, H. Pichler, W. W. Ho, S. Choi, S. Sachdev, M. Greiner, V. Vuletić, and M. D. Lukin, Quantum phases of matter on a 256-atom programmable quantum simulator, Nature 595, 227 (2021).
  • Blatt and Roos [2012] R. Blatt and C. F. Roos, Quantum simulations with trapped ions, Nat. Phys. 8, 277 (2012).
  • Madsen et al. [2022] L. S. Madsen, F. Laudenbach, M. F. Askarani, F. Rortais, T. Vincent, J. F. Bulmer, F. M. Miatto, L. Neuhaus, L. G. Helt, M. J. Collins, et al., Quantum computational advantage with a programmable photonic processor, Nature 606, 75 (2022).
  • Houck et al. [2012] A. A. Houck, H. E. Türeci, and J. Koch, On-chip quantum simulation with superconducting circuits, Nat. Phys. 8, 292 (2012).
  • Semeghini et al. [2021] G. Semeghini, H. Levine, A. Keesling, S. Ebadi, T. T. Wang, D. Bluvstein, R. Verresen, H. Pichler, M. Kalinowski, R. Samajdar, A. Omran, S. Sachdev, A. Vishwanath, M. Greiner, V. Vuletić, and M. D. Lukin, Probing topological spin liquids on a programmable quantum simulator, Science 374, 1242 (2021).
  • Satzinger et al. [2021] K. J. Satzinger, Y. Liu, A. Smith, C. Knapp, M. Newman, C. Jones, Z. Chen, C. Quintana, X. Mi, A. Dunsworth, C. Gidney, I. Aleiner, F. Arute, K. Arya, J. Atalaya, R. Babbush, J. C. Bardin, R. Barends, J. Basso, A. Bengtsson, A. Bilmes, M. Broughton, B. B. Buckley, D. A. Buell, B. Burkett, N. Bushnell, B. Chiaro, R. Collins, W. Courtney, S. Demura, A. R. Derk, D. Eppens, C. Erickson, E. Farhi, L. Foaro, A. G. Fowler, B. Foxen, M. Giustina, A. Greene, J. A. Gross, M. P. Harrigan, S. D. Harrington, J. Hilton, S. Hong, T. Huang, W. J. Huggins, L. B. Ioffe, S. V. Isakov, E. Jeffrey, Z. Jiang, D. Kafri, K. Kechedzhi, T. Khattar, S. Kim, P. V. Klimov, A. N. Korotkov, F. Kostritsa, D. Landhuis, P. Laptev, A. Locharla, E. Lucero, O. Martin, J. R. McClean, M. McEwen, K. C. Miao, M. Mohseni, S. Montazeri, W. Mruczkiewicz, J. Mutus, O. Naaman, M. Neeley, C. Neill, M. Y. Niu, T. E. O’Brien, A. Opremcak, B. Pató, A. Petukhov, N. C. Rubin, D. Sank, V. Shvarts, D. Strain, M. Szalay, B. Villalonga, T. C. White, Z. Yao, P. Yeh, J. Yoo, A. Zalcman, H. Neven, S. Boixo, A. Megrant, Y. Chen, J. Kelly, V. Smelyanskiy, A. Kitaev, M. Knap, F. Pollmann, and P. Roushan, Realizing topologically ordered states on a quantum processor, Science 374, 1237 (2021).
  • Wen [1991] X. G. Wen, Mean-field theory of spin-liquid states with finite energy gap and topological orders, Phys. Rev. B 44, 2664 (1991).
  • Sachdev [1992] S. Sachdev, Kagomé- and triangular-lattice Heisenberg antiferromagnets: Ordering from quantum fluctuations and quantum-disordered ground states with unconfined bosonic spinons, Phys. Rev. B 45, 12377 (1992).
  • Eisert et al. [2020] J. Eisert, D. Hangleiter, N. Walk, I. Roth, D. Markham, R. Parekh, U. Chabaud, and E. Kashefi, Quantum certification and benchmarking, Nature Rev. Phys. 2, 382 (2020).
  • Bluvstein et al. [2022] D. Bluvstein, H. Levine, G. Semeghini, T. T. Wang, S. Ebadi, M. Kalinowski, A. Keesling, N. Maskara, H. Pichler, M. Greiner, V. Vuletić, and M. D. Lukin, A quantum processor based on coherent transport of entangled atom arrays, Nature 604, 451 (2022).
  • Brydges et al. [2019] T. Brydges, A. Elben, P. Jurcevic, B. Vermersch, C. Maier, B. P. Lanyon, P. Zoller, R. Blatt, and C. F. Roos, Probing Rényi entanglement entropy via randomized measurements, Science 364, 260 (2019).
  • Hu et al. [2024] H.-Y. Hu, A. Gu, S. Majumder, H. Ren, Y. Zhang, D. S. Wang, Y.-Z. You, Z. Minev, S. F. Yelin, and A. Seif, Demonstration of robust and efficient quantum property learning with shallow shadows (2024), arXiv:2402.17911 .
  • Kirk et al. [2022] K. V. Kirk, J. Cotler, H.-Y. Huang, and M. D. Lukin, Hardware-efficient learning of quantum many-body states (2022), arXiv:2212.06084 .
  • Gross et al. [2010] D. Gross, Y.-K. Liu, S. T. Flammia, S. Becker, and J. Eisert, Quantum state tomography via compressed sensing, Phys. Rev. Lett. 105, 150401 (2010).
  • Haah et al. [2017] J. Haah, A. W. Harrow, Z. Ji, X. Wu, and N. Yu, Sample-optimal tomography of quantum states, IEEE Trans. Inf. Theory 63, 5628 (2017).
  • Cramer et al. [2010] M. Cramer, M. B. Plenio, S. T. Flammia, R. Somma, D. Gross, S. D. Bartlett, O. Landon-Cardinal, D. Poulin, and Y.-K. Liu, Efficient quantum state tomography, Nat. Commun. 1, 149 (2010).
  • Baumgratz et al. [2013] T. Baumgratz, D. Gross, M. Cramer, and M. B. Plenio, Scalable reconstruction of density matrices, Phys. Rev. Lett. 111, 020401 (2013).
  • Hübener et al. [2013] R. Hübener, A. Mari, and J. Eisert, Wick’s theorem for matrix product states, Phys. Rev. Lett. 110, 040401 (2013).
  • Perez-Garcia et al. [2007] D. Perez-Garcia, F. Verstraete, M. M. Wolf, and J. I. Cirac, Matrix product state representations (2007), arxiv:quant-ph/0608197 .
  • Eisert et al. [2010] J. Eisert, M. Cramer, and M. B. Plenio, Area laws for the entanglement entropy, Rev. Mod. Phys. 82, 277 (2010).
  • Torlai et al. [2018] G. Torlai, G. Mazzola, J. Carrasquilla, M. Troyer, R. Melko, and G. Carleo, Neural-network quantum state tomography, Nat. Phys. 14, 447 (2018).
  • Carrasquilla et al. [2019] J. Carrasquilla, G. Torlai, R. G. Melko, and L. Aolita, Reconstructing quantum states with generative models, Nature Mach. Intell. 1, 155 (2019).
  • Huang et al. [2020] H.-Y. Huang, R. Kueng, and J. Preskill, Predicting many properties of a quantum system from very few measurements, Nat. Phys. 16, 1050 (2020).
  • Wang et al. [2020] J. Wang, Z.-Y. Han, S.-B. Wang, Z. Li, L.-Z. Mu, H. Fan, and L. Wang, Scalable quantum tomography with fidelity estimation, Phys. Rev. A 101, 032321 (2020).
  • Gomez et al. [2022] A. M. Gomez, S. F. Yelin, and K. Najafi, Reconstructing Quantum States Using Basis-Enhanced Born Machines (2022), arXiv:2206.01273 [quant-ph] .
  • Kurmapu et al. [2023] M. K. Kurmapu, V. Tiunova, E. Tiunov, M. Ringbauer, C. Maier, R. Blatt, T. Monz, A. K. Fedorov, and A. Lvovsky, Reconstructing complex states of a 20202020-qubit quantum simulator, PRX Quantum 4, 040345 (2023).
  • Torlai et al. [2023] G. Torlai, C. J. Wood, A. Acharya, G. Carleo, J. Carrasquilla, and L. Aolita, Quantum process tomography with unsupervised learning and tensor networks, Nat. Commun. 14, 2858 (2023).
  • Huang et al. [2022] H.-Y. Huang, R. Kueng, G. Torlai, V. V. Albert, and J. Preskill, Provably efficient machine learning for quantum many-body problems, Science 377, eabk3333 (2022).
  • Bluvstein et al. [2023] D. Bluvstein, S. J. Evered, A. A. Geim, S. H. Li, H. Zhou, T. Manovitz, S. Ebadi, M. Cain, M. Kalinowski, D. Hangleiter, J. P. Bonilla Ataides, N. Maskara, I. Cong, X. Gao, P. Sales Rodriguez, T. Karolyshyn, G. Semeghini, M. J. Gullans, M. Greiner, V. Vuletić, and M. D. Lukin, Logical quantum processor based on reconfigurable atom arrays, Nature 626, 58–65 (2023).
  • Bravyi and Terhal [2010] S. Bravyi and B. Terhal, Complexity of stoquastic frustration-free Hamiltonians, SIAM J. Comput. 39, 1462 (2010).
  • Newey and McFadden [1994] W. K. Newey and D. McFadden, Chapter 36 Large sample estimation and hypothesis testing, in Handbook of Econometrics, Vol. 4 (Elsevier, 1994) pp. 2111–2245.
  • Wilde et al. [2022] F. Wilde, A. Kshetrimayum, I. Roth, D. Hangleiter, R. Sweke, and J. Eisert, Scalably learning quantum many-body Hamiltonians from dynamical data (2022), arxiv:2209.14328 .
  • Scholten and Blume-Kohout [2018] T. L. Scholten and R. Blume-Kohout, Behavior of the maximum likelihood in quantum state tomography, New Journal of Physics 20, 023050 (2018).
  • Hajri et al. [2017] H. Hajri, S. Said, and Y. Berthoumieu, \ORIGselectlanguageenglishMaximum likelihood estimators on manifolds, in \ORIGselectlanguageenglishGeometric Science of Information, Vol. 10589, edited by F. Nielsen and F. Barbaresco (Springer International Publishing, Cham, 2017) pp. 692–700, series Title: Lecture Notes in Computer Science.
  • Haegeman et al. [2014] J. Haegeman, M. Mariën, T. J. Osborne, and F. Verstraete, Geometry of matrix product states: Metric, parallel transport, and curvature, J. Math. Phys. 55, 021902 (2014).
  • Elben et al. [2022] A. Elben, S. T. Flammia, H.-Y. Huang, R. Kueng, J. Preskill, B. Vermersch, and P. Zoller, The randomized measurement toolbox, Nature Rev. Phys. 5, 9–24 (2022).
  • Hadfield et al. [2020] C. Hadfield, S. Bravyi, R. Raymond, and A. Mezzacapo, Measurements of quantum Hamiltonians with locally-biased classical shadows (2020), arXiv:2006.15788 .
  • Schollwöck [2011] U. Schollwöck, The density-matrix renormalization group in the age of matrix product states, Ann. Phys. 326, 96 (2011).
  • Shaw et al. [2024] A. L. Shaw, Z. Chen, J. Choi, D. K. Mark, P. Scholl, R. Finkelstein, A. Elben, S. Choi, and M. Endres, Benchmarking highly entangled states on a 60-atom analogue quantum simulator, Nature 628, 71 (2024).
  • Huang et al. [2024] H.-Y. Huang, J. Preskill, and M. Soleimanifar, Certifying almost all quantum states with few single-qubit measurements (2024), arXiv:2404.07281 .
  • Kitaev [2003] A. Y. Kitaev, Fault-tolerant quantum computation by anyons, Ann. Phys. 303, 2 (2003).
  • Anderson [1967] P. W. Anderson, Infrared catastrophe in Fermi gases with local scattering potentials, Phys. Rev. Lett. 18, 1049 (1967).
  • Zanardi and Paunković [2006] P. Zanardi and N. Paunković, Ground state overlap and quantum phase transitions, Phys. Rev. E 74, 031123 (2006).
  • de Léséleuc et al. [2019] S. de Léséleuc, V. Lienhard, P. Scholl, D. Barredo, S. Weber, N. Lang, H. P. Büchler, T. Lahaye, and A. Browaeys, Observation of a symmetry-protected topological phase of interacting bosons with Rydberg atoms, Science 365, 775 (2019).
  • Samajdar et al. [2020] R. Samajdar, W. W. Ho, H. Pichler, M. D. Lukin, and S. Sachdev, Complex density wave orders and quantum phase transitions in a model of square-lattice Rydberg atom arrays, Phys. Rev. Lett. 124, 103601 (2020).
  • Scholl et al. [2021] P. Scholl, M. Schuler, H. J. Williams, A. A. Eberharter, D. Barredo, K.-N. Schymik, V. Lienhard, L.-P. Henry, T. C. Lang, T. Lahaye, A. M. Läuchli, and A. Browaeys, Quantum simulation of 2D antiferromagnets with hundreds of Rydberg atoms, Nature 595, 233 (2021).
  • Moessner and Raman [2011] R. Moessner and K. S. Raman, Quantum dimer models, in Introduction to Frustrated Magnetism (Springer, 2011) pp. 437–479.
  • Misguich et al. [2002] G. Misguich, D. Serban, and V. Pasquier, Quantum dimer model on the Kagome lattice: Solvable dimer-liquid and Ising gauge theory, Phys. Rev. Lett. 89, 137202 (2002).
  • Samajdar et al. [2021] R. Samajdar, W. W. Ho, H. Pichler, M. D. Lukin, and S. Sachdev, Quantum phases of Rydberg atoms on a Kagome lattice, Proc. Natl. Acad. Sci. U.S.A. 118, e2015785118 (2021).
  • Verresen et al. [2021] R. Verresen, M. D. Lukin, and A. Vishwanath, Prediction of toric code topological order from Rydberg blockade, Phys. Rev. X 11, 031005 (2021).
  • Samajdar et al. [2023] R. Samajdar, D. G. Joshi, Y. Teng, and S. Sachdev, Emergent 2subscript2{\mathbb{Z}}_{2}blackboard_Z start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT gauge theories and topological excitations in Rydberg atom arrays, Phys. Rev. Lett. 130, 043601 (2023).
  • Cong et al. [2024] I. Cong, N. Maskara, M. C. Tran, H. Pichler, G. Semeghini, S. F. Yelin, S. Choi, and M. D. Lukin, Enhancing detection of topological order by local error correction, Nat. Commun. 15, 1527 (2024).
  • Kitaev and Preskill [2006] A. Kitaev and J. Preskill, Topological entanglement entropy, Phys. Rev. Lett. 96, 110404 (2006).
  • Levin and Wen [2006] M. Levin and X.-G. Wen, Detecting topological order in a ground state wave function, Phys. Rev. Lett. 96, 110405 (2006).
  • Glasser et al. [2019] I. Glasser, R. Sweke, N. Pancotti, J. Eisert, and I. Cirac, Expressive power of tensor-network factorizations for probabilistic modeling, in Advances in Neural Information Processing Systems, Vol. 32 (Curran Associates, Inc., 2019).
  • Verstraete and Cirac [2004] F. Verstraete and J. I. Cirac, Renormalization algorithms for quantum-many body systems in two and higher dimensions (2004), arXiv:cond-mat/0407066 .
  • Koepsell et al. [2021] J. Koepsell, D. Bourgund, P. Sompet, S. Hirthe, A. Bohrdt, Y. Wang, F. Grusdt, E. Demler, G. Salomon, C. Gross, and I. Bloch, Microscopic evolution of doped Mott insulators from polaronic metal to Fermi liquid, Science 374, 82–86 (2021).
  • Denzler et al. [2023] J. Denzler, A. A. Mele, E. Derbyshire, T. Guaita, and J. Eisert, Learning fermionic correlations by evolving with random translationally invariant Hamiltonians (2023), arXiv:2309.12933 [quant-ph] .
  • Robledo Moreno et al. [2022] J. Robledo Moreno, G. Carleo, A. Georges, and J. Stokes, Fermionic wave functions from neural-network constrained hidden states, Proc. Natl. Acad. Sci. U.S.A 119, e2122059119 (2022).
  • Shackleton and Zhang [2024] H. Shackleton and S. Zhang, Variational wavefunctions for topologically-ordered Fermi liquids, Bull. Am. Phys. Soc. , W08.00014 (March 2024).
  • Müller et al. [2024] T. Müller, R. Thomale, S. Sachdev, and Y. Iqbal, Polaronic correlations from optimized ancilla wave functions for the Fermi-Hubbard model, to appear  (2024).
  • Evered et al. [2023] S. J. Evered, D. Bluvstein, M. Kalinowski, S. Ebadi, T. Manovitz, H. Zhou, S. H. Li, A. A. Geim, T. T. Wang, N. Maskara, H. Levine, G. Semeghini, M. Greiner, V. Vuletić, and M. D. Lukin, High-fidelity parallel entangling gates on a neutral atom quantum computer, Nature 622, 268 (2023).
  • Lange et al. [2023] H. Lange, M. Kebrič, M. Buser, U. Schollwöck, F. Grusdt, and A. Bohrdt, Adaptive quantum state tomography with active learning, Quantum 7, 1129 (2023).
  • Gebhart et al. [2023] V. Gebhart, R. Santagati, A. A. Gentile, E. M. Gauger, D. Craig, N. Ares, L. Banchi, F. Marquardt, L. Pezzè, and C. Bonato, Learning quantum systems, Nature Rev. Phys. 5, 141 (2023).
  • Ohliger et al. [2013] M. Ohliger, V. Nesme, and J. Eisert, Efficient and feasible state tomography of quantum many-body systems, New J. Phys. 15, 015024 (2013).
  • Nietner [2023] A. Nietner, Unifying (quantum) statistical and parametrized (quantum) algorithms (2023), arXiv:2310.17716 .
  • Kufel et al. [2024] D. S. Kufel, J. Kemp, S. M. Linsel, C. R. Laumann, and N. Y. Yao, Approximately-symmetric neural networks for quantum spin liquids, arXiv preprint arXiv:2405.17541  (2024).
  • Wei et al. [2023] V. Wei, W. A. Coish, P. Ronagh, and C. A. Muschik, Neural-shadow quantum state tomography (2023), arXiv:2305.01078 .
  • Chia et al. [2024] N.-H. Chia, C.-Y. Lai, and H.-H. Lin, Efficient learning of t𝑡titalic_t-doped stabilizer states with single-copy measurements, Quantum 8, 1250 (2024).
  • Masot-Llima and Garcia-Saez [2024] S. Masot-Llima and A. Garcia-Saez, Stabilizer tensor networks: universal quantum simulator on a basis of stabilizer states (2024), arXiv:2403.08724 .
  • Chen et al. [2021] S. Chen, J. Cotler, H.-Y. Huang, and J. Li, A hierarchy for replica quantum advantage (2021), arXiv:2111.05874 .
  • Islam et al. [2015] R. Islam, R. Ma, P. M. Preiss, M. Eric Tai, A. Lukin, M. Rispoli, and M. Greiner, Measuring entanglement entropy in a quantum many-body system, Nature 528, 77 (2015).
  • Elben et al. [2018] A. Elben, B. Vermersch, M. Dalmonte, J. I. Cirac, and P. Zoller, Rényi entropies from random quenches in atomic Hubbard and spin models, Phys. Rev. Lett. 120, 050406 (2018).
  • Kokail et al. [2021] C. Kokail, R. van Bijnen, A. Elben, B. Vermersch, and P. Zoller, Entanglement Hamiltonian tomography in quantum simulation, Nat. Phys. 17, 936 (2021).
  • Elben et al. [2020] A. Elben, J. Yu, G. Zhu, M. Hafezi, F. Pollmann, P. Zoller, and B. Vermersch, Many-body topological invariants from randomized measurements in synthetic quantum matter, Sci. Adv. 6, eaaz3666 (2020).
  • Cian et al. [2021] Z.-P. Cian, H. Dehghani, A. Elben, B. Vermersch, G. Zhu, M. Barkeshli, P. Zoller, and M. Hafezi, Many-body Chern number from statistical correlations of randomized measurements, Phys. Rev. Lett. 126, 050501 (2021).
  • Werner et al. [2016] A. H. Werner, D. Jaschke, P. Silvi, M. Kliesch, T. Calarco, J. Eisert, and S. Montangero, Positive tensor network approach for simulating open quantum many-body systems, Physical review letters 116, 237201 (2016).
  • Kühnel [2008] W. Kühnel, Differentialgeometrie, 4th ed. (Vieweg, Wiesbaden, 2008).
  • Nakahara [2003] M. Nakahara, \ORIGselectlanguageenglishGeometry, Topology and Physics, 2nd ed. (Taylor & Francis, 2003).
  • Cheng [2013] R. Cheng, Quantum geometric tensor (Fubini-Study metric) in simple quantum system: A pedagogical introduction (2013), arXiv:1012.1337 .
  • Bengtsson and Zyczkowski [2020] I. Bengtsson and K. Zyczkowski, Geometry of quantum states (Cambridge University Press, Cambridge, 2020).
  • Hauru et al. [2021] M. Hauru, M. Van Damme, and J. Haegeman, Riemannian optimization of isometric tensor networks, SciPost Phys. 10, 040 (2021).
  • Luchnikov et al. [2021] I. A. Luchnikov, M. E. Krechetov, and S. N. Filippov, Riemannian geometry and automatic differentiation for optimization problems of quantum physics and quantum technologies, New J. Phys. 23, 073006 (2021).
  • Ippoliti et al. [2023] M. Ippoliti, Y. Li, T. Rakovszky, and V. Khemani, Operator relaxation and the optimal depth of classical shadows, Phys. Rev. Lett. 130, 230403 (2023).
  • Hu et al. [2023] H.-Y. Hu, S. Choi, and Y.-Z. You, Classical shadow tomography with locally scrambled quantum dynamics, Phys. Rev. Research 5, 023027 (2023).
  • Bertoni et al. [2024] C. Bertoni, J. Haferkamp, M. Hinsche, M. Ioannou, J. Eisert, and H. Pashayan, Shallow shadows: Expectation estimation using low-depth random Clifford circuits, Phys. Rev. Lett. 132, “in press” (2024).
  • Ebadi et al. [2022] S. Ebadi, A. Keesling, M. Cain, T. T. Wang, H. Levine, D. Bluvstein, G. Semeghini, A. Omran, J.-G. Liu, R. Samajdar, X.-Z. Luo, B. Nash, X. Gao, B. Barak, E. Farhi, S. Sachdev, N. Gemelke, L. Zhou, S. Choi, H. Pichler, S.-T. Wang, M. Greiner, V. Vuletić, and M. D. Lukin, Quantum optimization of maximum independent set using Rydberg atom arrays, Science 376, 1209 (2022).
  • Schiffer et al. [2024] B. F. Schiffer, D. S. Wild, N. Maskara, M. Cain, M. D. Lukin, and R. Samajdar, Circumventing superexponential runtimes for hard instances of quantum adiabatic optimization, Phys. Rev. Res. 6, 013271 (2024).
  • Tikhanovskaya et al. [2024] M. Tikhanovskaya, S. Sachdev, and R. Samajdar, Equilibrium dynamics of infinite-range quantum spin glasses in a field, PRX Quantum 5, 020313 (2024).
  • Troyer and Wiese [2005] M. Troyer and U.-J. Wiese, Computational complexity and fundamental limitations to fermionic quantum Monte Carlo simulations, Phys. Rev. Lett. 94, 170201 (2005).
  • Hangleiter et al. [2020] D. Hangleiter, I. Roth, D. Nagaj, and J. Eisert, Easing the Monte Carlo sign problem, Sci. Adv. 6, eabb8341 (2020).
  • Gray [2018] J. Gray, Quimb: A Python Package for Quantum Information and Many-Body Calculations, J. Open Source Softw. 3, 819 (2018).
\do@columngrid

oneΔ

Appendix


APPENDIX A Relation to previous works

This appendix aims to provide a comparison of our method to existing quantum state tomography techniques. We highlight the unique challenges and opportunities in characterizing complex quantum systems, particularly two-dimensional topological states. In addition to the references summarized below, we direct interested readers to Ref. [66] for a comprehensive review of quantum state tomography. Finally, having discussed advantages of our method for learning quasi-2d2𝑑2d2 italic_d quantum states with topological order, we summarize the limitations faced by our method discussed in the main text.

  1. 1.

    Matrix product state (MPS) tomography [18, 19, 67, 20] is an effective method for reconstructing injective one-dimensional quantum states with short-range correlations associated with a local parent Hamiltonian, by finding the best MPS consistent with local reduced density matrices from randomized measurements. In principle, these reduced density matrices can be estimated from Pauli measurements using the techniques of classical shadows [68]. Its application to 2d2𝑑2d2 italic_d topological states, however, is challenging: first of all, the size of the reduced density matrices required for uniquely determining a quantum state in 2d2𝑑2d2 italic_d is unclear, and secondly, due to nontrivial ground-state degeneracies of topological states, they cannot be uniquely determined by local operators and thus, cannot be efficiently determined with this method.

  2. 2.

    Variational MPS tomography [26, 27, 28] extends MPS tomography to one-dimensional states without the injectivity constraint. It uses a variational approach to optimize the MPS representation through maximum likelihood estimation (MLE). Prior work also considered using matrix product operator (MPO) for quantum process tomography [29]. Our work extends this protocol to 2d2𝑑2d2 italic_d states using a “snake path”, commonly adopted in cylindrical DMRG calculations for 2d2𝑑2d2 italic_d systems. We further enhance our method by incorporating regularization techniques inspired by classical shadow tomography, combining the benefits of both approaches to better characterize two-dimensional quantum systems. We refer the reader to Sec. II for the details of our protocol.

  3. 3.

    Neural network quantum state tomography (NNQST) [23] utilizes neural networks as variational ansätze within the MLE framework to model quantum states. Although particularly successful for states with fully positive amplitudes (for example, ground states of frustration-free stoquastic Hamiltonians [32]), this method requires careful consideration when the target states have nontrivial signs and phases. Using measurement outcomes from different bases is therefore essential—a requirement directly addressed in MPS-based approaches, which can inherently relate measurements post local unitary transformations to learn relative phases. In this context, it may be useful to explore a most recently proposed architecture that can accurately represent quantum spin liquids by incorporating physical symmetries [69].

  4. 4.

    Generative modeling for density matrix reconstruction [24] extends the use of neural networks to mixed-state tomography. This approach uses single-qubit positive operator-valued measure (POVM) measurements for information-complete reconstructions. While generative models can theoretically model measurement probability distributions across an exponentially large outcome space, practical challenges arise in ensuring convergence to the accurate density matrix, particularly given the sparse nature of such distributions in large systems. It is unclear how effective this approach is [25] for learning entangled quantum states such as the topological ordered states considered in this work.

  5. 5.

    Classical shadow tomography [25, 30] efficiently estimates properties of quantum states by measuring them post application of random unitary rotations, followed by post-processing. This technique has demonstrated that a number of samples independent of the system size is sufficient for estimating the fidelity of specific projectors, such as ground states of the toric code Hamiltonian [25], given prior knowledge of the target projector and the utilization of global Clifford measurements. A significant extension of this method [30, 15] employed a dimensional reduction technique along with a nonlinear kernel for post-processing to distinguish toric code states from trivial states using only random Pauli measurements. This unsupervised learning task, while innovative, requires the assembly of numerous distinct quantum states for shadow tomography, presenting a considerable experimental demand. Our method, utilizing random-XZ𝑋𝑍XZitalic_X italic_Z measurements, sidesteps these requirements, offering a more experimentally feasible approach for a broader range of quantum systems.

  6. 6.

    Shadow-NNQST [70] combines a neural-network ansatz used in NNQST with “regularization” of fidelity estimations from classical shadow tomography. This regularization step necessitates global Clifford measurements for efficient fidelity estimation [25], and an arbitrary Clifford unitary on n𝑛nitalic_n qubits has nontrivial depth. In contrast, our approach leverages random-XZ𝑋𝑍XZitalic_X italic_Z measurements, avoiding the need for such specific, experimentally costly measurement protocols. More recently, a quantity termed the “shadow overlap” has demonstrated faithful certification of fidelity requiring only polynomial sample complexity for a large set of quantum states [42], which could significantly improve the training process of NNQST or our variational tensor network tomography. Notably, this technique also only uses single-qubit measurements that are experimentally feasible.

  7. 7.

    Learning an unknown stabilizer state can be carried out efficiently using single-copy measurements [71]. Efficient learning of stabilizer states through single-copy measurements offers a direct method for those topological states that are also stabilizer states. Our protocol expands its applicability to nonstabilizer states amenable to representation as an MPS, which allows for broader quantum state characterization beyond the stabilizer formalism. A recent work has considered combining MPS with the stabilizer formalism [72], which allows for the simulations of states beyond the scope of either method individually—this framework could be also explored in the current context of quantum state tomography.

  8. 8.

    Predicting specific properties of quantum states. If one knows the targeted properties that one wants to learn, there are a plethora of alternative methods, of which we list the most relevant approaches below. We will not try to compare our protocol to all of these directly, since such methods differ in their flexibility with regard to experimental measurements as well as the number of samples required to achieve accurate predictions.

    1. (a)

      Rényi entropy:

      • A multicopy approach [73], based on measuring two copies of a quantum state, has been demonstrated experimentally [74, 12]. However, this requires the preparation of multiple identical copies of a state while demanding advanced control over the quantum system to ensure the fidelity of the copies.

      • Randomized measurements [13, 75, 14] leverage the statistical properties of the outcomes to estimate the Rényi entropy without needing multiple copies of a system. This is both similar to and encompasses classical shadow tomography.

    2. (b)

      Von Neumann entropy:

      • Entanglement Hamiltonian tomography [76] assumes that the reduced density matrices of a subsystem can be approximated by an ansatz involving local operators in the Hamiltonian. By learning the coefficients of this ansatz, the technique aims to reconstruct the entanglement Hamiltonian and thereby predict the von Neumann entropy.

    3. (c)

      Topological invariants:

      • Direct tomography can be applied to small subsystem sizes, from which the topological entanglement entropy can be extracted [55, 56]; this approach was demonstrated experimentally for a surface code [8]. Nevertheless, the applicability of this technique for general systems remains unclear, particularly with regard to the required size of the subsystem for accurate measurement of the entanglement entropy.

      • Randomized measurements can also be used to predict topological invariants of systems such as symmetry protected topological (SPT) phases [77] or fractional quantum Hall states [78].

A.1 Potential advantages and limitations

Our variational tensor network tomography, aimed at diagnostics and recovery of quantum states, offers potential for extracting various properties such as the Von Neumann entanglement entropy, difficult to learn using other techniques. This method combines maximum likelihood estimation of the quantum state with randomized measurements and is detailed in previous works such as [26, 28, 29]. We demonstrate this approach using simplified XZ𝑋𝑍XZitalic_X italic_Z measurements, which proves sufficient for the two-dimensional topological quantum states considered, with complex long-range entanglement. Such quantum states are represented by non-injective MPS, and can not be learned via MPS tomography [18, 19].

One significant limitation is that the target state must admit an efficient tensor network representation. In numerical experiments, we use an MPS ansatz, recognizing that while it effectively approximates one-dimensional states, it may not capture the full complexity of two-dimensional systems. This works sets the stage for future exploration of more sophisticated tensor networks. We also discuss the adaptation of statistical properties of MLE for quantum tomography, outlining both its potential and the challenges posed by the gauge freedom in MPS representations.

Another limitation is that real laboratory states can be mixed, which requires using generalizations of tensor networks such as locally purified matrix product operators [79]. Additionally, one generally do not have access to the ground truth fidelity with the lab state. This necessitates methods like cross-validation as discussed in Appendix G or efficient fidelity certifications [42] to validate the trained model against empirical data.

APPENDIX B Asymptotic MLE error via MPS manifold

Our variational algorithm in the framework of maximum likelihood estimation (MLE) aims to find the MPS representation |ψ(A)ket𝜓𝐴|\psi({A})\rangle| italic_ψ ( italic_A ) ⟩ that is most consistent with measurements of the target state vector |ϕketitalic-ϕ|\phi\rangle| italic_ϕ ⟩. Intuitively, we expect that the accuracy of the optimal MLE solution improves with more samples. The goal of this appendix is to analyze the accuracy of such a solution as a function of the number of samples N𝑁Nitalic_N. More precisely, we provide a probabilistic upper bound of the infidelity 1F1𝐹1-F1 - italic_F by deriving a concentration inequality using the statistical properties of MLE. First, we summarize our main finding of the probabilistic bound in the theorem below:

Theorem 2 (Probabilistic bound).

Let |ψ(A^(N))ket𝜓superscript^𝐴𝑁|\psi(\hat{A}^{(N)})\rangle| italic_ψ ( over^ start_ARG italic_A end_ARG start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT ) ⟩ be the MLE solution with a maximum bond dimension χmaxsubscript𝜒max\chi_{\mathrm{max}}italic_χ start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT for an n𝑛nitalic_n-qubit target state vector |ϕketitalic-ϕ|\phi\rangle| italic_ϕ ⟩. With probability greater than 1δ1𝛿1-\delta1 - italic_δ, the infidelity is upper bounded asymptotically in the limit of large number of samples N𝑁Nitalic_N as

[1|ϕ|ψ(A^(N))|ϵ(N)]1δ,delimited-[]1inner-productitalic-ϕ𝜓superscript^𝐴𝑁italic-ϵ𝑁1𝛿\mathbb{P}\left[1-\lvert\mbox{$\langle\phi|\psi(\hat{A}^{(N)})\rangle$}\rvert% \leq\epsilon(N)\right]\geq 1-\delta,blackboard_P [ 1 - | ⟨ italic_ϕ | italic_ψ ( over^ start_ARG italic_A end_ARG start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT ) ⟩ | ≤ italic_ϵ ( italic_N ) ] ≥ 1 - italic_δ , (8)

where ϵ(N)=O(nχmax2Nδ)italic-ϵ𝑁𝑂𝑛superscriptsubscript𝜒max2𝑁𝛿\epsilon(N)=O\left(\sqrt{\displaystyle\frac{n\chi_{\mathrm{max}}^{2}}{N\delta}% }\right)italic_ϵ ( italic_N ) = italic_O ( square-root start_ARG divide start_ARG italic_n italic_χ start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_N italic_δ end_ARG end_ARG ).

We provide a detailed and rigorous proof of Theorem 2 concerning the optimal MLE solution, focusing on conditions under which it represents a global minimum. Achieving such a global minimum may not always be practical, and our theorem does not address the complexity of reaching such a solution. Moreover, the theorem is specifically applicable in the asymptotic limit as N𝑁N\rightarrow\inftyitalic_N → ∞. For the validity of our theorem, we assume that the target state is within the variational space used for MLE, i.e., it is an MPS of known bond dimension. In practice this condition might not be satisfied.

Ideally, one would like to prove a theorem that applies to our variational MPS tomography in practice. There are a wide variety of technical obstacles to achieving this; thus, our goal in this work is to take the first steps toward this direction and provide a thorough introduction of the required mathematical tools. In this process, we (a) present an initial theorem under a specific set of assumptions, and (b) offer pedagogical insights into the technical tools and conceptual challenges involved in proving our theorem and potentially more comprehensive statements. The success of our numerical results, combined with these theoretical explorations, provides both motivation and foundational insights for develo** more specialized analyses.

To build some intuition, we first review the maximum likelihood estimation (MLE) within traditional settings. It turns out that these results cannot be directly applied to our context due to the gauge freedom of MPS. To generalize the MLE analysis to our MPS-estimation setting, we first elaborate on the gaps that we are filling in and how our statement differs from conventional results.

Let us begin with a standard MLE problem, where the goal is to estimate a single parameter. Considering a sequence of N𝑁Nitalic_N independent and identically distributed (i.i.d.) samples (xi)i=1Nsuperscriptsubscriptsubscript𝑥𝑖𝑖1𝑁(x_{i})_{i=1}^{N}( italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT drawn from a target Gaussian distribution 𝒫θ0subscript𝒫subscript𝜃0\mathcal{P}_{\theta_{0}}caligraphic_P start_POSTSUBSCRIPT italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT with mean θ0subscript𝜃0\theta_{0}italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and a known variance σ2superscript𝜎2\sigma^{2}italic_σ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT,

𝒫θ0(x)=12πσexp[(xθ0)22σ2].subscript𝒫subscript𝜃0𝑥12𝜋𝜎superscript𝑥subscript𝜃022superscript𝜎2\mathcal{P}_{\theta_{0}}(x)=\frac{1}{\sqrt{2\pi}\sigma}\exp\left[-\frac{(x-% \theta_{0})^{2}}{2\sigma^{2}}\right].caligraphic_P start_POSTSUBSCRIPT italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT ( italic_x ) = divide start_ARG 1 end_ARG start_ARG square-root start_ARG 2 italic_π end_ARG italic_σ end_ARG roman_exp [ - divide start_ARG ( italic_x - italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_σ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ] . (9)

Since the variance is known, our goal is to find the estimator θ^(N)superscript^𝜃𝑁\hat{\theta}^{(N)}over^ start_ARG italic_θ end_ARG start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT of the mean such that the likelihood of the observed samples is maximized. The negative log-likelihood (NLL) of observing the N𝑁Nitalic_N samples is

(θ)=1Ni=1Nlog(exp[(xiθ)22σ2])=1Ni=1N(xiθ)22σ2.𝜃1𝑁superscriptsubscript𝑖1𝑁superscriptsubscript𝑥𝑖𝜃22superscript𝜎21𝑁superscriptsubscript𝑖1𝑁superscriptsubscript𝑥𝑖𝜃22superscript𝜎2\mathcal{L}(\theta)=-\frac{1}{N}\sum_{i=1}^{N}\log\left(\exp\left[-\frac{(x_{i% }-\theta)^{2}}{2\sigma^{2}}\right]\right)=\frac{1}{N}\sum_{i=1}^{N}\frac{(x_{i% }-\theta)^{2}}{2\sigma^{2}}.caligraphic_L ( italic_θ ) = - divide start_ARG 1 end_ARG start_ARG italic_N end_ARG ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT roman_log ( roman_exp [ - divide start_ARG ( italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - italic_θ ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_σ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ] ) = divide start_ARG 1 end_ARG start_ARG italic_N end_ARG ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT divide start_ARG ( italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - italic_θ ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_σ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG . (10)

For this simple case of using a Gaussian distribution as our model prior, the minimization of the NLL \mathcal{L}caligraphic_L is the same as minimizing the least-squared error. Differentiating \mathcal{L}caligraphic_L with respect to θ𝜃\thetaitalic_θ we readily find the extremum θ^(N)=1Nixisuperscript^𝜃𝑁1𝑁subscript𝑖subscript𝑥𝑖\hat{\theta}^{(N)}=\frac{1}{N}\sum_{i}x_{i}over^ start_ARG italic_θ end_ARG start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT = divide start_ARG 1 end_ARG start_ARG italic_N end_ARG ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, which is the well-known empirical mean.

Note that the estimator θ^(N)superscript^𝜃𝑁\hat{\theta}^{(N)}over^ start_ARG italic_θ end_ARG start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT is a random variable, as it depends on the random samples (xi)i=1Nsuperscriptsubscriptsubscript𝑥𝑖𝑖1𝑁(x_{i})_{i=1}^{N}( italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT. Hence, we can describe the probability distribution of θ^(N)superscript^𝜃𝑁\hat{\theta}^{(N)}over^ start_ARG italic_θ end_ARG start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT. Under certain conditions this distribution tends to the normal distribution as the number of samples goes to infinity. This property is known as asymptotic normality and has been studied rigorously [33]. We will state the precise technical conditions for asymptotic normality later (in the context of MLE on manifolds) in Lemma 3, but emphasize here already that, in general, they entail two crucial requirements: 1) the MLE solution is consistent such that θ0subscript𝜃0\theta_{0}italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the unique solution—the unique minimizer of \mathcal{L}caligraphic_L in the infinite-sample limit—describing 𝒫θ0subscript𝒫subscript𝜃0\mathcal{P}_{\theta_{0}}caligraphic_P start_POSTSUBSCRIPT italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT, i.e., there exists no other θ𝜃\thetaitalic_θ for which 𝒫θ=𝒫θ0subscript𝒫𝜃subscript𝒫subscript𝜃0\mathcal{P}_{\theta}=\mathcal{P}_{\theta_{0}}caligraphic_P start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT = caligraphic_P start_POSTSUBSCRIPT italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT; 2) the landscape of the loss function \mathcal{L}caligraphic_L is well behaved for all N𝑁Nitalic_N, and in particular, in the limit N𝑁N\rightarrow\inftyitalic_N → ∞, the Hessian needs to be well behaved. To be precise, asymptotic normality means that in the limit of a large number of samples N𝑁Nitalic_N, the estimator θ^(N)superscript^𝜃𝑁\hat{\theta}^{(N)}over^ start_ARG italic_θ end_ARG start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT converges to a normal distribution in the sense

N(θ^(N)θ0)d.𝒩(0,σ2).\sqrt{N}(\hat{\theta}^{(N)}-\theta_{0})\overset{\mathrm{d.}}{\longrightarrow}% \mathcal{N}(0,\sigma^{2}).square-root start_ARG italic_N end_ARG ( over^ start_ARG italic_θ end_ARG start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT - italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) start_OVERACCENT roman_d . end_OVERACCENT start_ARG ⟶ end_ARG caligraphic_N ( 0 , italic_σ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) . (11)

Here, d.\overset{\mathrm{d.}}{\rightarrow}start_OVERACCENT roman_d . end_OVERACCENT start_ARG → end_ARG denotes convergence in distribution, i.e., as N𝑁Nitalic_N increases, the sequence of distributions of the random variables N(θ^(N)θ0)𝑁superscript^𝜃𝑁subscript𝜃0\sqrt{N}(\hat{\theta}^{(N)}-\theta_{0})square-root start_ARG italic_N end_ARG ( over^ start_ARG italic_θ end_ARG start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT - italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) for growing N𝑁Nitalic_N approximates the normal distribution increasingly well. With the normality condition, we can write down a concentration inequality to probabilistically upper bound the error |θ^(N)θ0|superscript^𝜃𝑁subscript𝜃0\lvert\hat{\theta}^{(N)}-\theta_{0}\rvert| over^ start_ARG italic_θ end_ARG start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT - italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT |. As N𝑁Nitalic_N increases, the error reduces with a rate of 1/N1𝑁1/\sqrt{N}1 / square-root start_ARG italic_N end_ARG so that in the infinite-N𝑁Nitalic_N limit, the convergence θ^(N)p.θ0\hat{\theta}^{(N)}\overset{\mathrm{p.}}{\longrightarrow}\theta_{0}over^ start_ARG italic_θ end_ARG start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT start_OVERACCENT roman_p . end_OVERACCENT start_ARG ⟶ end_ARG italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is guaranteed by the consistency condition. To illustrate how this condition could fail, let us consider a modified Gaussian distribution periodic in the parameter

𝒫θ0=12πσexp[(xsin(θ0))22σ2].subscript𝒫subscript𝜃012𝜋𝜎superscript𝑥subscript𝜃022superscript𝜎2\mathcal{P}_{\theta_{0}}=\frac{1}{\sqrt{2\pi}\sigma}\exp\left[-\frac{(x-\sin(% \theta_{0}))^{2}}{2\sigma^{2}}\right].caligraphic_P start_POSTSUBSCRIPT italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG square-root start_ARG 2 italic_π end_ARG italic_σ end_ARG roman_exp [ - divide start_ARG ( italic_x - roman_sin ( italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_σ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ] . (12)

Now, a parameter shifted by a multiples of 2π2𝜋2\pi2 italic_π still represents the same distribution 𝒫θ0=𝒫θ0+2πsubscript𝒫subscript𝜃0subscript𝒫subscript𝜃02𝜋\mathcal{P}_{\theta_{0}}=\mathcal{P}_{\theta_{0}+2\pi}caligraphic_P start_POSTSUBSCRIPT italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT = caligraphic_P start_POSTSUBSCRIPT italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + 2 italic_π end_POSTSUBSCRIPT. Therefore, θ^(N)superscript^𝜃𝑁\hat{\theta}^{(N)}over^ start_ARG italic_θ end_ARG start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT need not converge to θ0subscript𝜃0\theta_{0}italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. In physics, the phenomenon where multiple parameters describe the same function is known as gauge redundancy. In this simple scenario, we can explicitly account for the periodicity in the optimal estimator, or equivalently impose “gauge fixing” conditions to select a unique representative, by restricting the parameter domain to [0,2π)02𝜋[0,2\pi)[ 0 , 2 italic_π ). However, for distributions parametrized by MPS, it is challenging in practice to fix the gauge degrees of freedom consistently. Specifically, transforming MPS to canonical forms using singular value decomposition only partially fixes the gauge. Additional unitary-gauge degrees of freedom remain in such a decomposition. Moreover, degeneracies in the singular values introduce redundancy within the subspace spanned by degenerate eigenvectors.

Consequently, the MPS tensors that minimize our loss function do not satisfy the consistency condition. Nevertheless, if our measurement scheme is tomographically complete [see Appendix D], the physical state itself meets the consistency condition. This implies that even though multiple MPS tensor configurations may solve the MLE problem, they all correspond to the same physical state. Thus, our goal is to investigate the asymptotic properties of MLE estimators across the set of physical states represented by MPS. These states, represented by a well-defined class of MPS described further in this section, form a manifold. Therefore, our analysis aims to understand the asymptotic properties of MLE estimators on this manifold. Such a generalized asymptotic normality on a manifold has been established in previous work [36]. What remains is to relate the normality condition to infidelity, which we discuss by lifting the distance on a manifold to the state distance in the embedding Hilbert space via Lemma 4.

In the rest of this self-contained appendix, we provide the necessary background on MLE and differential geometry for readers interested in technical details and a proof of the theorem, provided in Appendix B.6. The first five sections review existing results, which we then use in our proof. First, we briefly review notations for MPS commonly used for the rest of the section in Appendix B.1. We provide a minimal list of technical definitions in differential geometry in Appendix B.2, which can be skipped for readers already familiar with the topic. Equipped with this knowledge of differential geometry, we review how the projective Hilbert space can be viewed as a manifold in Appendix B.3. Then, given a subspace of the projective Hilbert space defined by an MPS, we review properties of the projective MPS manifold in Appendix B.4. Finally, our main result relies on a theorem establishing asymptotic normality of the MLE estimator on a manifold, which we restate in Appendix B.5. We conclude this section with a summary of our main result as well as our interpretation of its relations to our numerical results in Appendix B.7.

B.1 Matrix product states

In this subsection, we introduce the notations and definitions used for matrix product states (MPS) throughout the rest of the subsection, following Ref. [37]. To represent a quantum state across n𝑛nitalic_n sites with physical dimension d𝑑ditalic_d, we define the MPS tensors over complex numbers as

𝔸χ=i=1nχi×d×χi+1l.subscript𝔸𝜒superscriptsubscriptdirect-sum𝑖1𝑛superscriptsubscript𝜒𝑖𝑑subscript𝜒𝑖1superscript𝑙\mathbb{A}_{\chi}=\bigoplus_{i=1}^{n}\mathbb{C}^{\chi_{i}\times d\times\chi_{i% +1}}\cong\mathbb{C}^{l}.blackboard_A start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT = ⨁ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT blackboard_C start_POSTSUPERSCRIPT italic_χ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT × italic_d × italic_χ start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT ≅ blackboard_C start_POSTSUPERSCRIPT italic_l end_POSTSUPERSCRIPT . (13)

Here, χ=(χ1,,χn+1)𝜒subscript𝜒1subscript𝜒𝑛1\chi=(\chi_{1},\ldots,\chi_{n+1})italic_χ = ( italic_χ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , italic_χ start_POSTSUBSCRIPT italic_n + 1 end_POSTSUBSCRIPT ) represents the bond dimensions and sums to a total tensor dimension of l=di=1nχiχi+1𝑙𝑑superscriptsubscript𝑖1𝑛subscript𝜒𝑖subscript𝜒𝑖1l=d\sum_{i=1}^{n}\chi_{i}\chi_{i+1}italic_l = italic_d ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT italic_χ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_χ start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT. Specifically, the tensor at the i𝑖iitalic_i-th site is denoted by A[i]superscript𝐴delimited-[]𝑖A^{[i]}italic_A start_POSTSUPERSCRIPT [ italic_i ] end_POSTSUPERSCRIPT. The tensor elements are Aα,β[i]δsuperscriptsubscript𝐴𝛼𝛽delimited-[]𝑖𝛿A_{\alpha,\beta}^{[i]\delta}italic_A start_POSTSUBSCRIPT italic_α , italic_β end_POSTSUBSCRIPT start_POSTSUPERSCRIPT [ italic_i ] italic_δ end_POSTSUPERSCRIPT, with 1αχi1𝛼subscript𝜒𝑖1\leq\alpha\leq\chi_{i}1 ≤ italic_α ≤ italic_χ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, 1βχi+11𝛽subscript𝜒𝑖11\leq\beta\leq\chi_{i+1}1 ≤ italic_β ≤ italic_χ start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT, and 1δd1𝛿𝑑1\leq\delta\leq d1 ≤ italic_δ ≤ italic_d. For this work, we limit the physical dimension to d=2𝑑2d=2italic_d = 2 (representing qubits) and consider open boundary conditions such that χ1=χn+1=1subscript𝜒1subscript𝜒𝑛11\chi_{1}=\chi_{n+1}=1italic_χ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_χ start_POSTSUBSCRIPT italic_n + 1 end_POSTSUBSCRIPT = 1. We define an MPS via a map from the space of MPS tensors 𝔸χsubscript𝔸𝜒\mathbb{A}_{\chi}blackboard_A start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT to the Hilbert space dnsuperscriptsuperscript𝑑𝑛\mathcal{H}\cong\mathbb{C}^{d^{n}}caligraphic_H ≅ blackboard_C start_POSTSUPERSCRIPT italic_d start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT of n𝑛nitalic_n sites.

ψχ:𝔸χ:subscript𝜓𝜒subscript𝔸𝜒\displaystyle\psi_{\chi}:\mathbb{A}_{\chi}italic_ψ start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT : blackboard_A start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT ,absent\displaystyle\longrightarrow\mathcal{H},⟶ caligraphic_H , (14)
A𝐴\displaystyle Aitalic_A b{0,1}nA[1]b1A[n]bn|b.absentsubscript𝑏superscript01𝑛superscript𝐴delimited-[]1subscript𝑏1superscript𝐴delimited-[]𝑛subscript𝑏𝑛ket𝑏\displaystyle\longmapsto\sum_{b\in{\{0,1\}}^{n}}A^{[1]b_{1}}\cdots A^{[n]b_{n}% }\mbox{$|b\rangle$}.⟼ ∑ start_POSTSUBSCRIPT italic_b ∈ { 0 , 1 } start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_POSTSUBSCRIPT italic_A start_POSTSUPERSCRIPT [ 1 ] italic_b start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT ⋯ italic_A start_POSTSUPERSCRIPT [ italic_n ] italic_b start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT end_POSTSUPERSCRIPT | italic_b ⟩ . (15)

The set {|b}ket𝑏\{\mbox{$|b\rangle$}\}{ | italic_b ⟩ } represents the computational basis, where the bit-string b{0,1}n𝑏superscript01𝑛b\in\{0,1\}^{n}italic_b ∈ { 0 , 1 } start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT, unless specified otherwise. Note that in the main text, we also use the notation |ψ(A)ket𝜓𝐴|\psi(A)\rangle| italic_ψ ( italic_A ) ⟩ to refer to the physical state, which is inferred from the context.

Physically distinct quantum states are identified up to a global phase and normalization factor, and are unique elements in the projective Hilbert space 𝖯()𝖯\mathsf{P}(\mathcal{H})sansserif_P ( caligraphic_H ). The elements in this space are equivalence classes, often called “rays” of vectors in \mathcal{H}caligraphic_H.

Definition 1 (Projective Hilbert space [37]).

Associated with a complex Hilbert space \mathcal{H}caligraphic_H, the projective Hilbert space is defined as

𝖯():=/GL(1,),assign𝖯GL1\mathsf{P}(\mathcal{H}):=\mathcal{H}/\operatorname{GL}(1,\mathbb{C}),sansserif_P ( caligraphic_H ) := caligraphic_H / roman_GL ( 1 , blackboard_C ) , (16)

which contains all rays represented by the equivalence classes [|ψ]delimited-[]ket𝜓[\mbox{$|\psi\rangle$}][ | italic_ψ ⟩ ]. Two nonzero vectors |ψ,|ϕket𝜓ketitalic-ϕ\mbox{$|\psi\rangle$},\mbox{$|\phi\rangle$}\in\mathcal{H}| italic_ψ ⟩ , | italic_ϕ ⟩ ∈ caligraphic_H are equivalent if and only if |ψ=λ|ϕket𝜓𝜆ketitalic-ϕ\mbox{$|\psi\rangle$}=\lambda\mbox{$|\phi\rangle$}| italic_ψ ⟩ = italic_λ | italic_ϕ ⟩ for some λGL(1,)𝜆GL1\lambda\in\operatorname{GL}(1,\mathbb{C})italic_λ ∈ roman_GL ( 1 , blackboard_C ), the group of nonzero complex numbers. For simplicity, we use the notation [ψ]delimited-[]𝜓[\psi][ italic_ψ ] to denote [|ψ]delimited-[]ket𝜓[\mbox{$|\psi\rangle$}][ | italic_ψ ⟩ ].

We can restrict MPS to the projective Hilbert space 𝖯()𝖯\mathsf{P}(\mathcal{H})sansserif_P ( caligraphic_H ) by defining a new map** ψ~χsubscript~𝜓𝜒\tilde{\psi}_{\chi}over~ start_ARG italic_ψ end_ARG start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT. This projection maps an MPS tensor onto a ray, effectively associating each MPS with a unique physical state representation. While it is possible to consider the manifold of MPS without this projection [37], using the projective MPS simplifies our MLE analysis by removing ambiguious normalization or phase factors that are not physically observable.

Definition 2 (Projective MPS [37] (Definition 13)).

The map ψ~χsubscript~𝜓𝜒\tilde{\psi}_{\chi}over~ start_ARG italic_ψ end_ARG start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT from the space of MPS tensors 𝔸χsubscript𝔸𝜒\mathbb{A}_{\chi}blackboard_A start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT to the projective Hilbert space 𝖯()𝖯\mathsf{P}(\mathcal{H})sansserif_P ( caligraphic_H ) is

ψ~χ:𝔸χ:subscript~𝜓𝜒subscript𝔸𝜒\displaystyle\tilde{\psi}_{\chi}:\mathbb{A}_{\chi}over~ start_ARG italic_ψ end_ARG start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT : blackboard_A start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT 𝖯(),absent𝖯\displaystyle\longrightarrow\mathsf{P}(\mathcal{H}),⟶ sansserif_P ( caligraphic_H ) , (17)
A𝐴\displaystyle Aitalic_A [ψχ(A)],absentdelimited-[]subscript𝜓𝜒𝐴\displaystyle\longmapsto[\psi_{\chi}(A)],⟼ [ italic_ψ start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT ( italic_A ) ] , (18)

where [ψχ(A)]delimited-[]subscript𝜓𝜒𝐴[\psi_{\chi}(A)][ italic_ψ start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT ( italic_A ) ] denotes the ray representing the equivalence class of the state vector ψχ(A)subscript𝜓𝜒𝐴\psi_{\chi}(A)italic_ψ start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT ( italic_A ) under rescaling by nonzero complex numbers.

Throughout the rest of this section, what we refer to by |ψket𝜓|\psi\rangle| italic_ψ ⟩ or [ψ]delimited-[]𝜓[\psi][ italic_ψ ] should be understood from the context as denoting either a normalized state vector or a ray in the projective Hilbert space, in distinction to the projective map** ψ~χsubscript~𝜓𝜒\tilde{\psi}_{\chi}over~ start_ARG italic_ψ end_ARG start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT. Note that the image ψ~χ(𝔸χ)subscript~𝜓𝜒subscript𝔸𝜒\tilde{\psi}_{\chi}(\mathbb{A}_{\chi})over~ start_ARG italic_ψ end_ARG start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT ( blackboard_A start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT ) represents only a subspace of 𝖯()𝖯\mathsf{P}(\mathcal{H})sansserif_P ( caligraphic_H ) and does not possess the structure of a vector space 222For instance, consider the sum of two MPS of bond dimension χ𝜒\chiitalic_χ as (|ψχ+|ϕχ)/2ketsubscript𝜓𝜒ketsubscriptitalic-ϕ𝜒2({\mbox{$|\psi_{\chi}\rangle$}+\mbox{$|\phi_{\chi}\rangle$}})/{\sqrt{2}}( | italic_ψ start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT ⟩ + | italic_ϕ start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT ⟩ ) / square-root start_ARG 2 end_ARG. This new state has bond dimension up to 2χ2𝜒2\chi2 italic_χ, which is beyond the original space defined for bond dimension χ𝜒\chiitalic_χ.. In the following sections, we will provide constraints on the domain of MPS tensors such that their image under ψ~χsubscript~𝜓𝜒\tilde{\psi}_{\chi}over~ start_ARG italic_ψ end_ARG start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT (and ψχsubscript𝜓𝜒\psi_{\chi}italic_ψ start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT) forms a differentiable manifold within 𝖯()𝖯\mathsf{P}(\mathcal{H})sansserif_P ( caligraphic_H ). Equip** the set of MPS with such an additional structure, we will be able to rigorously analyze the accuracy of the MLE solution on the space of a manifold, as detailed in Appendix B.5.

B.2 Differential geometry

This subsection introduces the essential concepts of differential geometry required to navigate through the subsequent discussions in Appendix B.3, B.4, and B.5. Assuming no prior expertise in differential geometry, we aim to outline here the key definitions and provide intuitive explanations for these concepts that can be found in standard textbooks [80]. While this section is in principle self-consistent, we highly encourage readers who would like to fully understand this material to consult Refs. [80, 81] for more context. Readers who are well familiar with differential geometry of Riemannian manifolds may choose to skip most of this section and only briefly review Def. 8 and Fact 1, which introduces terminologies used specifically in the context of our theorem.

Let us begin by briefly recalling the concept of an r𝑟ritalic_r-dimensional differentiable manifold M𝑀Mitalic_M. We use θM𝜃𝑀\theta\in Mitalic_θ ∈ italic_M to denote elements on the manifold, where, in the subsequent sections, θ𝜃\thetaitalic_θ parametrizes some underlying probability distribution. Formally, any local neighbourhood MiMsubscript𝑀𝑖𝑀M_{i}\subset Mitalic_M start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ⊂ italic_M can be related to Euclidean space rsuperscript𝑟\mathbb{R}^{r}blackboard_R start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT. This relationship is established through an injective map φi:Mir:subscript𝜑𝑖subscript𝑀𝑖superscript𝑟\varphi_{i}:M_{i}\rightarrow\mathbb{R}^{r}italic_φ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT : italic_M start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT → blackboard_R start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT, known as a chart. A chart provides a local coordinate system near each point in its neighbourhood. Intuitively, the manifold is covered by a collection of neighbourhoods Misubscript𝑀𝑖M_{i}italic_M start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT such that the union iIMi=Msubscript𝑖𝐼subscript𝑀𝑖𝑀\bigcup_{i\in I}M_{i}=M⋃ start_POSTSUBSCRIPT italic_i ∈ italic_I end_POSTSUBSCRIPT italic_M start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_M represents the entire manifold.

The collection of charts, known as an atlas, allows us to study functions defined on M𝑀Mitalic_M using the usual machinery from calculus defined on rsuperscript𝑟\mathbb{R}^{r}blackboard_R start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT. Moreover, for M𝑀Mitalic_M to be called differentiable, any composition φi1φjsuperscriptsubscript𝜑𝑖1subscript𝜑𝑗\varphi_{i}^{-1}\circ\varphi_{j}italic_φ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ∘ italic_φ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT defined on any two overlap** patches MiMjsubscript𝑀𝑖subscript𝑀𝑗M_{i}\cap M_{j}\neq\emptysetitalic_M start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ∩ italic_M start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ≠ ∅ must be differentiable. This allows us to define the tangent vectors X𝑋Xitalic_X at any point θM𝜃𝑀\theta\in Mitalic_θ ∈ italic_M. In alignment with Ref. [37], we adopt a geometric perspective of tangent vectors by representing them as column vectors (X1,,Xr)superscriptsubscript𝑋1subscript𝑋𝑟top(X_{1},\cdots,X_{r})^{\top}( italic_X start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , ⋯ , italic_X start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT. The components of the vector X𝑋Xitalic_X depend on the local coordinates in rsuperscript𝑟\mathbb{R}^{r}blackboard_R start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT, defined by the chosen chart φ𝜑\varphiitalic_φ.333Tangent vectors to M𝑀Mitalic_M at points θ𝜃\thetaitalic_θ can be naturally seen as directional derivatives. For a smooth function f:M:𝑓𝑀f:M\rightarrow\mathbb{R}italic_f : italic_M → blackboard_R defined on the manifold, if X𝑋Xitalic_X is a tangent vector to the manifold M𝑀Mitalic_M at θ𝜃\thetaitalic_θ, then the directional derivative (defined in Def. 6) is given by DX(f)=X(f)subscript𝐷𝑋𝑓𝑋𝑓D_{X}(f)=X(f)italic_D start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT ( italic_f ) = italic_X ( italic_f ). For what follows, it is convenient to directly refer to tangent vectors. The space of all tangent vectors at θ𝜃\thetaitalic_θ is called the tangent space TθMsubscriptT𝜃𝑀\mathrm{T}_{\theta}Mroman_T start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT italic_M and it is isomorphic to rsuperscript𝑟\mathbb{R}^{r}blackboard_R start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT. Consequently, each tangent vector can be expressed as X=a=1rXaea𝑋superscriptsubscript𝑎1𝑟subscript𝑋𝑎subscript𝑒𝑎X=\sum_{a=1}^{r}X_{a}e_{a}italic_X = ∑ start_POSTSUBSCRIPT italic_a = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT italic_X start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT where {ea}subscript𝑒𝑎\{e_{a}\}{ italic_e start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT } is a basis of TθMsubscriptT𝜃𝑀\mathrm{T}_{\theta}Mroman_T start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT italic_M.

To compute the lengths and angles of the tangent vectors, it is essential to define an inner product for each tangent space, called metric. In general, a metric provides an inner product for each tangent space TθMsubscriptT𝜃𝑀\mathrm{T}_{\theta}Mroman_T start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT italic_M which explicitly depends on the point θ𝜃\thetaitalic_θ, i.e., as one moves across the manifold, the inner product changes. When the inner product is real-valued and its change is gradual, such that it is differentiable (as precisely defined below), the metric is called a Riemannian metric. This allows us to track the change of lengths and angles between tangent vectors as one moves across the manifold so that we can define distances between points on the manifold.

Definition 3 (Riemannian metric [80]).

A Riemannian metric g𝑔gitalic_g is a smooth family of inner products {gθ|θM}conditional-setsubscript𝑔𝜃𝜃𝑀\{g_{\theta}|\theta\in M\}{ italic_g start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT | italic_θ ∈ italic_M } on the tangent spaces of M𝑀Mitalic_M. The metric gθsubscript𝑔𝜃g_{\theta}italic_g start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT at each point θ𝜃\thetaitalic_θ, defined by

gθ:TθM×TθM,:subscript𝑔𝜃subscriptT𝜃𝑀subscriptT𝜃𝑀g_{\theta}:\mathrm{T}_{\theta}M\times\mathrm{T}_{\theta}M\longrightarrow% \mathbb{R},italic_g start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT : roman_T start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT italic_M × roman_T start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT italic_M ⟶ blackboard_R , (19)

is bilinear and positive definite. Represented in the basis {ea}subscript𝑒𝑎\{e_{a}\}{ italic_e start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT }, its matrix elements (gθ)a,b=gθ(ea,eb)subscriptsubscript𝑔𝜃𝑎𝑏subscript𝑔𝜃subscript𝑒𝑎subscript𝑒𝑏(g_{\theta})_{a,b}=g_{\theta}(e_{a},e_{b})( italic_g start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT italic_a , italic_b end_POSTSUBSCRIPT = italic_g start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT ( italic_e start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT , italic_e start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ) vary smoothly with respect to θ𝜃\thetaitalic_θ.

In the discussion that follows, we will use g𝑔gitalic_g to denote the metric gθsubscript𝑔𝜃g_{\theta}italic_g start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT, as it is understood that the metric is defined for the tangent space at each point θM𝜃𝑀\theta\in Mitalic_θ ∈ italic_M. Additionally, we adopt the notation |g\mbox{$\langle\bullet|\bullet\rangle$}_{g}⟨ ∙ | ∙ ⟩ start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT to represent the inner product of any two tangent vectors, which is computed according to the metric g𝑔gitalic_g.

Definition 4 (Riemannian manifold).

A differentiable manifold M𝑀Mitalic_M together with a Riemannian metric g𝑔gitalic_g forms a Riemannian manifold (M,g)𝑀𝑔(M,g)( italic_M , italic_g ).

To be able to discuss curves connecting two points on a manifold (Def. 8), we introduce the concept of a Riemannian connection, which generalizes the notion of the covariant derivative.

Definition 5 (Riemannian connection [80]).

Let Γ(TM)ΓT𝑀\Gamma(\mathrm{T}M)roman_Γ ( roman_T italic_M ) be the set of differentiable vector fields θXTθMmaps-to𝜃𝑋subscriptT𝜃𝑀\theta\mapsto X\in\mathrm{T}_{\theta}Mitalic_θ ↦ italic_X ∈ roman_T start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT italic_M on a Riemannian manifold (M,g)𝑀𝑔(M,g)( italic_M , italic_g ). A map :Γ(TM)×Γ(TM)Γ(TM):ΓT𝑀ΓT𝑀ΓT𝑀\nabla:\Gamma(\mathrm{T}M)\times\Gamma(\mathrm{T}M)\rightarrow\Gamma(\mathrm{T% }M)∇ : roman_Γ ( roman_T italic_M ) × roman_Γ ( roman_T italic_M ) → roman_Γ ( roman_T italic_M ) is called a Riemannian connection, typically expressed as XYsubscript𝑋𝑌\nabla_{X}Y∇ start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT italic_Y representing (X,Y)𝑋𝑌\nabla(X,Y)∇ ( italic_X , italic_Y ). It holds that

  1. 1.

    fX+hYZ=fXZ+hYZsubscript𝑓𝑋𝑌𝑍𝑓subscript𝑋𝑍subscript𝑌𝑍\nabla_{fX+hY}Z=f\nabla_{X}Z+h\nabla_{Y}Z∇ start_POSTSUBSCRIPT italic_f italic_X + italic_h italic_Y end_POSTSUBSCRIPT italic_Z = italic_f ∇ start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT italic_Z + italic_h ∇ start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT italic_Z,

  2. 2.

    X(Y+Z)=XY+XZsubscript𝑋𝑌𝑍subscript𝑋𝑌subscript𝑋𝑍\nabla_{X}(Y+Z)=\nabla_{X}Y+\nabla_{X}Z∇ start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT ( italic_Y + italic_Z ) = ∇ start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT italic_Y + ∇ start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT italic_Z,

  3. 3.

    XfZ=fXZ+(DXf)Zsubscript𝑋𝑓𝑍𝑓subscript𝑋𝑍subscriptD𝑋𝑓𝑍\nabla_{X}fZ=f\nabla_{X}Z+(\mathrm{D}_{X}f)Z∇ start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT italic_f italic_Z = italic_f ∇ start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT italic_Z + ( roman_D start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT italic_f ) italic_Z,

  4. 4.

    DXg(Y,Z)=g(XY,Z)+g(Y,XZ)subscriptD𝑋𝑔𝑌𝑍𝑔subscript𝑋𝑌𝑍𝑔𝑌subscript𝑋𝑍\mathrm{D}_{X}g(Y,Z)=g(\nabla_{X}Y,Z)+g(Y,\nabla_{X}Z)roman_D start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT italic_g ( italic_Y , italic_Z ) = italic_g ( ∇ start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT italic_Y , italic_Z ) + italic_g ( italic_Y , ∇ start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT italic_Z ),

  5. 5.

    XYYX[X,Y]=0subscript𝑋𝑌subscript𝑌𝑋𝑋𝑌0\nabla_{X}Y-\nabla_{Y}X-[X,Y]=0∇ start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT italic_Y - ∇ start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT italic_X - [ italic_X , italic_Y ] = 0,

for all X,Y,ZΓ(TM)𝑋𝑌𝑍ΓT𝑀X,Y,Z\in\Gamma(\mathrm{T}M)italic_X , italic_Y , italic_Z ∈ roman_Γ ( roman_T italic_M ) and all smooth functions f,h:M:𝑓𝑀f,h:M\rightarrow\mathbb{R}italic_f , italic_h : italic_M → blackboard_R defined on the manifold. [,][\bullet,\bullet][ ∙ , ∙ ] is the Lie bracket defined by [X,Y](f):=DXDYfDYDXfassign𝑋𝑌𝑓subscriptD𝑋subscriptD𝑌𝑓subscriptD𝑌subscriptD𝑋𝑓[X,Y](f):=\mathrm{D}_{X}\mathrm{D}_{Y}f-\mathrm{D}_{Y}\mathrm{D}_{X}f[ italic_X , italic_Y ] ( italic_f ) := roman_D start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT roman_D start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT italic_f - roman_D start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT roman_D start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT italic_f.

Here we use DXfsubscriptD𝑋𝑓\mathrm{D}_{X}froman_D start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT italic_f to denote the directional derivative of a scalar-valued function f𝑓fitalic_f, also commonly denoted by X(f)𝑋𝑓X(f)italic_X ( italic_f ) in the literature. The directional derivative has a direct correspondence with the tangent vectors, and can be formally defined as follows:

Definition 6 (Directional derivative).

Let M𝑀Mitalic_M be a differentiable manifold, θM𝜃𝑀\theta\in Mitalic_θ ∈ italic_M and XTθM𝑋subscriptT𝜃𝑀X\in\mathrm{T}_{\theta}Mitalic_X ∈ roman_T start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT italic_M. Given a differentiable function f:M:𝑓𝑀f:M\rightarrow\mathbb{R}italic_f : italic_M → blackboard_R, the directional derivative DXfsubscriptD𝑋𝑓\mathrm{D}_{X}froman_D start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT italic_f (also denoted X(f)𝑋𝑓X(f)italic_X ( italic_f )) at θ𝜃\thetaitalic_θ is defined as

DXf|θ=dmissingdtf(γ(t))|t=0,evaluated-atsubscriptD𝑋𝑓𝜃evaluated-atdmissingd𝑡𝑓𝛾𝑡𝑡0\mathrm{D}_{X}f\big{|}_{\theta}=\frac{\mathrm{d}missing}{\mathrm{d}t}f(\gamma(% t))\Big{|}_{t=0},roman_D start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT italic_f | start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT = divide start_ARG roman_dmissing end_ARG start_ARG roman_d italic_t end_ARG italic_f ( italic_γ ( italic_t ) ) | start_POSTSUBSCRIPT italic_t = 0 end_POSTSUBSCRIPT , (20)

where γ:M:𝛾𝑀\gamma:\mathbb{R}\rightarrow Mitalic_γ : blackboard_R → italic_M is a differentiable curve such that γ(0)=θ𝛾0𝜃\gamma(0)=\thetaitalic_γ ( 0 ) = italic_θ and the velocity γ˙(0)=X˙𝛾0𝑋\dot{\gamma}(0)=Xover˙ start_ARG italic_γ end_ARG ( 0 ) = italic_X.

The Riemannian connection is sometimes also referred to as the Levi-Civita connection and for each Riemannian manifold, there exists exactly one Riemannian connection [80]. As mentioned above and as the notation suggests, the Riemannian connection provides us with a notion of directional (or covariant, to be more precise) derivatives on a manifold. This allows us to describe how vectors change as they are moved along a curve on the manifold. Specifically, we can now define a notion of parallelism along curves, a concept that is very intuitive in Euclidean space but much less so in curved spaces.

Definition 7 (Parallel transport).

Let α:[0,1]M:𝛼01𝑀\alpha:[0,1]\rightarrow Mitalic_α : [ 0 , 1 ] → italic_M be a smooth curve on the Riemannian manifold (M,g)𝑀𝑔(M,g)( italic_M , italic_g ) and α˙(t)˙𝛼𝑡\dot{\alpha}(t)over˙ start_ARG italic_α end_ARG ( italic_t ) be its velocity vector in the tangent space Tα(t)MsubscriptT𝛼𝑡𝑀\mathrm{T}_{\alpha(t)}Mroman_T start_POSTSUBSCRIPT italic_α ( italic_t ) end_POSTSUBSCRIPT italic_M. The parallel transport of a tangent vector X0Tα(0)Msubscript𝑋0subscriptT𝛼0𝑀X_{0}\in\mathrm{T}_{\alpha(0)}Mitalic_X start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∈ roman_T start_POSTSUBSCRIPT italic_α ( 0 ) end_POSTSUBSCRIPT italic_M along the curve α𝛼\alphaitalic_α under a Riemannian connection \nabla is defined by the equation

α˙(t)X(t)=0,subscript˙𝛼𝑡𝑋𝑡0\nabla_{\dot{\alpha}(t)}X(t)=0,∇ start_POSTSUBSCRIPT over˙ start_ARG italic_α end_ARG ( italic_t ) end_POSTSUBSCRIPT italic_X ( italic_t ) = 0 , (21)

where X(t)Tα(t)M𝑋𝑡subscriptT𝛼𝑡𝑀X(t)\in\mathrm{T}_{\alpha(t)}Mitalic_X ( italic_t ) ∈ roman_T start_POSTSUBSCRIPT italic_α ( italic_t ) end_POSTSUBSCRIPT italic_M for all t[0,1]𝑡01t\in[0,1]italic_t ∈ [ 0 , 1 ] and X(0)=X0𝑋0subscript𝑋0X(0)=X_{0}italic_X ( 0 ) = italic_X start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. The vector field X𝑋Xitalic_X is said to be parallel along α𝛼\alphaitalic_α if it satisfies the above condition.

A special case of parallel vector fields is given by geodesic curves. The velocity vector γ˙˙𝛾\dot{\gamma}over˙ start_ARG italic_γ end_ARG of a geodesic curve γ:IM:𝛾𝐼𝑀\gamma:I\rightarrow Mitalic_γ : italic_I → italic_M (with I𝐼I\subseteq\mathbb{R}italic_I ⊆ blackboard_R some parameter interval) is parallel along γ𝛾\gammaitalic_γ itself. For simplicity, we will take the unit interval I=[0,1]𝐼01I=[0,1]italic_I = [ 0 , 1 ] because a curve with any other interval can always be reparameterized. The notion of a geodesic extends the concept of a straight line from Euclidean space to curved spaces like manifolds. In general, there can be multiple geodesic curves connecting two points on a manifold444For example, the great circle on the sphere S2superscript𝑆2S^{2}italic_S start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT is a geodesic curve. Two points on the great circle separate the circle into two arcs. The shorter arc is the shortest connecting geodesic, and the longer arc is the longer connecting geodesic.. For our purpose, we only consider the curves that represent the shortest path between two points on the manifold, parametrized by arc length on the unit interval.

Definition 8 (Shortest connecting geodesic).

Let (M,g)𝑀𝑔(M,g)( italic_M , italic_g ) be a Riemannian manifold and :Γ(TM)×Γ(TM)Γ(TM):ΓT𝑀ΓT𝑀ΓT𝑀\nabla:\Gamma(\mathrm{T}M)\times\Gamma(\mathrm{T}M)\rightarrow\Gamma(\mathrm{T% }M)∇ : roman_Γ ( roman_T italic_M ) × roman_Γ ( roman_T italic_M ) → roman_Γ ( roman_T italic_M ) the associated Riemannian connection. Given two points θ1,θ2Wsubscript𝜃1subscript𝜃2𝑊\theta_{1},\theta_{2}\in Witalic_θ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_θ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ∈ italic_W where WM𝑊𝑀W\subset Mitalic_W ⊂ italic_M is a sufficiently small neighborhood, the curve

γ:[0,1]:𝛾01\displaystyle\gamma:[0,1]italic_γ : [ 0 , 1 ] W,absent𝑊\displaystyle\longrightarrow W,⟶ italic_W , (22)
t𝑡\displaystyle titalic_t γ(t),absent𝛾𝑡\displaystyle\longmapsto\gamma(t),⟼ italic_γ ( italic_t ) , (23)

is the shortest connecting geodesic of θ1subscript𝜃1\theta_{1}italic_θ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and θ2subscript𝜃2\theta_{2}italic_θ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT if γ(0)=θ1𝛾0subscript𝜃1\gamma(0)=\theta_{1}italic_γ ( 0 ) = italic_θ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and γ(1)=θ2𝛾1subscript𝜃2\gamma(1)=\theta_{2}italic_γ ( 1 ) = italic_θ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, and for all t[0,1]𝑡01t\in[0,1]italic_t ∈ [ 0 , 1 ] the tangent vectors γ˙˙𝛾\dot{\gamma}over˙ start_ARG italic_γ end_ARG satisfy

γ˙γ˙=0.subscript˙𝛾˙𝛾0\nabla_{\dot{\gamma}}\dot{\gamma}=0.∇ start_POSTSUBSCRIPT over˙ start_ARG italic_γ end_ARG end_POSTSUBSCRIPT over˙ start_ARG italic_γ end_ARG = 0 . (24)

This implies that γ˙˙𝛾\dot{\gamma}over˙ start_ARG italic_γ end_ARG is parallel transported along the geodesic curve γ𝛾\gammaitalic_γ.

Note, that in general, one can loosen the condition for geodesics to γ˙γ˙=λ(t)γ˙subscript˙𝛾˙𝛾𝜆𝑡˙𝛾\nabla_{\dot{\gamma}}\dot{\gamma}=\lambda(t)\dot{\gamma}∇ start_POSTSUBSCRIPT over˙ start_ARG italic_γ end_ARG end_POSTSUBSCRIPT over˙ start_ARG italic_γ end_ARG = italic_λ ( italic_t ) over˙ start_ARG italic_γ end_ARG where λ𝜆\lambdaitalic_λ is some scalar function and γ𝛾\gammaitalic_γ is required to be regular (γ˙(t)0˙𝛾𝑡0\dot{\gamma}(t)\neq 0over˙ start_ARG italic_γ end_ARG ( italic_t ) ≠ 0 for all t𝑡titalic_t). However, by setting γ˙γ˙=0subscript˙𝛾˙𝛾0\nabla_{\dot{\gamma}}\dot{\gamma}=0∇ start_POSTSUBSCRIPT over˙ start_ARG italic_γ end_ARG end_POSTSUBSCRIPT over˙ start_ARG italic_γ end_ARG = 0, we enforce that the velocity vector γ˙˙𝛾\dot{\gamma}over˙ start_ARG italic_γ end_ARG of the geodesic cannot change in the direction of the curve, so that the velocity vector has a constant length throughout the curve. This is formally stated in Fact 1 below.

Fact 1 (Velocity vector).

Consider a geodesic γ𝛾\gammaitalic_γ (defined in Def. 8) on a Riemannian manifold (M,g)𝑀𝑔(M,g)( italic_M , italic_g ) of dimension r𝑟ritalic_r. If {ea(t)}a=1rsuperscriptsubscriptsubscript𝑒𝑎𝑡𝑎1𝑟\{e_{a}(t)\}_{a=1}^{r}{ italic_e start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ( italic_t ) } start_POSTSUBSCRIPT italic_a = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT represents the set of basis vectors parallel transported along γ𝛾\gammaitalic_γ, the velocity vector γ˙˙𝛾\dot{\gamma}over˙ start_ARG italic_γ end_ARG can be expressed as

γ˙=a=1rΔaea(t),˙𝛾superscriptsubscript𝑎1𝑟subscriptΔ𝑎subscript𝑒𝑎𝑡\dot{\gamma}=\sum_{a=1}^{r}\Delta_{a}e_{a}(t),over˙ start_ARG italic_γ end_ARG = ∑ start_POSTSUBSCRIPT italic_a = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT roman_Δ start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ( italic_t ) , (25)

where the decomposition coefficients {Δa}a=1rsuperscriptsubscriptsubscriptΔ𝑎𝑎1𝑟\{\Delta_{a}\}_{a=1}^{r}{ roman_Δ start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT } start_POSTSUBSCRIPT italic_a = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT are constant due to the constraint γ˙γ˙=0subscript˙𝛾˙𝛾0\nabla_{\dot{\gamma}}\dot{\gamma}=0∇ start_POSTSUBSCRIPT over˙ start_ARG italic_γ end_ARG end_POSTSUBSCRIPT over˙ start_ARG italic_γ end_ARG = 0. Moreover, the length of the velocity vector is given by

γ˙(t)norm˙𝛾𝑡\displaystyle\|\dot{\gamma}(t)\|∥ over˙ start_ARG italic_γ end_ARG ( italic_t ) ∥ :=γ˙|γ˙gassignabsentsubscriptinner-product˙𝛾˙𝛾𝑔\displaystyle:=\sqrt{\mbox{$\langle\dot{\gamma}|\dot{\gamma}\rangle$}_{g}}:= square-root start_ARG ⟨ over˙ start_ARG italic_γ end_ARG | over˙ start_ARG italic_γ end_ARG ⟩ start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT end_ARG (26)
=a,bΔaga,bΔb,absentsubscript𝑎𝑏subscriptΔ𝑎subscript𝑔𝑎𝑏subscriptΔ𝑏\displaystyle=\sqrt{\sum_{a,b}\Delta_{a}\,g_{a,b}\,\Delta_{b}},= square-root start_ARG ∑ start_POSTSUBSCRIPT italic_a , italic_b end_POSTSUBSCRIPT roman_Δ start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT italic_g start_POSTSUBSCRIPT italic_a , italic_b end_POSTSUBSCRIPT roman_Δ start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT end_ARG ,

which is a constant for all t[0,1]𝑡01t\in[0,1]italic_t ∈ [ 0 , 1 ], due to the aforementioned constraint γ˙γ˙=0subscript˙𝛾˙𝛾0\nabla_{\dot{\gamma}}\dot{\gamma}=0∇ start_POSTSUBSCRIPT over˙ start_ARG italic_γ end_ARG end_POSTSUBSCRIPT over˙ start_ARG italic_γ end_ARG = 0 representing the constant speed at which γ𝛾\gammaitalic_γ traverses the manifold. The metric tensor components ga,b(t)=ea(t)|eb(t)gsubscript𝑔𝑎𝑏𝑡subscriptinner-productsubscript𝑒𝑎𝑡subscript𝑒𝑏𝑡𝑔g_{a,b}(t)=\mbox{$\langle e_{a}(t)|e_{b}(t)\rangle$}_{g}italic_g start_POSTSUBSCRIPT italic_a , italic_b end_POSTSUBSCRIPT ( italic_t ) = ⟨ italic_e start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ( italic_t ) | italic_e start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ( italic_t ) ⟩ start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT in the basis {ea(t)}a=1rsuperscriptsubscriptsubscript𝑒𝑎𝑡𝑎1𝑟\{e_{a}(t)\}_{a=1}^{r}{ italic_e start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ( italic_t ) } start_POSTSUBSCRIPT italic_a = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT along the curve are constant as well.

Proof.

We begin by expanding the condition for the geodesics for a given coordinate system

00\displaystyle 0 =γ˙γ˙absentsubscript˙𝛾˙𝛾\displaystyle=\nabla_{\dot{\gamma}}\dot{\gamma}= ∇ start_POSTSUBSCRIPT over˙ start_ARG italic_γ end_ARG end_POSTSUBSCRIPT over˙ start_ARG italic_γ end_ARG (27)
=aγ˙Δaeaabsentsubscript𝑎subscript˙𝛾subscriptΔ𝑎subscript𝑒𝑎\displaystyle=\sum_{a}\nabla_{\dot{\gamma}}\Delta_{a}e_{a}= ∑ start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ∇ start_POSTSUBSCRIPT over˙ start_ARG italic_γ end_ARG end_POSTSUBSCRIPT roman_Δ start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT
=aΔaγ˙ea+a(Dγ˙Δa)ea.absentsubscript𝑎subscriptΔ𝑎subscript˙𝛾subscript𝑒𝑎subscript𝑎subscriptD˙𝛾subscriptΔ𝑎subscript𝑒𝑎\displaystyle=\sum_{a}\Delta_{a}\nabla_{\dot{\gamma}}e_{a}+\sum_{a}(\mathrm{D}% _{\dot{\gamma}}\Delta_{a})e_{a}.= ∑ start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT roman_Δ start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ∇ start_POSTSUBSCRIPT over˙ start_ARG italic_γ end_ARG end_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT + ∑ start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ( roman_D start_POSTSUBSCRIPT over˙ start_ARG italic_γ end_ARG end_POSTSUBSCRIPT roman_Δ start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ) italic_e start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT .

Using the definition of the parallel transport of the basis vectors, γ˙ea(t)=0subscript˙𝛾subscript𝑒𝑎𝑡0\nabla_{\dot{\gamma}}e_{a}(t)=0∇ start_POSTSUBSCRIPT over˙ start_ARG italic_γ end_ARG end_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ( italic_t ) = 0, we conclude that a(Dγ˙Δi)ea=0subscript𝑎subscriptD˙𝛾subscriptΔ𝑖subscript𝑒𝑎0\sum_{a}(\mathrm{D}_{\dot{\gamma}}\Delta_{i})e_{a}=0∑ start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ( roman_D start_POSTSUBSCRIPT over˙ start_ARG italic_γ end_ARG end_POSTSUBSCRIPT roman_Δ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) italic_e start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT = 0. Since the basis vectors {ea}subscript𝑒𝑎\{e_{a}\}{ italic_e start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT } are linearly independent, the coefficients Dγ˙ΔasubscriptD˙𝛾subscriptΔ𝑎\mathrm{D}_{\dot{\gamma}}\Delta_{a}roman_D start_POSTSUBSCRIPT over˙ start_ARG italic_γ end_ARG end_POSTSUBSCRIPT roman_Δ start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT must all be zero individually, hence ΔasubscriptΔ𝑎\Delta_{a}roman_Δ start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT is a constant in t𝑡titalic_t for all a𝑎aitalic_a. Moreover, any Riemannian connection preserves the metric (Property 4 in Def. 5) and thus, we have

Dγ˙ga,b(t)=γ˙ea(t)|eb(t)+ea(t)|γ˙eb(t),subscriptD˙𝛾subscript𝑔𝑎𝑏𝑡inner-productsubscript˙𝛾subscript𝑒𝑎𝑡subscript𝑒𝑏𝑡inner-productsubscript𝑒𝑎𝑡subscript˙𝛾subscript𝑒𝑏𝑡\mathrm{D}_{\dot{\gamma}}g_{a,b}(t)=\mbox{$\langle\nabla_{\dot{\gamma}}e_{a}(t% )|e_{b}(t)\rangle$}+\mbox{$\langle e_{a}(t)|\nabla_{\dot{\gamma}}e_{b}(t)% \rangle$},roman_D start_POSTSUBSCRIPT over˙ start_ARG italic_γ end_ARG end_POSTSUBSCRIPT italic_g start_POSTSUBSCRIPT italic_a , italic_b end_POSTSUBSCRIPT ( italic_t ) = ⟨ ∇ start_POSTSUBSCRIPT over˙ start_ARG italic_γ end_ARG end_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ( italic_t ) | italic_e start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ( italic_t ) ⟩ + ⟨ italic_e start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ( italic_t ) | ∇ start_POSTSUBSCRIPT over˙ start_ARG italic_γ end_ARG end_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ( italic_t ) ⟩ , (28)

which is zero since γ˙ea(t)=0subscript˙𝛾subscript𝑒𝑎𝑡0\nabla_{\dot{\gamma}}e_{a}(t)=0∇ start_POSTSUBSCRIPT over˙ start_ARG italic_γ end_ARG end_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ( italic_t ) = 0 for all a=1,,r𝑎1𝑟a=1,\ldots,ritalic_a = 1 , … , italic_r, by definition. ∎

Given a connecting geodesic curve on a Riemannian manifold, the length of the curve can serve as a distance measure between the two end points θ1subscript𝜃1\theta_{1}italic_θ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and θ2subscript𝜃2\theta_{2}italic_θ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT. The length of the curve depends on the metric along the path. Let γ𝛾\gammaitalic_γ be a connecting geodesic as in Def. 8. The arc length Lγ(0,1)subscript𝐿𝛾01L_{\gamma}(0,1)italic_L start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT ( 0 , 1 ) of any segment of of the curve between γ(0)𝛾0\gamma(0)italic_γ ( 0 ) and γ(1)𝛾1\gamma(1)italic_γ ( 1 ) is given by

Lγ(0,1)subscript𝐿𝛾01\displaystyle L_{\gamma}(0,1)italic_L start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT ( 0 , 1 ) =01dtγ˙|γ˙gabsentsuperscriptsubscript01differential-d𝑡subscriptinner-product˙𝛾˙𝛾𝑔\displaystyle=\int_{0}^{1}\mathrm{d}t\sqrt{\mbox{$\langle\dot{\gamma}|\dot{% \gamma}\rangle$}_{g}}= ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT roman_d italic_t square-root start_ARG ⟨ over˙ start_ARG italic_γ end_ARG | over˙ start_ARG italic_γ end_ARG ⟩ start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT end_ARG (29)
=γ˙|γ˙g,absentsubscriptinner-product˙𝛾˙𝛾𝑔\displaystyle=\sqrt{\mbox{$\langle\dot{\gamma}|\dot{\gamma}\rangle$}_{g}},= square-root start_ARG ⟨ over˙ start_ARG italic_γ end_ARG | over˙ start_ARG italic_γ end_ARG ⟩ start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT end_ARG ,

which is the constant length of the velocity vector along γ𝛾\gammaitalic_γ.

B.3 Projective Hilbert space as a manifold

In this subsection, we apply the previously discussed concepts of differential geometry to describe quantum systems. In particular, we review how the projective Hilbert space 𝖯()𝖯\mathsf{P}(\mathcal{H})sansserif_P ( caligraphic_H ) can be understood as a manifold [37]. These descriptions allow us to relate the length of geodesics within this manifold to the quantum state overlap—the fidelity—that we aim to analyze.

Definition 9 (Projective Hilbert space as a manifold [37]).

The projective Hilbert space 𝖯()2n1similar-to-or-equals𝖯superscriptsuperscript2𝑛1\mathsf{P}(\mathcal{H})\simeq\mathbb{C}\mathbb{P}^{2^{n}-1}sansserif_P ( caligraphic_H ) ≃ blackboard_C blackboard_P start_POSTSUPERSCRIPT 2 start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT is a differentiable manifold, whose elements correspond to vectors in the Hilbert space modulo a global phase and normalization.

Note that 𝖯()𝖯\mathsf{P}(\mathcal{H})sansserif_P ( caligraphic_H ) can be naturally seen as a complex manifold of complex dimension 2n1superscript2𝑛12^{n}-12 start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT - 1, although it could, at the same time, also be defined as a real manifold of dimension 2(2n1)2superscript2𝑛12(2^{n}-1)2 ( 2 start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT - 1 ). Following Ref. [81] (Sec. 8.2) and Ref. [37] (Sec. IIA), we can take the tangent space of the complex manifold 𝖯()𝖯\mathsf{P}(\mathcal{H})sansserif_P ( caligraphic_H ) at some point [ψ]𝖯()delimited-[]𝜓𝖯[\psi]\in\mathsf{P}(\mathcal{H})[ italic_ψ ] ∈ sansserif_P ( caligraphic_H ) as the real vector space T[ψ]𝖯()subscriptTdelimited-[]𝜓𝖯\mathrm{T}_{[\psi]}\mathsf{P}(\mathcal{H})roman_T start_POSTSUBSCRIPT [ italic_ψ ] end_POSTSUBSCRIPT sansserif_P ( caligraphic_H ) of dimension 2(2n1)2superscript2𝑛12(2^{n}-1)2 ( 2 start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT - 1 ). To avoid confusion, we remark that it can be useful to consider the complexification T[ψ]𝖯()subscriptsuperscriptTdelimited-[]𝜓𝖯\mathrm{T}^{\mathbb{C}}_{[\psi]}\mathsf{P}(\mathcal{H})roman_T start_POSTSUPERSCRIPT blackboard_C end_POSTSUPERSCRIPT start_POSTSUBSCRIPT [ italic_ψ ] end_POSTSUBSCRIPT sansserif_P ( caligraphic_H ) (i.e., the complex span of any basis of T[ψ]𝖯()subscriptTdelimited-[]𝜓𝖯\mathrm{T}_{[\psi]}\mathsf{P}(\mathcal{H})roman_T start_POSTSUBSCRIPT [ italic_ψ ] end_POSTSUBSCRIPT sansserif_P ( caligraphic_H )) of this vector space to define a set of so-called holomorphic tangent vectors. Note, however, that T[ψ]𝖯()subscriptsuperscriptTdelimited-[]𝜓𝖯\mathrm{T}^{\mathbb{C}}_{[\psi]}\mathsf{P}(\mathcal{H})roman_T start_POSTSUPERSCRIPT blackboard_C end_POSTSUPERSCRIPT start_POSTSUBSCRIPT [ italic_ψ ] end_POSTSUBSCRIPT sansserif_P ( caligraphic_H ) has complex dimension 2(2n1)2superscript2𝑛12(2^{n}-1)2 ( 2 start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT - 1 ) and is therefore over-parametrized.

For our result, we only rely on the manifold 𝖯()𝖯\mathsf{P}(\mathcal{H})sansserif_P ( caligraphic_H ) having the structure of a Riemannian manifold and therefore, we will only work with T[ψ]𝖯()subscriptTdelimited-[]𝜓𝖯\mathrm{T}_{[\psi]}\mathsf{P}(\mathcal{H})roman_T start_POSTSUBSCRIPT [ italic_ψ ] end_POSTSUBSCRIPT sansserif_P ( caligraphic_H ). However, as we will see, T[ψ]𝖯()subscriptTdelimited-[]𝜓𝖯\mathrm{T}_{[\psi]}\mathsf{P}(\mathcal{H})roman_T start_POSTSUBSCRIPT [ italic_ψ ] end_POSTSUBSCRIPT sansserif_P ( caligraphic_H ) is closely related to the Hilbert space \mathcal{H}caligraphic_H itself. Hence, it is more natural to view the 2(2n1)2superscript2𝑛12(2^{n}-1)2 ( 2 start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT - 1 )-dimensional real tangent vectors as 2n1superscript2𝑛12^{n}-12 start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT - 1-dimensional complex vectors. Those correspond to vectors in \mathcal{H}caligraphic_H that are orthogonal to [ψ]delimited-[]𝜓[\psi][ italic_ψ ]. We stress that these vectors are not elements of the complexified tangent space (which would be complex vectors of dimension 2(2n1)2superscript2𝑛12(2^{n}-1)2 ( 2 start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT - 1 )).

Definition 10 (Tangent space of projective Hilbert space [37]).

The tangent space T[ψ]𝖯()=/\mathrm{T}_{[\psi]}\mathsf{P}(\mathcal{H})=\mathcal{H}/\!\simroman_T start_POSTSUBSCRIPT [ italic_ψ ] end_POSTSUBSCRIPT sansserif_P ( caligraphic_H ) = caligraphic_H / ∼ at a base point denoted by [ψ]delimited-[]𝜓[\psi][ italic_ψ ] is the space of all tangent vectors that satisfy the following equivalence relation. Let |ψ1,|ψ2ketsubscript𝜓1ketsubscript𝜓2\mbox{$|\psi_{1}\rangle$},\mbox{$|\psi_{2}\rangle$}\in\mathcal{H}| italic_ψ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ⟩ , | italic_ψ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⟩ ∈ caligraphic_H be two state vectors in the Hilbert space,

|ψ1|ψ2:|ψ1|ψ2=α|ψ,\mbox{$|\psi_{1}\rangle$}\sim\mbox{$|\psi_{2}\rangle$}:\!\iff\mbox{$|\psi_{1}% \rangle$}-\mbox{$|\psi_{2}\rangle$}=\alpha\mbox{$|\psi\rangle$},| italic_ψ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ⟩ ∼ | italic_ψ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⟩ : ⇔ | italic_ψ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ⟩ - | italic_ψ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⟩ = italic_α | italic_ψ ⟩ , (30)

for some constant α𝛼\alpha\in\mathbb{C}italic_α ∈ blackboard_C.

The equivalence relation removes one dimension from the Hilbert space so that the tangent space is isomorphic to the Euclidean complex space and the Eucliean real space of twice the dimension, i.e., T[ψ]𝖯()2n12(2n1)similar-to-or-equalssubscriptTdelimited-[]𝜓𝖯superscriptsuperscript2𝑛1similar-to-or-equalssuperscript2superscript2𝑛1\mathrm{T}_{[\psi]}\mathsf{P}(\mathcal{H})\simeq\mathbb{C}^{2^{n}-1}\simeq% \mathbb{R}^{2(2^{n}-1)}roman_T start_POSTSUBSCRIPT [ italic_ψ ] end_POSTSUBSCRIPT sansserif_P ( caligraphic_H ) ≃ blackboard_C start_POSTSUPERSCRIPT 2 start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ≃ blackboard_R start_POSTSUPERSCRIPT 2 ( 2 start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT - 1 ) end_POSTSUPERSCRIPT.

For the purpose of our result, we only rely on the minimal structure of a Riemannian manifold. 𝖯()𝖯\mathsf{P}(\mathcal{H})sansserif_P ( caligraphic_H ) can be equipped with the Fubini-Study metric in the context of quantum geometry. Such a choice of the Fubini-Study metric is natural because it represents the inner product between two state vectors in the Hilbert space.

Definition 11 (Fubini-Study metric [82]).

Let |ψ1ketsubscript𝜓1|\psi_{1}\rangle| italic_ψ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ⟩ and |ψ2ketsubscript𝜓2|\psi_{2}\rangle| italic_ψ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⟩ be two tangent vectors in T[ψ]𝖯()subscriptTdelimited-[]𝜓𝖯\mathrm{T}_{[\psi]}\mathsf{P}(\mathcal{H})roman_T start_POSTSUBSCRIPT [ italic_ψ ] end_POSTSUBSCRIPT sansserif_P ( caligraphic_H ) and |ψket𝜓|\psi\rangle| italic_ψ ⟩ be the normalized representation of [ψ]delimited-[]𝜓[\psi][ italic_ψ ]. The Fubini-Study metric is defined as

g:T[ψ]𝖯()×T[ψ]𝖯():𝑔subscriptTdelimited-[]𝜓𝖯subscriptTdelimited-[]𝜓𝖯\displaystyle g:\mathrm{T}_{[\psi]}\mathsf{P}(\mathcal{H})\times\mathrm{T}_{[% \psi]}\mathsf{P}(\mathcal{H})italic_g : roman_T start_POSTSUBSCRIPT [ italic_ψ ] end_POSTSUBSCRIPT sansserif_P ( caligraphic_H ) × roman_T start_POSTSUBSCRIPT [ italic_ψ ] end_POSTSUBSCRIPT sansserif_P ( caligraphic_H ) ,absent\displaystyle\longrightarrow\mathbb{R},⟶ blackboard_R , (31)
(|ψ1,|ψ2)ketsubscript𝜓1ketsubscript𝜓2\displaystyle(\mbox{$|{\psi_{1}}\rangle$},\mbox{$|{\psi_{2}}\rangle$})( | italic_ψ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ⟩ , | italic_ψ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⟩ ) Re[ψ1|ψ2ψ1|ψψ|ψ2].\displaystyle\longmapsto\mathop{\mathrm{Re}}\big{[}\mbox{$\langle\psi_{1}|\psi% _{2}\rangle$}-\langle\psi_{1}\lvert\psi\rangle\langle\psi\rvert\psi_{2}\rangle% \big{]}.⟼ roman_Re [ ⟨ italic_ψ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT | italic_ψ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⟩ - ⟨ italic_ψ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT | italic_ψ ⟩ ⟨ italic_ψ | italic_ψ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⟩ ] . (32)

Note that here we have defined the Fubini-Study metric by only taking the real part of what is defined in Ref. [37]. This is a more common definition used in other literature [82], and suffices for our purpose of defining the length of a geodesic.

In the following, we denote the Fubini-Study metric by g𝑔gitalic_g, since the choice of metric for 𝖯()𝖯\mathsf{P}(\mathcal{H})sansserif_P ( caligraphic_H ) is unambiguous. Using the Fubini-Study metric, we can calculate the distance between two rays in projective Hilbert space.

Definition 12 (State distance defined by Fubini-Study metric in projective Hilbert space).

Let γ:[0,1]𝖯():𝛾01𝖯\gamma:[0,1]\rightarrow\mathsf{P}(\mathcal{H})italic_γ : [ 0 , 1 ] → sansserif_P ( caligraphic_H ) be the shortest geodesic (see Def. 8) connecting two rays γ(0)=[ϕ]𝛾0delimited-[]italic-ϕ\gamma(0)=[\phi]italic_γ ( 0 ) = [ italic_ϕ ] and γ(1)=[ψ]𝛾1delimited-[]𝜓\gamma(1)=[\psi]italic_γ ( 1 ) = [ italic_ψ ] on the Riemannian manifold (𝖯(),g)𝖯𝑔(\mathsf{P}(\mathcal{H}),g)( sansserif_P ( caligraphic_H ) , italic_g ). The Fubini-Study distance between these two quantum states is defined as the geodesic curve length

dFS([ϕ],[ψ]):=01dtγ˙|γ˙g.assignsubscript𝑑FSdelimited-[]italic-ϕdelimited-[]𝜓superscriptsubscript01differential-d𝑡subscriptinner-product˙𝛾˙𝛾𝑔d_{\mathrm{FS}}([\phi],[\psi]):=\int_{0}^{1}\mathrm{d}t\sqrt{\mbox{$\langle% \dot{\gamma}|\dot{\gamma}\rangle$}_{g}}.italic_d start_POSTSUBSCRIPT roman_FS end_POSTSUBSCRIPT ( [ italic_ϕ ] , [ italic_ψ ] ) := ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT roman_d italic_t square-root start_ARG ⟨ over˙ start_ARG italic_γ end_ARG | over˙ start_ARG italic_γ end_ARG ⟩ start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT end_ARG . (33)

Note that this Fubini-Study distance is known to be related to the angle between the two states [ϕ]delimited-[]italic-ϕ[\phi][ italic_ϕ ] and [ψ]delimited-[]𝜓[\psi][ italic_ψ ]. This is often stated without any proof, but rather the intuitive motivation is that the shortest path in projective space corresponds to the arc length of the enclosed segment of the unit circle containing [ϕ]delimited-[]italic-ϕ[\phi][ italic_ϕ ] and [ψ]delimited-[]𝜓[\psi][ italic_ψ ]. Since we define the state distance via the curve length of the shortest connecting geodesic, we state an explicit expression of the Fubini-Study distance below.

Lemma 1 (Fubini-Study distance).

The Fubini-Study distance can be calculated as

dFS([ϕ],[ψ])=arccos|ϕ|ψ|.subscript𝑑FSdelimited-[]italic-ϕdelimited-[]𝜓inner-productitalic-ϕ𝜓d_{\mathrm{FS}}([\phi],[\psi])=\arccos\left|\mbox{$\langle\phi|\psi\rangle$}% \right|.italic_d start_POSTSUBSCRIPT roman_FS end_POSTSUBSCRIPT ( [ italic_ϕ ] , [ italic_ψ ] ) = roman_arccos | ⟨ italic_ϕ | italic_ψ ⟩ | . (34)

Ref. [83]  (Sec. 4.5) contains a proof by defining the Fubini-Study distance as given by Eq. (34) and showing that the Fubini-Study metric in in Eq. (31) is consistent.

B.4 Manifold of matrix product states

Building on the notations established for MPS in Appendix B.1 and the concepts from differential geometry outlined in Appendix B.2 and B.3, we now proceed to introduce the variational set of projective MPS, as described in Ref. [37]. It turns out that the variational set ψχ(𝔸χ)subscript𝜓𝜒subscript𝔸𝜒\psi_{\chi}(\mathbb{A}_{\chi})italic_ψ start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT ( blackboard_A start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT ) (with 𝔸χsubscript𝔸𝜒\mathbb{A}_{\chi}blackboard_A start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT defined in Eq. (13)) does not have the structure of a manifold. In essence, points in 𝔸χsubscript𝔸𝜒\mathbb{A}_{\chi}blackboard_A start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT where one of the Schmidt ranks is lower than specified by the bond dimensions χ1,,χn+1subscript𝜒1subscript𝜒𝑛1\chi_{1},\ldots,\chi_{n+1}italic_χ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , italic_χ start_POSTSUBSCRIPT italic_n + 1 end_POSTSUBSCRIPT constitute self intersections555 Technically, a self intersection at a point θM𝜃𝑀\theta\in Mitalic_θ ∈ italic_M is a point which has two neighborhoods M1subscript𝑀1M_{1}italic_M start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and M2subscript𝑀2M_{2}italic_M start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT such that the images under the charts φ1(M1)subscript𝜑1subscript𝑀1\varphi_{1}(M_{1})italic_φ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_M start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) and φ2(M2)subscript𝜑2subscript𝑀2\varphi_{2}(M_{2})italic_φ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_M start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) are open (required by the definition of a differentiable manifold), but the images φ1(M1M2)subscript𝜑1subscript𝑀1subscript𝑀2\varphi_{1}(M_{1}\cap M_{2})italic_φ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_M start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ∩ italic_M start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) and φ2(M1M2)subscript𝜑2subscript𝑀1subscript𝑀2\varphi_{2}(M_{1}\cap M_{2})italic_φ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_M start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ∩ italic_M start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) are not open. A simple example is the figure eight, which is the set of points in 2superscript2\mathbb{R}^{2}blackboard_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT satisfying y2=x2x4superscript𝑦2superscript𝑥2superscript𝑥4y^{2}=x^{2}-x^{4}italic_y start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_x start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT. This is a 1-dimensional manifold if one excludes the self-intersection point (x,y)=(0,0)𝑥𝑦00(x,y)=(0,0)( italic_x , italic_y ) = ( 0 , 0 ). To see how this phenomenon manifests itself in the variational set of MPS ψχ(𝔸χ)subscript𝜓𝜒subscript𝔸𝜒\psi_{\chi}(\mathbb{A}_{\chi})italic_ψ start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT ( blackboard_A start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT ), we can consider the example of n=2𝑛2n=2italic_n = 2 qudits of local dimension d=3𝑑3d=3italic_d = 3 and bond dimension χ=2𝜒2\chi=2italic_χ = 2. The (complex) dimension of the corresponding manifold M=ψχ=2(𝒜χ=2)𝑀subscript𝜓𝜒2subscript𝒜𝜒2M=\psi_{\chi=2}(\mathcal{A}_{\chi=2})italic_M = italic_ψ start_POSTSUBSCRIPT italic_χ = 2 end_POSTSUBSCRIPT ( caligraphic_A start_POSTSUBSCRIPT italic_χ = 2 end_POSTSUBSCRIPT ) is dim(M)=8,dimension𝑀8\dim(M)=8,roman_dim ( italic_M ) = 8 , (35) since the dimension of 𝔸χ=2subscript𝔸𝜒2\mathbb{A}_{\chi=2}blackboard_A start_POSTSUBSCRIPT italic_χ = 2 end_POSTSUBSCRIPT (and therefore also 𝒜χ=2subscript𝒜𝜒2\mathcal{A}_{\chi=2}caligraphic_A start_POSTSUBSCRIPT italic_χ = 2 end_POSTSUBSCRIPT) is 12 and the dimension of the structure group is 4 (see Theorem 14 in Ref. [37]). Therefore, the tangent space of M𝑀Mitalic_M at any point has complex dimension 8. However, if the state vector |0,0ket00|0,0\rangle| 0 , 0 ⟩—which has Schmidt rank 1—was an element of M𝑀Mitalic_M, there would be 9 linearly independent “tangent vectors” at |0,0ket00|0,0\rangle| 0 , 0 ⟩. This can be easily seen by considering the curves γi,j(t)=|0,0+t|i,j,subscript𝛾𝑖𝑗𝑡ket00𝑡ket𝑖𝑗\gamma_{i,j}(t)=\mbox{$|0,0\rangle$}+t\mbox{$|i,j\rangle$},italic_γ start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT ( italic_t ) = | 0 , 0 ⟩ + italic_t | italic_i , italic_j ⟩ , (36) for any i=0,1,2𝑖012i=0,1,2italic_i = 0 , 1 , 2 and any j=0,1,2𝑗012j=0,1,2italic_j = 0 , 1 , 2. For any t𝑡t\in\mathbb{R}italic_t ∈ blackboard_R, any i=0,1,2𝑖012i=0,1,2italic_i = 0 , 1 , 2, and any j=0,1,2𝑗012j=0,1,2italic_j = 0 , 1 , 2, the Schmidt rank of γi,j(t)subscript𝛾𝑖𝑗𝑡\gamma_{i,j}(t)italic_γ start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT ( italic_t ) is at most 2 and indeed, all velocity vectors γ˙i,jsubscript˙𝛾𝑖𝑗\dot{\gamma}_{i,j}over˙ start_ARG italic_γ end_ARG start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT are linearly independent. This would require the tangent space of M𝑀Mitalic_M to be 9-dimensional at |0,0ket00|0,0\rangle| 0 , 0 ⟩, which contradicts the fact that the manifold is 8-dimensional, as discussed above. in ψχ(𝔸χ)subscript𝜓𝜒subscript𝔸𝜒\psi_{\chi}(\mathbb{A}_{\chi})italic_ψ start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT ( blackboard_A start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT ). We refer to Ref. [37] for more details on this issue and only mention here that one needs to introduce additional constraints in order to define a variational set which forms a valid Riemannian manifold.

Let us begin with the MPS tensors A[1],,A[n]superscript𝐴delimited-[]1superscript𝐴delimited-[]𝑛A^{[1]},\ldots,A^{[n]}italic_A start_POSTSUPERSCRIPT [ 1 ] end_POSTSUPERSCRIPT , … , italic_A start_POSTSUPERSCRIPT [ italic_n ] end_POSTSUPERSCRIPT as defined in Eq. (14) and denote the transfer matrices by T(α,α),(β,β)[i]=δAα,β[i]δA¯α,β[i]δsubscriptsuperscript𝑇delimited-[]𝑖𝛼superscript𝛼𝛽superscript𝛽subscript𝛿superscriptsubscript𝐴𝛼𝛽delimited-[]𝑖𝛿superscriptsubscript¯𝐴superscript𝛼superscript𝛽delimited-[]𝑖𝛿T^{[i]}_{(\alpha,\alpha^{\prime}),(\beta,\beta^{\prime})}=\sum_{\delta}A_{% \alpha,\beta}^{[i]\delta}\bar{A}_{\alpha^{\prime},\beta^{\prime}}^{[i]\delta}italic_T start_POSTSUPERSCRIPT [ italic_i ] end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ( italic_α , italic_α start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) , ( italic_β , italic_β start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT italic_A start_POSTSUBSCRIPT italic_α , italic_β end_POSTSUBSCRIPT start_POSTSUPERSCRIPT [ italic_i ] italic_δ end_POSTSUPERSCRIPT over¯ start_ARG italic_A end_ARG start_POSTSUBSCRIPT italic_α start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_β start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT [ italic_i ] italic_δ end_POSTSUPERSCRIPT, where A¯¯𝐴\bar{A}over¯ start_ARG italic_A end_ARG represents the complex conjugation of A𝐴Aitalic_A. Furthermore, we denote the left and right virtual density matrices by

l[i](A)superscript𝑙delimited-[]𝑖𝐴\displaystyle l^{[i]}(A)italic_l start_POSTSUPERSCRIPT [ italic_i ] end_POSTSUPERSCRIPT ( italic_A ) =T[1]T[i1],absentsuperscript𝑇delimited-[]1superscript𝑇delimited-[]𝑖1\displaystyle=T^{[1]}\cdots T^{[i-1]},= italic_T start_POSTSUPERSCRIPT [ 1 ] end_POSTSUPERSCRIPT ⋯ italic_T start_POSTSUPERSCRIPT [ italic_i - 1 ] end_POSTSUPERSCRIPT , (37)
r[i](A)superscript𝑟delimited-[]𝑖𝐴\displaystyle r^{[i]}(A)italic_r start_POSTSUPERSCRIPT [ italic_i ] end_POSTSUPERSCRIPT ( italic_A ) =T[i1]T[n],absentsuperscript𝑇delimited-[]𝑖1superscript𝑇delimited-[]𝑛\displaystyle=T^{[i-1]}\cdots T^{[n]},= italic_T start_POSTSUPERSCRIPT [ italic_i - 1 ] end_POSTSUPERSCRIPT ⋯ italic_T start_POSTSUPERSCRIPT [ italic_n ] end_POSTSUPERSCRIPT , (38)

while we define l[0](A)=r[n](A)=1superscript𝑙delimited-[]0𝐴superscript𝑟delimited-[]𝑛𝐴1l^{[0]}(A)=r^{[n]}(A)=1italic_l start_POSTSUPERSCRIPT [ 0 ] end_POSTSUPERSCRIPT ( italic_A ) = italic_r start_POSTSUPERSCRIPT [ italic_n ] end_POSTSUPERSCRIPT ( italic_A ) = 1. As mentioned above, we need to exclude MPS tensors for which the Schmidt rank between sites i𝑖iitalic_i and i+1𝑖1i+1italic_i + 1 is not equal to χi+1subscript𝜒𝑖1\chi_{i+1}italic_χ start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT, i.e., where any of the left or right density matrices are not positive definite.

Definition 13 (Full-rank MPS tensors ([37] Sec III.B Definition 10)).

The subset 𝒜χ𝔸χsubscript𝒜𝜒subscript𝔸𝜒\mathcal{A}_{\chi}\subset\mathbb{A}_{\chi}caligraphic_A start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT ⊂ blackboard_A start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT of full-rank MPS tensors contains all tensors characterized by having positive-definite virtual left and right density matrices l[i](A)superscript𝑙delimited-[]𝑖𝐴l^{[i]}(A)italic_l start_POSTSUPERSCRIPT [ italic_i ] end_POSTSUPERSCRIPT ( italic_A ) and r[i](A)superscript𝑟delimited-[]𝑖𝐴r^{[i]}(A)italic_r start_POSTSUPERSCRIPT [ italic_i ] end_POSTSUPERSCRIPT ( italic_A ), at every site i𝑖iitalic_i:

𝒜χ={A𝔸χ|l[i](A)>0,r[i](A)>0,i}.subscript𝒜𝜒conditional-set𝐴subscript𝔸𝜒formulae-sequencesuperscript𝑙delimited-[]𝑖𝐴0superscript𝑟delimited-[]𝑖𝐴0for-all𝑖\mathcal{A}_{\chi}=\{A\in\mathbb{A}_{\chi}\,|\,l^{[i]}(A)>0,\,\,r^{[i]}(A)>0,% \forall i\}.caligraphic_A start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT = { italic_A ∈ blackboard_A start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT | italic_l start_POSTSUPERSCRIPT [ italic_i ] end_POSTSUPERSCRIPT ( italic_A ) > 0 , italic_r start_POSTSUPERSCRIPT [ italic_i ] end_POSTSUPERSCRIPT ( italic_A ) > 0 , ∀ italic_i } . (39)

We can now define the set of projective full-rank MPS

M~:=ψ~χ(𝒜χ)assign~𝑀subscript~𝜓𝜒subscript𝒜𝜒\tilde{M}:=\tilde{\psi}_{\chi}(\mathcal{A}_{\chi})over~ start_ARG italic_M end_ARG := over~ start_ARG italic_ψ end_ARG start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT ( caligraphic_A start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT ) (40)

to represent physical states within the projective Hilbert space that are described by full-rank MPS tensors. Having removed the self intersections, this variational set has the structure of a Riemannian manifold.

Lemma 2 (Projective MPS manifold ([37] Theorem 15)).

The set M~~𝑀\tilde{M}over~ start_ARG italic_M end_ARG, defined as the image of the projective MPS map** ψ~χsubscript~𝜓𝜒\tilde{\psi}_{\chi}over~ start_ARG italic_ψ end_ARG start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT over the full-rank MPS tensors 𝒜χsubscript𝒜𝜒\mathcal{A}_{\chi}caligraphic_A start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT, forms a differentiable manifold within the projective Hilbert space 𝖯()𝖯\mathsf{P}(\mathcal{H})sansserif_P ( caligraphic_H ), known as the projective MPS manifold.

Note that in Ref. [37], it is demonstrated that M~~𝑀\tilde{M}over~ start_ARG italic_M end_ARG is actually a Kähler manifold when one endows it with the complex-valued Fubini-Study metric, which is shown to be a Hermitian metric. This introduces additional structure—studied in complex geometry—to the manifold. Since we only rely on the variational set being a well-defined Riemannian manifold, we ignore the additional structure here and only define the real-valued Fubini-Study metric in alignment with Def. 11 by inducing the Fubini-Study metric from the embedding space 𝖯()𝖯\mathsf{P}(\mathcal{H})sansserif_P ( caligraphic_H ). The projective MPS manifold is a submanifold M~𝖯()~𝑀𝖯\tilde{M}\subset\mathsf{P}(\mathcal{H})over~ start_ARG italic_M end_ARG ⊂ sansserif_P ( caligraphic_H ) of the projective Hilbert space, which is equipped with the Fubini-Study metric defined in Def. 11. For each point [ψ]M~delimited-[]𝜓~𝑀[\psi]\in\tilde{M}[ italic_ψ ] ∈ over~ start_ARG italic_M end_ARG, the associated tangent space T[ψ]M~subscriptTdelimited-[]𝜓~𝑀\mathrm{T}_{[\psi]}\tilde{M}roman_T start_POSTSUBSCRIPT [ italic_ψ ] end_POSTSUBSCRIPT over~ start_ARG italic_M end_ARG is a subspace of T[ψ]𝖯()subscriptTdelimited-[]𝜓𝖯\mathrm{T}_{[\psi]}\mathsf{P}(\mathcal{H})roman_T start_POSTSUBSCRIPT [ italic_ψ ] end_POSTSUBSCRIPT sansserif_P ( caligraphic_H ). As a result, one can define the induced Fubini-Study metric g~~𝑔\tilde{g}over~ start_ARG italic_g end_ARG on M~~𝑀\tilde{M}over~ start_ARG italic_M end_ARG such that for any [ψ]M~delimited-[]𝜓~𝑀[\psi]\in\tilde{M}[ italic_ψ ] ∈ over~ start_ARG italic_M end_ARG and any |ϕ1,|ϕ2T[ψ]M~ketsubscriptitalic-ϕ1ketsubscriptitalic-ϕ2subscriptTdelimited-[]𝜓~𝑀\mbox{$|\phi_{1}\rangle$},\mbox{$|\phi_{2}\rangle$}\in\mathrm{T}_{[\psi]}% \tilde{M}| italic_ϕ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ⟩ , | italic_ϕ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⟩ ∈ roman_T start_POSTSUBSCRIPT [ italic_ψ ] end_POSTSUBSCRIPT over~ start_ARG italic_M end_ARG, one has

ϕ1|ϕ2g~=ϕ1|ϕ2g,subscriptinner-productsubscriptitalic-ϕ1subscriptitalic-ϕ2~𝑔subscriptinner-productsubscriptitalic-ϕ1subscriptitalic-ϕ2𝑔\mbox{$\langle\phi_{1}|\phi_{2}\rangle$}_{\tilde{g}}=\mbox{$\langle\phi_{1}|% \phi_{2}\rangle$}_{g},⟨ italic_ϕ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT | italic_ϕ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT over~ start_ARG italic_g end_ARG end_POSTSUBSCRIPT = ⟨ italic_ϕ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT | italic_ϕ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT , (41)

as done in Sec. IIIE of Ref. [37].

B.5 Maximum likelihood estimation on manifold

In this subsection, we review and provide intuitions for existing results on the asymptotic normality of maximum likelihood estimation (MLE) on a manifold [36] in Lemma 3. More concretely, Ref. [36] defines an appropriate notion of asymptotic normality of the distance between the estimator and the target on a manifold. Our main result of Theorem 2 follows directly by applying Lemma 3 to our analysis of MLE on the manifold of projective MPS, as defined in Appendix B.4.

Let (M,g)𝑀𝑔(M,g)( italic_M , italic_g ) be a Riemannian manifold of dimension r:=dim(M)assign𝑟dimension𝑀r:=\dim(M)italic_r := roman_dim ( italic_M ) with a Riemannian connection \nabla and let 𝒫θsubscript𝒫𝜃\mathcal{P}_{\theta}caligraphic_P start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT be a family of distributions parametrized by elements θM𝜃𝑀\theta\in Mitalic_θ ∈ italic_M. Let 𝒮𝒮\mathcal{S}caligraphic_S represent the sample space of the distributions, i.e., 𝒫θ:𝒮[0,1]:subscript𝒫𝜃𝒮01\mathcal{P}_{\theta}:\mathcal{S}\rightarrow[0,1]caligraphic_P start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT : caligraphic_S → [ 0 , 1 ]. Our goal is to identify the target point θ0Msubscript𝜃0𝑀\theta_{0}\in Mitalic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∈ italic_M on the manifold. To achieve this, we assume access to a dataset, which is formally defined as a random variable D(N)superscript𝐷𝑁{D}^{(N)}italic_D start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT consisting of N𝑁Nitalic_N samples drawn i.i.d. from the target distribution 𝒫θ0subscript𝒫subscript𝜃0\mathcal{P}_{\theta_{0}}caligraphic_P start_POSTSUBSCRIPT italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT. Given a realization of the dataset D=(xi𝒫θ0)i=1N𝐷superscriptsubscriptsimilar-tosubscript𝑥𝑖subscript𝒫subscript𝜃0𝑖1𝑁D=(x_{i}\sim\mathcal{P}_{\theta_{0}})_{i=1}^{N}italic_D = ( italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ∼ caligraphic_P start_POSTSUBSCRIPT italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT, MLE aims to identify the element on the manifold θDsuperscript𝜃𝐷\theta^{D}italic_θ start_POSTSUPERSCRIPT italic_D end_POSTSUPERSCRIPT for which the associated distribution 𝒫θDsubscript𝒫superscript𝜃𝐷\mathcal{P}_{\theta^{D}}caligraphic_P start_POSTSUBSCRIPT italic_θ start_POSTSUPERSCRIPT italic_D end_POSTSUPERSCRIPT end_POSTSUBSCRIPT is most likely to have generated the dataset D𝐷Ditalic_D. This is achieved by minimizing the negative log-likelihood (NLL). More specifically, we define the NLL for the dataset instance D𝐷Ditalic_D, denoted by NLLDsubscriptsuperscript𝐷NLL\mathcal{L}^{D}_{\text{NLL}}caligraphic_L start_POSTSUPERSCRIPT italic_D end_POSTSUPERSCRIPT start_POSTSUBSCRIPT NLL end_POSTSUBSCRIPT, as a map from the manifold M𝑀Mitalic_M to the real numbers,

NLLD:M:subscriptsuperscript𝐷NLL𝑀\displaystyle\mathcal{L}^{D}_{\text{NLL}}:M\leavevmode\nobreak\ caligraphic_L start_POSTSUPERSCRIPT italic_D end_POSTSUPERSCRIPT start_POSTSUBSCRIPT NLL end_POSTSUBSCRIPT : italic_M ,absent\displaystyle\longrightarrow\mathbb{R},⟶ blackboard_R , (42)
θ𝜃\displaystyle\thetaitalic_θ 1Ni=1Nlog𝒫θ[xi],absent1𝑁superscriptsubscript𝑖1𝑁subscript𝒫𝜃delimited-[]subscript𝑥𝑖\displaystyle\longmapsto-\frac{1}{N}\sum_{i=1}^{N}\log\mathcal{P}_{\theta}[x_{% i}],⟼ - divide start_ARG 1 end_ARG start_ARG italic_N end_ARG ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT roman_log caligraphic_P start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT [ italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ] , (43)

which measures how likely the observed samples are under the distribution 𝒫θsubscript𝒫𝜃\mathcal{P}_{\theta}caligraphic_P start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT. By minimizing the NLL for the dataset D𝐷Ditalic_D we find the estimator

θD=argminθMNLLD(θ).superscript𝜃𝐷subscriptargmin𝜃𝑀subscriptsuperscript𝐷NLL𝜃\theta^{D}=\text{argmin}_{\theta\in M}\mathcal{L}^{D}_{\text{NLL}}(\theta).italic_θ start_POSTSUPERSCRIPT italic_D end_POSTSUPERSCRIPT = argmin start_POSTSUBSCRIPT italic_θ ∈ italic_M end_POSTSUBSCRIPT caligraphic_L start_POSTSUPERSCRIPT italic_D end_POSTSUPERSCRIPT start_POSTSUBSCRIPT NLL end_POSTSUBSCRIPT ( italic_θ ) . (44)

For clarity, we use NLLD(θ)subscriptsuperscript𝐷NLL𝜃\mathcal{L}^{D}_{\text{NLL}}(\theta)caligraphic_L start_POSTSUPERSCRIPT italic_D end_POSTSUPERSCRIPT start_POSTSUBSCRIPT NLL end_POSTSUBSCRIPT ( italic_θ ) and θDsuperscript𝜃𝐷\theta^{D}italic_θ start_POSTSUPERSCRIPT italic_D end_POSTSUPERSCRIPT to denote the NLL and MLE estimator for the specific dataset D𝐷Ditalic_D consisting of N𝑁Nitalic_N samples, whereas we use the notation θ^(N)superscript^𝜃𝑁\hat{\theta}^{(N)}over^ start_ARG italic_θ end_ARG start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT for the random variable corresponding to the minimizer of NLL associated to the N𝑁Nitalic_N-sample dataset D(N)superscript𝐷𝑁{D}^{(N)}italic_D start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT drawn randomly. In Euclidean space, minimization can be achieved by finding points where the gradient of the loss function vanishes. On a manifold M𝑀Mitalic_M, we can define an estimating form ω𝜔\omegaitalic_ω, which plays the role of the gradient.

Definition 14 (Estimating form ([36] Definition 1)).

A map

ω:𝒮×MTM,:𝜔𝒮𝑀T𝑀\omega:\mathcal{S}\times M\rightarrow\mathrm{T}M,italic_ω : caligraphic_S × italic_M → roman_T italic_M , (45)

is an estimating form if 𝔼x𝒫θ[ω(x,θ)]=0subscript𝔼similar-to𝑥subscript𝒫𝜃delimited-[]𝜔𝑥𝜃0\mathbb{E}_{x\sim\mathcal{P}_{\theta}}[\omega(x,\theta)]=0blackboard_E start_POSTSUBSCRIPT italic_x ∼ caligraphic_P start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT end_POSTSUBSCRIPT [ italic_ω ( italic_x , italic_θ ) ] = 0 for all θM𝜃𝑀\theta\in Mitalic_θ ∈ italic_M. TMT𝑀\mathrm{T}Mroman_T italic_M is the tangent bundle, which is a collection of tangent spaces over all points in the manifold.

A more concrete expression of the estimating form follows from the covariant derivative of the NLL.

Fact 2 (Differentiable estimating form).

The gradient (in local coordinates) of the NLL is a differentiable estimating form

ω(x,θ)=(De1log(𝒫θ[x]),,Derlog(𝒫θ[x])),𝜔𝑥𝜃superscriptsubscriptDsubscript𝑒1subscript𝒫𝜃delimited-[]𝑥subscriptDsubscript𝑒𝑟subscript𝒫𝜃delimited-[]𝑥top\omega(x,\theta)=-\big{(}\mathrm{D}_{e_{1}}\log(\mathcal{P}_{\theta}[x]),% \ldots,\mathrm{D}_{e_{r}}\log(\mathcal{P}_{\theta}[x])\big{)}^{\top},italic_ω ( italic_x , italic_θ ) = - ( roman_D start_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUBSCRIPT roman_log ( caligraphic_P start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT [ italic_x ] ) , … , roman_D start_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT end_POSTSUBSCRIPT roman_log ( caligraphic_P start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT [ italic_x ] ) ) start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT , (46)

where {ea}a=1rsuperscriptsubscriptsubscript𝑒𝑎𝑎1𝑟\{e_{a}\}_{a=1}^{r}{ italic_e start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT } start_POSTSUBSCRIPT italic_a = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT is a basis of the tangent space TθMsubscriptT𝜃𝑀\mathrm{T}_{\theta}Mroman_T start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT italic_M, and DXsubscriptD𝑋\mathrm{D}_{X}roman_D start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT denotes the directional derivative with respect to a tangent vector XTθM𝑋subscriptT𝜃𝑀X\in\mathrm{T}_{\theta}Mitalic_X ∈ roman_T start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT italic_M as introduced in Def. 6.

Proof.

To show 𝔼x𝒫θ[ω(x,θ)]=0subscript𝔼similar-to𝑥subscript𝒫𝜃delimited-[]𝜔𝑥𝜃0\mathbb{E}_{x\sim\mathcal{P}_{\theta}}[\omega(x,\theta)]=0blackboard_E start_POSTSUBSCRIPT italic_x ∼ caligraphic_P start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT end_POSTSUBSCRIPT [ italic_ω ( italic_x , italic_θ ) ] = 0, we write down its definition

𝔼x𝒫θ[ω(x,θ)]subscript𝔼similar-to𝑥subscript𝒫𝜃delimited-[]𝜔𝑥𝜃\displaystyle\mathbb{E}_{x\sim\mathcal{P}_{\theta}}[\omega(x,\theta)]blackboard_E start_POSTSUBSCRIPT italic_x ∼ caligraphic_P start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT end_POSTSUBSCRIPT [ italic_ω ( italic_x , italic_θ ) ] =1|𝒮|x𝒮𝒫θ[x](De1log(𝒫θ[x]),,Derlog(𝒫θ[x])),absent1𝒮subscript𝑥𝒮subscript𝒫𝜃delimited-[]𝑥superscriptsubscriptDsubscript𝑒1subscript𝒫𝜃delimited-[]𝑥subscriptDsubscript𝑒𝑟subscript𝒫𝜃delimited-[]𝑥top\displaystyle=-\frac{1}{\lvert\mathcal{S}\rvert}\sum_{x\in\mathcal{S}}\mathcal% {P}_{\theta}[x]\Big{(}\mathrm{D}_{e_{1}}\log(\mathcal{P}_{\theta}[x]),\ldots,% \mathrm{D}_{e_{r}}\log(\mathcal{P}_{\theta}[x])\Big{)}^{\top},= - divide start_ARG 1 end_ARG start_ARG | caligraphic_S | end_ARG ∑ start_POSTSUBSCRIPT italic_x ∈ caligraphic_S end_POSTSUBSCRIPT caligraphic_P start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT [ italic_x ] ( roman_D start_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUBSCRIPT roman_log ( caligraphic_P start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT [ italic_x ] ) , … , roman_D start_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT end_POSTSUBSCRIPT roman_log ( caligraphic_P start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT [ italic_x ] ) ) start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT , (47)
=1|𝒮|x𝒮(De1𝒫θ[x],,Der𝒫θ[x]),absent1𝒮subscript𝑥𝒮superscriptsubscriptDsubscript𝑒1subscript𝒫𝜃delimited-[]𝑥subscriptDsubscript𝑒𝑟subscript𝒫𝜃delimited-[]𝑥top\displaystyle=-\frac{1}{\lvert\mathcal{S}\rvert}\sum_{x\in\mathcal{S}}\Big{(}% \mathrm{D}_{e_{1}}\mathcal{P}_{\theta}[x],\ldots,\mathrm{D}_{e_{r}}\mathcal{P}% _{\theta}[x]\Big{)}^{\top},= - divide start_ARG 1 end_ARG start_ARG | caligraphic_S | end_ARG ∑ start_POSTSUBSCRIPT italic_x ∈ caligraphic_S end_POSTSUBSCRIPT ( roman_D start_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUBSCRIPT caligraphic_P start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT [ italic_x ] , … , roman_D start_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT end_POSTSUBSCRIPT caligraphic_P start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT [ italic_x ] ) start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT ,
=(De11|𝒮|x𝒮𝒫θ[x]=1,,Der1|𝒮|x𝒮𝒫θ[x]),absentsuperscriptsubscriptDsubscript𝑒1subscript1𝒮subscript𝑥𝒮subscript𝒫𝜃delimited-[]𝑥absent1subscriptDsubscript𝑒𝑟1𝒮subscript𝑥𝒮subscript𝒫𝜃delimited-[]𝑥top\displaystyle=-\Big{(}\mathrm{D}_{e_{1}}\underbrace{\frac{1}{\lvert\mathcal{S}% \rvert}\sum_{x\in\mathcal{S}}\mathcal{P}_{\theta}[x]}_{=1},\ldots,\mathrm{D}_{% e_{r}}\frac{1}{\lvert\mathcal{S}\rvert}\sum_{x\in\mathcal{S}}\mathcal{P}_{% \theta}[x]\Big{)}^{\top},= - ( roman_D start_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUBSCRIPT under⏟ start_ARG divide start_ARG 1 end_ARG start_ARG | caligraphic_S | end_ARG ∑ start_POSTSUBSCRIPT italic_x ∈ caligraphic_S end_POSTSUBSCRIPT caligraphic_P start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT [ italic_x ] end_ARG start_POSTSUBSCRIPT = 1 end_POSTSUBSCRIPT , … , roman_D start_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT end_POSTSUBSCRIPT divide start_ARG 1 end_ARG start_ARG | caligraphic_S | end_ARG ∑ start_POSTSUBSCRIPT italic_x ∈ caligraphic_S end_POSTSUBSCRIPT caligraphic_P start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT [ italic_x ] ) start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT ,
=0.absent0\displaystyle=0.= 0 .

In this definition, the directional derivative is well defined as long as 𝒫θsubscript𝒫𝜃\mathcal{P}_{\theta}caligraphic_P start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT is differentiable, while terms with probability 𝒫θ[x]=0subscript𝒫𝜃delimited-[]𝑥0\mathcal{P}_{\theta}[x]=0caligraphic_P start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT [ italic_x ] = 0 are to be omitted in the sum, by the limit ylogy0𝑦𝑦0y\log y\rightarrow 0italic_y roman_log italic_y → 0 as y0𝑦0y\rightarrow 0italic_y → 0. Consequently, the set of estimating forms {ω(x,θ),θM}𝜔𝑥𝜃𝜃𝑀\{\omega(x,\theta),\theta\in M\}{ italic_ω ( italic_x , italic_θ ) , italic_θ ∈ italic_M } are differentiable vector fields. ∎

Additionally, the asymptotic normality of an MLE estimator in Euclidean space requires the Hessian of the loss function to be well behaved around the target parameters. In the manifold setting, we need a similar condition and define the Hessian matrix below.

Definition 15 (Hessian matrix [36]).

Let ω𝜔\omegaitalic_ω be a differentiable estimating form. Choose a basis set {ea}a=1rTθMsuperscriptsubscriptsubscript𝑒𝑎𝑎1𝑟subscript𝑇𝜃𝑀\{e_{a}\}_{a=1}^{r}\in T_{\theta}M{ italic_e start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT } start_POSTSUBSCRIPT italic_a = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT ∈ italic_T start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT italic_M in the tangent space of the manifold at θ𝜃\thetaitalic_θ. The Hessian matrix, in this basis, is defined through the Riemannian connection of ω𝜔\omegaitalic_ω, and is given by its matrix elements:

Ha,b(x,θ)=eaω(x,θ)|ebg.subscriptH𝑎𝑏𝑥𝜃subscriptinner-productsubscriptsubscript𝑒𝑎𝜔𝑥𝜃subscript𝑒𝑏𝑔\mathrm{H}_{a,b}(x,\theta)=\mbox{$\langle\nabla_{e_{a}}\omega(x,\theta)|e_{b}% \rangle$}_{g}.roman_H start_POSTSUBSCRIPT italic_a , italic_b end_POSTSUBSCRIPT ( italic_x , italic_θ ) = ⟨ ∇ start_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_ω ( italic_x , italic_θ ) | italic_e start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT . (48)

In subsequent discussions, we often consider the expectation of the Hessian matrix elements over a distribution 𝒫θsubscript𝒫𝜃\mathcal{P}_{\theta}caligraphic_P start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT given by

Ha,b(θ)=𝔼x𝒫θ[eaω(x,θ)|ebg].subscriptH𝑎𝑏𝜃subscript𝔼similar-to𝑥subscript𝒫𝜃delimited-[]subscriptinner-productsubscriptsubscript𝑒𝑎𝜔𝑥𝜃subscript𝑒𝑏𝑔\mathrm{H}_{a,b}(\theta)=\mathbb{E}_{x\sim\mathcal{P}_{\theta}}\big{[}\mbox{$% \langle\nabla_{e_{a}}\omega(x,\theta)|e_{b}\rangle$}_{g}\big{]}.roman_H start_POSTSUBSCRIPT italic_a , italic_b end_POSTSUBSCRIPT ( italic_θ ) = blackboard_E start_POSTSUBSCRIPT italic_x ∼ caligraphic_P start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT end_POSTSUBSCRIPT [ ⟨ ∇ start_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_ω ( italic_x , italic_θ ) | italic_e start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT ] . (49)

Having reviewed the relevant concepts, we are now prepared to state the theorem concerning asymptotic normality of MLE on a manifold following Ref. [36]. Let us adopt the following notation: given an estimate θM𝜃𝑀\theta\in Mitalic_θ ∈ italic_M, we denote by γθ(t)subscript𝛾𝜃𝑡\gamma_{\theta}(t)italic_γ start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT ( italic_t ) the shortest connecting geodesic (see Def. 8) such that γ(0)=θ0𝛾0subscript𝜃0\gamma(0)=\theta_{0}italic_γ ( 0 ) = italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and γ(1)=θ𝛾1𝜃\gamma(1)=\thetaitalic_γ ( 1 ) = italic_θ, i.e., γθsubscript𝛾𝜃\gamma_{\theta}italic_γ start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT is the shortest path connecting the MLE estimate to the target. Moreover, let {ea}a=1rsuperscriptsubscriptsubscript𝑒𝑎𝑎1𝑟\{e_{a}\}_{a=1}^{r}{ italic_e start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT } start_POSTSUBSCRIPT italic_a = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT be a basis of the target tangent space Tθ0MsubscriptTsubscript𝜃0𝑀\mathrm{T}_{\theta_{0}}Mroman_T start_POSTSUBSCRIPT italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_M and given a shortest connecting curve γθsubscript𝛾𝜃\gamma_{\theta}italic_γ start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT, denote the parallel transport (see Def. 7) of the basis vectors by {ea(t)}a=1rsuperscriptsubscriptsubscript𝑒𝑎𝑡𝑎1𝑟\{e_{a}(t)\}_{a=1}^{r}{ italic_e start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ( italic_t ) } start_POSTSUBSCRIPT italic_a = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT, i.e., ea(0)=easubscript𝑒𝑎0subscript𝑒𝑎e_{a}(0)=e_{a}italic_e start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ( 0 ) = italic_e start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT for all a𝑎aitalic_a. This allows us to decompose tangent vectors with respect to well-defined basis vectors along the entire curve γθsubscript𝛾𝜃\gamma_{\theta}italic_γ start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT.

Lemma 3 (Asymptotic normality ([36] Theorem 2)).

Let ω𝜔\omegaitalic_ω be the differentiable estimating form as defined in Def. 14 and HH\mathrm{H}roman_H be the Hessian matrix defined in Def. 15. Let (xi)i=1Nsuperscriptsubscriptsubscript𝑥𝑖𝑖1𝑁(x_{i})_{i=1}^{N}( italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT be a sequence of i.i.d. samples from 𝒫θ0subscript𝒫subscript𝜃0\mathcal{P}_{\theta_{0}}caligraphic_P start_POSTSUBSCRIPT italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT, where θ0Msubscript𝜃0𝑀\theta_{0}\in Mitalic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∈ italic_M is the target element. Denote by θ^(N)superscript^𝜃𝑁\hat{\theta}^{(N)}over^ start_ARG italic_θ end_ARG start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT the MLE estimator, which is the minimizer of the NLL, and let γ𝛾\gammaitalic_γ be the connecting geodesic between γ(0)=θ0𝛾0subscript𝜃0\gamma(0)=\theta_{0}italic_γ ( 0 ) = italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and γ(1)=θ^(N)𝛾1superscript^𝜃𝑁\gamma(1)=\hat{\theta}^{(N)}italic_γ ( 1 ) = over^ start_ARG italic_θ end_ARG start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT. Furthermore, consider the following conditions:

  1. 1.

    The estimator θ^(N)superscript^𝜃𝑁\hat{\theta}^{(N)}over^ start_ARG italic_θ end_ARG start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT satisfies i=1Nω(xi,θ^(N))=0superscriptsubscript𝑖1𝑁𝜔subscript𝑥𝑖superscript^𝜃𝑁0\sum_{i=1}^{N}\omega(x_{i},\hat{\theta}^{(N)})=0∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT italic_ω ( italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , over^ start_ARG italic_θ end_ARG start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT ) = 0 (first-order condition) for all N𝑁Nitalic_N and converges in distribution to the target θ^(N)p.θ0\hat{\theta}^{(N)}\overset{\mathrm{p}.}{\longrightarrow}\theta_{0}over^ start_ARG italic_θ end_ARG start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT start_OVERACCENT roman_p . end_OVERACCENT start_ARG ⟶ end_ARG italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT as N𝑁N\rightarrow\inftyitalic_N → ∞.

  2. 2.

    The expected Hessian matrix Ha,b(θ0)subscriptH𝑎𝑏subscript𝜃0\mathrm{H}_{a,b}(\theta_{0})roman_H start_POSTSUBSCRIPT italic_a , italic_b end_POSTSUBSCRIPT ( italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) at the target point, defined in a basis {ea}a=1rsuperscriptsubscriptsubscript𝑒𝑎𝑎1𝑟\{e_{a}\}_{a=1}^{r}{ italic_e start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT } start_POSTSUBSCRIPT italic_a = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT, is finite and invertible.

  3. 3.

    The Hessian H(x,θ)H𝑥𝜃\mathrm{H}(x,\theta)roman_H ( italic_x , italic_θ ) in a neighborhood of θ0subscript𝜃0\theta_{0}italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT converges uniformly to H(x,θ0)H𝑥subscript𝜃0\mathrm{H}(x,\theta_{0})roman_H ( italic_x , italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) as θθ0𝜃subscript𝜃0\theta\rightarrow\theta_{0}italic_θ → italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT in the following sense:

    𝔼x𝒫θ0[supθBθ0(δ)|ea(1)ω(x,θ)|eb(1)eaω(x,θ0)|eb|]p.0\mathbb{E}_{x\sim\mathcal{P}_{\theta_{0}}}\left[\underset{\theta\in B_{\theta_% {0}}(\delta)}{\sup}\Big{|}\mbox{$\langle\nabla_{e_{a}(1)}\omega(x,\theta)|e_{b% }(1)\rangle$}-\mbox{$\langle\nabla_{e_{a}}\omega(x,\theta_{0})|e_{b}\rangle$}% \Big{|}\right]\overset{\mathrm{p.}}{\longrightarrow}0blackboard_E start_POSTSUBSCRIPT italic_x ∼ caligraphic_P start_POSTSUBSCRIPT italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT end_POSTSUBSCRIPT [ start_UNDERACCENT italic_θ ∈ italic_B start_POSTSUBSCRIPT italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT ( italic_δ ) end_UNDERACCENT start_ARG roman_sup end_ARG | ⟨ ∇ start_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ( 1 ) end_POSTSUBSCRIPT italic_ω ( italic_x , italic_θ ) | italic_e start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ( 1 ) ⟩ - ⟨ ∇ start_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_ω ( italic_x , italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) | italic_e start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ⟩ | ] start_OVERACCENT roman_p . end_OVERACCENT start_ARG ⟶ end_ARG 0 (50)

    as δ0𝛿0\delta\rightarrow 0italic_δ → 0, where Bθ0(δ)subscript𝐵subscript𝜃0𝛿B_{\theta_{0}}(\delta)italic_B start_POSTSUBSCRIPT italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT ( italic_δ ) is the set of all points with distance at most δ𝛿\deltaitalic_δ from θ0subscript𝜃0\theta_{0}italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT (the “δ𝛿\deltaitalic_δ-Ball” around θ0subscript𝜃0\theta_{0}italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT).

When these conditions are satisfied the components Δ1,,ΔrsubscriptΔ1subscriptΔ𝑟\Delta_{1},\ldots,\Delta_{r}roman_Δ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , roman_Δ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT of the velocity vector γ˙(t)=aΔaea(t)˙𝛾𝑡subscript𝑎subscriptΔ𝑎subscript𝑒𝑎𝑡\dot{\gamma}(t)=\sum_{a}\Delta_{a}e_{a}(t)over˙ start_ARG italic_γ end_ARG ( italic_t ) = ∑ start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT roman_Δ start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ( italic_t ) converge in distribution to a multivariate normal distribution

N(Δ1,,Δr)d.𝒩(0,Σ),\sqrt{N}(\Delta_{1},\ldots,\Delta_{r})^{\top}\overset{\mathrm{d.}}{% \longrightarrow}\mathcal{N}(0,\Sigma),square-root start_ARG italic_N end_ARG ( roman_Δ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , roman_Δ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT start_OVERACCENT roman_d . end_OVERACCENT start_ARG ⟶ end_ARG caligraphic_N ( 0 , roman_Σ ) , (51)

where Σ=(H)1ΓHΣsuperscriptsuperscriptH1ΓH\Sigma=(\mathrm{H}^{\dagger})^{-1}\Gamma\mathrm{H}roman_Σ = ( roman_H start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_Γ roman_H and the Gram matrix Γa,b=𝔼x𝒫θ0[ω(x,θ0)|eaω(x,θ0)|eb]subscriptΓ𝑎𝑏subscript𝔼similar-to𝑥subscript𝒫subscript𝜃0delimited-[]ω(x,θ0)|eaω(x,θ0)|eb\Gamma_{a,b}=\mathbb{E}_{x\sim\mathcal{P}_{\theta_{0}}}\big{[}\mbox{$\langle% \omega(x,\theta_{0})|e_{a}\rangle$}\mbox{$\langle\omega(x,\theta_{0})|e_{b}% \rangle$}\big{]}roman_Γ start_POSTSUBSCRIPT italic_a , italic_b end_POSTSUBSCRIPT = blackboard_E start_POSTSUBSCRIPT italic_x ∼ caligraphic_P start_POSTSUBSCRIPT italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT end_POSTSUBSCRIPT [ italic_⟨ω(x,θ0)|ea⟩ italic_⟨ω(x,θ0)|eb⟩ ].

B.6 Main result: A concentration inequality for infidelity

In this subsection, we prove our main result that the infidelity between an MLE estimator and the target state can be probabilistically upper bounded when the number of samples N𝑁Nitalic_N is asymptotically large. Theorem 2 states that such an upper bound depends on the properties of the target state and converges to zero with a rate depending on the number of samples as 1/N1𝑁1/\sqrt{N}1 / square-root start_ARG italic_N end_ARG. A central component of our proof relies on the asymptotic normality of the MLE estimator on a manifold discussed in Lemma 3. Before rigorously proving our theorem, we first integrate the concepts from previous sections and restate the normality condition in the context of the projective MPS manifold M~~𝑀\tilde{M}over~ start_ARG italic_M end_ARG, as discussed in Appendix B.4 and B.5.

Consider the target state [ϕ]M~delimited-[]italic-ϕ~𝑀[{\phi}]\in\tilde{M}[ italic_ϕ ] ∈ over~ start_ARG italic_M end_ARG, with a normalized representative |ϕketitalic-ϕ\mbox{$|\phi\rangle$}\in\mathcal{H}| italic_ϕ ⟩ ∈ caligraphic_H. Given a unitary ensemble 𝒰𝒰\mathcal{U}caligraphic_U, the sample space 𝒮𝒮\mathcal{S}caligraphic_S consists of all unitary-bit-string pairs 𝒮=𝒰×{0,1}n𝒮𝒰superscript01𝑛\mathcal{S}=\mathcal{U}\times\{0,1\}^{n}caligraphic_S = caligraphic_U × { 0 , 1 } start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT. Let 𝒟𝒟\mathcal{D}caligraphic_D denote the data probability distribution corresponding to the target state [ϕ]delimited-[]italic-ϕ[{\phi}][ italic_ϕ ] given by

𝒟[(U,b),[ϕ]]=1|𝒰|×b|U|ϕϕ|U|b.𝒟𝑈𝑏delimited-[]italic-ϕ1𝒰delimited-⟨⟩𝑏𝑈italic-ϕdelimited-⟨⟩italic-ϕsuperscript𝑈𝑏\mathcal{D}[(U,b),[\phi]]=\frac{1}{\lvert\mathcal{U}\rvert}\times\langle b% \lvert U\rvert\phi\rangle\!\langle\phi\lvert U^{\dagger}\rvert b\rangle.caligraphic_D [ ( italic_U , italic_b ) , [ italic_ϕ ] ] = divide start_ARG 1 end_ARG start_ARG | caligraphic_U | end_ARG × ⟨ italic_b | italic_U | italic_ϕ ⟩ ⟨ italic_ϕ | italic_U start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT | italic_b ⟩ . (52)

For a dataset D={(Ui,bi)}i=1N𝐷superscriptsubscriptsubscript𝑈𝑖subscript𝑏𝑖𝑖1𝑁D=\{(U_{i},b_{i})\}_{i=1}^{N}italic_D = { ( italic_U start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) } start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT drawn i.i.d. from 𝒟𝒟\mathcal{D}caligraphic_D, its corresponding negative log-likelihood (NLL) is defined as

NLLD:M~:subscriptsuperscript𝐷NLL~𝑀\displaystyle\mathcal{L}^{D}_{\text{NLL}}:\tilde{M}\leavevmode\nobreak\ caligraphic_L start_POSTSUPERSCRIPT italic_D end_POSTSUPERSCRIPT start_POSTSUBSCRIPT NLL end_POSTSUBSCRIPT : over~ start_ARG italic_M end_ARG ,absent\displaystyle\longrightarrow\mathbb{R},⟶ blackboard_R , (53)
[ψ]delimited-[]𝜓\displaystyle[{\psi}][ italic_ψ ] 1Ni=1Nlog|bi|Ui|ψ|2,\displaystyle\longmapsto-\frac{1}{N}\sum_{i=1}^{N}\log\lvert\langle b_{i}% \lvert U_{i}\rvert\psi\rangle\rvert^{2},⟼ - divide start_ARG 1 end_ARG start_ARG italic_N end_ARG ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT roman_log | ⟨ italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | italic_U start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | italic_ψ ⟩ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , (54)

Notice that by introducing a scaling factor c>1𝑐1c>1italic_c > 1, the loss associated with any vector can be arbitrarily reduced through the transformation |ψc|ψmaps-toket𝜓𝑐ket𝜓\mbox{$|\psi\rangle$}\mapsto c\mbox{$|\psi\rangle$}| italic_ψ ⟩ ↦ italic_c | italic_ψ ⟩. However, such rescaling does not affect physical observables and is effectively managed by constraining the analysis to the projective Hilbert space. This highlights the practical value of defining the projective MPS and its associated projective manifold. The MLE estimate is given by the minimizer of the NLL

[ψD]=argmin[ψ]M~NLLD([ψ]).delimited-[]superscript𝜓𝐷subscriptargmindelimited-[]𝜓~𝑀subscriptsuperscript𝐷NLLdelimited-[]𝜓[{\psi}^{D}]=\text{argmin}_{[\psi]\in\tilde{M}}\mathcal{L}^{D}_{\text{NLL}}% \left([\psi]\right).[ italic_ψ start_POSTSUPERSCRIPT italic_D end_POSTSUPERSCRIPT ] = argmin start_POSTSUBSCRIPT [ italic_ψ ] ∈ over~ start_ARG italic_M end_ARG end_POSTSUBSCRIPT caligraphic_L start_POSTSUPERSCRIPT italic_D end_POSTSUPERSCRIPT start_POSTSUBSCRIPT NLL end_POSTSUBSCRIPT ( [ italic_ψ ] ) . (55)

Using the notation established in Appendix B.5, we use [ψ^(N)]delimited-[]superscript^𝜓𝑁[\hat{\psi}^{(N)}][ over^ start_ARG italic_ψ end_ARG start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT ] to denote the random variable of the MLE estimator, which is the minimizer of the N𝑁Nitalic_N-sample NLL for a dataset random variable D(N)superscript𝐷𝑁{D}^{(N)}italic_D start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT. Using Fact 2 we find a concrete expression of the estimating form by taking covariant derivative of the NLL as follows.

Corollary 1 (Differentiable estimating form).

ω((U,b),[ψ])=(De1log|b|U|ψ|2,,Derlog|b|U|ψ|2)\omega((U,b),[\psi])=-\big{(}\mathrm{D}_{e_{1}}\log\lvert\langle b\lvert U% \rvert\psi\rangle\rvert^{2},\ldots,\mathrm{D}_{e_{r}}\log\lvert\langle b\lvert U% \rvert\psi\rangle\rvert^{2}\big{)}^{\top}italic_ω ( ( italic_U , italic_b ) , [ italic_ψ ] ) = - ( roman_D start_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUBSCRIPT roman_log | ⟨ italic_b | italic_U | italic_ψ ⟩ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , … , roman_D start_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT end_POSTSUBSCRIPT roman_log | ⟨ italic_b | italic_U | italic_ψ ⟩ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT is a differentiable estimating form.

In order to establish the essential property of asymptotic normality of the projective MPS estimator, as required to prove our main theorem, we first adapt the necessary conditions in Lemma 3 to our MPS manifold setting as follows.

Proposition 1 (Consistent estimator).

Let D(N)={(Ui,bi)}i=1Nsuperscript𝐷𝑁superscriptsubscriptsubscript𝑈𝑖subscript𝑏𝑖𝑖1𝑁{D}^{(N)}=\{(U_{i},b_{i})\}_{i=1}^{N}italic_D start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT = { ( italic_U start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) } start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT be a sequence of i.i.d. samples from 𝒟𝒟\mathcal{D}caligraphic_D defined by the target state [ϕ]delimited-[]italic-ϕ[\phi][ italic_ϕ ] and a unitary ensemble 𝒰𝒰\mathcal{U}caligraphic_U, tomographically complete in the space of variational states on M~~𝑀\tilde{M}over~ start_ARG italic_M end_ARG. There exists a sequence of N𝑁Nitalic_N-sample NLL minimizers ([ψ^(N)])N=1superscriptsubscriptdelimited-[]superscript^𝜓𝑁𝑁1([\hat{\psi}^{(N)}])_{N=1}^{\infty}( [ over^ start_ARG italic_ψ end_ARG start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT ] ) start_POSTSUBSCRIPT italic_N = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT such that i=1Nω((Ui,bi),[ψ^(N)])=0superscriptsubscript𝑖1𝑁𝜔subscript𝑈𝑖subscript𝑏𝑖delimited-[]superscript^𝜓𝑁0\sum_{i=1}^{N}\omega((U_{i},b_{i}),[\hat{\psi}^{(N)}])=0∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT italic_ω ( ( italic_U start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) , [ over^ start_ARG italic_ψ end_ARG start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT ] ) = 0 for all N𝑁Nitalic_N and the estimator converges to the target state in probability [ψ^(N)]p.[ϕ][\hat{\psi}^{(N)}]\overset{\mathrm{p.}}{\longrightarrow}[\phi][ over^ start_ARG italic_ψ end_ARG start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT ] start_OVERACCENT roman_p . end_OVERACCENT start_ARG ⟶ end_ARG [ italic_ϕ ].

Proof.

In the infinite-sample limit, he NLL is given by

limNNLL([ψ])=1|𝒰|(U,b)𝒰×{0,1}n|b|U|ϕ|2log(|b|U|ψ|2/|𝒰|),subscript𝑁subscriptNLLdelimited-[]𝜓1𝒰subscript𝑈𝑏𝒰superscript01𝑛superscriptdelimited-⟨⟩𝑏𝑈italic-ϕ2superscriptdelimited-⟨⟩𝑏𝑈𝜓2𝒰\lim_{N\rightarrow\infty}\mathcal{L}_{\text{NLL}}([\psi])=-\frac{1}{\lvert% \mathcal{U}\rvert}\sum_{(U,b)\in\mathcal{U}\times\{0,1\}^{n}}\lvert\langle b% \lvert U\rvert\phi\rangle\rvert^{2}\log\Big{(}\lvert\langle b\lvert U\rvert% \psi\rangle\rvert^{2}/\lvert\mathcal{U}\rvert\Big{)},roman_lim start_POSTSUBSCRIPT italic_N → ∞ end_POSTSUBSCRIPT caligraphic_L start_POSTSUBSCRIPT NLL end_POSTSUBSCRIPT ( [ italic_ψ ] ) = - divide start_ARG 1 end_ARG start_ARG | caligraphic_U | end_ARG ∑ start_POSTSUBSCRIPT ( italic_U , italic_b ) ∈ caligraphic_U × { 0 , 1 } start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_POSTSUBSCRIPT | ⟨ italic_b | italic_U | italic_ϕ ⟩ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_log ( | ⟨ italic_b | italic_U | italic_ψ ⟩ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / | caligraphic_U | ) , (56)

where |ϕketitalic-ϕ|\phi\rangle| italic_ϕ ⟩ and |ψket𝜓|\psi\rangle| italic_ψ ⟩ are normalized representations of their rays. The NLL loss function achieves the minimum denoted by [ψ^]delimited-[]^𝜓[\hat{\psi}][ over^ start_ARG italic_ψ end_ARG ] 666The minimum is given by the Shannon entropy of the target distribution 𝒟𝒟\mathcal{D}caligraphic_D, calculated as (U,b)𝒟[(U,b),[ϕ]]log𝒟[(U,b),[ϕ]]subscript𝑈𝑏𝒟𝑈𝑏delimited-[]italic-ϕ𝒟𝑈𝑏delimited-[]italic-ϕ-\sum_{(U,b)}\mathcal{D}[(U,b),[\phi]]\log\mathcal{D}[(U,b),[\phi]]- ∑ start_POSTSUBSCRIPT ( italic_U , italic_b ) end_POSTSUBSCRIPT caligraphic_D [ ( italic_U , italic_b ) , [ italic_ϕ ] ] roman_log caligraphic_D [ ( italic_U , italic_b ) , [ italic_ϕ ] ]. This follows from Gibbs’ inequality, which, for two discrete probability distributions {p1,,pN}subscript𝑝1subscript𝑝𝑁\{p_{1},\cdots,p_{N}\}{ italic_p start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , ⋯ , italic_p start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT } and {q1,,qN}subscript𝑞1subscript𝑞𝑁\{q_{1},\cdots,q_{N}\}{ italic_q start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , ⋯ , italic_q start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT }, states that i=1Npilogpii=1Npilogqisuperscriptsubscript𝑖1𝑁subscript𝑝𝑖subscript𝑝𝑖superscriptsubscript𝑖1𝑁subscript𝑝𝑖subscript𝑞𝑖-\sum_{i=1}^{N}p_{i}\log p_{i}\leq-\sum_{i=1}^{N}p_{i}\log q_{i}- ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT italic_p start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT roman_log italic_p start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ≤ - ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT italic_p start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT roman_log italic_q start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. The equality is achieved when two distributions are the same pi=qisubscript𝑝𝑖subscript𝑞𝑖p_{i}=q_{i}italic_p start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_q start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. if the probabilities agree for all measurement outcomes

|b|U|ψ^|2=|b|U|ϕ|2,(b,U).superscriptdelimited-⟨⟩𝑏𝑈^𝜓2superscriptdelimited-⟨⟩𝑏𝑈italic-ϕ2for-all𝑏𝑈\lvert\langle b\lvert U\rvert\hat{\psi}\rangle\rvert^{2}=\lvert\langle b\lvert U% \rvert\phi\rangle\rvert^{2},\quad\forall(b,U).| ⟨ italic_b | italic_U | over^ start_ARG italic_ψ end_ARG ⟩ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = | ⟨ italic_b | italic_U | italic_ϕ ⟩ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , ∀ ( italic_b , italic_U ) . (57)

Since the unitary ensemble 𝒰𝒰\mathcal{U}caligraphic_U is tomographically complete, Eq. (57) implies |ψ^=c|ϕket^𝜓𝑐ketitalic-ϕ\mbox{$|\hat{\psi}\rangle$}=c\mbox{$|\phi\rangle$}| over^ start_ARG italic_ψ end_ARG ⟩ = italic_c | italic_ϕ ⟩ up to a global phase c𝑐citalic_c, so that [ψ^]=[ϕ]delimited-[]^𝜓delimited-[]italic-ϕ[\hat{\psi}]=[\phi][ over^ start_ARG italic_ψ end_ARG ] = [ italic_ϕ ]. ∎

Assumption 1.

The expected Hessian at the target state [ϕ]delimited-[]italic-ϕ[\phi][ italic_ϕ ] is invertible and finite.

Assumption 2.

The Hessian at [ψ^(N)]delimited-[]superscript^𝜓𝑁[\hat{\psi}^{(N)}][ over^ start_ARG italic_ψ end_ARG start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT ] in a small neighbourhood near the target converges uniformly to the Hessian at [ϕ]delimited-[]italic-ϕ[\phi][ italic_ϕ ] in expectation with respect to 𝒟𝒟\mathcal{D}caligraphic_D.

Having established or assumed the conditions outlined in Lemma 3, we are now prepared to formally apply the normality result to the MPS-manifold setting. This application is detailed in Corollary 2, which extends the lemma to our specific context.

Corollary 2 (Asymptotic normality on the projective MPS manifold).

Let 𝒟𝒟\mathcal{D}caligraphic_D be the distribution corresponding to the target state [ϕ]delimited-[]italic-ϕ[\phi][ italic_ϕ ] (defined in Eq. (52)), and D(N)={(Ui,bi)}i=1Nsuperscript𝐷𝑁superscriptsubscriptsubscript𝑈𝑖subscript𝑏𝑖𝑖1𝑁{D}^{(N)}=\{(U_{i},b_{i})\}_{i=1}^{N}italic_D start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT = { ( italic_U start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) } start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT be a sequence of i.i.d. samples from 𝒟𝒟\mathcal{D}caligraphic_D. Let [ψ^(N)]delimited-[]superscript^𝜓𝑁[\hat{\psi}^{(N)}][ over^ start_ARG italic_ψ end_ARG start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT ] be the minimizer of the negative log-likelihood function and ω𝜔\omegaitalic_ω its assoicated estimating form as defined in Corollary 1. Define γ𝛾\gammaitalic_γ as the shortest connecting geodesic such that γ(0)=[ϕ]𝛾0delimited-[]italic-ϕ\gamma(0)=[\phi]italic_γ ( 0 ) = [ italic_ϕ ], a point γ(1)=[ψ^(N)]𝛾1delimited-[]superscript^𝜓𝑁\gamma(1)=[\hat{\psi}^{(N)}]italic_γ ( 1 ) = [ over^ start_ARG italic_ψ end_ARG start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT ], and γ˙˙𝛾\dot{\gamma}over˙ start_ARG italic_γ end_ARG as the velocity vector along the curve. Let the velocity vector be decomposed in the parallel transporte—along γ𝛾\gammaitalic_γ—basis vectors {ea(t)}subscript𝑒𝑎𝑡\{e_{a}(t)\}{ italic_e start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ( italic_t ) } as γ˙=aΔaea(t)˙𝛾subscript𝑎subscriptΔ𝑎subscript𝑒𝑎𝑡\dot{\gamma}=\sum_{a}\Delta_{a}e_{a}(t)over˙ start_ARG italic_γ end_ARG = ∑ start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT roman_Δ start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ( italic_t ). Suppose the conditions of Proposition 1, Assumption 1, and Assumption 2 are satisfied. By Lemma 3, the velocity vector’s components {Δa}a=1rsuperscriptsubscriptsubscriptΔ𝑎𝑎1𝑟\{\Delta_{a}\}_{a=1}^{r}{ roman_Δ start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT } start_POSTSUBSCRIPT italic_a = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT converge in distribution to a multivariate normal distribution

N(Δ1,,Δr)d.𝒩(0,Σ),\sqrt{N}(\Delta_{1},\ldots,\Delta_{r})^{\top}\overset{\mathrm{d.}}{% \longrightarrow}\mathcal{N}(0,\Sigma),square-root start_ARG italic_N end_ARG ( roman_Δ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , roman_Δ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT start_OVERACCENT roman_d . end_OVERACCENT start_ARG ⟶ end_ARG caligraphic_N ( 0 , roman_Σ ) , (58)

where Σ=(H)1ΓHΣsuperscriptsuperscriptH1ΓH\Sigma=(\mathrm{H}^{\dagger})^{-1}\Gamma\mathrm{H}roman_Σ = ( roman_H start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_Γ roman_H and the Gram matrix Γa,b=𝔼(U,b)𝒟[ω((U,b),[ϕ])|eaω((U,b),[ϕ])|eb]subscriptΓ𝑎𝑏subscript𝔼similar-to𝑈𝑏𝒟delimited-[]ω((U,b),[ϕ])|eaω((U,b),[ϕ])|eb\Gamma_{a,b}=\mathbb{E}_{(U,b)\sim\mathcal{D}}\big{[}\mbox{$\langle\omega((U,b% ),[\phi])|e_{a}\rangle$}\mbox{$\langle\omega((U,b),[\phi])|e_{b}\rangle$}\big{]}roman_Γ start_POSTSUBSCRIPT italic_a , italic_b end_POSTSUBSCRIPT = blackboard_E start_POSTSUBSCRIPT ( italic_U , italic_b ) ∼ caligraphic_D end_POSTSUBSCRIPT [ italic_⟨ω((U,b),[ϕ])|ea⟩ italic_⟨ω((U,b),[ϕ])|eb⟩ ].

Corollary 2 establishes the asymptotic normality of the velocity vector’s components along the shortest connecting geodesic γ𝛾\gammaitalic_γ in the projective MPS manifold M~~𝑀\tilde{M}over~ start_ARG italic_M end_ARG, which is crucial for determining the geodesic length (see Fact 1 and Eq. (29)). Since M~~𝑀\tilde{M}over~ start_ARG italic_M end_ARG is a submanifold of the projective Hilbert space 𝖯()𝖯\mathsf{P}(\mathcal{H})sansserif_P ( caligraphic_H ) with an induced metric g~~𝑔\tilde{g}over~ start_ARG italic_g end_ARG, lifting γ:[0,1]M~:𝛾01~𝑀\gamma:[0,1]\rightarrow\tilde{M}italic_γ : [ 0 , 1 ] → over~ start_ARG italic_M end_ARG to this embedding space to γ:[0,1]𝖯():superscript𝛾01𝖯\gamma^{\prime}:[0,1]\rightarrow\mathsf{P}(\mathcal{H})italic_γ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT : [ 0 , 1 ] → sansserif_P ( caligraphic_H ) reveals that its length is always longer than or equal to that of a shortest connecting geodesic in 𝖯()𝖯\mathsf{P}(\mathcal{H})sansserif_P ( caligraphic_H ); equality occurs only when the geodesic curve in 𝖯()𝖯\mathsf{P}(\mathcal{H})sansserif_P ( caligraphic_H ) is exactly γsuperscript𝛾\gamma^{\prime}italic_γ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT. As the geodesic curve in 𝖯()𝖯\mathsf{P}(\mathcal{H})sansserif_P ( caligraphic_H ) defines the Fubini-Study distance, the length of the geodesic thus relates to the state overlap using Lemma 1. This relation allows us to upper bound the infidelity between quantum states by the length of γ𝛾\gammaitalic_γ within M~~𝑀\tilde{M}over~ start_ARG italic_M end_ARG. Using the properties and asymptotic behavior of these geodesic paths, we provide a bounded estimation of the state infidelity in Lemma 4 below.

Lemma 4 (Infidelity bounded by geodesic curve length).

Given two states [ψ],[ϕ]delimited-[]𝜓delimited-[]italic-ϕ[\psi],[\phi][ italic_ψ ] , [ italic_ϕ ] in M~~𝑀\tilde{M}over~ start_ARG italic_M end_ARG, the projective space of MPS (see Lemma 2), denote the shortest connecting geodesic by γ:[0,1]M~:𝛾01~𝑀\gamma:[0,1]\rightarrow\tilde{M}italic_γ : [ 0 , 1 ] → over~ start_ARG italic_M end_ARG, such that γ(0)=[ψ]𝛾0delimited-[]𝜓\gamma(0)=[\psi]italic_γ ( 0 ) = [ italic_ψ ] and γ(1)=[ϕ]𝛾1delimited-[]italic-ϕ\gamma(1)=[\phi]italic_γ ( 1 ) = [ italic_ϕ ], as defined in Def. 8. The infidelity of [ψ]delimited-[]𝜓[\psi][ italic_ψ ] and [ϕ]delimited-[]italic-ϕ[\phi][ italic_ϕ ] is bounded by the length of the geodesic:

1|ψ|ϕ|01dtγ˙|γ˙g~,1inner-product𝜓italic-ϕsuperscriptsubscript01differential-d𝑡subscriptinner-product˙𝛾˙𝛾~𝑔1-|\mbox{$\langle\psi|\phi\rangle$}|\leq\int_{0}^{1}\mathrm{d}t\sqrt{\mbox{$% \langle\dot{\gamma}|\dot{\gamma}\rangle$}_{\tilde{g}}},1 - | ⟨ italic_ψ | italic_ϕ ⟩ | ≤ ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT roman_d italic_t square-root start_ARG ⟨ over˙ start_ARG italic_γ end_ARG | over˙ start_ARG italic_γ end_ARG ⟩ start_POSTSUBSCRIPT over~ start_ARG italic_g end_ARG end_POSTSUBSCRIPT end_ARG , (59)

where |g~\mbox{$\langle\bullet|\bullet\rangle$}_{\tilde{g}}⟨ ∙ | ∙ ⟩ start_POSTSUBSCRIPT over~ start_ARG italic_g end_ARG end_POSTSUBSCRIPT denotes the induced Fubini-Study metric on M~~𝑀\tilde{M}over~ start_ARG italic_M end_ARG.

Proof.

By Lemma 1, the infidelity between two state vectors |ψket𝜓|\psi\rangle| italic_ψ ⟩ and |ϕketitalic-ϕ|\phi\rangle| italic_ϕ ⟩ is upper bounded by their Fubini-Study distance in the projective Hilbert space as

1|ψ|ϕ|1inner-product𝜓italic-ϕ\displaystyle 1-|\mbox{$\langle\psi|\phi\rangle$}|1 - | ⟨ italic_ψ | italic_ϕ ⟩ | =1cosdFS([ϕ],[ψ]),absent1subscript𝑑FSdelimited-[]italic-ϕdelimited-[]𝜓\displaystyle=1-\cos d_{\mathrm{FS}}([\phi],[\psi]),= 1 - roman_cos italic_d start_POSTSUBSCRIPT roman_FS end_POSTSUBSCRIPT ( [ italic_ϕ ] , [ italic_ψ ] ) , (60)
dFS([ϕ],[ψ]).absentsubscript𝑑FSdelimited-[]italic-ϕdelimited-[]𝜓\displaystyle\leq d_{\mathrm{FS}}([\phi],[\psi]).≤ italic_d start_POSTSUBSCRIPT roman_FS end_POSTSUBSCRIPT ( [ italic_ϕ ] , [ italic_ψ ] ) .

Recall that the Fubini-Study distance is defined as the length of the shortest curve between two points in projective Hilbert space (see Def. 12). Hence, for any differentiable curve α:[0,1]𝖯():𝛼01𝖯\alpha:[0,1]\rightarrow\mathsf{P}(\mathcal{H})italic_α : [ 0 , 1 ] → sansserif_P ( caligraphic_H ), with α(0)=[ψ]𝛼0delimited-[]𝜓\alpha(0)=[\psi]italic_α ( 0 ) = [ italic_ψ ] and α(1)=[ϕ]𝛼1delimited-[]italic-ϕ\alpha(1)=[\phi]italic_α ( 1 ) = [ italic_ϕ ], the length of α𝛼\alphaitalic_α also upper bounds the infidelity as

1|ψ|ϕ|01dtα˙|α˙g,1inner-product𝜓italic-ϕsuperscriptsubscript01differential-d𝑡subscriptinner-product˙𝛼˙𝛼𝑔1-|\mbox{$\langle\psi|\phi\rangle$}|\leq\int_{0}^{1}\mathrm{d}t\sqrt{\mbox{$% \langle\dot{\alpha}|\dot{\alpha}\rangle$}_{g}},1 - | ⟨ italic_ψ | italic_ϕ ⟩ | ≤ ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT roman_d italic_t square-root start_ARG ⟨ over˙ start_ARG italic_α end_ARG | over˙ start_ARG italic_α end_ARG ⟩ start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT end_ARG , (61)

where |g\mbox{$\langle\bullet|\bullet\rangle$}_{g}⟨ ∙ | ∙ ⟩ start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT denotes the Fubini-Study metric on 𝖯()𝖯\mathsf{P}(\mathcal{H})sansserif_P ( caligraphic_H ). Note that the choice of the parameter interval [0,1]01[0,1][ 0 , 1 ] is without loss of generality because we can reparametrize any connecting curve to an arbitrary closed interval of our choice. Since M~~𝑀\tilde{M}over~ start_ARG italic_M end_ARG is a submanifold of 𝖯()𝖯\mathsf{P}(\mathcal{H})sansserif_P ( caligraphic_H ), we can lift the shortest connecting geodesic γ𝛾\gammaitalic_γ from the projective MPS manifold M~~𝑀\tilde{M}over~ start_ARG italic_M end_ARG into projective Hilbert space 𝖯()𝖯\mathsf{P}(\mathcal{H})sansserif_P ( caligraphic_H ). To be precise, we can formally distinguish the curve γ𝛾\gammaitalic_γ in the two abstract manifolds M~~𝑀\tilde{M}over~ start_ARG italic_M end_ARG and 𝖯()𝖯\mathsf{P}(\mathcal{H})sansserif_P ( caligraphic_H ). For this, we formally identify each point γ(t)M~𝛾𝑡~𝑀\gamma(t)\in\tilde{M}italic_γ ( italic_t ) ∈ over~ start_ARG italic_M end_ARG with its corresponding point in γ(t)𝖯()superscript𝛾𝑡𝖯\gamma^{\prime}(t)\in\mathsf{P}(\mathcal{H})italic_γ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t ) ∈ sansserif_P ( caligraphic_H ). Since g~~𝑔\tilde{g}over~ start_ARG italic_g end_ARG is induced by g𝑔gitalic_g, the lengths of any tangent vectors are the same in both the submanifold and the embedding manifold γ˙|γ˙g~=γ˙|γ˙gsubscriptinner-product˙𝛾˙𝛾~𝑔subscriptinner-productsuperscript˙𝛾superscript˙𝛾𝑔\mbox{$\langle\dot{\gamma}|\dot{\gamma}\rangle$}_{\tilde{g}}=\mbox{$\langle% \dot{\gamma}^{\prime}|\dot{\gamma}^{\prime}\rangle$}_{g}⟨ over˙ start_ARG italic_γ end_ARG | over˙ start_ARG italic_γ end_ARG ⟩ start_POSTSUBSCRIPT over~ start_ARG italic_g end_ARG end_POSTSUBSCRIPT = ⟨ over˙ start_ARG italic_γ end_ARG start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT | over˙ start_ARG italic_γ end_ARG start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT (see Eqs. (103) and (108) in Ref. [37]). Therefore, the lengths of γ𝛾\gammaitalic_γ and γsuperscript𝛾\gamma^{\prime}italic_γ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT are equal. Choosing the curve to be α=γ𝛼superscript𝛾\alpha=\gamma^{\prime}italic_α = italic_γ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT in Eq. (61), we have proved the result stated in Eq. (59). ∎

Finally, using Corollary 2 and Lemma 4, we provide an upper bound on the infidelity, which is our main result below.

Theorem 2 (Probabilistic bound).

Let M~~𝑀\tilde{M}over~ start_ARG italic_M end_ARG be the manifold of full-rank MPS of a maximum bond dimensions χmaxsubscript𝜒\chi_{\max}italic_χ start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT, as per Eq. (40), and let [ϕ]M~delimited-[]italic-ϕ~𝑀[\phi]\in\tilde{M}[ italic_ϕ ] ∈ over~ start_ARG italic_M end_ARG be some target state. Let D(N)={(Ui,bi)}i=1Nsuperscript𝐷𝑁superscriptsubscriptsubscript𝑈𝑖subscript𝑏𝑖𝑖1𝑁D^{(N)}=\{(U_{i},\,b_{i})\}_{i=1}^{N}italic_D start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT = { ( italic_U start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) } start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT be the random variable representing a dataset of N𝑁Nitalic_N samples drawn from the distribution 𝒟[(U,b),[ϕ]]𝒟𝑈𝑏delimited-[]italic-ϕ\mathcal{D}[(U,b),[\phi]]caligraphic_D [ ( italic_U , italic_b ) , [ italic_ϕ ] ] defined in Eq. (52). Denote by [ψ^(N)]delimited-[]superscript^𝜓𝑁[\hat{\psi}^{(N)}][ over^ start_ARG italic_ψ end_ARG start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT ] the random variable representing the minimizer, in M~~𝑀\tilde{M}over~ start_ARG italic_M end_ARG, of the N𝑁Nitalic_N-sample negative log-likelihood defined by D(N)superscript𝐷𝑁D^{(N)}italic_D start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT, and let ω𝜔\omegaitalic_ω be its associated estimating form as defined in Corollary 1. Assume the conditions of Proposition 1, Assumption 1, and Assumption 2 are satisfied. Then, for any δ(0,1]𝛿01\delta\in(0,1]italic_δ ∈ ( 0 , 1 ], there exists N0subscript𝑁0N_{0}italic_N start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT such that for all N>N0𝑁subscript𝑁0N>N_{0}italic_N > italic_N start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, one has

[1|ψ^(N)|ϕ|ϵ(N)]1δ,delimited-[]1inner-productsuperscript^𝜓𝑁italic-ϕitalic-ϵ𝑁1𝛿\mathbb{P}\left[1-|\langle\hat{\psi}^{(N)}|\phi\rangle|\leq\epsilon(N)\right]% \geq 1-\delta,blackboard_P [ 1 - | ⟨ over^ start_ARG italic_ψ end_ARG start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT | italic_ϕ ⟩ | ≤ italic_ϵ ( italic_N ) ] ≥ 1 - italic_δ , (62)

where

ϵ(N)=O(nχmax2Nδ).italic-ϵ𝑁𝑂𝑛superscriptsubscript𝜒2𝑁𝛿\epsilon(N)=O\left(\sqrt{\frac{n\chi_{\max}^{2}}{N\delta}}\right).italic_ϵ ( italic_N ) = italic_O ( square-root start_ARG divide start_ARG italic_n italic_χ start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_N italic_δ end_ARG end_ARG ) . (63)
Proof.

For simplicity, we assume that |ψ^(N)ketsuperscript^𝜓𝑁|\hat{\psi}^{(N)}\rangle| over^ start_ARG italic_ψ end_ARG start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT ⟩ and |ϕketitalic-ϕ|\phi\rangle| italic_ϕ ⟩ are normalized elements of the equivalence classes [ψ^(N)]delimited-[]superscript^𝜓𝑁[\hat{\psi}^{(N)}][ over^ start_ARG italic_ψ end_ARG start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT ] and [ϕ]delimited-[]italic-ϕ[\phi][ italic_ϕ ], respectively. Let g~~𝑔\tilde{g}over~ start_ARG italic_g end_ARG be the Fubini-Study metric of the projective MPS manifold M~~𝑀\tilde{M}over~ start_ARG italic_M end_ARG induced by the Fubini-Study metric of the projective Hilbert space via Eq. (41). On this manifold, γ𝛾\gammaitalic_γ is the shortest connecting geodesic between γ(0)=[ϕ]𝛾0delimited-[]italic-ϕ\gamma(0)=[\phi]italic_γ ( 0 ) = [ italic_ϕ ] and γ(1)=[ψ^(N)]𝛾1delimited-[]superscript^𝜓𝑁\gamma(1)=[\hat{\psi}^{(N)}]italic_γ ( 1 ) = [ over^ start_ARG italic_ψ end_ARG start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT ]. By Proposition. 1, the estimator converges to the target state in the infinite-sample limit

[ψ^(N)]p.[ϕ],[\hat{\psi}^{(N)}]\overset{\mathrm{p.}}{\longrightarrow}[\phi],[ over^ start_ARG italic_ψ end_ARG start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT ] start_OVERACCENT roman_p . end_OVERACCENT start_ARG ⟶ end_ARG [ italic_ϕ ] , (64)

as N𝑁N\rightarrow\inftyitalic_N → ∞. To quantify the error between the N𝑁Nitalic_N-sample estimator and the target state, we use Lemma. 4 to upper bound the N𝑁Nitalic_N-sample infidelity by the length of the geodesic curve. More specifically, according to Fact 1, this geodesic length is the same as the length of the velocity vector, which is constant throughout the curve, so that

1|ψ^(N)|ϕ|γ˙|γ˙g~.1inner-productsuperscript^𝜓𝑁italic-ϕsubscriptinner-product˙𝛾˙𝛾~𝑔1-|\langle\hat{\psi}^{(N)}|\phi\rangle|\leq\sqrt{\mbox{$\langle\dot{\gamma}|% \dot{\gamma}\rangle$}_{\tilde{g}}}.1 - | ⟨ over^ start_ARG italic_ψ end_ARG start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT | italic_ϕ ⟩ | ≤ square-root start_ARG ⟨ over˙ start_ARG italic_γ end_ARG | over˙ start_ARG italic_γ end_ARG ⟩ start_POSTSUBSCRIPT over~ start_ARG italic_g end_ARG end_POSTSUBSCRIPT end_ARG . (65)

Therefore, to obtain an upper bound for the infidelity, it suffices to analyze the velocity vector defined at an arbitrary point along the curve, which we choose to be γ(0)=[ϕ]𝛾0delimited-[]italic-ϕ\gamma(0)=[\phi]italic_γ ( 0 ) = [ italic_ϕ ]. The velocity vector, in the basis {ea}a=1rsuperscriptsubscriptsubscript𝑒𝑎𝑎1𝑟\{e_{a}\}_{a=1}^{r}{ italic_e start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT } start_POSTSUBSCRIPT italic_a = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT of the tangent space T[ϕ]M~subscript𝑇delimited-[]italic-ϕ~𝑀T_{[\phi]}\tilde{M}italic_T start_POSTSUBSCRIPT [ italic_ϕ ] end_POSTSUBSCRIPT over~ start_ARG italic_M end_ARG, can be written as

γ˙=a=1rΔa(N)ea.˙𝛾superscriptsubscript𝑎1𝑟subscriptsuperscriptΔ𝑁𝑎subscript𝑒𝑎\dot{\gamma}=\sum_{a=1}^{r}{\Delta}^{(N)}_{a}e_{a}.over˙ start_ARG italic_γ end_ARG = ∑ start_POSTSUBSCRIPT italic_a = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT roman_Δ start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT . (66)

Expressing the Fubini-Study metric in the same basis as g~a,bsubscript~𝑔𝑎𝑏\tilde{g}_{a,b}over~ start_ARG italic_g end_ARG start_POSTSUBSCRIPT italic_a , italic_b end_POSTSUBSCRIPT, the squared length of the velocity vector is explicitly given by

γ˙|γ˙g~=a,bΔa(N)g~a,bΔb(N).subscriptinner-product˙𝛾˙𝛾~𝑔subscript𝑎𝑏subscriptsuperscriptΔ𝑁𝑎subscript~𝑔𝑎𝑏subscriptsuperscriptΔ𝑁𝑏{\mbox{$\langle\dot{\gamma}|\dot{\gamma}\rangle$}_{\tilde{g}}}={\sum_{a,b}{% \Delta}^{(N)}_{a}\,\tilde{g}_{a,b}\,{\Delta}^{(N)}_{b}}.⟨ over˙ start_ARG italic_γ end_ARG | over˙ start_ARG italic_γ end_ARG ⟩ start_POSTSUBSCRIPT over~ start_ARG italic_g end_ARG end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_a , italic_b end_POSTSUBSCRIPT roman_Δ start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT over~ start_ARG italic_g end_ARG start_POSTSUBSCRIPT italic_a , italic_b end_POSTSUBSCRIPT roman_Δ start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT . (67)

For simplicity, we represent the velocity vector’s components as a column vector Δ(N)=(Δ1(N),,Δr(N))superscriptΔ𝑁superscriptsubscriptsuperscriptΔ𝑁1subscriptsuperscriptΔ𝑁𝑟top{\Delta}^{(N)}=({\Delta}^{(N)}_{1},\ldots,{\Delta}^{(N)}_{r})^{\top}roman_Δ start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT = ( roman_Δ start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , roman_Δ start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT, so that its squared length is

γ˙|γ˙g~=Δ(N)g~Δ(N),subscriptinner-product˙𝛾˙𝛾~𝑔superscriptsuperscriptΔ𝑁top~𝑔superscriptΔ𝑁{\mbox{$\langle\dot{\gamma}|\dot{\gamma}\rangle$}_{\tilde{g}}}={{{\Delta}^{(N)% }}^{\top}\tilde{g}\,{\Delta}^{(N)}},⟨ over˙ start_ARG italic_γ end_ARG | over˙ start_ARG italic_γ end_ARG ⟩ start_POSTSUBSCRIPT over~ start_ARG italic_g end_ARG end_POSTSUBSCRIPT = roman_Δ start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT over~ start_ARG italic_g end_ARG roman_Δ start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT , (68)

where, in slight abuse of notation, we write g~=(g~a,b)a,b=1r~𝑔superscriptsubscriptsubscript~𝑔𝑎𝑏𝑎𝑏1𝑟\tilde{g}=(\tilde{g}_{a,b})_{a,b=1}^{r}over~ start_ARG italic_g end_ARG = ( over~ start_ARG italic_g end_ARG start_POSTSUBSCRIPT italic_a , italic_b end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT italic_a , italic_b = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT. By virtue of Assumptions 1 and 2, together with the consistency of the estimator in Eq. (64), Corollary 2 states that the N𝑁Nitalic_N-sample estimator NΔ(N)𝑁superscriptΔ𝑁\sqrt{N}{\Delta}^{(N)}square-root start_ARG italic_N end_ARG roman_Δ start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT converges in distribution to the normal distribution as

NΔ(N)dΔ()𝒩(0,Σ).similar-to𝑁superscriptΔ𝑁dsuperscriptΔ𝒩0Σ\sqrt{N}\Delta^{(N)}\overset{\mathrm{d}}{\rightarrow}\Delta^{(\infty)}\sim% \mathcal{N}(0,\Sigma).square-root start_ARG italic_N end_ARG roman_Δ start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT overroman_d start_ARG → end_ARG roman_Δ start_POSTSUPERSCRIPT ( ∞ ) end_POSTSUPERSCRIPT ∼ caligraphic_N ( 0 , roman_Σ ) . (69)

Therefore, in the infinite-sample limit N𝑁N\rightarrow\inftyitalic_N → ∞, we can upper bound the probability of the norm of Δ()superscriptΔ\Delta^{(\infty)}roman_Δ start_POSTSUPERSCRIPT ( ∞ ) end_POSTSUPERSCRIPT (with respect to the metric g~~𝑔\tilde{g}over~ start_ARG italic_g end_ARG) greater than an error ε𝜀\varepsilonitalic_ε using Markov’s inequality as

[Δ()g~Δ()ε]𝔼(Δ()g~Δ())ε.delimited-[]superscriptsuperscriptΔtop~𝑔superscriptΔ𝜀𝔼superscriptsuperscriptΔtop~𝑔superscriptΔ𝜀\displaystyle\mathbb{P}\left[{\Delta^{(\infty)}}^{\top}\tilde{g}\,\Delta^{(% \infty)}\geq\varepsilon\right]\leq\frac{\mathbb{E}\left({\Delta^{(\infty)}}^{% \top}\tilde{g}\,\Delta^{(\infty)}\right)}{\varepsilon}.blackboard_P [ roman_Δ start_POSTSUPERSCRIPT ( ∞ ) end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT over~ start_ARG italic_g end_ARG roman_Δ start_POSTSUPERSCRIPT ( ∞ ) end_POSTSUPERSCRIPT ≥ italic_ε ] ≤ divide start_ARG blackboard_E ( roman_Δ start_POSTSUPERSCRIPT ( ∞ ) end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT over~ start_ARG italic_g end_ARG roman_Δ start_POSTSUPERSCRIPT ( ∞ ) end_POSTSUPERSCRIPT ) end_ARG start_ARG italic_ε end_ARG . (70)

To compute the expectation of the Δ()superscriptΔ\Delta^{(\infty)}roman_Δ start_POSTSUPERSCRIPT ( ∞ ) end_POSTSUPERSCRIPT-inner product on the r.h.s., we use the following identity

𝔼(Δ()g~Δ())𝔼superscriptsuperscriptΔtop~𝑔superscriptΔ\displaystyle\mathbb{E}\left({{\Delta^{(\infty)}}^{\top}\tilde{g}\,\Delta^{(% \infty)}}\right)blackboard_E ( roman_Δ start_POSTSUPERSCRIPT ( ∞ ) end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT over~ start_ARG italic_g end_ARG roman_Δ start_POSTSUPERSCRIPT ( ∞ ) end_POSTSUPERSCRIPT ) =𝔼(Tr(g~Δ()Δ())),absent𝔼Tr~𝑔superscriptΔsuperscriptsuperscriptΔtop\displaystyle=\mathbb{E}\left({\mathop{\mathrm{Tr}}(\tilde{g}\,\Delta^{(\infty% )}{\Delta^{(\infty)}}^{\top})}\right),= blackboard_E ( roman_Tr ( over~ start_ARG italic_g end_ARG roman_Δ start_POSTSUPERSCRIPT ( ∞ ) end_POSTSUPERSCRIPT roman_Δ start_POSTSUPERSCRIPT ( ∞ ) end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT ) ) , (71)
=Tr(g~𝔼(Δ()Δ())),absentTr~𝑔𝔼superscriptΔsuperscriptsuperscriptΔtop\displaystyle=\mathop{\mathrm{Tr}}\Big{(}\tilde{g}\,\mathbb{E}\left({\Delta^{(% \infty)}{\Delta^{(\infty)}}^{\top}}\right)\Big{)},= roman_Tr ( over~ start_ARG italic_g end_ARG blackboard_E ( roman_Δ start_POSTSUPERSCRIPT ( ∞ ) end_POSTSUPERSCRIPT roman_Δ start_POSTSUPERSCRIPT ( ∞ ) end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT ) ) ,
=Tr(g~Σ).absentTr~𝑔Σ\displaystyle=\mathop{\mathrm{Tr}}\left(\tilde{g}\,\Sigma\right).= roman_Tr ( over~ start_ARG italic_g end_ARG roman_Σ ) .

Going from the second to the third line, first we have used the definition of the covariance matrix

Σ=𝔼(Δ()Δ())𝔼(Δ())𝔼(Δ()).Σ𝔼superscriptΔsuperscriptsuperscriptΔtop𝔼superscriptΔ𝔼superscriptsuperscriptΔtop\Sigma=\mathbb{E}\left({\Delta^{(\infty)}}{\Delta^{(\infty)}}^{\top}\right)-% \mathbb{E}\left({\Delta^{(\infty)}}\right)\mathbb{E}\left({\Delta^{(\infty)}}^% {\top}\right).roman_Σ = blackboard_E ( roman_Δ start_POSTSUPERSCRIPT ( ∞ ) end_POSTSUPERSCRIPT roman_Δ start_POSTSUPERSCRIPT ( ∞ ) end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT ) - blackboard_E ( roman_Δ start_POSTSUPERSCRIPT ( ∞ ) end_POSTSUPERSCRIPT ) blackboard_E ( roman_Δ start_POSTSUPERSCRIPT ( ∞ ) end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT ) . (72)

Second, using the zero-mean property of 𝔼(Δ())=0𝔼superscriptΔ0\mathbb{E}\left({\Delta^{(\infty)}}\right)=0blackboard_E ( roman_Δ start_POSTSUPERSCRIPT ( ∞ ) end_POSTSUPERSCRIPT ) = 0, we arrive at an explicit expression for the expectation of the Δ()superscriptΔ\Delta^{(\infty)}roman_Δ start_POSTSUPERSCRIPT ( ∞ ) end_POSTSUPERSCRIPT-inner product as

𝔼(Δ()g~Δ())=Tr(g~Σ).𝔼superscriptsuperscriptΔtop~𝑔superscriptΔTr~𝑔Σ\displaystyle\mathbb{E}\left({{\Delta^{(\infty)}}^{\top}\tilde{g}\,\Delta^{(% \infty)}}\right)=\mathop{\mathrm{Tr}}\left(\tilde{g}\,\Sigma\right).blackboard_E ( roman_Δ start_POSTSUPERSCRIPT ( ∞ ) end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT over~ start_ARG italic_g end_ARG roman_Δ start_POSTSUPERSCRIPT ( ∞ ) end_POSTSUPERSCRIPT ) = roman_Tr ( over~ start_ARG italic_g end_ARG roman_Σ ) . (73)

Plugging the expectation values derived in Eq. (73) into the r.h.s. of Eq. (70), we have the following bound

[Δ()g~Δ()ε]Tr(g~Σ)ε.delimited-[]superscriptsuperscriptΔtop~𝑔superscriptΔ𝜀Tr~𝑔Σ𝜀\mathbb{P}\left[{\Delta^{(\infty)}}^{\top}\tilde{g}\,\Delta^{(\infty)}\geq% \varepsilon\right]\leq\frac{\mathop{\mathrm{Tr}}\left(\tilde{g}\,\Sigma\right)% }{\varepsilon}.blackboard_P [ roman_Δ start_POSTSUPERSCRIPT ( ∞ ) end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT over~ start_ARG italic_g end_ARG roman_Δ start_POSTSUPERSCRIPT ( ∞ ) end_POSTSUPERSCRIPT ≥ italic_ε ] ≤ divide start_ARG roman_Tr ( over~ start_ARG italic_g end_ARG roman_Σ ) end_ARG start_ARG italic_ε end_ARG . (74)

To analyze the finite-N𝑁Nitalic_N Δ(N)superscriptΔ𝑁{\Delta}^{(N)}roman_Δ start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT-inner product estimator, we first notice that the estimator converges in probability to its infinite-N𝑁Nitalic_N limit as

[NΔ(N)g~Δ(N)ε]N[Δ()g~Δ()ε].delimited-[]𝑁superscriptsuperscriptΔ𝑁top~𝑔superscriptΔ𝑁𝜀𝑁delimited-[]superscriptsuperscriptΔtop~𝑔superscriptΔ𝜀\displaystyle\mathbb{P}\left[N{\Delta^{(N)}}^{\top}\tilde{g}\,\Delta^{(N)}\geq% \varepsilon\right]\overset{N\rightarrow\infty}{\longrightarrow}\mathbb{P}\left% [{\Delta^{(\infty)}}^{\top}\tilde{g}\,\Delta^{(\infty)}\geq\varepsilon\right].blackboard_P [ italic_N roman_Δ start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT over~ start_ARG italic_g end_ARG roman_Δ start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT ≥ italic_ε ] start_OVERACCENT italic_N → ∞ end_OVERACCENT start_ARG ⟶ end_ARG blackboard_P [ roman_Δ start_POSTSUPERSCRIPT ( ∞ ) end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT over~ start_ARG italic_g end_ARG roman_Δ start_POSTSUPERSCRIPT ( ∞ ) end_POSTSUPERSCRIPT ≥ italic_ε ] . (75)

Once we choose a convergence error tolerance η𝜂\etaitalic_η, there exists an N0subscript𝑁0N_{0}italic_N start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT such that the finite-sample estimator is close to the infinite-sample limit upper bounded by η𝜂\etaitalic_η for all N>N0𝑁subscript𝑁0N>N_{0}italic_N > italic_N start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT:

[NΔ(N)g~Δ(N)ε][Δ()g~Δ()ε]η.delimited-[]𝑁superscriptsuperscriptΔ𝑁top~𝑔superscriptΔ𝑁𝜀delimited-[]superscriptsuperscriptΔtop~𝑔superscriptΔ𝜀𝜂\displaystyle\mathbb{P}\left[N{\Delta^{(N)}}^{\top}\tilde{g}\,\Delta^{(N)}\geq% \varepsilon\right]-\mathbb{P}\left[{\Delta^{(\infty)}}^{\top}\tilde{g}\,\Delta% ^{(\infty)}\geq\varepsilon\right]\leq\eta.blackboard_P [ italic_N roman_Δ start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT over~ start_ARG italic_g end_ARG roman_Δ start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT ≥ italic_ε ] - blackboard_P [ roman_Δ start_POSTSUPERSCRIPT ( ∞ ) end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT over~ start_ARG italic_g end_ARG roman_Δ start_POSTSUPERSCRIPT ( ∞ ) end_POSTSUPERSCRIPT ≥ italic_ε ] ≤ italic_η . (76)

Therefore, we can bound the finite-sample estimator by

[NΔ(N)g~Δ(N)ε]delimited-[]𝑁superscriptsuperscriptΔ𝑁top~𝑔superscriptΔ𝑁𝜀\displaystyle\mathbb{P}\left[N{\Delta^{(N)}}^{\top}\tilde{g}\,\Delta^{(N)}\geq% \varepsilon\right]blackboard_P [ italic_N roman_Δ start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT over~ start_ARG italic_g end_ARG roman_Δ start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT ≥ italic_ε ] Tr(g~Σ)ε+η.absentTr~𝑔Σ𝜀𝜂\displaystyle\leq\frac{\mathop{\mathrm{Tr}}(\tilde{g}\,\Sigma)}{\varepsilon}+\eta.≤ divide start_ARG roman_Tr ( over~ start_ARG italic_g end_ARG roman_Σ ) end_ARG start_ARG italic_ε end_ARG + italic_η . (77)

Notice that in the l.h.s., there is a factor of N𝑁Nitalic_N; in order to bound the squared length, we can rescale ε=ε/Nsuperscript𝜀𝜀𝑁\varepsilon^{\prime}=\varepsilon/Nitalic_ε start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = italic_ε / italic_N, so that the following inequality is satisfied for all NN0𝑁subscript𝑁0N\geq N_{0}italic_N ≥ italic_N start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT

[Δ(N)g~Δ(N)ε]Tr(gΣ)Nε+η.delimited-[]superscriptsuperscriptΔ𝑁top~𝑔superscriptΔ𝑁superscript𝜀Tr𝑔Σ𝑁superscript𝜀𝜂\displaystyle\mathbb{P}\left[{\Delta^{(N)}}^{\top}\tilde{g}\,\Delta^{(N)}\geq% \varepsilon^{\prime}\right]\leq\frac{\mathop{\mathrm{Tr}}(g\Sigma)}{N% \varepsilon^{\prime}}+\eta.blackboard_P [ roman_Δ start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT over~ start_ARG italic_g end_ARG roman_Δ start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT ≥ italic_ε start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ] ≤ divide start_ARG roman_Tr ( italic_g roman_Σ ) end_ARG start_ARG italic_N italic_ε start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_ARG + italic_η . (78)

By choosing a tolerance probability δ𝛿\deltaitalic_δ such that the lower bound for the squared length is satisfied,

Tr(g~Σ)Nε(N)+η=δ,Tr~𝑔Σ𝑁superscript𝜀𝑁𝜂𝛿\displaystyle\frac{\mathop{\mathrm{Tr}}\left(\tilde{g}\,\Sigma\right)}{N% \varepsilon^{\prime}(N)}+\eta=\delta,divide start_ARG roman_Tr ( over~ start_ARG italic_g end_ARG roman_Σ ) end_ARG start_ARG italic_N italic_ε start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_N ) end_ARG + italic_η = italic_δ , (79)

we can solve for the N𝑁Nitalic_N-sample error ε(N)superscript𝜀𝑁\varepsilon^{\prime}(N)italic_ε start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_N ) as

ε(N)=Tr(g~Σ)N(δη).superscript𝜀𝑁Tr~𝑔Σ𝑁𝛿𝜂\varepsilon^{\prime}(N)=\frac{\mathop{\mathrm{Tr}}(\tilde{g}\,\Sigma)}{N(% \delta-\eta)}.italic_ε start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_N ) = divide start_ARG roman_Tr ( over~ start_ARG italic_g end_ARG roman_Σ ) end_ARG start_ARG italic_N ( italic_δ - italic_η ) end_ARG . (80)

Finally, we can upper bound the squared length by taking the complement of the inequality. With probability 1δ1𝛿1-\delta1 - italic_δ, the squared-length estimator is upper bounded by ε(N)superscript𝜀𝑁\varepsilon^{\prime}(N)italic_ε start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_N ), i.e.

[Δ(N)g~Δ(N)ε(N)]1δ.delimited-[]superscriptsuperscriptΔ𝑁top~𝑔superscriptΔ𝑁superscript𝜀𝑁1𝛿\mathbb{P}\left[{\Delta^{(N)}}^{\top}\tilde{g}\,\Delta^{(N)}\leq\varepsilon^{% \prime}(N)\right]\geq 1-\delta.blackboard_P [ roman_Δ start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT over~ start_ARG italic_g end_ARG roman_Δ start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT ≤ italic_ε start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_N ) ] ≥ 1 - italic_δ . (81)

To achieve a bound for the infidelity, we take the square root of both sides of the inequality and arrive at an upper bound of the length of the velocity vector, which bounds the infidelity by Eq. (65) and Eq. (68), so that

[1|ψ^(N)|ϕ|ε(N)]1δ.delimited-[]1inner-productsuperscript^𝜓𝑁italic-ϕsuperscript𝜀𝑁1𝛿\mathbb{P}\left[1-|\langle\hat{\psi}^{(N)}|\phi\rangle|\leq\sqrt{\varepsilon^{% \prime}(N)}\right]\geq 1-\delta.blackboard_P [ 1 - | ⟨ over^ start_ARG italic_ψ end_ARG start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT | italic_ϕ ⟩ | ≤ square-root start_ARG italic_ε start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_N ) end_ARG ] ≥ 1 - italic_δ . (82)

Finally, we can upper bound the N𝑁Nitalic_N-sample error threshold as

ε(N)superscript𝜀𝑁\displaystyle\varepsilon^{\prime}(N)italic_ε start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_N ) dim(M~)g~ΣN(δη),absentdimension~𝑀subscriptnorm~𝑔Σ𝑁𝛿𝜂\displaystyle\leq\frac{\dim(\tilde{M})\|\tilde{g}\,\Sigma\|_{\infty}}{N(\delta% -\eta)},≤ divide start_ARG roman_dim ( over~ start_ARG italic_M end_ARG ) ∥ over~ start_ARG italic_g end_ARG roman_Σ ∥ start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT end_ARG start_ARG italic_N ( italic_δ - italic_η ) end_ARG , (83)

where \|\bullet\|_{\infty}∥ ∙ ∥ start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT is the \infty-Schatten norm. The dimension of the projective MPS manifold can be upper bounded as dim(M~)=O(nχmax2)dimension~𝑀𝑂𝑛superscriptsubscript𝜒2\dim(\tilde{M})=O(n\chi_{\max}^{2})roman_dim ( over~ start_ARG italic_M end_ARG ) = italic_O ( italic_n italic_χ start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) where χmaxsubscript𝜒\chi_{\max}italic_χ start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT is the maximum bond dimension χmax=max(χ)subscript𝜒𝜒\chi_{\max}=\max(\chi)italic_χ start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = roman_max ( italic_χ ).

The η𝜂\etaitalic_η-term in the denominator is a consequence of the fact that the estimator [ψ^(N)]delimited-[]superscript^𝜓𝑁[\hat{\psi}^{(N)}][ over^ start_ARG italic_ψ end_ARG start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT ] is only asymptotically normal. The smaller one wants to make η𝜂\etaitalic_η, the larger N0subscript𝑁0N_{0}italic_N start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT becomes so that we have to move “closer” to the asymptotic regime. Asymptotically, i.e., in the limit N𝑁N\rightarrow\inftyitalic_N → ∞, we can choose η0𝜂0\eta\rightarrow 0italic_η → 0 to arrive at our main result in Theorem 2

[1|ψ^(N)|ϕ|ϵ(N)]1δ,delimited-[]1inner-productsuperscript^𝜓𝑁italic-ϕitalic-ϵ𝑁1𝛿\mathbb{P}\left[1-|\langle\hat{\psi}^{(N)}|\phi\rangle|\leq\epsilon(N)\right]% \geq 1-\delta,blackboard_P [ 1 - | ⟨ over^ start_ARG italic_ψ end_ARG start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT | italic_ϕ ⟩ | ≤ italic_ϵ ( italic_N ) ] ≥ 1 - italic_δ , (84)

where the infidelity error threshold is given by

ϵ(N)=O(nχmax2Nδ).italic-ϵ𝑁𝑂𝑛superscriptsubscript𝜒2𝑁𝛿\epsilon(N)=O\left(\sqrt{\frac{n\chi_{\max}^{2}}{N\delta}}\right).italic_ϵ ( italic_N ) = italic_O ( square-root start_ARG divide start_ARG italic_n italic_χ start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_N italic_δ end_ARG end_ARG ) . (85)

Refer to caption
Supplementary Figure 1: Infidelity scaling for interpolation between a random and GHZ 3333-qubit state. The numerical results illustrate the polynomial scaling of the infidelity with the number of samples N𝑁Nitalic_N across a family of states, parametrized by x𝑥xitalic_x. Here, x𝑥xitalic_x ranges from 0 (darkest color, representing the GHZ state) to 1 (lightest color, representing the random state). Dashed lines indicate polynomial fits of the form 1F=c×Nα1𝐹𝑐superscript𝑁𝛼1-F=c\times N^{-\alpha}1 - italic_F = italic_c × italic_N start_POSTSUPERSCRIPT - italic_α end_POSTSUPERSCRIPT, with values for c𝑐citalic_c and α𝛼\alphaitalic_α provided in the legend.

B.7 Discussions and interpretations

In this subsection, we discuss Theorem 2 and its connection to the numerical results presented in the main text. This theorem demonstrates that the MPS solution of maximum likelihood estimate (MLE) approaches the target state when the number of samples N𝑁Nitalic_N is sufficiently large, with the infidelity 1F1𝐹1-F1 - italic_F converging at a rate proportional to 1/N1𝑁1/\sqrt{N}1 / square-root start_ARG italic_N end_ARG. This provides a theoretical foundation for understanding the scaling behavior observed in our numerical experiments.

First, we comment on how this theoretical result relates to our numerics in practice. In the derivation of our theorem, we have assumed that our variational space forms a projective MPS manifold. However, in our numerical simulations, we fix the bond dimension across sites, including variational states that are not always full rank, although they do include a projective MPS manifold as a subspace. In addition, several assumptions underpinning our result could be compromised: 1) The Hessian of the negative log-likelihood (NLL) loss function may not behave as expected. 2) The unitary ensemble is not tomographically complete in our variational space. Notably, while Theorem 1 confirms that our random-XZ𝑋𝑍XZitalic_X italic_Z measurements are tomographically complete for real and pure states, our variational approach involves complex numbers for enhanced stability in optimization, leading to a space of complex pure states. To overcome this discrepancy, incorporating random-Pauli measurements, which are known to be tomographically complete, could provide a solution. Alternatively, our analysis of the complex projective MPS manifold could be adapted to the setting of a real projective MPS, and one can verify that the necessary conditions on the manifold structure are satisfied in the subspace of real states.

Focusing on our numerical observations, we empirically find that the infidelity decreases polynomially with the number of samples N𝑁Nitalic_N as 1F1Nαproportional-to1𝐹1superscript𝑁𝛼1-F\propto\frac{1}{N^{\alpha}}1 - italic_F ∝ divide start_ARG 1 end_ARG start_ARG italic_N start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT end_ARG. Theorem 2 generally predicts that α=0.5𝛼0.5\alpha=0.5italic_α = 0.5. Specifically, a strong alignment with the polynomial functional form indicates that N𝑁Nitalic_N is sufficiently large to approach the asymptotic limit. In practice, we see that some of our numerical results for n=9𝑛9n=9italic_n = 9 qubits exhibit such a polynomial scaling of the infidelity with α1𝛼1\alpha\approx 1italic_α ≈ 1 in Fig. 2(b), while others do not (see Fig. 3(b,c)). The violation of our prediction could be explained by the invalidity of one or more of our assumptions. Intriguingly, α=1𝛼1\alpha=1italic_α = 1 is actually better than the upper bound provided in our theorem, pointing to additional complexities.

To address discrepancies between our numerical observations and our theoretical framework, we have performed additional numerical experiments using random states to validate our theorem’s predictions. In Supplementary Fig. 1, we illustrate how the infidelity scales with the number of samples N𝑁Nitalic_N for states that interpolate between a GHZ state |GHZketGHZ|\mathrm{GHZ}\rangle| roman_GHZ ⟩ and a random state vector |ψχ=2ketsubscript𝜓𝜒2|\psi_{\chi=2}\rangle| italic_ψ start_POSTSUBSCRIPT italic_χ = 2 end_POSTSUBSCRIPT ⟩ with bond dimension 2222. The level of randomness is controlled by the parameter x𝑥xitalic_x, as defined by

|ϕ=1x|GHZ+x|ψχ=2.ketitalic-ϕ1𝑥ketGHZ𝑥ketsubscript𝜓𝜒2\mbox{$|\phi\rangle$}=\sqrt{1-x}\mbox{$|\mathrm{GHZ}\rangle$}+\sqrt{x}\mbox{$|% \psi_{\chi=2}\rangle$}.| italic_ϕ ⟩ = square-root start_ARG 1 - italic_x end_ARG | roman_GHZ ⟩ + square-root start_ARG italic_x end_ARG | italic_ψ start_POSTSUBSCRIPT italic_χ = 2 end_POSTSUBSCRIPT ⟩ . (86)

Our findings reveal that for a purely random state (x=1𝑥1x=1italic_x = 1), the observed scaling exponent α=0.53𝛼0.53\alpha=0.53italic_α = 0.53 aligns perfectly with our theoretical prediction. Conversely, at x=0𝑥0x=0italic_x = 0, corresponding to the GHZ state, an unexpected deviation with α=1𝛼1\alpha=1italic_α = 1 mirrors the anomaly previously observed in Fig. 2(b). This suggests that the inherent structures of stabilizer states, like the GHZ and the surface code state, might facilitate a better scaling than our current upper bound.

Finally, we note that in our practical numerics, we perform MLE in the space of MPS tensors, and not directly on the manifold. Theoretically, the trajectory towards the minimum would differ between these two approaches. Nonetheless, our theorem focuses on the properties of the optimal MLE estimator, acknowledging that practically achieving this solution may not always be efficient. Note, however, that once we have found MPS tensors corresponding to the minimizer of the loss function, they correspond to a unique physical state in Hilbert space. Thus, the theorem still makes a meaningful statement about the practical optimizer with respect to the MPS tensors, under the condition that the optimizer finds a global minimizer of the loss function. We leave potential improvement of the optimization by incorporating insights from the manifold structure [84, 85] for future work.

APPENDIX C Random-XZ𝑋𝑍XZitalic_X italic_Z classical shadow tomography

This self-contained appendix fully characterizes random-XZ𝑋𝑍XZitalic_X italic_Z classical shadow tomography and derives the necessary tools to implement it. We use the results outlined in this section when estimating observables in the regularization step of Eq. (5); see also Appendix E.2 for the details of regularization using random-XZ𝑋𝑍XZitalic_X italic_Z classical shadows. We are also motivated to characterize random-XZ𝑋𝑍XZitalic_X italic_Z classical shadows because measurements of this form allow one to fully characterize real, pure states, which are the target states in this work (refer to Appendix D for a proof). Random-XZ𝑋𝑍XZitalic_X italic_Z classical shadows may also be of interest more broadly due to the ease of their experimental implementation across different platforms. For instance, with superconducting qubits, measurement in different bases only requires single-qubit rotation which are cheap to implement. In the context of Rydberg atom arrays, this holds for both physical and logical qubits. Single-qubit rotations on physical qubits have high fidelities [64]. For logical qubits, transversal Hadamard rotates the logical to the X𝑋Xitalic_X basis by using single-qubit physical rotations, and even without these transversal Hadamard gates, logical qubits can still be effectively measured in the X𝑋Xitalic_X basis [31]. In this latter case, performing XZ𝑋𝑍XZitalic_X italic_Z measurements is not only doable but also provides the ability to perform error detection (e.g., consider the [[15,1,3]] Reed-Muller 3D color code).

Random-XZ𝑋𝑍XZitalic_X italic_Z classical shadows involve preparing some (unknown) state, randomly measuring some n𝑛nitalic_n-qubit Pauli string composed of only X𝑋Xitalic_X and Z𝑍Zitalic_Z single-site Paulis, and then post-processing to predict properties of interest. We assume that the chosen Pauli is uniformly sampled from the set of all n𝑛nitalic_n-qubit strings composed of only X𝑋Xitalic_Xs and Z𝑍Zitalic_Zs. As this set of measurements is not informationally complete, random-XZ𝑋𝑍XZitalic_X italic_Z classical shadows do not allow us to predict any arbitrary observable. However, we can predict some observables of interest, namely, those that contribute to our regularization step. In this appendix, we will discuss which observables can be estimated with this technique, discuss how to post-process the data to estimate observables, and how to calculate the sample complexity required for learning Pauli strings. We specifically discuss the sample complexity for learning Pauli strings because these are the observables that we estimate in our regularization step. In the first section of this appendix, we introduce these ideas pedagogically, with the rigorous backings for all our claims provided in the second section.

C.1 Performing random-XZ𝑋𝑍XZitalic_X italic_Z classical shadow tomography

Here, we derive an expression for the classical shadow of the quantum state measured by the randomized XZ𝑋𝑍XZitalic_X italic_Z measurements discussed in Appendix D. Random-XZ𝑋𝑍XZitalic_X italic_Z classical shadows involve taking many random measurements on some prepared state. Each measurement involves two steps. First, the state is rotated under some unitary U𝑈Uitalic_U which is sampled from U𝒰XZsimilar-to𝑈subscript𝒰𝑋𝑍U\sim\mathcal{U}_{XZ}italic_U ∼ caligraphic_U start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT, and then, in the second step, the state is measured in the computational basis states {|bb|}b{0,1}n\{\lvert b\rangle\!\langle b\rvert\}_{b\in\{0,1\}^{n}}{ | italic_b ⟩ ⟨ italic_b | } start_POSTSUBSCRIPT italic_b ∈ { 0 , 1 } start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_POSTSUBSCRIPT. This ensemble of unitaries 𝒰XZsubscript𝒰𝑋𝑍\mathcal{U}_{XZ}caligraphic_U start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT exhibits a uniform probability distribution over unitaries U𝑈Uitalic_U which take the form

U=i=1nUαi where αi{X,Z}.U=\otimes_{i=1}^{n}U^{\alpha_{i}}\hskip 2.84526pt\text{ where }\alpha_{i}\in\{% X,Z\}.italic_U = ⊗ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT italic_U start_POSTSUPERSCRIPT italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUPERSCRIPT where italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ∈ { italic_X , italic_Z } . (87)

Here, Uαisuperscript𝑈subscript𝛼𝑖U^{\alpha_{i}}italic_U start_POSTSUPERSCRIPT italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUPERSCRIPT is the single-qubit unitary which rotates from the computational basis to the (different) basis of the Pauli operator σiαsuperscriptsubscript𝜎𝑖𝛼\sigma_{i}^{\alpha}italic_σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT acting on site i𝑖iitalic_i. Therefore, acting on the prepared state, the ensemble 𝒰XZsubscript𝒰𝑋𝑍\mathcal{U}_{XZ}caligraphic_U start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT allows one to randomly measure any n𝑛nitalic_n-qubit Pauli string composed of only X𝑋Xitalic_X and Z𝑍Zitalic_Z single-site Paulis. Again, this set of measurements is not informationally complete, and as such, we cannot predict any arbitrary observable with random-XZ𝑋𝑍XZitalic_X italic_Z classical shadows.

In this appendix, we will use the nomenclature developed in Ref. [15]. In particular, we define the visible space of our random-XZ𝑋𝑍XZitalic_X italic_Z measurement scheme as the space of all operators O𝑂Oitalic_O for which we can estimate the expectation value Tr(ρO)Tr𝜌𝑂\mathop{\mathrm{Tr}}(\rho O)roman_Tr ( italic_ρ italic_O ); we denote this as VisibleSpace(𝒰𝒰\mathcal{U}caligraphic_U). Similarly, the invisible space is the space of operators that are invisible to us under the implementable unitaries 𝒰𝒰\mathcal{U}caligraphic_U, and we label this as InvisibleSpace(𝒰𝒰\mathcal{U}caligraphic_U). Note that the full operator space ()\mathcal{B}(\mathcal{H})caligraphic_B ( caligraphic_H ) associated to the Hilbert space \mathcal{H}caligraphic_H can be expressed with the following tensor sum:

()VisibleSpace(𝒰)InvisibleSpace(𝒰),similar-to-or-equalsdirect-sumVisibleSpace𝒰InvisibleSpace𝒰\mathcal{B}(\mathcal{H})\simeq\textnormal{{VisibleSpace}}(\mathcal{U})\oplus% \textnormal{{InvisibleSpace}}(\mathcal{U})\,,caligraphic_B ( caligraphic_H ) ≃ VisibleSpace ( caligraphic_U ) ⊕ InvisibleSpace ( caligraphic_U ) , (88)

where the symbol similar-to-or-equals\simeq stands for “isomorphic to”. As a general rule, a more limited set of implementable unitaries 𝒰𝒰\mathcal{U}caligraphic_U corresponds to a smaller VisibleSpace(𝒰)VisibleSpace𝒰\textnormal{{VisibleSpace}}(\mathcal{U})VisibleSpace ( caligraphic_U ). In Corollary 4, we find that the random-XZ𝑋𝑍XZitalic_X italic_Z visible space is spanned by Pauli strings constructed from single-site 𝟙1\mathds{1}blackboard_1, X𝑋Xitalic_X, and Z𝑍Zitalic_Z Paulis. We estimate these Paulis and then plug them into our regularization step.

We can estimate observables OVisibleSpace(𝒰XZ)𝑂VisibleSpacesubscript𝒰𝑋𝑍O\in\textnormal{{VisibleSpace}}(\mathcal{U}_{XZ})italic_O ∈ VisibleSpace ( caligraphic_U start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT ) by post-processing the data to construct an estimate ρ^^𝜌\hat{\rho}over^ start_ARG italic_ρ end_ARG of the true density matrix ρ𝜌\rhoitalic_ρ. This estimate is called the “classical shadow” of the true density matrix, and we can use it to compute Tr(ρ^O)Tr(ρO)Tr^𝜌𝑂Tr𝜌𝑂\mathop{\mathrm{Tr}}(\hat{\rho}O)\approx\mathop{\mathrm{Tr}}(\rho O)roman_Tr ( over^ start_ARG italic_ρ end_ARG italic_O ) ≈ roman_Tr ( italic_ρ italic_O ). We construct the classical shadow using the data from our measurements. Each measurement snapshot can be characterized by the unitary applied, U=iUαiU=\otimes_{i}\hskip 2.84526ptU^{\alpha_{i}}italic_U = ⊗ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_U start_POSTSUPERSCRIPT italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUPERSCRIPT, and the computational basis state measured, |b=i|bi\mbox{$|b\rangle$}=\otimes_{i}\mbox{$|b_{i}\rangle$}| italic_b ⟩ = ⊗ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ⟩. Here the i𝑖iitalic_i indexes the qubits. We can construct a single-shot estimate of our density matrix, ρ^1subscript^𝜌1\hat{\rho}_{1}over^ start_ARG italic_ρ end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, from random-XZ𝑋𝑍XZitalic_X italic_Z data as

ρ^1subscript^𝜌1\displaystyle\hat{\rho}_{1}over^ start_ARG italic_ρ end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT =\displaystyle== XZ1(U|bb|U)\displaystyle\mathcal{M}_{XZ}^{-1}(U^{\dagger}\rvert b\rangle\!\langle b\lvert U)caligraphic_M start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ( italic_U start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT | italic_b ⟩ ⟨ italic_b | italic_U )
=\displaystyle== i=1n(Tr(Uαi|bibi|UαiX)X+Tr(Uαi|bibi|UαiZ)Z+12𝕀)\displaystyle\otimes_{i=1}^{n}\left(\mathop{\mathrm{Tr}}(U^{\alpha_{i}\dagger}% \rvert b_{i}\rangle\!\langle b_{i}\lvert U^{\alpha_{i}}X)X+\mathop{\mathrm{Tr}% }(U^{\alpha_{i}\dagger}\rvert b_{i}\rangle\!\langle b_{i}\lvert U^{\alpha_{i}}% Z)Z+\frac{1}{2}\mathbb{I}\right)⊗ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ( roman_Tr ( italic_U start_POSTSUPERSCRIPT italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT † end_POSTSUPERSCRIPT | italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ⟩ ⟨ italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | italic_U start_POSTSUPERSCRIPT italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_X ) italic_X + roman_Tr ( italic_U start_POSTSUPERSCRIPT italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT † end_POSTSUPERSCRIPT | italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ⟩ ⟨ italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | italic_U start_POSTSUPERSCRIPT italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_Z ) italic_Z + divide start_ARG 1 end_ARG start_ARG 2 end_ARG blackboard_I )
=\displaystyle== i=1n(2Uαi|bibi|Uαi𝕀2).\displaystyle\otimes_{i=1}^{n}\left(2U^{\alpha_{i}\dagger}\rvert b_{i}\rangle% \!\langle b_{i}\lvert U^{\alpha_{i}}-\frac{\mathbb{I}}{2}\right).⊗ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ( 2 italic_U start_POSTSUPERSCRIPT italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT † end_POSTSUPERSCRIPT | italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ⟩ ⟨ italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | italic_U start_POSTSUPERSCRIPT italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUPERSCRIPT - divide start_ARG blackboard_I end_ARG start_ARG 2 end_ARG ) .

The subscript 1111 above indicates that the classical shadow was constructed with data from 1111 measurement. The channel XZ1subscriptsuperscript1𝑋𝑍\mathcal{M}^{-1}_{XZ}caligraphic_M start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT is the Moore-Penrose pseudoinverse of the measurement channel XZsubscript𝑋𝑍\mathcal{M}_{XZ}caligraphic_M start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT derived in Proposition 2 in the next subsection. This channel models the measurement step of the randomized-measurement procedure, and for an infinite set of informationally complete measurements, it guarantees that we will recover our original density matrix:

ρ=limNρ^N,𝜌subscript𝑁subscript^𝜌𝑁\rho=\lim_{N\rightarrow\infty}\hat{\rho}_{N},italic_ρ = roman_lim start_POSTSUBSCRIPT italic_N → ∞ end_POSTSUBSCRIPT over^ start_ARG italic_ρ end_ARG start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT , (90)

where ρ^Nsubscript^𝜌𝑁\hat{\rho}_{N}over^ start_ARG italic_ρ end_ARG start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT is the classical shadow constructed from N𝑁Nitalic_N measurements. One example, of an informationally complete measurement scheme is that of random Pauli measurements, where each qubit is randomly measured in either the X𝑋Xitalic_X, Y𝑌Yitalic_Y, or Z𝑍Zitalic_Z basis. Intuitively, since the Pauli strings form a basis of the operator space ()\mathcal{B}(\mathcal{H})caligraphic_B ( caligraphic_H ), this allows one to predict any observable. However, in our particular setting, the random-XZ𝑋𝑍XZitalic_X italic_Z measurements are not informationally complete, and as such, the infinite-measurement XZ𝑋𝑍XZitalic_X italic_Z classical shadow will not exactly recover every ρ𝜌\rhoitalic_ρ. Instead, with an infinite number of XZ𝑋𝑍XZitalic_X italic_Z measurements, we will recover the exact density matrix projected onto the visible space. Therefore, we can only learn the observables in the visible space because we only learn this portion of ρ𝜌\rhoitalic_ρ.

The sample complexity of classical shadow protocols is defined in terms of the shadow norm Oshadow2subscriptsuperscriptnorm𝑂2shadow\|O\|^{2}_{\text{shadow}}∥ italic_O ∥ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT shadow end_POSTSUBSCRIPT, an upper bound on the variance of predicting Tr(ρO)Tr𝜌𝑂\mathop{\mathrm{Tr}}(\rho O)roman_Tr ( italic_ρ italic_O ) with this measurement-and-reconstruction technique. Of course, to get an actual number of samples needed to predict Tr(ρO)Tr𝜌𝑂\mathop{\mathrm{Tr}}(\rho O)roman_Tr ( italic_ρ italic_O ) to some precision ϵitalic-ϵ\epsilonitalic_ϵ, one would need to employ a concentration inequality. While the original classical shadows proposal demonstrated a median-of-means estimation technique [25], this concentration inequality has a large prefactor, which increases the number of samples by almost two orders of magnitude. Instead, we recommend that the interested reader employs Bernstein’s inequality, as has been used in Ref. [15], which uses an empirical average over the data from N𝑁Nitalic_N shots

ρ^N=1Nj=1NXZ1(U(j)|b(j)b(j)|U(j)).\hat{\rho}_{N}=\frac{1}{N}\sum_{j=1}^{N}\mathcal{M}_{XZ}^{-1}\left(U^{(j)% \dagger}\rvert b^{(j)}\rangle\!\langle b^{(j)}\lvert U^{(j)}\right).over^ start_ARG italic_ρ end_ARG start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG italic_N end_ARG ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT caligraphic_M start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ( italic_U start_POSTSUPERSCRIPT ( italic_j ) † end_POSTSUPERSCRIPT | italic_b start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT ⟩ ⟨ italic_b start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT | italic_U start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT ) . (91)

Here, we represent the unitary applied in the j𝑗jitalic_jth random measurement and the outcome bit-string of that measurement as U(j)superscript𝑈𝑗U^{(j)}italic_U start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT and b(j)superscript𝑏𝑗b^{(j)}italic_b start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT, respectively.

Just as is the case with random Pauli classical shadows (measuring any Pauli string in a given shot), random-XZ𝑋𝑍XZitalic_X italic_Z classical shadows are most effective at learning local observables, the intuition being that we more often measure in bases that tell us something about that observable. Consider, for example, the observables X𝟙n1tensor-product𝑋superscript1tensor-productabsent𝑛1X\otimes\mathds{1}^{\otimes n-1}italic_X ⊗ blackboard_1 start_POSTSUPERSCRIPT ⊗ italic_n - 1 end_POSTSUPERSCRIPT and Xnsuperscript𝑋tensor-productabsent𝑛X^{\otimes n}italic_X start_POSTSUPERSCRIPT ⊗ italic_n end_POSTSUPERSCRIPT. As long as we measure the first qubit in the X𝑋Xitalic_X basis, the measurement will provide information about Tr(ρX𝟙n1)Trtensor-product𝜌𝑋superscript1tensor-productabsent𝑛1\mathop{\mathrm{Tr}}(\rho\hskip 2.84526ptX\otimes\mathds{1}^{\otimes n-1})roman_Tr ( italic_ρ italic_X ⊗ blackboard_1 start_POSTSUPERSCRIPT ⊗ italic_n - 1 end_POSTSUPERSCRIPT ). This happens with probability 1/2121/21 / 2: there is a 1/2121/21 / 2 chance of measuring the first qubit in the X𝑋Xitalic_X basis and a 1/2121/21 / 2 chance of measuring it in the Z𝑍Zitalic_Z basis. Now, let us consider the second observable: Xnsuperscript𝑋tensor-productabsent𝑛X^{\otimes n}italic_X start_POSTSUPERSCRIPT ⊗ italic_n end_POSTSUPERSCRIPT. In order to learn about the expectation value of Xnsuperscript𝑋tensor-productabsent𝑛X^{\otimes n}italic_X start_POSTSUPERSCRIPT ⊗ italic_n end_POSTSUPERSCRIPT with our state, every qubit must be measured in the X𝑋Xitalic_X basis, and this happens with probability 1/2n1superscript2𝑛1/2^{n}1 / 2 start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT since there are n𝑛nitalic_n qubits and each has a 1/2121/21 / 2 chance of being measured in X𝑋Xitalic_X. Therefore, it is much easier to learn X𝟙n1tensor-product𝑋superscript1tensor-productabsent𝑛1X\otimes\mathds{1}^{\otimes n-1}italic_X ⊗ blackboard_1 start_POSTSUPERSCRIPT ⊗ italic_n - 1 end_POSTSUPERSCRIPT than Xnsuperscript𝑋tensor-productabsent𝑛X^{\otimes n}italic_X start_POSTSUPERSCRIPT ⊗ italic_n end_POSTSUPERSCRIPT. In the next subsection of this appendix, we provide sample complexity upper bounds on Pauli strings in the form of the shadow norm of Pauli strings. This result is consistent with our intuition that random-XZ𝑋𝑍XZitalic_X italic_Z classical shadows can more easily learn local Pauli strings than nonlocal Pauli strings. In particular, a k𝑘kitalic_k-local Pauli string Pksubscript𝑃𝑘P_{k}italic_P start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT (k𝑘kitalic_k non-identity single-site Paulis) exhibits a shadow norm of Pkshadow2=2ksubscriptsuperscriptnormsubscript𝑃𝑘2shadowsuperscript2𝑘\|P_{k}\|^{2}_{\text{shadow}}=2^{k}∥ italic_P start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ∥ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT shadow end_POSTSUBSCRIPT = 2 start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT.

C.2 Derivation of tools for random-XZ𝑋𝑍XZitalic_X italic_Z classical shadows

In the remainder of this appendix, we derive the tools and claims discussed above for performing random-XZ𝑋𝑍XZitalic_X italic_Z classical shadow tomography. First, we will derive the classical shadow measurement channel XZsubscript𝑋𝑍\mathcal{M}_{XZ}caligraphic_M start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT for the learning procedure described in the previous subsection. Once a set of implementable unitaries has been specified, the measurement scheme can be defined in the language of probability density functions. For the random-XZ𝑋𝑍XZitalic_X italic_Z measurement scheme considered above, we define our ensemble 𝒰XZsubscript𝒰𝑋𝑍\mathcal{U}_{XZ}caligraphic_U start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT to have a uniform probability density function in the space of all possible unitaries rotating into X𝑋Xitalic_X and Z𝑍Zitalic_Z single-qubit bases. This allows us to use Observation 4 from Ref. [15] in order to define the XZ𝑋𝑍XZitalic_X italic_Z visible space. Specifically, the image of the XZ𝑋𝑍XZitalic_X italic_Z measurement channel defines the visible space. While the visible space is technically independent of the underlying probability density function (being an indicator of what can be learned with the measurement unitaries), the uniform distribution allows us to characterize the visible space since each unitary has a nonzero probability of being sampled [15]. In addition to characterizing the visible space, we will also use the measurement channel to define the shadow norm of Pauli strings. As the unitaries in the ensemble are Cliffords, the eigenoperators of the measurement channel are Pauli strings, and the eigenvalues are the inverse shadow norms of the corresponding eigenoperators [86, 87].

Proposition 2 (Random-XZ𝑋𝑍XZitalic_X italic_Z classical shadow measurement channel).

Assume a measurement primitive 𝒰XZsubscript𝒰𝑋𝑍\mathcal{U}_{XZ}caligraphic_U start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT such that the associated probability density function is uniform over random-XZ𝑋𝑍XZitalic_X italic_Z unitaries. The random-XZ𝑋𝑍XZitalic_X italic_Z measurement channel for n𝑛nitalic_n qubits can be written as

XZ(A)=PPauli𝟙,X,Z(n)12k(P)tr(AP)P,subscriptXZ𝐴subscript𝑃subscriptPauli1𝑋𝑍𝑛1superscript2𝑘𝑃tr𝐴𝑃𝑃\mathcal{M}_{\text{\rm XZ}}(A)=\sum_{P\in\text{\rm Pauli}_{\mathds{1},X,Z}(n)}% \frac{1}{2^{k(P)}}\text{\rm tr}(AP)P\,,caligraphic_M start_POSTSUBSCRIPT XZ end_POSTSUBSCRIPT ( italic_A ) = ∑ start_POSTSUBSCRIPT italic_P ∈ Pauli start_POSTSUBSCRIPT blackboard_1 , italic_X , italic_Z end_POSTSUBSCRIPT ( italic_n ) end_POSTSUBSCRIPT divide start_ARG 1 end_ARG start_ARG 2 start_POSTSUPERSCRIPT italic_k ( italic_P ) end_POSTSUPERSCRIPT end_ARG tr ( italic_A italic_P ) italic_P , (92)

where Pauli𝟙,X,Z(n)subscriptPauli1𝑋𝑍𝑛\text{\rm Pauli}_{\mathds{1},X,Z}(n)Pauli start_POSTSUBSCRIPT blackboard_1 , italic_X , italic_Z end_POSTSUBSCRIPT ( italic_n ) is the set of n𝑛nitalic_n-qubit Pauli strings constructed from only 𝟙1\mathds{1}blackboard_1s, X𝑋Xitalic_Xs, and Z𝑍Zitalic_Zs, and k(P)𝑘𝑃k(P)italic_k ( italic_P ) is the locality of Pauli string P𝑃Pitalic_P. Note that we assume all Pauli strings P𝑃Pitalic_P above are normalized with respect to the Hilbert-Schmidt inner product: tr(PP)=1tr𝑃superscript𝑃1\text{\rm tr}(PP^{\dagger})=1tr ( italic_P italic_P start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ) = 1.

Proof.

For the measurement primitive [25] assumed above, the random unitaries are uniformly sampled from 𝒰XZsubscript𝒰𝑋𝑍\mathcal{U}_{XZ}caligraphic_U start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT, which is a subset of the Clifford group. As such, by Lemma 1 of Ref. [88], the eigenoperators of the corresponding measurement channel XZsubscript𝑋𝑍\mathcal{M}_{XZ}caligraphic_M start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT are Pauli strings. Therefore, we can express the action of the measurement channel as MXZ(P)=λPPsubscript𝑀𝑋𝑍𝑃subscript𝜆𝑃𝑃M_{XZ}(P)=\lambda_{P}Pitalic_M start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT ( italic_P ) = italic_λ start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT italic_P, where λPsubscript𝜆𝑃\lambda_{P}italic_λ start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT denotes the eigenvalue corresponding to the Pauli string eigenoperator P𝑃Pitalic_P. It suffices to show that a k(P)𝑘𝑃k(P)italic_k ( italic_P )-local Pauli string PPauli𝟙,X,Z(n)𝑃subscriptPauli1𝑋𝑍𝑛P\in\text{\rm Pauli}_{\mathds{1},X,Z}(n)italic_P ∈ Pauli start_POSTSUBSCRIPT blackboard_1 , italic_X , italic_Z end_POSTSUBSCRIPT ( italic_n ) has eigenvalue λP=1/2k(P)subscript𝜆𝑃1superscript2𝑘𝑃\lambda_{P}=1/2^{k(P)}italic_λ start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT = 1 / 2 start_POSTSUPERSCRIPT italic_k ( italic_P ) end_POSTSUPERSCRIPT, and all other Pauli strings PPauli𝟙,X,Z(n)𝑃subscriptPauli1𝑋𝑍𝑛P\notin\text{\rm Pauli}_{\mathds{1},X,Z}(n)italic_P ∉ Pauli start_POSTSUBSCRIPT blackboard_1 , italic_X , italic_Z end_POSTSUBSCRIPT ( italic_n ) have eigenvalue 0.

To show this, we will derive an expression for λPsubscript𝜆𝑃\lambda_{P}italic_λ start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT. Let us start by rearranging the measurement channel. We can start with its original definition from Ref. [25], to get

(P)𝑃\displaystyle\mathcal{M}(P)caligraphic_M ( italic_P ) =\displaystyle== 𝔼U𝒰XZb{0,1}nU|bb|UTr(U|bb|UP)\displaystyle\operatorname{\mathbb{E}}_{U\sim\mathcal{U}_{XZ}}\sum_{b\in\{0,1% \}^{n}}U^{\dagger}\lvert b\rangle\!\langle b\rvert U\hskip 2.84526pt\mathrm{Tr% }\left(U^{\dagger}\lvert b\rangle\!\langle b\rvert UP\right)blackboard_E start_POSTSUBSCRIPT italic_U ∼ caligraphic_U start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_b ∈ { 0 , 1 } start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_POSTSUBSCRIPT italic_U start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT | italic_b ⟩ ⟨ italic_b | italic_U roman_Tr ( italic_U start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT | italic_b ⟩ ⟨ italic_b | italic_U italic_P )
=\displaystyle== 𝔼U𝒰XZTrB[b{0,1}nU|bb|U(U|bb|UP)B]\displaystyle\operatorname{\mathbb{E}}_{U\sim\mathcal{U}_{XZ}}\mathrm{Tr}_{B}% \left[\sum_{b\in\{0,1\}^{n}}U^{\dagger}\lvert b\rangle\!\langle b\rvert U% \hskip 2.84526pt\otimes\hskip 2.84526pt(U^{\dagger}\lvert b\rangle\!\langle b% \rvert U\hskip 2.84526ptP)_{B}\right]blackboard_E start_POSTSUBSCRIPT italic_U ∼ caligraphic_U start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT end_POSTSUBSCRIPT roman_Tr start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT [ ∑ start_POSTSUBSCRIPT italic_b ∈ { 0 , 1 } start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_POSTSUBSCRIPT italic_U start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT | italic_b ⟩ ⟨ italic_b | italic_U ⊗ ( italic_U start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT | italic_b ⟩ ⟨ italic_b | italic_U italic_P ) start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT ]
=\displaystyle== 𝔼U𝒰XZTrB[U2b|bbb,b|U2(𝟙P)]\displaystyle\operatorname{\mathbb{E}}_{U\sim\mathcal{U}_{XZ}}\mathrm{Tr}_{B}% \left[U^{\dagger\otimes 2}\hskip 2.84526pt\sum_{b}\lvert bb\rangle\!\langle b,% b\rvert\hskip 2.84526ptU^{\otimes 2}\hskip 2.84526pt(\mathds{1}\otimes P)\right]blackboard_E start_POSTSUBSCRIPT italic_U ∼ caligraphic_U start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT end_POSTSUBSCRIPT roman_Tr start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT [ italic_U start_POSTSUPERSCRIPT † ⊗ 2 end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT | italic_b italic_b ⟩ ⟨ italic_b , italic_b | italic_U start_POSTSUPERSCRIPT ⊗ 2 end_POSTSUPERSCRIPT ( blackboard_1 ⊗ italic_P ) ]
=\displaystyle== 𝔼U𝒰XZP~Pauli𝟙,Z(n)UP~UTr(PUP~U).subscript𝔼similar-to𝑈subscript𝒰𝑋𝑍subscript~𝑃subscriptPauli1𝑍𝑛superscript𝑈~𝑃𝑈Tr𝑃superscript𝑈~𝑃𝑈\displaystyle\operatorname{\mathbb{E}}_{U\sim\mathcal{U}_{XZ}}\sum_{\tilde{P}% \in\text{\rm Pauli}_{\mathds{1},Z}(n)}U^{\dagger}\tilde{P}U\hskip 2.84526pt% \mathop{\mathrm{Tr}}(P\hskip 2.84526ptU^{\dagger}\tilde{P}U).blackboard_E start_POSTSUBSCRIPT italic_U ∼ caligraphic_U start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT over~ start_ARG italic_P end_ARG ∈ Pauli start_POSTSUBSCRIPT blackboard_1 , italic_Z end_POSTSUBSCRIPT ( italic_n ) end_POSTSUBSCRIPT italic_U start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT over~ start_ARG italic_P end_ARG italic_U roman_Tr ( italic_P italic_U start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT over~ start_ARG italic_P end_ARG italic_U ) .

In the second line, we take the trace over subsystem “B𝐵Bitalic_B”, which we define to be the second subsystem (i.e., the one with P𝑃Pitalic_P in the term 𝟙Ptensor-product1𝑃\mathds{1}\otimes Pblackboard_1 ⊗ italic_P). In the third line, Pauli𝟙,Z(n)subscriptPauli1𝑍𝑛\text{\rm Pauli}_{\mathds{1},Z}(n)Pauli start_POSTSUBSCRIPT blackboard_1 , italic_Z end_POSTSUBSCRIPT ( italic_n ) is the set of all normalized Pauli strings containing only 𝟙1\mathds{1}blackboard_1s and Z𝑍Zitalic_Zs. One can begin to see the transition from line two to line three by considering the single-qubit measurement channel. For a single qubit, we have b{0,1}|b,bb,b|=12(𝟙𝟙+ZZ)subscript𝑏01𝑏𝑏𝑏𝑏1211𝑍𝑍\sum_{b\in\{0,1\}}\lvert b,b\rangle\!\langle b,b\rvert=\frac{1}{2}(\mathds{1}% \mathds{1}+ZZ)∑ start_POSTSUBSCRIPT italic_b ∈ { 0 , 1 } end_POSTSUBSCRIPT | italic_b , italic_b ⟩ ⟨ italic_b , italic_b | = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( blackboard_11 + italic_Z italic_Z ). This can be extended to n𝑛nitalic_n qubits, where the n𝑛nitalic_n-qubit version b{0,1}n|b,bb,b|subscript𝑏superscript01𝑛|b,bb,b|\sum_{b\in\{0,1\}^{n}}\mbox{$|b,b\rangle$}\mbox{$\langle b,b|$}∑ start_POSTSUBSCRIPT italic_b ∈ { 0 , 1 } start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_POSTSUBSCRIPT |b,b⟩ ⟨b,b| becomes a sum of all length-n𝑛nitalic_n 𝟙1\mathds{1}blackboard_1,Z𝑍Zitalic_Z strings tensored twice, normalized by a factor of 1/2n1superscript2𝑛1/2^{n}1 / 2 start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT [15]. This normalization does not appear in the final expression above because we absorb it into the two P~~𝑃\tilde{P}over~ start_ARG italic_P end_ARGs such that they are all normalized with respect to the Hilbert-Schmidt inner product.

We can further simplify this expression by noticing that Tr(PUP~U)Tr𝑃superscript𝑈~𝑃𝑈\mathop{\mathrm{Tr}}(P\hskip 2.84526ptU^{\dagger}\tilde{P}U)roman_Tr ( italic_P italic_U start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT over~ start_ARG italic_P end_ARG italic_U ) will be nonzero only when P=UP~U𝑃superscript𝑈~𝑃𝑈P=U^{\dagger}\tilde{P}Uitalic_P = italic_U start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT over~ start_ARG italic_P end_ARG italic_U. With this observation, the expression becomes

(P)=(𝔼U𝒰XZP~Pauli𝟙,Z(n),s.t.P~=UPU1)P.𝑃subscript𝔼similar-to𝑈subscript𝒰𝑋𝑍subscript~𝑃subscriptPauli1𝑍𝑛s.t.~𝑃𝑈𝑃superscript𝑈1𝑃\mathcal{M}(P)=\left(\operatorname{\mathbb{E}}_{U\sim\mathcal{U}_{XZ}}\sum_{% \begin{subarray}{c}\tilde{P}\in\text{\rm Pauli}_{\mathds{1},Z}(n),\\ \text{s.t.}\tilde{P}=UPU^{\dagger}\end{subarray}}1\right)\hskip 2.84526ptP.caligraphic_M ( italic_P ) = ( blackboard_E start_POSTSUBSCRIPT italic_U ∼ caligraphic_U start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT start_ARG start_ROW start_CELL over~ start_ARG italic_P end_ARG ∈ Pauli start_POSTSUBSCRIPT blackboard_1 , italic_Z end_POSTSUBSCRIPT ( italic_n ) , end_CELL end_ROW start_ROW start_CELL s.t. over~ start_ARG italic_P end_ARG = italic_U italic_P italic_U start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT end_CELL end_ROW end_ARG end_POSTSUBSCRIPT 1 ) italic_P . (94)

The quantity in parenthesis is simply an expression for λPsubscript𝜆𝑃\lambda_{P}italic_λ start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT, the eigenvalue corresponding to Pauli string P𝑃Pitalic_P. It is the probability that one of our random unitaries U𝒰XZ𝑈subscript𝒰𝑋𝑍U\in\mathcal{U}_{XZ}italic_U ∈ caligraphic_U start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT rotates P𝑃Pitalic_P into a Pauli string UPU𝑈𝑃superscript𝑈UPU^{\dagger}italic_U italic_P italic_U start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT composed of only 𝟙1\mathds{1}blackboard_1s and Z𝑍Zitalic_Zs. We will separate the 4nsuperscript4𝑛4^{n}4 start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT Pauli strings into two sets and consider their eigenvalues separately. First, let’s consider the Pauli strings P𝑃Pitalic_P in Pauli𝟙,X,Z(n)subscriptPauli1𝑋𝑍𝑛\text{\rm Pauli}_{\mathds{1},X,Z}(n)Pauli start_POSTSUBSCRIPT blackboard_1 , italic_X , italic_Z end_POSTSUBSCRIPT ( italic_n ). All sites in P𝑃Pitalic_P which have a 𝟙1\mathds{1}blackboard_1 will be trivially rotated back to 𝟙1\mathds{1}blackboard_1 because U𝑈Uitalic_U in UPU𝑈𝑃superscript𝑈UPU^{\dagger}italic_U italic_P italic_U start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT separates into U=iUiU=\otimes_{i}U_{i}italic_U = ⊗ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_U start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. All that we must consider are the non-identity sites of P𝑃Pitalic_P, and there are k(P)𝑘𝑃k(P)italic_k ( italic_P ) of them. Regardless of whether each site has an X𝑋Xitalic_X or a Z𝑍Zitalic_Z, there is a 1/2121/21 / 2 probability that each site is measured in the corresponding basis. Therefore, the eigenvalue of a k(P)𝑘𝑃k(P)italic_k ( italic_P )-local Pauli string PPauli𝟙,X,Z(n)𝑃subscriptPauli1𝑋𝑍𝑛P\in\text{\rm Pauli}_{\mathds{1},X,Z}(n)italic_P ∈ Pauli start_POSTSUBSCRIPT blackboard_1 , italic_X , italic_Z end_POSTSUBSCRIPT ( italic_n ) is λP=1/2k(P)subscript𝜆𝑃1superscript2𝑘𝑃\lambda_{P}=1/2^{k(P)}italic_λ start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT = 1 / 2 start_POSTSUPERSCRIPT italic_k ( italic_P ) end_POSTSUPERSCRIPT.

Next, let us consider Pauli strings that are not in Pauli𝟙,X,Z(n)subscriptPauli1𝑋𝑍𝑛\text{\rm Pauli}_{\mathds{1},X,Z}(n)Pauli start_POSTSUBSCRIPT blackboard_1 , italic_X , italic_Z end_POSTSUBSCRIPT ( italic_n ). In other words, such Pauli strings P𝑃Pitalic_P contain at least one single-site Y𝑌Yitalic_Y Pauli by definition. Since the random-XZ𝑋𝑍XZitalic_X italic_Z unitaries measure each qubit in either the X𝑋Xitalic_X or the Z𝑍Zitalic_Z basis, there exists no U𝑈Uitalic_U in 𝒰XZsubscript𝒰𝑋𝑍\mathcal{U}_{XZ}caligraphic_U start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT that would rotate P𝑃Pitalic_P to a string in Pauli𝟙,Z(n)subscriptPauli1𝑍𝑛\text{\rm Pauli}_{\mathds{1},Z}(n)Pauli start_POSTSUBSCRIPT blackboard_1 , italic_Z end_POSTSUBSCRIPT ( italic_n ). Therefore, there is no P~Pauli𝟙,Z(n)~𝑃subscriptPauli1𝑍𝑛\tilde{P}\in\text{\rm Pauli}_{\mathds{1},Z}(n)over~ start_ARG italic_P end_ARG ∈ Pauli start_POSTSUBSCRIPT blackboard_1 , italic_Z end_POSTSUBSCRIPT ( italic_n ) such that P~=UPU~𝑃𝑈𝑃superscript𝑈\tilde{P}=UPU^{\dagger}over~ start_ARG italic_P end_ARG = italic_U italic_P italic_U start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT, so the eigenvalue of any Pauli string P𝑃Pitalic_P containing at least one single-site Y𝑌Yitalic_Y Pauli is λP=0subscript𝜆𝑃0\lambda_{P}=0italic_λ start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT = 0.

Corollary 3 (Random-XZ𝑋𝑍XZitalic_X italic_Z shadow norm of Pauli strings).

Assume a measurement primitive 𝒰XZsubscript𝒰𝑋𝑍\mathcal{U}_{XZ}caligraphic_U start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT such that the probability density function p(U)𝑝𝑈p(U)italic_p ( italic_U ) is uniform over random-XZ𝑋𝑍XZitalic_X italic_Z unitaries. Then, the shadow norm of a Pauli string PPauli𝟙,X,Z(n)𝑃subscriptPauli1𝑋𝑍𝑛P\in\text{\rm Pauli}_{\mathds{1},X,Z}(n)italic_P ∈ Pauli start_POSTSUBSCRIPT blackboard_1 , italic_X , italic_Z end_POSTSUBSCRIPT ( italic_n ) is

Pshadow2=2k(P),subscriptsuperscriptnorm𝑃2shadowsuperscript2𝑘𝑃\|P\|^{2}_{\text{shadow}}=2^{k(P)},∥ italic_P ∥ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT shadow end_POSTSUBSCRIPT = 2 start_POSTSUPERSCRIPT italic_k ( italic_P ) end_POSTSUPERSCRIPT , (95)

where P𝑃Pitalic_P is normalized with respect to the Hilbert-Schmidt inner product and k(P)𝑘𝑃k(P)italic_k ( italic_P ) is the locality of P𝑃Pitalic_P. Otherwise, for all Pauli strings PPauli𝟙,X,Z(n)𝑃subscriptPauli1𝑋𝑍𝑛P\notin\text{\rm Pauli}_{\mathds{1},X,Z}(n)italic_P ∉ Pauli start_POSTSUBSCRIPT blackboard_1 , italic_X , italic_Z end_POSTSUBSCRIPT ( italic_n ), the shadow norm is Pshadow2=subscriptsuperscriptnorm𝑃2shadow\|P\|^{2}_{\text{shadow}}=\infty∥ italic_P ∥ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT shadow end_POSTSUBSCRIPT = ∞, and, therefore, the observable cannot be learned.

Proof.

The shadow norm of an eigenoperator O𝑂Oitalic_O of the measurement channel \mathcal{M}caligraphic_M is Oshadow2=1/λOsubscriptsuperscriptnorm𝑂2shadow1subscript𝜆𝑂\|O\|^{2}_{\text{shadow}}=1/\lambda_{O}∥ italic_O ∥ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT shadow end_POSTSUBSCRIPT = 1 / italic_λ start_POSTSUBSCRIPT italic_O end_POSTSUBSCRIPT, where λOsubscript𝜆𝑂\lambda_{O}italic_λ start_POSTSUBSCRIPT italic_O end_POSTSUBSCRIPT is the eigenvalue of \mathcal{M}caligraphic_M corresponding to O𝑂Oitalic_O: (O)=λOO𝑂subscript𝜆𝑂𝑂\mathcal{M}(O)=\lambda_{O}Ocaligraphic_M ( italic_O ) = italic_λ start_POSTSUBSCRIPT italic_O end_POSTSUBSCRIPT italic_O (see Ref. [86] for a proof). The eigenoperators of our XZ𝑋𝑍XZitalic_X italic_Z measurement channel XZsubscript𝑋𝑍\mathcal{M}_{XZ}caligraphic_M start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT are the Pauli strings (by Lemma 1 of Ref. [88]). Therefore, the result follows from Proposition 2. ∎

The intuition behind this result is as follows. Our measurment channel XZsubscript𝑋𝑍\mathcal{M}_{XZ}caligraphic_M start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT has a nontrivial null space, and as such, there are eigenoperator(s) of this channel whose eigenvalue is zero. By definition, these eigenoperators can never be learned and are invisible under these measurement settings [15]. In the language of probabilities, we do not have access to information about these eigenoperators because the probability of measuring in the eigenbases of these operators is zero. This is further elucidated by considering the shadow norm: the shadow norm of an operator in the null space of the measurement channel will be infinite. Therefore, no matter how many measurements are taken, it can never be learned.

Corollary 4 (Visible space of random-XZ𝑋𝑍XZitalic_X italic_Z measurements).

For the set of unitaries in the ensemble 𝒰XZsubscript𝒰𝑋𝑍\mathcal{U}_{XZ}caligraphic_U start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT defined above, the VisibleSpace(𝒰XZ)VisibleSpacesubscript𝒰𝑋𝑍\textnormal{{VisibleSpace}}(\mathcal{U}_{XZ})VisibleSpace ( caligraphic_U start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT ) on n𝑛nitalic_n qubits is the span of Pauli𝟙,X,Z(n)subscriptPauli1𝑋𝑍𝑛\text{\rm Pauli}_{\mathds{1},X,Z}(n)Pauli start_POSTSUBSCRIPT blackboard_1 , italic_X , italic_Z end_POSTSUBSCRIPT ( italic_n ), where Pauli𝟙,X,Z(n)subscriptPauli1𝑋𝑍𝑛\text{\rm Pauli}_{\mathds{1},X,Z}(n)Pauli start_POSTSUBSCRIPT blackboard_1 , italic_X , italic_Z end_POSTSUBSCRIPT ( italic_n ) is the set of n𝑛nitalic_n-qubit Pauli strings constructed from only 𝟙1\mathds{1}blackboard_1s, X𝑋Xitalic_Xs, and Z𝑍Zitalic_Zs.

Proof.

Observation 4 of Ref. [15] states that if unitaries are sampled uniformly at random in the ensemble 𝒰𝒰\mathcal{U}caligraphic_U, then VisibleSpace(𝒰)VisibleSpace𝒰\text{\rm{VisibleSpace}}(\mathcal{U})VisibleSpace ( caligraphic_U ) is the image of the associated measurement channel \mathcal{M}caligraphic_M. Following Proposition 2, the image of the random-XZ𝑋𝑍XZitalic_X italic_Z measurement channel is the span of Pauli𝟙,X,Z(n)subscriptPauli1𝑋𝑍𝑛\text{\rm Pauli}_{\mathds{1},X,Z}(n)Pauli start_POSTSUBSCRIPT blackboard_1 , italic_X , italic_Z end_POSTSUBSCRIPT ( italic_n ): all Pauli strings composed of single-site X𝑋Xitalic_X, Z𝑍Zitalic_Z and 𝟙1\mathds{1}blackboard_1s. Therefore, the VisibleSpace(𝒰XZ)VisibleSpacesubscript𝒰𝑋𝑍\textnormal{{VisibleSpace}}(\mathcal{U}_{XZ})VisibleSpace ( caligraphic_U start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT ) is the span of Pauli𝟙,X,Z(n)subscriptPauli1𝑋𝑍𝑛\text{\rm Pauli}_{\mathds{1},X,Z}(n)Pauli start_POSTSUBSCRIPT blackboard_1 , italic_X , italic_Z end_POSTSUBSCRIPT ( italic_n ). ∎

This result also informs our intuition about the XZ𝑋𝑍XZitalic_X italic_Z visible space, the space of observables that we can estimate using the unitaries 𝒰XZsubscript𝒰𝑋𝑍\mathcal{U}_{XZ}caligraphic_U start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT. Recall that each classical shadow takes the form ρ^=XZ1(U|bb|U)\hat{\rho}=\mathcal{M}_{XZ}^{-1}(U^{\dagger}\lvert b\rangle\!\langle b\rvert U)over^ start_ARG italic_ρ end_ARG = caligraphic_M start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ( italic_U start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT | italic_b ⟩ ⟨ italic_b | italic_U ). Definition 3 in Ref. [15] states that U|bb|UVisibleSpace(𝒰XZ)U^{\dagger}\lvert b\rangle\!\langle b\rvert U\in\textnormal{{VisibleSpace}}(% \mathcal{U}_{XZ})italic_U start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT | italic_b ⟩ ⟨ italic_b | italic_U ∈ VisibleSpace ( caligraphic_U start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT ), and per the observations above, the \mathcal{M}caligraphic_M channel is effectively block diagonal with respect to the visible and invisible spaces. Therefore, ρ^VisibleSpace(𝒰XZ)^𝜌VisibleSpacesubscript𝒰𝑋𝑍\hat{\rho}\in\textnormal{{VisibleSpace}}(\mathcal{U}_{XZ})over^ start_ARG italic_ρ end_ARG ∈ VisibleSpace ( caligraphic_U start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT ), and in the infinite-measurement limit, the classical shadow will exactly be the projection of ρ𝜌\rhoitalic_ρ onto VisibleSpace(𝒰XZ)VisibleSpacesubscript𝒰𝑋𝑍\textnormal{{VisibleSpace}}(\mathcal{U}_{XZ})VisibleSpace ( caligraphic_U start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT ). As such, we can define the projected density matrix ρXZsubscript𝜌𝑋𝑍\rho_{XZ}italic_ρ start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT, as the portion of the true density matrix ρ𝜌\rhoitalic_ρ that lives in the VisibleSpace(𝒰XZ)VisibleSpacesubscript𝒰𝑋𝑍\textnormal{{VisibleSpace}}(\mathcal{U}_{XZ})VisibleSpace ( caligraphic_U start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT )

ρXZ=PPauli𝟙,X,Z(n)Tr(Pρ)P.subscript𝜌𝑋𝑍subscript𝑃subscriptPauli1𝑋𝑍𝑛Tr𝑃𝜌𝑃\rho_{XZ}=\sum_{P\in\text{\rm Pauli}_{\mathds{1},X,Z}(n)}\mathop{\mathrm{Tr}}(% P\rho)P.italic_ρ start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_P ∈ Pauli start_POSTSUBSCRIPT blackboard_1 , italic_X , italic_Z end_POSTSUBSCRIPT ( italic_n ) end_POSTSUBSCRIPT roman_Tr ( italic_P italic_ρ ) italic_P . (96)

This projected density matrix represents the part of ρ𝜌\rhoitalic_ρ that is learnable using random-XZ𝑋𝑍XZitalic_X italic_Z measurements. Consequently, when we use classical shadows generated by the XZ𝑋𝑍XZitalic_X italic_Z unitaries 𝒰XZsubscript𝒰𝑋𝑍\mathcal{U}_{XZ}caligraphic_U start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT to estimate Odelimited-⟨⟩𝑂\langle O\rangle⟨ italic_O ⟩ used in our regularization step in Appendix E, we find that we can only estimate operators O𝑂Oitalic_O that live in the visible space.

Finally, we conclude this appendix by using the results of the derivations above to find the expression for the inverse of the measurement channel XZsubscript𝑋𝑍\mathcal{M}_{XZ}caligraphic_M start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT. This inverse measurement channel is a crucial tool because, as discussed at the start of this appendix, it is used to compute the classical shadow. Here we will show that this inverse measurement channel XZ1superscriptsubscript𝑋𝑍1\mathcal{M}_{XZ}^{-1}caligraphic_M start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT takes the form stated in Eq. (C.1) in the previous section. First, we can re-express the forward channel XZsubscript𝑋𝑍\mathcal{M}_{XZ}caligraphic_M start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT in Proposition 2 in terms of independent, single-site channels XZ(1)superscriptsubscript𝑋𝑍1\mathcal{M}_{XZ}^{(1)}caligraphic_M start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT. This channel factorizes as

XZ(A)=(XZ(1))n(A), where XZ(1)(B)=14(Tr(BX)X+Tr(BZ)Z)+12Tr(B)𝕀formulae-sequencesubscript𝑋𝑍𝐴superscriptsuperscriptsubscript𝑋𝑍1tensor-productabsent𝑛𝐴 where superscriptsubscript𝑋𝑍1𝐵14Tr𝐵𝑋𝑋Tr𝐵𝑍𝑍12Tr𝐵𝕀\mathcal{M}_{XZ}(A)=\left(\mathcal{M}_{XZ}^{(1)}\right)^{\otimes n}(A),\hskip 1% 4.22636pt\textnormal{ where }\hskip 2.84526pt\mathcal{M}_{XZ}^{(1)}(B)=\frac{1% }{4}\left(\mathop{\mathrm{Tr}}(BX)X+\mathop{\mathrm{Tr}}(BZ)Z\right)+\frac{1}{% 2}\mathop{\mathrm{Tr}}(B)\mathbb{I}\hskip 8.53581ptcaligraphic_M start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT ( italic_A ) = ( caligraphic_M start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT ⊗ italic_n end_POSTSUPERSCRIPT ( italic_A ) , where caligraphic_M start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ( italic_B ) = divide start_ARG 1 end_ARG start_ARG 4 end_ARG ( roman_Tr ( italic_B italic_X ) italic_X + roman_Tr ( italic_B italic_Z ) italic_Z ) + divide start_ARG 1 end_ARG start_ARG 2 end_ARG roman_Tr ( italic_B ) blackboard_I (97)

for all A𝐴\hskip 2.84526ptAitalic_A and B𝐵Bitalic_B in the n𝑛nitalic_n-qubit and single-qubit visible spaces, respectively. Explicitly, an operator in the single-qubit visible subspace can be decomposed as follows:

B=12(Tr(BX)X+Tr(BZ)Z+Tr(B)𝕀).𝐵12Tr𝐵𝑋𝑋Tr𝐵𝑍𝑍Tr𝐵𝕀B=\frac{1}{2}\left(\mathop{\mathrm{Tr}}(BX)X+\mathop{\mathrm{Tr}}(BZ)Z+\mathop% {\mathrm{Tr}}(B)\mathbb{I}\right).italic_B = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( roman_Tr ( italic_B italic_X ) italic_X + roman_Tr ( italic_B italic_Z ) italic_Z + roman_Tr ( italic_B ) blackboard_I ) . (98)

Next, we notice that each single-site channel XZ(1)superscriptsubscript𝑋𝑍1\mathcal{M}_{XZ}^{(1)}caligraphic_M start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT is not only independent, but also the same on every site. Thus, all that is left is to take the Moore-Penrose pseudoinverse of this single qubit channel. This yields the following result:

1XZ(1)(B)=(Tr(BX)X+Tr(BZ)Z+12Tr(B)𝕀),superscriptsubscriptsuperscript1𝑋𝑍1𝐵Tr𝐵𝑋𝑋Tr𝐵𝑍𝑍12Tr𝐵𝕀{\mathcal{M}^{-1}}_{XZ}^{(1)}(B)=\left(\mathop{\mathrm{Tr}}(BX)X+\mathop{% \mathrm{Tr}}(BZ)Z+\frac{1}{2}\mathop{\mathrm{Tr}}(B)\mathbb{I}\right),caligraphic_M start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ( italic_B ) = ( roman_Tr ( italic_B italic_X ) italic_X + roman_Tr ( italic_B italic_Z ) italic_Z + divide start_ARG 1 end_ARG start_ARG 2 end_ARG roman_Tr ( italic_B ) blackboard_I ) , (99)

which matches our result in equation Eq. (C.1).

APPENDIX D Tomographic completeness of randomized XZ𝑋𝑍XZitalic_X italic_Z measurements

The target states considered in this work are real pure states. We are interested in such states because they are eigenstates of a Hamiltonian whose matrix elements are purely real, e.g., a Hamiltonian that is equivalent to a composition of X𝑋Xitalic_X and Z𝑍Zitalic_Z operators up to a global rotation. For example, if we have a (complex) Hamiltonian H=i,jZiZj+iYi𝐻subscript𝑖𝑗subscript𝑍𝑖subscript𝑍𝑗subscript𝑖subscript𝑌𝑖H=\sum_{i,j}Z_{i}Z_{j}+\sum_{i}Y_{i}italic_H = ∑ start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT italic_Z start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_Z start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT + ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_Y start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, then we can perform a rotation identically at each site such that ZiZisubscript𝑍𝑖subscript𝑍𝑖Z_{i}\rightarrow Z_{i}italic_Z start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT → italic_Z start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and YiXisubscript𝑌𝑖subscript𝑋𝑖Y_{i}\rightarrow X_{i}italic_Y start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT → italic_X start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. These Hamiltonians, which are interesting from the perspective of both condensed matter physics and quantum optimization [89, 90, 91], can suffer from a sign problem [92], so finding their ground states is a highly nontrivial task [93].

In this appendix, we will explore how and when we can fully characterize real pure states. These states can be expressed as a pure-state vector |ψ=bcb|bket𝜓subscript𝑏subscript𝑐𝑏ket𝑏\mbox{$|\psi\rangle$}=\sum_{b}c_{b}\mbox{$|b\rangle$}| italic_ψ ⟩ = ∑ start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT | italic_b ⟩ with real coefficients cb=cbsuperscriptsubscript𝑐𝑏subscript𝑐𝑏c_{b}^{*}=c_{b}italic_c start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = italic_c start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT. For simplicity, we will assume that the basis {|b}b{0,1}nsubscriptket𝑏𝑏superscript01𝑛\{\mbox{$|b\rangle$}\}_{b\in\{0,1\}^{n}}{ | italic_b ⟩ } start_POSTSUBSCRIPT italic_b ∈ { 0 , 1 } start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_POSTSUBSCRIPT is the computational basis, unless indicated otherwise. We will demonstrate that we can fully characterize these states by only measuring each qubit in either the X𝑋Xitalic_X or the Z𝑍Zitalic_Z basis (Lemma 5). While our numerical demonstrations are not always limited to measurements only in the X𝑋Xitalic_X or Z𝑍Zitalic_Z bases, our motivation for specifically pursuing measurement schemes of this type is that randomized-XZ𝑋𝑍XZitalic_X italic_Z measurements are tomographically complete on the subspace of real and pure states (Theorem 1).

Below, this appendix is separated into two parts. First, we will provide an overarching discussion of our formal statements in order to develop our intuition and contextualize them in our broader work. We will then provide rigorous proofs of our claims.

D.1 Discussion of results

As elaborated on in the previous appendix, we can parametrize the learning power of a measurement protocol by looking at the visible space. Recall that the visible space is the space of all operators O𝑂Oitalic_O for which we can estimate the expectation value Tr(ρO)Tr𝜌𝑂\mathop{\mathrm{Tr}}(\rho O)roman_Tr ( italic_ρ italic_O ). In particular, in Corollary 4, we defined the visible space corresponding to the unitaries in the random-XZ𝑋𝑍XZitalic_X italic_Z ensemble 𝒰XZsubscript𝒰𝑋𝑍\mathcal{U}_{XZ}caligraphic_U start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT. We found that the visible space is not the entire operator space ()\mathcal{B}(\mathcal{H})caligraphic_B ( caligraphic_H ). In other words, there are some observables that are invisible to us—they can never be learned. With random-XZ𝑋𝑍XZitalic_X italic_Z measurements, we can measure any Pauli string in Pauli𝟙,X,Z(n)subscriptPauli1𝑋𝑍𝑛\text{\rm Pauli}_{\mathds{1},X,Z}(n)Pauli start_POSTSUBSCRIPT blackboard_1 , italic_X , italic_Z end_POSTSUBSCRIPT ( italic_n ), the set of n𝑛nitalic_n-qubit Pauli strings constructed from only single-site 𝟙1\mathds{1}blackboard_1, X𝑋Xitalic_X, and Z𝑍Zitalic_Z Paulis. Therefore, the visible space is the span of all such Pauli strings. It turns out that this XZ𝑋𝑍XZitalic_X italic_Z measurement scheme is sufficient for learning states that live in the visible space as well as those with some components outside the invisible space, as long as certain assumptions are satisfied. One such example of this is a major result of this work: randomized-XZ𝑋𝑍XZitalic_X italic_Z measurements are tomographically complete on the subspace of real and pure states (Theorem 1).

Real, pure states do not live entirely in VisibleSpace(𝒰XZ)VisibleSpacesubscript𝒰𝑋𝑍\textnormal{{VisibleSpace}}(\mathcal{U}_{XZ})VisibleSpace ( caligraphic_U start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT ). For example, the GHZ states have nontrivial support on Pauli strings with an even number of single-site Y𝑌Yitalic_Y Paulis. Even though these states do not live entirely in VisibleSpace(𝒰XZ)VisibleSpacesubscript𝒰𝑋𝑍\textnormal{{VisibleSpace}}(\mathcal{U}_{XZ})VisibleSpace ( caligraphic_U start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT ), we can still fully characterize them with access to the visible XZ𝑋𝑍XZitalic_X italic_Z observables. This is because we can rely on the assumption that the state is both real and pure. Therefore, even though real, pure states are not fully encapsulated in the visible space, the XZ𝑋𝑍XZitalic_X italic_Z measurement scheme is sufficient for fully characterizing these states. To gain some intuition, let us consider a real, pure two-qubit state vector |ψket𝜓|\psi\rangle| italic_ψ ⟩. We can write this state as a density matrix and decompose it in terms of its support on all Pauli strings PPauli𝟙,X,Y,Z(n)𝑃subscriptPauli1𝑋𝑌𝑍𝑛P\in\text{\rm Pauli}_{\mathds{1},X,Y,Z}(n)italic_P ∈ Pauli start_POSTSUBSCRIPT blackboard_1 , italic_X , italic_Y , italic_Z end_POSTSUBSCRIPT ( italic_n ) as

|ψψ|\displaystyle\lvert\psi\rangle\!\langle\psi\rvert| italic_ψ ⟩ ⟨ italic_ψ | =\displaystyle== 12n(PPauli𝟙,X,Y,Z(n)cPP)1superscript2𝑛subscript𝑃subscriptPauli1𝑋𝑌𝑍𝑛subscript𝑐𝑃𝑃\displaystyle\frac{1}{2^{n}}\left(\sum_{P\in\text{\rm Pauli}_{\mathds{1},X,Y,Z% }(n)}c_{P}P\right)divide start_ARG 1 end_ARG start_ARG 2 start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_ARG ( ∑ start_POSTSUBSCRIPT italic_P ∈ Pauli start_POSTSUBSCRIPT blackboard_1 , italic_X , italic_Y , italic_Z end_POSTSUBSCRIPT ( italic_n ) end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT italic_P )
=\displaystyle== 12n(PPauli𝟙,X,Z(n)cPP+cYYYY).1superscript2𝑛subscriptsuperscript𝑃subscriptPauli1𝑋𝑍𝑛subscript𝑐superscript𝑃superscript𝑃subscript𝑐𝑌𝑌𝑌𝑌\displaystyle\frac{1}{2^{n}}\left(\sum_{P^{\prime}\in\text{\rm Pauli}_{\mathds% {1},X,Z}(n)}c_{P^{\prime}}P^{\prime}+c_{YY}YY\right).divide start_ARG 1 end_ARG start_ARG 2 start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_ARG ( ∑ start_POSTSUBSCRIPT italic_P start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∈ Pauli start_POSTSUBSCRIPT blackboard_1 , italic_X , italic_Z end_POSTSUBSCRIPT ( italic_n ) end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT italic_P start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT italic_P start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT + italic_c start_POSTSUBSCRIPT italic_Y italic_Y end_POSTSUBSCRIPT italic_Y italic_Y ) .

Since the state is real, its support must be zero on strings with an odd number of Y𝑌Yitalic_Ys, and therefore, there is only one term remaining that contains single-site Y𝑌Yitalic_Y Paulis: the two-qubit string YYtensor-product𝑌𝑌Y\otimes Yitalic_Y ⊗ italic_Y. As a result, this two-qubit state does not live entirely in the visible space because YYVisibleSpace(𝒰XZ)tensor-product𝑌𝑌VisibleSpacesubscript𝒰𝑋𝑍Y\otimes Y\notin\textnormal{{VisibleSpace}}(\mathcal{U}_{XZ})italic_Y ⊗ italic_Y ∉ VisibleSpace ( caligraphic_U start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT ). However, we can still estimate the coefficient cYYsubscript𝑐𝑌𝑌c_{YY}italic_c start_POSTSUBSCRIPT italic_Y italic_Y end_POSTSUBSCRIPT using the purity condition: Tr(|ψψ|2)=1\mathrm{Tr}(\lvert\psi\rangle\!\langle\psi\rvert^{2})=1roman_Tr ( | italic_ψ ⟩ ⟨ italic_ψ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) = 1. In other words, since our XZ𝑋𝑍XZitalic_X italic_Z measurements allow us to learn all coefficients except cYYsubscript𝑐𝑌𝑌c_{YY}italic_c start_POSTSUBSCRIPT italic_Y italic_Y end_POSTSUBSCRIPT, we can solve for cYYsubscript𝑐𝑌𝑌c_{YY}italic_c start_POSTSUBSCRIPT italic_Y italic_Y end_POSTSUBSCRIPT using this extra constraint. While, so far, we have been considering a real, pure n=2𝑛2n=2italic_n = 2 qubit state, it turns out that this holds true for any n𝑛nitalic_n-qubit real, pure state. The information from the XZ𝑋𝑍XZitalic_X italic_Z measurements allows us to extract the support of |ψψ|\lvert\psi\rangle\!\langle\psi\rvert| italic_ψ ⟩ ⟨ italic_ψ | on Pauli strings containing an even number of Y𝑌Yitalic_Ys. This will be shown rigorously in the following subsection.

D.2 Proof of Theorem 1

In this subsection, we prove that XZ𝑋𝑍XZitalic_X italic_Z measurements, where each qubit is measured in either the X𝑋Xitalic_X or the Z𝑍Zitalic_Z basis, are tomographically complete on the space of real pure states. Tomographic completeness is similar to informational completeness but defined with respect to state characterization. A measurement primitive, which samples U𝒰similar-to𝑈𝒰U\sim\mathcal{U}italic_U ∼ caligraphic_U and subsequently measures the rotated state in the computational basis {|b:b{0,1}n}conditional-setket𝑏𝑏superscript01𝑛\{\mbox{$|b\rangle$}:b\in\{0,1\}^{n}\}{ | italic_b ⟩ : italic_b ∈ { 0 , 1 } start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT }, is tomographically complete if and only if for each ρσ𝜌𝜎\rho\neq\sigmaitalic_ρ ≠ italic_σ, there exist U𝒰𝑈𝒰U\in\mathcal{U}italic_U ∈ caligraphic_U and b𝑏bitalic_b such that b|UρU|bb|UσU|bbra𝑏𝑈𝜌superscript𝑈ket𝑏bra𝑏𝑈𝜎superscript𝑈ket𝑏\mbox{$\langle b|$}U\rho U^{\dagger}\mbox{$|b\rangle$}\neq\mbox{$\langle b|$}U% \sigma U^{\dagger}\mbox{$|b\rangle$}⟨ italic_b | italic_U italic_ρ italic_U start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT | italic_b ⟩ ≠ ⟨ italic_b | italic_U italic_σ italic_U start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT | italic_b ⟩ [25]. For our purposes, we are only interested in establishing tomographic completeness on a subspace of the density matrices, namely, all states that are both real and pure.

Theorem 1 states that real, pure states are tomographically complete under XZ𝑋𝑍XZitalic_X italic_Z measurements. While our theorem might at first be surprising, it demonstrates that there is always a unitary U𝒰XZsimilar-to𝑈subscript𝒰𝑋𝑍U\sim\mathcal{U}_{XZ}italic_U ∼ caligraphic_U start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT such that its measurement will allow us to differentiate between real, pure states ρ𝜌\rhoitalic_ρ and σ𝜎\sigmaitalic_σ. Furthermore, it elucidates that the visible space does not immediately render an understanding of tomographic completeness in the presence of extra assumptions.

Before providing the rigorous details, we first outline a sketch of the proof. Theorem 1 follows immediately from Lemma 5, which shows that we can fully characterize a real state vector |ψket𝜓|\psi\rangle| italic_ψ ⟩ up to a global phase using only XZ𝑋𝑍XZitalic_X italic_Z measurements. We prove Lemma 5 by induction on the number of qubits k𝑘kitalic_k. Our base case is a real, pure state on k=1𝑘1k=1italic_k = 1 qubits. This can be immediately learned by measuring in the X𝑋Xitalic_X and Z𝑍Zitalic_Z bases. The inductive step (k𝑘kitalic_k qubits) involves first learning all the amplitudes |cb|subscript𝑐𝑏|c_{b}|| italic_c start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT | of |ψket𝜓|\psi\rangle| italic_ψ ⟩, which can again be expressed as

|ψ=b{0,1}ncb|b.ket𝜓subscript𝑏superscript01𝑛subscript𝑐𝑏ket𝑏\mbox{$|\psi\rangle$}=\sum_{b\in\{0,1\}^{n}}c_{b}\mbox{$|b\rangle$}.| italic_ψ ⟩ = ∑ start_POSTSUBSCRIPT italic_b ∈ { 0 , 1 } start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT | italic_b ⟩ . (101)

We learn the amplitudes by measuring every qubit in the Z𝑍Zitalic_Z basis. Then, we use our inductive assumption to argue that we can always characterize all relative signs using XZ𝑋𝑍XZitalic_X italic_Z measurements. In particular, our technique is more powerful than simply proving learnability, because we explicitly provide an algorithm by which one can use these measurements to fully characterize a real, pure state up to a global sign. Therefore, we show that XZ𝑋𝑍XZitalic_X italic_Z measurements are tomographically complete on the space of real pure states. We leave the details of this algorithm to the interested reader (see below).

Lemma 5 (Random-XZ𝑋𝑍XZitalic_X italic_Z measurements are sufficient for characterizing a real, pure state).

Up to a global sign, we can uniquely characterize any real state vector |ψket𝜓|\psi\rangle| italic_ψ ⟩ using XZ measurements, formally defined as measuring any qubit in either the X or the Z basis. We represent our real pure state in the computational basis {|b}b{0,1}nsubscriptket𝑏𝑏superscript01𝑛\{\mbox{$|b\rangle$}\}_{b\in\{0,1\}^{n}}{ | italic_b ⟩ } start_POSTSUBSCRIPT italic_b ∈ { 0 , 1 } start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_POSTSUBSCRIPT as |ψ=bcb|bket𝜓subscript𝑏subscript𝑐𝑏ket𝑏\mbox{$|\psi\rangle$}=\sum_{b}c_{b}\mbox{$|b\rangle$}| italic_ψ ⟩ = ∑ start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT | italic_b ⟩, where cb=cbsuperscriptsubscript𝑐𝑏subscript𝑐𝑏c_{b}^{*}=c_{b}italic_c start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = italic_c start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT and b|cb|2=1subscript𝑏superscriptsubscript𝑐𝑏21\sum_{b}|c_{b}|^{2}=1∑ start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT | italic_c start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = 1. We uniquely characterize the state by learning each cbsubscript𝑐𝑏c_{b}italic_c start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT up to a global sign.

Proof.

To fully characterize the state vector |ψket𝜓|\psi\rangle| italic_ψ ⟩, we must learn all coefficients {cb}subscript𝑐𝑏\{c_{b}\}{ italic_c start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT } up to a global sign. Each coefficient is characterized by its amplitude and sign: cb=sgn(cb)×|cb|subscript𝑐𝑏sgnsubscript𝑐𝑏subscript𝑐𝑏c_{b}=\mathop{\mathrm{sgn}}(c_{b})\times|c_{b}|italic_c start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT = roman_sgn ( italic_c start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ) × | italic_c start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT |. We can extract the amplitudes by measuring all qubits in the Z𝑍Zitalic_Z basis – whereafter, it remains to learn the relative signs. The remainder of this proof will show how we can learn the relative signs with XZ𝑋𝑍XZitalic_X italic_Z measurements. We will proceed by induction on the number of qubits k𝑘kitalic_k.

Base case (k=1𝑘1k=1italic_k = 1 qubits). We extract the relative sign between |0ket0|0\rangle| 0 ⟩ and |1ket1|1\rangle| 1 ⟩ by learning the sign of the expectation value ψ|X|ψdelimited-⟨⟩𝜓𝑋𝜓\langle\psi\lvert X\rvert\psi\rangle⟨ italic_ψ | italic_X | italic_ψ ⟩: sgn(ψ|X|ψ)sgndelimited-⟨⟩𝜓𝑋𝜓\mathop{\mathrm{sgn}}(\langle\psi\lvert X\rvert\psi\rangle)roman_sgn ( ⟨ italic_ψ | italic_X | italic_ψ ⟩ ).

Inductive step (k𝑘kitalic_k qubits). Assume we can learn all the relative signs of a real, pure state on k1𝑘1k-1italic_k - 1 qubits with XZ𝑋𝑍XZitalic_X italic_Z measurements. We have a k𝑘kitalic_k-qubit state. If we measure our k𝑘kitalic_kth qubit in the computational Z𝑍Zitalic_Z basis, then, by the inductive assumption, we can learn all the relative signs in the {|0|b~}b~{0,1}k1subscripttensor-productket0ket~𝑏~𝑏superscript01𝑘1\{\mbox{$|0\rangle$}\otimes\mbox{$|\tilde{b}\rangle$}\}_{\tilde{b}\in\{0,1\}^{% k-1}}{ | 0 ⟩ ⊗ | over~ start_ARG italic_b end_ARG ⟩ } start_POSTSUBSCRIPT over~ start_ARG italic_b end_ARG ∈ { 0 , 1 } start_POSTSUPERSCRIPT italic_k - 1 end_POSTSUPERSCRIPT end_POSTSUBSCRIPT subspace as well as in the {|1|b~}b~{0,1}k1subscripttensor-productket1ket~𝑏~𝑏superscript01𝑘1\{\mbox{$|1\rangle$}\otimes\mbox{$|\tilde{b}\rangle$}\}_{\tilde{b}\in\{0,1\}^{% k-1}}{ | 1 ⟩ ⊗ | over~ start_ARG italic_b end_ARG ⟩ } start_POSTSUBSCRIPT over~ start_ARG italic_b end_ARG ∈ { 0 , 1 } start_POSTSUPERSCRIPT italic_k - 1 end_POSTSUPERSCRIPT end_POSTSUBSCRIPT subspace. Here, the tilde on b~~𝑏\tilde{b}over~ start_ARG italic_b end_ARG indicates that the bit-string is on k1𝑘1k-1italic_k - 1 qubits instead of k𝑘kitalic_k. Also note that when we project the k𝑘kitalic_kth qubit of our state vector |ψket𝜓|\psi\rangle| italic_ψ ⟩ onto |0ket0|0\rangle| 0 ⟩ or |1ket1|1\rangle| 1 ⟩, the state on the remaining k1𝑘1k-1italic_k - 1 qubits is projected onto a real pure state.

What is, therefore, left to be shown is that we can learn the relative sign between these two subspaces. If the amplitudes of all states in either of the subspaces are zero, then the relative sign is inconsequential because we already know the state up to a global phase. Otherwise, there must exist at least one state in each subspace that has nonzero amplitude. Let us choose two computational basis states: b(0)superscript𝑏0b^{(0)}italic_b start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT from the {|0|b~}b~subscripttensor-productket0ket~𝑏~𝑏\{\mbox{$|0\rangle$}\otimes\mbox{$|\tilde{b}\rangle$}\}_{\tilde{b}}{ | 0 ⟩ ⊗ | over~ start_ARG italic_b end_ARG ⟩ } start_POSTSUBSCRIPT over~ start_ARG italic_b end_ARG end_POSTSUBSCRIPT subspace and b(1)superscript𝑏1b^{(1)}italic_b start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT from the {|1|b~}b~subscripttensor-productket1ket~𝑏~𝑏\{\mbox{$|1\rangle$}\otimes\mbox{$|\tilde{b}\rangle$}\}_{\tilde{b}}{ | 1 ⟩ ⊗ | over~ start_ARG italic_b end_ARG ⟩ } start_POSTSUBSCRIPT over~ start_ARG italic_b end_ARG end_POSTSUBSCRIPT subspace, respectively. We choose them such that they that have nonzero amplitude and, among all possible choices, they have the smallest Hamming weight 𝐝(b(0),b(1))𝐝superscript𝑏0superscript𝑏1\mathbf{d}(b^{(0)},b^{(1)})bold_d ( italic_b start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT , italic_b start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ); this Hamming weight must be greater than or equal to 1 because the k𝑘kitalic_kth bit will always differ. We can extract the relative sign of the two spaces by estimating the observable OVisibleSpace(𝒰XZ)𝑂VisibleSpacesubscript𝒰𝑋𝑍O\in\textnormal{{VisibleSpace}}(\mathcal{U}_{XZ})italic_O ∈ VisibleSpace ( caligraphic_U start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT ) given by

O=i=1kOi;Oi={|++|if i=k,|bi(0)bi(0)|if bi(0)=bi(1),Xotherwise,O=\otimes_{i=1}^{k}O_{i};\hskip 11.38109ptO_{i}=\begin{cases}\lvert+\rangle\!% \langle+\rvert&\quad\text{if }i=k,\\ \left\rvert b^{(0)}_{i}\right\rangle\left\langle b^{(0)}_{i}\right\lvert&\quad% \text{if }b^{(0)}_{i}=b^{(1)}_{i},\\ X&\quad\text{otherwise},\\ \end{cases}italic_O = ⊗ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT italic_O start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ; italic_O start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = { start_ROW start_CELL | + ⟩ ⟨ + | end_CELL start_CELL if italic_i = italic_k , end_CELL end_ROW start_ROW start_CELL | italic_b start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ⟩ ⟨ italic_b start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | end_CELL start_CELL if italic_b start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_b start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , end_CELL end_ROW start_ROW start_CELL italic_X end_CELL start_CELL otherwise , end_CELL end_ROW (102)

where bi(0)subscriptsuperscript𝑏0𝑖b^{(0)}_{i}italic_b start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is the i𝑖iitalic_ith bit of bit-string b(0)superscript𝑏0b^{(0)}italic_b start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT, and likewise bi(1)subscriptsuperscript𝑏1𝑖b^{(1)}_{i}italic_b start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is the i𝑖iitalic_ith bit of b(1)superscript𝑏1b^{(1)}italic_b start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT.

Consider first the case when the Hamming weight is 𝐝(b(0),b(1))=1𝐝superscript𝑏0superscript𝑏11\mathbf{d}(b^{(0)},b^{(1)})=1bold_d ( italic_b start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT , italic_b start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ) = 1. Then, since the k𝑘kitalic_kth qubit must always differ between the two states, all of the remaining bits are the same: ik,bi(0)=bi(1)formulae-sequencefor-all𝑖𝑘subscriptsuperscript𝑏0𝑖subscriptsuperscript𝑏1𝑖\forall i\neq k,b^{(0)}_{i}=b^{(1)}_{i}∀ italic_i ≠ italic_k , italic_b start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_b start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. Therefore, ψ|O|ψdelimited-⟨⟩𝜓𝑂𝜓\langle\psi\lvert O\rvert\psi\rangle⟨ italic_ψ | italic_O | italic_ψ ⟩ takes the form |cb(0)+cb(1)|2superscriptsubscript𝑐superscript𝑏0subscript𝑐superscript𝑏12|c_{b^{(0)}}+c_{b^{(1)}}|^{2}| italic_c start_POSTSUBSCRIPT italic_b start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT end_POSTSUBSCRIPT + italic_c start_POSTSUBSCRIPT italic_b start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. Since we already know |cb(0)|subscript𝑐superscript𝑏0|c_{b^{(0)}}|| italic_c start_POSTSUBSCRIPT italic_b start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT end_POSTSUBSCRIPT | and |cb(1)|subscript𝑐superscript𝑏1|c_{b^{(1)}}|| italic_c start_POSTSUBSCRIPT italic_b start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT end_POSTSUBSCRIPT |, and by assumption they are nonzero, we can uniquely identify the relative sign between the subspaces. Next, we consider the case when the Hamming weight is 𝐝(b(0),b(1))>1𝐝superscript𝑏0superscript𝑏11\mathbf{d}(b^{(0)},b^{(1)})>1bold_d ( italic_b start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT , italic_b start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ) > 1 and learn ψ|O|ψdelimited-⟨⟩𝜓𝑂𝜓\langle\psi\lvert O\rvert\psi\rangle⟨ italic_ψ | italic_O | italic_ψ ⟩. Notice that O𝑂Oitalic_O’s projection of |ψket𝜓|\psi\rangle| italic_ψ ⟩ onto the matching bits in b(0)superscript𝑏0b^{(0)}italic_b start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT and b(1)superscript𝑏1b^{(1)}italic_b start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT leaves only two (nonzero amplitude) basis state components:

(i s.t. bi(0)=bi(1)bi(0)|)|ψ=cb(0)(i s.t. bi(0)=bi(1)bi(0)|)|b(0)+cb(1)(i s.t. bi(0)=bi(1)bi(0)|)|b(1).\left(\otimes_{i\textnormal{ s.t. }b^{(0)}_{i}=b^{(1)}_{i}}\left\langle b^{(0)% }_{i}\right\lvert\right)\mbox{$|\psi\rangle$}=c_{b^{(0)}}\left(\otimes_{i% \textnormal{ s.t. }b^{(0)}_{i}=b^{(1)}_{i}}\left\langle b^{(0)}_{i}\right% \lvert\right)\mbox{$|b^{(0)}\rangle$}+c_{b^{(1)}}\left(\otimes_{i\textnormal{ % s.t. }b^{(0)}_{i}=b^{(1)}_{i}}\left\langle b^{(0)}_{i}\right\lvert\right)\mbox% {$|b^{(1)}\rangle$}.( ⊗ start_POSTSUBSCRIPT italic_i s.t. italic_b start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_b start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT ⟨ italic_b start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | ) | italic_ψ ⟩ = italic_c start_POSTSUBSCRIPT italic_b start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ( ⊗ start_POSTSUBSCRIPT italic_i s.t. italic_b start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_b start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT ⟨ italic_b start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | ) | italic_b start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT ⟩ + italic_c start_POSTSUBSCRIPT italic_b start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ( ⊗ start_POSTSUBSCRIPT italic_i s.t. italic_b start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_b start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT ⟨ italic_b start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | ) | italic_b start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ⟩ . (103)

This is because we chose the pair b(0)superscript𝑏0b^{(0)}italic_b start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT and b(1)superscript𝑏1b^{(1)}italic_b start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT such that they have the smallest Hamming weight among all possible pairings. Then, the X𝑋Xitalic_X operator flips any remaining bits at ik𝑖𝑘i\neq kitalic_i ≠ italic_k, and we obtain a state with precisely swapped amplitudes

(i s.t. bi(0)=bi(1)bi(0)|)(cb(0)|b(1)+cb(1)|b(0)).\left(\otimes_{i\textnormal{ s.t. }b^{(0)}_{i}=b^{(1)}_{i}}\left\langle b^{(0)% }_{i}\right\lvert\right)\left(c_{b^{(0)}}\mbox{$|b^{(1)}\rangle$}+c_{b^{(1)}}% \mbox{$|b^{(0)}\rangle$}\right).( ⊗ start_POSTSUBSCRIPT italic_i s.t. italic_b start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_b start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT ⟨ italic_b start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | ) ( italic_c start_POSTSUBSCRIPT italic_b start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT end_POSTSUBSCRIPT | italic_b start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ⟩ + italic_c start_POSTSUBSCRIPT italic_b start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT end_POSTSUBSCRIPT | italic_b start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT ⟩ ) . (104)

Consequently, evaluating the expectation value yields ψ|O|ψ=cb(0)×cb(1)delimited-⟨⟩𝜓𝑂𝜓subscript𝑐superscript𝑏0subscript𝑐superscript𝑏1\langle\psi\lvert O\rvert\psi\rangle=c_{b^{(0)}}\times c_{b^{(1)}}⟨ italic_ψ | italic_O | italic_ψ ⟩ = italic_c start_POSTSUBSCRIPT italic_b start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT end_POSTSUBSCRIPT × italic_c start_POSTSUBSCRIPT italic_b start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT end_POSTSUBSCRIPT. Since |cb(0)|subscript𝑐superscript𝑏0|c_{b^{(0)}}|| italic_c start_POSTSUBSCRIPT italic_b start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT end_POSTSUBSCRIPT | and |cb(1)|subscript𝑐superscript𝑏1|c_{b^{(1)}}|| italic_c start_POSTSUBSCRIPT italic_b start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT end_POSTSUBSCRIPT | are known and (by assumption) nonzero, we can uniquely identify the relative sign between the subspaces.

Theorem 1 (Tomographic completeness of random-XZ𝑋𝑍XZitalic_X italic_Z measurements on the space of real, pure states).

A measurement primitive which samples U𝒰XZsimilar-to𝑈subscript𝒰𝑋𝑍U\sim\mathcal{U}_{XZ}italic_U ∼ caligraphic_U start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT and subsequently measures the rotated state in the computational basis {|b:b{0,1}n}conditional-setket𝑏𝑏superscript01𝑛\{\mbox{$|b\rangle$}:b\in\{0,1\}^{n}\}{ | italic_b ⟩ : italic_b ∈ { 0 , 1 } start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT } is tomographically complete [25] on the space of real, pure states. Namely, for all states |ψψ||ϕϕ|\lvert\psi\rangle\!\langle\psi\rvert\neq\lvert\phi\rangle\!\langle\phi\rvert| italic_ψ ⟩ ⟨ italic_ψ | ≠ | italic_ϕ ⟩ ⟨ italic_ϕ |, there exist U𝒰XZ𝑈subscript𝒰𝑋𝑍U\in\mathcal{U}_{XZ}italic_U ∈ caligraphic_U start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT and b𝑏bitalic_b such that b|UρU|bb|UσU|bbra𝑏𝑈𝜌superscript𝑈ket𝑏bra𝑏𝑈𝜎superscript𝑈ket𝑏\mbox{$\langle b|$}U\rho U^{\dagger}\mbox{$|b\rangle$}\neq\mbox{$\langle b|$}U% \sigma U^{\dagger}\mbox{$|b\rangle$}⟨ italic_b | italic_U italic_ρ italic_U start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT | italic_b ⟩ ≠ ⟨ italic_b | italic_U italic_σ italic_U start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT | italic_b ⟩.

Proof.

We proceed by contradiction. Assume that this measurement primitive is not tomographically complete on the space of real, pure states. Then, there exist real, pure states |ψψ||ϕϕ|\lvert\psi\rangle\!\langle\psi\rvert\neq\lvert\phi\rangle\!\langle\phi\rvert| italic_ψ ⟩ ⟨ italic_ψ | ≠ | italic_ϕ ⟩ ⟨ italic_ϕ | such that

U,bb|U|ψψ|U|b=b|U|ϕϕ|U|b.\forall\,U,b\hskip 5.69054pt\mbox{$\langle b|$}U\lvert\psi\rangle\!\langle\psi% \rvert U^{\dagger}\mbox{$|b\rangle$}=\mbox{$\langle b|$}U\lvert\phi\rangle\!% \langle\phi\rvert U^{\dagger}\mbox{$|b\rangle$}.∀ italic_U , italic_b ⟨ italic_b | italic_U | italic_ψ ⟩ ⟨ italic_ψ | italic_U start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT | italic_b ⟩ = ⟨ italic_b | italic_U | italic_ϕ ⟩ ⟨ italic_ϕ | italic_U start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT | italic_b ⟩ . (105)

In other words, after the application of any U𝑈Uitalic_U, the resulting computational basis state probabilities will always be the same: we cannot differentiate between |ψψ|\lvert\psi\rangle\!\langle\psi\rvert| italic_ψ ⟩ ⟨ italic_ψ | and |ϕϕ|\lvert\phi\rangle\!\langle\phi\rvert| italic_ϕ ⟩ ⟨ italic_ϕ | with this measurement primitive. However, Lemma 5 shows that this same measurement primitive can uniquely characterize, up to a global phase, both |ψket𝜓|\psi\rangle| italic_ψ ⟩ and |ϕketitalic-ϕ|\phi\rangle| italic_ϕ ⟩. Therefore, since these states differ by more than just some global phase (|ψψ||ϕϕ|\lvert\psi\rangle\!\langle\psi\rvert\neq\lvert\phi\rangle\!\langle\phi\rvert| italic_ψ ⟩ ⟨ italic_ψ | ≠ | italic_ϕ ⟩ ⟨ italic_ϕ | by definition), we arrive at a contradiction. ∎

APPENDIX E Loss function

This appendix presents the details of the loss function in Eq. (5). In the first subsection, we provide some background for maximum likelihood estimation and derive its contribution as the first term in our loss function. In the next subsection, we explicitly state the physical observables that we use for regularization in our numerical results. While kee** most of the section self-contained, some of the notations and terminologies used have been set up in Appendix C.

E.1 Maximum likelihood estimation: Randomized Kullback–Leibler divergence

First, we review the standard framework of maximum likelihood estimation (MLE), before applying such a loss function to our specific setting of quantum measurements. In general, we observe data from an underlying probability distribution 𝒟:X:𝒟𝑋\mathcal{D}:X\rightarrow\mathbb{R}caligraphic_D : italic_X → blackboard_R on the dataset domain X𝑋Xitalic_X, and our goal is to find the most likely model with parameters A𝐴Aitalic_A denoted as 𝒬A:X:subscript𝒬𝐴𝑋\mathcal{Q}_{A}:X\rightarrow\mathbb{R}caligraphic_Q start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT : italic_X → blackboard_R that is consistent with our observations. To find such a model, we would want to variationally optimize for the model’s parameters A𝐴Aitalic_A by minimizing the Kullback–Leibler (KL) divergence

K(𝒟|𝒬A)𝐾conditional𝒟subscript𝒬𝐴\displaystyle K(\mathcal{D}|\mathcal{Q}_{A})italic_K ( caligraphic_D | caligraphic_Q start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ) =xX𝒟(x)log𝒬A(x)𝒟(x).absentsubscript𝑥𝑋𝒟𝑥subscript𝒬𝐴𝑥𝒟𝑥\displaystyle=-\sum_{x\in X}\mathcal{D}(x)\log{\frac{\mathcal{Q}_{A}(x)}{% \mathcal{D}(x)}}.= - ∑ start_POSTSUBSCRIPT italic_x ∈ italic_X end_POSTSUBSCRIPT caligraphic_D ( italic_x ) roman_log divide start_ARG caligraphic_Q start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ( italic_x ) end_ARG start_ARG caligraphic_D ( italic_x ) end_ARG . (106)

Since the data distribution 𝒟𝒟\mathcal{D}caligraphic_D is a constant, we just need to minimize the cross-entropy contribution to the KL divergence

(𝒟|𝒬A)conditional𝒟subscript𝒬𝐴\displaystyle\mathds{H}(\mathcal{D}|\mathcal{Q}_{A})blackboard_H ( caligraphic_D | caligraphic_Q start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ) =xX𝒟(x)log𝒬A(x).absentsubscript𝑥𝑋𝒟𝑥subscript𝒬𝐴𝑥\displaystyle=-\sum_{x\in X}\mathcal{D}(x)\log{\mathcal{Q}_{A}(x)}.= - ∑ start_POSTSUBSCRIPT italic_x ∈ italic_X end_POSTSUBSCRIPT caligraphic_D ( italic_x ) roman_log caligraphic_Q start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ( italic_x ) . (107)

More specifically, for our purposes, the dataset consists of bit-string measurements of a target state |ϕϕ|\rvert\phi\rangle\!\langle\phi\lvert| italic_ϕ ⟩ ⟨ italic_ϕ | in the computational basis after applying a unitary U𝑈Uitalic_U. In a randomized measurement scheme, such as the XZ𝑋𝑍XZitalic_X italic_Z-shadow measurements discussed in Appendix C, we first draw a unitary from an ensemble 𝒰𝒰\mathcal{U}caligraphic_U with a uniform probability over a set {U}𝑈\{U\}{ italic_U }. Then, we measure the bit-string b𝑏bitalic_b in the computational basis. Taken together, we sample a data point, which is a tuple of the sampled unitary and the measured bit-string xi=(Ui,bi)subscript𝑥𝑖subscript𝑈𝑖subscript𝑏𝑖x_{i}=(U_{i},b_{i})italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = ( italic_U start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ), with probability

𝒟[(U,b)]=1|{U}|×b|U|ϕϕ|U|b.𝒟delimited-[]𝑈𝑏1𝑈delimited-⟨⟩𝑏𝑈italic-ϕdelimited-⟨⟩italic-ϕsuperscript𝑈𝑏\mathcal{D}[(U,b)]=\frac{1}{\lvert\{U\}\rvert}\times\langle b\lvert U\rvert% \phi\rangle\!\langle\phi\lvert U^{\dagger}\rvert b\rangle.caligraphic_D [ ( italic_U , italic_b ) ] = divide start_ARG 1 end_ARG start_ARG | { italic_U } | end_ARG × ⟨ italic_b | italic_U | italic_ϕ ⟩ ⟨ italic_ϕ | italic_U start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT | italic_b ⟩ . (108)

Similarly, we define the probability of observing this data point from a model parametrized by A𝐴Aitalic_A as 𝒬Asubscript𝒬𝐴\mathcal{Q}_{A}caligraphic_Q start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT. Then, the randomized KL divergence is

K(𝒟|𝒬A)𝐾conditional𝒟subscript𝒬𝐴\displaystyle K(\mathcal{D}|\mathcal{Q}_{A})italic_K ( caligraphic_D | caligraphic_Q start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ) =𝔼U𝒰b{0,1}n𝒟(U,b)log𝒬A(U,b)𝒟(U,b).absentsubscript𝔼similar-to𝑈𝒰subscript𝑏superscript01𝑛𝒟𝑈𝑏subscript𝒬𝐴𝑈𝑏𝒟𝑈𝑏\displaystyle=-\operatorname{\mathbb{E}}_{U\sim\mathcal{U}}\sum_{b\in\{0,1\}^{% n}}\mathcal{D}(U,b)\log{\frac{\mathcal{Q}_{A}(U,b)}{\mathcal{D}(U,b)}}.= - blackboard_E start_POSTSUBSCRIPT italic_U ∼ caligraphic_U end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_b ∈ { 0 , 1 } start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_POSTSUBSCRIPT caligraphic_D ( italic_U , italic_b ) roman_log divide start_ARG caligraphic_Q start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ( italic_U , italic_b ) end_ARG start_ARG caligraphic_D ( italic_U , italic_b ) end_ARG . (109a)

As in Eq. (107), we can drop the constant factor from the data distribution and obtain the randomized cross entropy or the infinite-sample loss function

(𝒟|𝒬A)=𝔼U𝒰b{0,1}n𝒟(U,b)log[𝒬A(U,b)].conditional𝒟subscript𝒬𝐴subscript𝔼similar-to𝑈𝒰subscript𝑏superscript01𝑛𝒟𝑈𝑏subscript𝒬𝐴𝑈𝑏\mathds{H}(\mathcal{D}|\mathcal{Q}_{A})=-\operatorname{\mathbb{E}}_{U\sim% \mathcal{U}}\sum_{b\in\{0,1\}^{n}}\mathcal{D}(U,b)\log\left[\mathcal{Q}_{A}(U,% b)\right].blackboard_H ( caligraphic_D | caligraphic_Q start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ) = - blackboard_E start_POSTSUBSCRIPT italic_U ∼ caligraphic_U end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_b ∈ { 0 , 1 } start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_POSTSUBSCRIPT caligraphic_D ( italic_U , italic_b ) roman_log [ caligraphic_Q start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ( italic_U , italic_b ) ] . (110)

For N𝑁Nitalic_N-samples i.i.d. drawn from 𝒟𝒟\mathcal{D}caligraphic_D as our dataset denoted by D={(Ui,bi)}i=1N𝐷superscriptsubscriptsubscript𝑈𝑖subscript𝑏𝑖𝑖1𝑁D=\{(U_{i},b_{i})\}_{i=1}^{N}italic_D = { ( italic_U start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) } start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT, this leads to the negative log-likelihood (NLL) contribution to our loss function which is

LNLLD(A)=1Ni=1Nlog𝒬A(bi,Ui).subscriptsuperscript𝐿𝐷NLL𝐴1𝑁superscriptsubscript𝑖1𝑁subscript𝒬𝐴subscript𝑏𝑖subscript𝑈𝑖L^{D}_{\textrm{NLL}}(A)=-\frac{1}{N}\sum_{i=1}^{N}\log\mathcal{Q}_{A}(b_{i},U_% {i}).italic_L start_POSTSUPERSCRIPT italic_D end_POSTSUPERSCRIPT start_POSTSUBSCRIPT NLL end_POSTSUBSCRIPT ( italic_A ) = - divide start_ARG 1 end_ARG start_ARG italic_N end_ARG ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT roman_log caligraphic_Q start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ( italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_U start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) . (111)

In this work, we focus on a model distribution given by a MPS |ψχ(A)ketsubscript𝜓𝜒𝐴|\psi_{\chi}(A)\rangle| italic_ψ start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT ( italic_A ) ⟩ with tensors A𝐴Aitalic_A and bond dimension χ𝜒\chiitalic_χ such that

𝒬A(U,b)=1|{U}|×|b|U|ψχ(A)|2.subscript𝒬𝐴𝑈𝑏1𝑈superscriptbra𝑏𝑈ketsubscript𝜓𝜒𝐴2\mathcal{Q}_{A}(U,b)=\frac{1}{\lvert\{U\}\rvert}\times\lvert\mbox{$\langle b|$% }U\mbox{$|\psi_{\chi}(A)\rangle$}\rvert^{2}.caligraphic_Q start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ( italic_U , italic_b ) = divide start_ARG 1 end_ARG start_ARG | { italic_U } | end_ARG × | ⟨ italic_b | italic_U | italic_ψ start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT ( italic_A ) ⟩ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT . (112)

Drop** the constant normalization factor from the uniform distribution over the unitary ensemble gives us the MLE contribution described in Eq. (5) of the main text.

E.2 Regularization

In this section, we discuss the second contribution to our loss function, which additionally regularizes the optimization landscape of our training process. Note that these estimations from finite-N𝑁Nitalic_N samples have deviations from the true physical expectation values; in general, we do not want our regularization strength to be dominant.

In the main text, we numerically explore different regularization strengths. Moreover, we have focused on two measurement schemes: global-XZ𝑋𝑍XZitalic_X italic_Z measurements, and random-XZ𝑋𝑍XZitalic_X italic_Z measurements. For the former, we measure bit-strings in two fixed bases Xnsuperscript𝑋tensor-productabsent𝑛X^{\otimes n}italic_X start_POSTSUPERSCRIPT ⊗ italic_n end_POSTSUPERSCRIPT and Znsuperscript𝑍tensor-productabsent𝑛Z^{\otimes n}italic_Z start_POSTSUPERSCRIPT ⊗ italic_n end_POSTSUPERSCRIPT, and we can estimate certain physical observables using empirical means. For the latter, using the toolbox of randomized measurements, we apply random unitaries from an ensemble and measure bit-strings in the computational basis. Depending on the choice of the unitary architectures, different physical properties can be estimated. Performing classical shadow tomography using random-XZ𝑋𝑍XZitalic_X italic_Z measurements—as detailed in Appendix C—allows us to estimate any observables in the visible space of our measurement. We now discuss our choices of regularizers for the two systems considered in the main text:

Perturbed surface code. Here, we use the stabilizers 𝒮𝒮\mathcal{S}caligraphic_S of the exact surface code defined in Appendix F.1. The intuition behind such a choice is that we know that the stabilizer expectations will change continuously with the perturbation and that these operators would have large weight in the density matrix. The regularization contribution to the loss function is

R(A)=S𝒮|SNψ(A)|S|ψ(A)|2,𝑅𝐴subscript𝑆𝒮superscriptsubscriptdelimited-⟨⟩𝑆𝑁delimited-⟨⟩𝜓𝐴𝑆𝜓𝐴2R(A)=\sum_{\begin{subarray}{c}S\in\mathcal{S}\end{subarray}}\lvert\langle S% \rangle_{N}-\langle\psi(A)\lvert S\rvert\psi(A)\rangle\rvert^{2},italic_R ( italic_A ) = ∑ start_POSTSUBSCRIPT start_ARG start_ROW start_CELL italic_S ∈ caligraphic_S end_CELL end_ROW end_ARG end_POSTSUBSCRIPT | ⟨ italic_S ⟩ start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT - ⟨ italic_ψ ( italic_A ) | italic_S | italic_ψ ( italic_A ) ⟩ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , (113)

where SNsubscriptdelimited-⟨⟩𝑆𝑁\langle S\rangle_{N}⟨ italic_S ⟩ start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT are estimates from the measurements and the second term is the model’s prediction. The regularizer R(A)𝑅𝐴R(A)italic_R ( italic_A ) would be modulated by β𝛽\betaitalic_β, which is a hyperparameter controlling its strength. The knowledge of these specific observables allows us to estimate their expectation values in two measurement schemes.

With global-XZ𝑋𝑍XZitalic_X italic_Z measurements in the bases Xnsuperscript𝑋tensor-productabsent𝑛X^{\otimes n}italic_X start_POSTSUPERSCRIPT ⊗ italic_n end_POSTSUPERSCRIPT and Znsuperscript𝑍tensor-productabsent𝑛Z^{\otimes n}italic_Z start_POSTSUPERSCRIPT ⊗ italic_n end_POSTSUPERSCRIPT, we estimate the stabilizer expectations using empirical means. For instance, to estimate SZ,3=Z1Z2subscript𝑆𝑍3subscript𝑍1subscript𝑍2S_{Z,3}=Z_{1}Z_{2}italic_S start_POSTSUBSCRIPT italic_Z , 3 end_POSTSUBSCRIPT = italic_Z start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_Z start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT from fixed Znsuperscript𝑍tensor-productabsent𝑛Z^{\otimes n}italic_Z start_POSTSUPERSCRIPT ⊗ italic_n end_POSTSUPERSCRIPT basis measurements {Z1(i)Z2(i)Zn(i)}i=1Nsuperscriptsubscriptsuperscriptsubscript𝑍1𝑖superscriptsubscript𝑍2𝑖superscriptsubscript𝑍𝑛𝑖𝑖1𝑁\{Z_{1}^{(i)}Z_{2}^{(i)}\cdots Z_{n}^{(i)}\}_{i=1}^{N}{ italic_Z start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT italic_Z start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT ⋯ italic_Z start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT } start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT, we average over the measurement outcomes

Z1Z2N,Global=1Ni=1NZ1(i)Z2(i).subscriptdelimited-⟨⟩subscript𝑍1subscript𝑍2𝑁Global1𝑁superscriptsubscript𝑖1𝑁superscriptsubscript𝑍1𝑖superscriptsubscript𝑍2𝑖\langle Z_{1}Z_{2}\rangle_{N,\,\text{Global}}=-\frac{1}{N}\sum_{i=1}^{N}Z_{1}^% {(i)}Z_{2}^{(i)}.⟨ italic_Z start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_Z start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_N , Global end_POSTSUBSCRIPT = - divide start_ARG 1 end_ARG start_ARG italic_N end_ARG ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT italic_Z start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT italic_Z start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT . (114)

In the scheme of random-XZ𝑋𝑍XZitalic_X italic_Z measurements, we use classical shadow tomography for such estimation. In particular, we can estimate any k𝑘kitalic_k-local observable SZ,lsubscript𝑆𝑍𝑙S_{Z,\,l}italic_S start_POSTSUBSCRIPT italic_Z , italic_l end_POSTSUBSCRIPT using the classical shadows ρ^Nsubscript^𝜌𝑁\hat{\rho}_{N}over^ start_ARG italic_ρ end_ARG start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT defined in Eq. (91) as

SZ,lN,Random=Tr(SZ,lρ^N).subscriptdelimited-⟨⟩subscript𝑆𝑍𝑙𝑁RandomTrsubscript𝑆𝑍𝑙subscript^𝜌𝑁\langle S_{Z,\,l}\rangle_{N,\,\text{Random}}=\mathop{\mathrm{Tr}}(S_{Z,\,l}% \hat{\rho}_{N}).⟨ italic_S start_POSTSUBSCRIPT italic_Z , italic_l end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_N , Random end_POSTSUBSCRIPT = roman_Tr ( italic_S start_POSTSUBSCRIPT italic_Z , italic_l end_POSTSUBSCRIPT over^ start_ARG italic_ρ end_ARG start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT ) . (115)

As discussed in Appendix C, the number of samples required to estimate such observables accurately scales exponentially as 𝒪(2k)𝒪superscript2𝑘\mathcal{O}(2^{k})caligraphic_O ( 2 start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT ). Here, a key surprise of our protocol is that even though we only regularize with physically local observables knmuch-less-than𝑘𝑛k\ll nitalic_k ≪ italic_n, we are able to see an improvement in global quantities as the infidelity is reduced.

Rydberg atom arrays. More generally, if we do not know a priori which operators are important—as is indeed the case for non-fixed-point states realizable in quantum simulators such as Rydberg atom arrays—we would need to consider all the operators that one can estimate using a finite number of samples. The intuition behind the regularization contribution is that capturing more physical observables accurately is consistent with improvement of the state fidelity. In fact, describing a subsystem with a size comparable to the correlation length guarantees an efficient learning of a one-dimensional injective quantum state using MPS tomography [18, 19]. Here, using random-XZ𝑋𝑍XZitalic_X italic_Z measurements, the space of the operators we can estimate is spanned by all combinations of 𝟙,X,Z1𝑋𝑍\mathds{1},X,Zblackboard_1 , italic_X , italic_Z strings normalized with respect to the Hilbert-Schmidt inner product Pauli𝟙,X,Z(n)subscriptPauli1𝑋𝑍𝑛\text{\rm Pauli}_{\mathds{1},X,Z}(n)Pauli start_POSTSUBSCRIPT blackboard_1 , italic_X , italic_Z end_POSTSUBSCRIPT ( italic_n ). This motivates us to define the projected density matrix

ρXZ=PPauli𝟙,X,Z(n)Tr(Pρ)P.subscript𝜌𝑋𝑍subscript𝑃subscriptPauli1𝑋𝑍𝑛Tr𝑃𝜌𝑃\rho_{XZ}=\sum_{P\in\text{\rm Pauli}_{\mathds{1},X,Z}(n)}\mathop{\mathrm{Tr}}(% P\rho)P.italic_ρ start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_P ∈ Pauli start_POSTSUBSCRIPT blackboard_1 , italic_X , italic_Z end_POSTSUBSCRIPT ( italic_n ) end_POSTSUBSCRIPT roman_Tr ( italic_P italic_ρ ) italic_P . (116)

Now, the classical shadows learnable using random-XZ𝑋𝑍XZitalic_X italic_Z measurements [cf. Eq. (C.1)] converge to this projection in the infinite-data limit, i.e.,

ρXZ=limNρ^m.subscript𝜌𝑋𝑍subscript𝑁subscript^𝜌𝑚\rho_{XZ}=\lim_{N\rightarrow\infty}\hat{\rho}_{m}.italic_ρ start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT = roman_lim start_POSTSUBSCRIPT italic_N → ∞ end_POSTSUBSCRIPT over^ start_ARG italic_ρ end_ARG start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT . (117)

In the finite-sample regime, we regularize on observables that are supported on subsystems B𝐵Bitalic_B of 6666 qubits, described by a classical shadow

ρ^NB=iB(2Uαi|bibi|Uαi𝕀2).\hat{\rho}_{N}^{B}=\otimes_{i\in B}\left(2U^{\alpha_{i}\dagger}\rvert b_{i}% \rangle\!\langle b_{i}\lvert U^{\alpha_{i}}-\frac{\mathbb{I}}{2}\right).over^ start_ARG italic_ρ end_ARG start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_B end_POSTSUPERSCRIPT = ⊗ start_POSTSUBSCRIPT italic_i ∈ italic_B end_POSTSUBSCRIPT ( 2 italic_U start_POSTSUPERSCRIPT italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT † end_POSTSUPERSCRIPT | italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ⟩ ⟨ italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | italic_U start_POSTSUPERSCRIPT italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUPERSCRIPT - divide start_ARG blackboard_I end_ARG start_ARG 2 end_ARG ) . (118)

Such a projection can also be computed from our MPS model |ψ(A)ket𝜓𝐴|\psi(A)\rangle| italic_ψ ( italic_A ) ⟩. First, we trace out the complement B¯¯𝐵\bar{B}over¯ start_ARG italic_B end_ARG

ρ~B(A)=TrB¯(|ψ(A)ψ(A)|).\tilde{\rho}^{B}(A)=\mathop{\mathrm{Tr}}_{\bar{B}}(\rvert\psi(A)\rangle\!% \langle\psi(A)\lvert).over~ start_ARG italic_ρ end_ARG start_POSTSUPERSCRIPT italic_B end_POSTSUPERSCRIPT ( italic_A ) = roman_Tr start_POSTSUBSCRIPT over¯ start_ARG italic_B end_ARG end_POSTSUBSCRIPT ( | italic_ψ ( italic_A ) ⟩ ⟨ italic_ψ ( italic_A ) | ) . (119)

Next, we project the reduced density matrix onto VisibleSpace(𝒰XZ)VisibleSpacesubscript𝒰𝑋𝑍\textnormal{{VisibleSpace}}(\mathcal{U}_{XZ})VisibleSpace ( caligraphic_U start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT ) as

ρ~XZB(A)=PPauli𝟙,X,Z(n)Tr(Pρ~B(A))P.superscriptsubscript~𝜌𝑋𝑍𝐵𝐴subscript𝑃subscriptPauli1𝑋𝑍𝑛Tr𝑃superscript~𝜌𝐵𝐴𝑃\tilde{\rho}_{XZ}^{B}(A)=\sum_{P\in\text{\rm Pauli}_{\mathds{1},X,Z}(n)}% \mathop{\mathrm{Tr}}(P\tilde{\rho}^{B}(A))P.over~ start_ARG italic_ρ end_ARG start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_B end_POSTSUPERSCRIPT ( italic_A ) = ∑ start_POSTSUBSCRIPT italic_P ∈ Pauli start_POSTSUBSCRIPT blackboard_1 , italic_X , italic_Z end_POSTSUBSCRIPT ( italic_n ) end_POSTSUBSCRIPT roman_Tr ( italic_P over~ start_ARG italic_ρ end_ARG start_POSTSUPERSCRIPT italic_B end_POSTSUPERSCRIPT ( italic_A ) ) italic_P . (120)

Putting all the pieces together, we regularize over these projected density matrices—supported on all the unit cells that cover the lattice—using the second term in our loss function

R(A)=ρ^NBρ~XZB(A)2.𝑅𝐴subscriptdelimited-∥∥superscriptsubscript^𝜌𝑁𝐵superscriptsubscript~𝜌𝑋𝑍𝐵𝐴2\displaystyle R(A)=\lVert\hat{\rho}_{N}^{B}-\tilde{\rho}_{XZ}^{B}(A)\rVert_{2}.italic_R ( italic_A ) = ∥ over^ start_ARG italic_ρ end_ARG start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_B end_POSTSUPERSCRIPT - over~ start_ARG italic_ρ end_ARG start_POSTSUBSCRIPT italic_X italic_Z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_B end_POSTSUPERSCRIPT ( italic_A ) ∥ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT . (121)

Here, the matrix norm is given by the Schatten-2222 norm (or Frobenius norm)

O2=[i,jOi,j2]12.subscriptdelimited-∥∥𝑂2superscriptdelimited-[]subscript𝑖𝑗superscriptsubscript𝑂𝑖𝑗212\lVert O\rVert_{2}=\left[\sum_{i,j}O_{i,j}^{2}\right]^{\frac{1}{2}}.∥ italic_O ∥ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = [ ∑ start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT italic_O start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] start_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT . (122)

APPENDIX F Dataset and state generation with DMRG

In this appendix, we discuss the generation of our numerical dataset using Gibbs sampling from the target MPS. The target MPS is variationally optimized using the density-matrix renormalization group (DMRG) algorithm. Here, we outline a snippet of the numerical code written in the Python language. The calculations are performed using an open source library quimb [94]. All of the code used to generate our numerical results is available at: https://github.com/teng10/tn-shadow-qst.

F.1 Perturbed surface code

In this section, we discuss the details of our surface code numerics and provide the relevant code snippets used to generate our dataset. The perturbed surface code Hamiltonian for a system of n𝑛nitalic_n qubits is defined as

Hsc=lSZ,llSX,lhziZi,SZ,l=ilZi,SX,l=ilXi,formulae-sequencesubscript𝐻scsubscript𝑙subscript𝑆𝑍𝑙subscript𝑙subscript𝑆𝑋𝑙subscript𝑧subscript𝑖subscript𝑍𝑖formulae-sequencesubscript𝑆𝑍𝑙subscriptproduct𝑖subscript𝑙subscript𝑍𝑖subscript𝑆𝑋𝑙subscriptproduct𝑖subscript𝑙subscript𝑋𝑖H_{\text{sc}}=-\sum_{l}S_{Z,\,l}-\sum_{l}S_{X,\,l}-h_{z}\sum_{i}Z_{i},\quad S_% {Z,\,l}=\prod_{i\in\square_{l}}Z_{i},\quad S_{X,\,l}=\prod_{i\in\square_{l}}X_% {i},italic_H start_POSTSUBSCRIPT sc end_POSTSUBSCRIPT = - ∑ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT italic_S start_POSTSUBSCRIPT italic_Z , italic_l end_POSTSUBSCRIPT - ∑ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT italic_S start_POSTSUBSCRIPT italic_X , italic_l end_POSTSUBSCRIPT - italic_h start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_Z start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_S start_POSTSUBSCRIPT italic_Z , italic_l end_POSTSUBSCRIPT = ∏ start_POSTSUBSCRIPT italic_i ∈ □ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_Z start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_S start_POSTSUBSCRIPT italic_X , italic_l end_POSTSUBSCRIPT = ∏ start_POSTSUBSCRIPT italic_i ∈ □ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_X start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , (123)

where 𝒮={SZ,l,SX,l,l=1,,n12}\mathcal{S}=\{S_{Z,\,l},S_{X,\,l},\,l=1,\cdots,\frac{n-1}{2}\}caligraphic_S = { italic_S start_POSTSUBSCRIPT italic_Z , italic_l end_POSTSUBSCRIPT , italic_S start_POSTSUBSCRIPT italic_X , italic_l end_POSTSUBSCRIPT , italic_l = 1 , ⋯ , divide start_ARG italic_n - 1 end_ARG start_ARG 2 end_ARG } is the set of n1𝑛1n-1italic_n - 1 stabilizers around a square, as illustrated in Fig. 2(a). To construct the Hamiltonian for an (Lx,Ly)=(3,3)subscript𝐿𝑥subscript𝐿𝑦33(L_{x},L_{y})=(3,3)( italic_L start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT , italic_L start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT ) = ( 3 , 3 ) system, the following lines create an object called PhysicalSystem and generate its matrix product operator representation.

1surface_code = physical_systems.SurfaceCode(3, 3)
2surface_code_mpo = surface_code.get_ham()

To run the DMRG algorithm, first, a random MPS is selected as an initial state. Then, the MPO and this initial state are used to solve for an MPS approximation of the ground state with a given bond dimension.

1import quimb.tensor as qtn
2size = 9
3bond_dim = 10
4mps = qtn.MPS_rand_state(size, bond_dim)
5dmrg = qtn.DMRG2(surface_code_mpo, bond_dims=bond_dim, p0=mps)
6dmrg.solve()
7result_mps = dmrg.state

Finally, to generate the dataset for training, we first select a measurement scheme (e.g., random-XZ𝑋𝑍XZitalic_X italic_Z [see Appendix D]), and specify the number of samples required.

1init_seed = 42 # for reproducibility
2sampling_method = x_or_z_basis_sampler # random XZ measurements
3num_samples = 1000
4ds = run_data_generation._run_data_generation(init_seed, num_samples, sampling_method, mps)

F.2 Ruby-lattice quantum spin liquid

In this section, we include the details of our DMRG calculation for preparing ground states of the Rydberg Hamiltonian on a ruby lattice [7]. Specifically, we consider a system of atoms arrayed on a ruby lattice with an aspect ratio of ρ=3𝜌3\rho=\sqrt{3}italic_ρ = square-root start_ARG 3 end_ARG [Supplementary Fig. 2] and cylindrical boundary conditions. The Hamiltonian is given by

HRyd=Ω2XΔ1+Z2+12,V,4(1+Z)(1+Z).subscript𝐻RydΩ2subscriptsubscript𝑋Δsubscript1subscript𝑍212subscriptsuperscriptsubscript𝑉superscript41subscript𝑍1subscript𝑍superscript\displaystyle H_{\rm Ryd}=\frac{\Omega}{2}\sum_{\ell}X_{\ell}-\Delta\sum_{\ell% }\frac{1+Z_{\ell}}{2}+\frac{1}{2}\sum_{\ell,\ell^{\prime}}\frac{V_{\ell,\ell^{% \prime}}}{4}(1+Z_{\ell})(1+Z_{\ell^{\prime}}).italic_H start_POSTSUBSCRIPT roman_Ryd end_POSTSUBSCRIPT = divide start_ARG roman_Ω end_ARG start_ARG 2 end_ARG ∑ start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT italic_X start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT - roman_Δ ∑ start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT divide start_ARG 1 + italic_Z start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG + divide start_ARG 1 end_ARG start_ARG 2 end_ARG ∑ start_POSTSUBSCRIPT roman_ℓ , roman_ℓ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT divide start_ARG italic_V start_POSTSUBSCRIPT roman_ℓ , roman_ℓ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT end_ARG start_ARG 4 end_ARG ( 1 + italic_Z start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT ) ( 1 + italic_Z start_POSTSUBSCRIPT roman_ℓ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ) . (124)

Denoting by |b=|±ket𝑏ketplus-or-minus\mbox{$|b\rangle$}=\mbox{$|\pm\rangle$}| italic_b ⟩ = | ± ⟩ the eigenstate of the Pauli Z𝑍Zitalic_Z operator in the computational basis (i.e., Z|b=b|b𝑍ket𝑏𝑏ket𝑏Z\mbox{$|b\rangle$}=b\mbox{$|b\rangle$}italic_Z | italic_b ⟩ = italic_b | italic_b ⟩), we work with a sign convention such that b=+1𝑏1b=+1italic_b = + 1 (11-1- 1) corresponds to an atom being in the Rydberg (ground) state.

Depending on the ratio of the detuning to the Rabi frequency, Δ/ΩΔΩ\Delta/\Omegaroman_Δ / roman_Ω, different quantum phases can be prepared [51, 52]. The native van der Waals interaction potential decays rapidly with the distance as V,=V0/|rr|6subscript𝑉superscriptsubscript𝑉0superscriptsubscript𝑟subscript𝑟superscript6V_{\ell,\ell^{\prime}}={V_{0}}/{\lvert r_{\ell}-r_{\ell^{\prime}}\rvert^{6}}italic_V start_POSTSUBSCRIPT roman_ℓ , roman_ℓ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT = italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / | italic_r start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT - italic_r start_POSTSUBSCRIPT roman_ℓ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT. Choosing the interaction strength such that ΩV,much-less-thanΩsubscript𝑉superscript\Omega\ll V_{\ell,\ell^{\prime}}roman_Ω ≪ italic_V start_POSTSUBSCRIPT roman_ℓ , roman_ℓ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT, the strong repulsion between neighboring atoms implements the Rydberg blockade effect, such that no two atoms are allowed to be simultaneously excited within a blockade radius of Rb=(V0/Ω)16subscript𝑅𝑏superscriptsubscript𝑉0Ω16R_{b}=({V_{0}}/{\Omega})^{\frac{1}{6}}italic_R start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT = ( italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / roman_Ω ) start_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG 6 end_ARG end_POSTSUPERSCRIPT. While, in principle, the interactions act between all pairs of atoms, in practice, an approximated potential obtained by truncating the long-range tails of the interaction is known to capture the essential physics of the spin-liquid phase [52]. In our DMRG simulations, we only retain interactions up to third-nearest neighbors (r=2a𝑟2𝑎r=2aitalic_r = 2 italic_a, where a𝑎aitalic_a is the lattice spacing), as illustrated in Fig. 3(a). Setting Ω=1Ω1\Omega=1roman_Ω = 1, we approximate the potential as a step function

V,={47Ω,|rr|2a,0,otherwise.subscript𝑉superscriptcases47Ωsubscript𝑟subscript𝑟superscript2𝑎0otherwiseV_{\ell,\ell^{\prime}}=\begin{cases}47\Omega,&\quad\lvert r_{\ell}-r_{\ell^{% \prime}}\rvert\leq 2a,\\ 0,&\quad\text{otherwise}.\end{cases}italic_V start_POSTSUBSCRIPT roman_ℓ , roman_ℓ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT = { start_ROW start_CELL 47 roman_Ω , end_CELL start_CELL | italic_r start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT - italic_r start_POSTSUBSCRIPT roman_ℓ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT | ≤ 2 italic_a , end_CELL end_ROW start_ROW start_CELL 0 , end_CELL start_CELL otherwise . end_CELL end_ROW (125)

Note that unlike the so-called PXP model [52], in which the Hilbert space is constrained to forbid simultaneous excitations of neighboring atoms, here, the blockade is only imposed as an energetic penalty. The interaction scale in Eq. (125) follows from the choice of the blockade radius, Rb=3.8asubscript𝑅𝑏3.8𝑎R_{b}=3.8aitalic_R start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT = 3.8 italic_a. Moreover, the Rabi frequency ΩΩ\Omegaroman_Ω is positive, so the present Hamiltonian is nonstoquastic [32], without guarantees of positive wavefunction amplitudes in the computational basis.

In the present work, we focus on the ruby-lattice Rydberg Hamiltonian with periodic boundary conditions along the y𝑦yitalic_y-direction and open boundary conditions along the x𝑥xitalic_x-axis. We perform numerical simulations for a lattice of size (Lx,Ly)=(4,2)subscript𝐿𝑥subscript𝐿𝑦42(L_{x},L_{y})=(4,2)( italic_L start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT , italic_L start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT ) = ( 4 , 2 ); the unit cell has 6666 basis atoms, so the total number of qubits is n=48𝑛48n=48italic_n = 48. To mitigate finite-size effects, we compensate for the fewer neighboring atoms of any site along the boundary by adding a local field hbd=0.6Ωsubscriptbd0.6Ωh_{\text{bd}}=-0.6\Omegaitalic_h start_POSTSUBSCRIPT bd end_POSTSUBSCRIPT = - 0.6 roman_Ω to the regions shaded by the blue boxes in Supplementary Fig. 2:

HRyd, bd=HRydhbd1+Z2.subscript𝐻Ryd, bdsubscript𝐻Rydsubscriptbdsubscript1subscript𝑍2H_{\text{Ryd, bd}}=H_{\text{Ryd}}-h_{\text{bd}}\sum_{\ell\in{\partial}}\frac{1% +Z_{\ell}}{2}.italic_H start_POSTSUBSCRIPT Ryd, bd end_POSTSUBSCRIPT = italic_H start_POSTSUBSCRIPT Ryd end_POSTSUBSCRIPT - italic_h start_POSTSUBSCRIPT bd end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT roman_ℓ ∈ ∂ end_POSTSUBSCRIPT divide start_ARG 1 + italic_Z start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG . (126)
1delta = 1.7 # detuning frequency for spin liquid state
2boundary_z_field = -0.6 # adding an onsite field to mitigate boundary effects
3rydberg = physical_systems.RubyRydbergPXP(4, 2)
4rydberg_mpo = rydberg.get_ham(delta, boundary_z_field)
Refer to caption
Supplementary Figure 2: Rydberg atom array on a cylinder. The ruby lattice is placed on a cylinder, imposing periodic (open) boundary conditions along the y𝑦yitalic_y (x𝑥xitalic_x)-axis. The numbers label the snake-like path for the tensors in the MPS representation of the ground state. An onsite field applied to the sites in the blue boxes compensates for the reduced coordination number of atoms situated along the open boundaries, thereby preventing them from getting pinned.

APPENDIX G Training details

In this appendix, we discuss our approach to optimizing matrix product state (MPS) tensors through simultaneous updates of all tensor components by gradient descent. We then highlight the selection process for models likely to converge in the absence of direct access to the infidelity. This can be done by partitioning datasets into training and test datasets and post selecting models that successfully generalize on the test dataset. This method’s efficacy is illustrated through the analysis of training and test negative log-likelihood (NLL) metrics. As evidenced in the surface code example, we evaluate the model performance’s with sample size to ensure robust generalization.

G.1 Optimization scheme via gradient descent

We approach the optimization of matrix product state (MPS) tensors by updating all tensors simultaneously. While the target states we are learning are real states, we use MPS tensors with complex variables because we have found that optimization over complex tensors is more robust, as also observed in Ref.  [27]. The minimization of our loss function involves the logarithm of sums over expectation values of Ui|bibi|UiU_{i}^{\dagger}\rvert b_{i}\rangle\!\langle b_{i}\lvert U_{i}italic_U start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT | italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ⟩ ⟨ italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | italic_U start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT defined by the measured unitary-bit-strings D={(Ui,bi)}i=1N𝐷superscriptsubscriptsubscript𝑈𝑖subscript𝑏𝑖𝑖1𝑁D=\{(U_{i},b_{i})\}_{i=1}^{N}italic_D = { ( italic_U start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) } start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT,

LNLLD(A)superscriptsubscript𝐿NLL𝐷𝐴\displaystyle L_{\text{NLL}}^{D}(A)italic_L start_POSTSUBSCRIPT NLL end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_D end_POSTSUPERSCRIPT ( italic_A ) =1Ni=1Nlog(ψ(A)|Ui|bibi|Ui|ψ(A)).absent1𝑁superscriptsubscript𝑖1𝑁delimited-⟨⟩𝜓superscript𝐴subscriptsuperscript𝑈𝑖subscript𝑏𝑖delimited-⟨⟩subscript𝑏𝑖subscript𝑈𝑖𝜓𝐴\displaystyle=-\frac{1}{N}\sum_{i=1}^{N}\log\left(\langle\psi(A^{*})\lvert U^{% \dagger}_{i}\rvert b_{i}\rangle\langle b_{i}\lvert U_{i}\rvert\psi(A)\rangle% \right).= - divide start_ARG 1 end_ARG start_ARG italic_N end_ARG ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT roman_log ( ⟨ italic_ψ ( italic_A start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) | italic_U start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ⟩ ⟨ italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | italic_U start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | italic_ψ ( italic_A ) ⟩ ) . (127)

For simplicity, we have considered the loss function without a regularization with β=0𝛽0\beta=0italic_β = 0. While Ui|bibi|UiU_{i}^{\dagger}\rvert b_{i}\rangle\!\langle b_{i}\lvert U_{i}italic_U start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT | italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ⟩ ⟨ italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | italic_U start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT can be represented as a matrix product operator (MPO), this loss function is different from the traditional eigenvalue problem of finding the ground state of a local Hamiltonian typically considered in DMRG calculations. Given the presence of the logarithmic term in our cost function, previous work has considered a “DMRG-like” swee** algorithm [26]. However, to simplify the training process, we have adopted global updates of MPS tensors over local swee** updates and find that this strategy suffices for our purpose. Future improvements are possible by adopting optimization techniques on the MPS manifold [85, 84] as discussed in B.7, particularly focusing on the projection of gauge degrees of freedom to refine tensor updates.

Note that in order to minimize the loss function, we calculate gradients for all MPS tensors in the direction opposite to that dictated by the derivative of the complex conjugate 777To see this, let f(z,z¯)::𝑓𝑧¯𝑧f(z,\bar{z}):\mathbb{C}\rightarrow\mathbb{R}italic_f ( italic_z , over¯ start_ARG italic_z end_ARG ) : blackboard_C → blackboard_R be a real-valued function of variables z=x+iy𝑧𝑥𝑖𝑦z=x+iyitalic_z = italic_x + italic_i italic_y and z¯=xiy¯𝑧𝑥𝑖𝑦\bar{z}=x-iyover¯ start_ARG italic_z end_ARG = italic_x - italic_i italic_y. Note that the loss function could be nonholomorphic. An update to minimize the loss function is given by (x+iy)=xδfx+i(yδfx)superscript𝑥𝑖superscript𝑦𝑥𝛿𝑓𝑥𝑖𝑦𝛿𝑓𝑥(x^{\prime}+iy^{\prime})=x-\delta\frac{\partial f}{\partial x}+i(y-\delta\frac% {\partial f}{\partial x})( italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT + italic_i italic_y start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) = italic_x - italic_δ divide start_ARG ∂ italic_f end_ARG start_ARG ∂ italic_x end_ARG + italic_i ( italic_y - italic_δ divide start_ARG ∂ italic_f end_ARG start_ARG ∂ italic_x end_ARG ). This is the same as z=zδ2fz¯superscript𝑧𝑧𝛿2𝑓¯𝑧z^{\prime}=z-\frac{\delta}{2}\frac{\partial f}{\partial\bar{z}}italic_z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = italic_z - divide start_ARG italic_δ end_ARG start_ARG 2 end_ARG divide start_ARG ∂ italic_f end_ARG start_ARG ∂ over¯ start_ARG italic_z end_ARG end_ARG. Numerically, we compute LA𝐿𝐴\frac{\partial L}{\partial A}divide start_ARG ∂ italic_L end_ARG start_ARG ∂ italic_A end_ARG using JAX and take the direction given by its complex conjugate (LA)superscript𝐿𝐴(\frac{\partial L}{\partial A})^{*}( divide start_ARG ∂ italic_L end_ARG start_ARG ∂ italic_A end_ARG ) start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT.

A=AδdLdA¯.superscript𝐴𝐴𝛿d𝐿d¯𝐴\displaystyle A^{\prime}=A-\delta\frac{\mathrm{d}L}{\mathrm{d}\bar{A}}.italic_A start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = italic_A - italic_δ divide start_ARG roman_d italic_L end_ARG start_ARG roman_d over¯ start_ARG italic_A end_ARG end_ARG . (128)

To integrate with machine-learning optimization techniques, our procedure employs a two-step strategy: initial descent towards an approximate global minimum using first-order stochastic gradient descent (SGD), followed by refinement through a second-order quasi-Newton’s method, the limited-memory Broyden–Fletcher–Goldfarb–Shanno (L-BFGS) algorithm. This combination allows for efficient preliminary optimization followed by detailed adjustments.

G.2 Model selection

In applying our protocol to a laboratory state, in general, we would not have access to the infidelity. Here, we outline the steps to select models that are likely to converge without access to the fidelity. In the context of machine learning, a common approach is to divide the full dataset into the training and test datasets. First, we train the MPS model on only the training dataset partition and then evaluate the negative log-likelihood (NLL) in Eq. (111) using the test dataset. For the model to not be “overfitting” the training dataset, we would want both the training and test NLLs to be small. Specifically, we consider the surface code numerics as an example below. In Supplementary Fig. 3, we plot both the training NLL and the test NLL for different numbers of samples. We observe that for a small number of samples, while we achieve a lower training NLL, the test NLL is higher, indicating overfitting of the MPS to the training measurements. On the other hand, an MPS trained with a large number of samples achieves a lower infidelity and hence, better generalization.

A different approach is to estimate the fidelity of the learned MPS model with the laboratory state [26] by using classical shadow tomography with global random unitary ensembles, which could be hard to implement in practice. Finally, we remark that a recent protocol would also allow us to efficiently certify the fidelity using single-qubit measurements given a query model [42], in our case, an MPS.

Refer to caption
Supplementary Figure 3: Model selection by comparing training and test losses for a surface code with different system sizes and measurement schemes. On the dashed lines, the training NLL === test NLL; ideally, the model should be selected along the dashed line and with a small training NLL. The models which have small NLLs and are close to the dashed line yield high fidelities (darker colors).