HTML conversions sometimes display errors due to content that did not convert correctly from the source. This paper uses the following packages that are not yet supported by the HTML conversion tool. Feedback on these issues are not necessary; they are known and are being worked on.

  • failed: qcircuit
  • failed: pagecolor
  • failed: tikzscale

Authors: achieve the best HTML results from your LaTeX submissions by following these best practices.

License: arXiv.org perpetual non-exclusive license
arXiv:2402.18809v1 [quant-ph] 29 Feb 2024
thanks: These authors contributed equally to this work: C.O. ([email protected]); S.C. ([email protected]).thanks: These authors contributed equally to this work: C.O. ([email protected]); S.C. ([email protected]).

Entanglement-enabled advantage for learning a bosonic random displacement channel

Changhun Oh Pritzker School of Molecular Engineering, The University of Chicago, Chicago, Illinois 60637, USA    Senrui Chen Pritzker School of Molecular Engineering, The University of Chicago, Chicago, Illinois 60637, USA    Yat Wong Pritzker School of Molecular Engineering, The University of Chicago, Chicago, Illinois 60637, USA    Sisi Zhou Perimeter Institute for Theoretical Physics, Waterloo, Ontario N2L 2Y5, Canada Institute for Quantum Information and Matter, California Institute of Technology, Pasadena, CA 91125, USA Department of Physics and Astronomy and Institute for Quantum Computing, University of Waterloo, Ontario N2L 2Y5, Canada    Hsin-Yuan Huang Institute for Quantum Information and Matter, California Institute of Technology, Pasadena, CA 91125, USA Center for Theoretical Physics, Massachusetts Institute of Technology, Cambridge, MA 02139, USA Google Quantum AI, Venice, CA, USA    Jens A.H. Nielsen    Zheng-Hao Liu    Jonas S. Neergaard-Nielsen    Ulrik L. Andersen Center for Macroscopic Quantum States (bigQ), Department of Physics, Technical University of Denmark, Building 307, Fysikvej, 2800 Kgs. Lyngby, Denmark    Liang Jiang [email protected] Pritzker School of Molecular Engineering, The University of Chicago, Chicago, Illinois 60637, USA    John Preskill [email protected] Institute for Quantum Information and Matter, California Institute of Technology, Pasadena, CA 91125, USA
(February 29, 2024)
Abstract

We show that quantum entanglement can provide an exponential advantage in learning properties of a bosonic continuous-variable (CV) system. The task we consider is estimating a probabilistic mixture of displacement operators acting on n𝑛nitalic_n bosonic modes, called a random displacement channel. We prove that if the n𝑛nitalic_n modes are not entangled with an ancillary quantum memory, then the channel must be sampled a number of times exponential in n𝑛nitalic_n in order to estimate its characteristic function to reasonable precision; this lower bound on sample complexity applies even if the channel inputs and measurements performed on channel outputs are chosen adaptively. On the other hand, we present a simple entanglement-assisted scheme that only requires a number of samples independent of n𝑛nitalic_n, given a sufficient amount of squeezing. This establishes an exponential separation in sample complexity. We then analyze the effect of photon loss and show that the entanglement-assisted scheme is still significantly more efficient than any lossless entanglement-free scheme under mild experimental conditions. Our work illuminates the role of entanglement in learning continuous-variable systems and points toward experimentally feasible demonstrations of provable entanglement-enabled advantage using CV quantum platforms.

Quantum science and technology holds promise to revolutionize how we understand and interact with nature, enabling computational speedups [1], classically impossible communication tasks [2, 3], and measurements with unprecedented sensitivity [4, 5, 6]. Rapid progress during the noisy intermediate-scale quantum (NISQ) era [7] has brought these promises closer to reality, but the challenge remains to demonstrate rigorous quantum advantage for practical problems.

Over the past few years, there has been ongoing theoretical and experimental progress in exploring quantum computational advantage [8, 9, 10, 11, 12, 13, 14, 15, 16]. Another recent line of research seeks quantum advantage in learning [17, 18, 19, 20, 21, 22, 23, 24], revealing that access to quantum memory enables us to learn properties of nature more efficiently. Specifically, Refs. [18, 19] establish a framework for proving exponential separation in sample complexity between learning with and without a coherently controllable quantum memory. In contrast to its computational counterpart, this entanglement-enabled advantage in learning can be proven without invoking computational assumptions and can sometimes be more experimentally accessible. A proof-of-principle experiment has been conducted on Google’s superconducting quantum processor using 40404040 qubits [18].

Most learning tasks studied so far are restricted to discrete-variable (DV) systems. It is natural to ask whether entanglement-enabled advantage can also be realized for learning properties of bosonic continuous-variable (CV) systems. This is particularly interesting and important because CV systems are ubiquitous in nature and have many applications in quantum information science, such as quantum sensing [25, 26, 27, 6, 8]. However, generalizing the results in DV systems to CV systems can be difficult because bosonic systems have infinite-dimensional Hilbert spaces, making it challenging to formulate rigorous results concerning the complexity of learning properties of these systems. Recent progress has been achieved in studies of entanglement-enhanced learning of CV-state characteristic functions [28]; however, the lower bounds obtained so far apply to a restricted class of learning strategies rather than to general entanglement-free schemes.

In this work, we rigorously establish an entanglement-enabled advantage in learning a probabilistic mixture of n𝑛nitalic_n-mode displacement operations, called a bosonic random displacement channel. Specifically, we show that any schemes without ancillary quantum memory require a number of samples exponential in n𝑛nitalic_n to learn the characteristic functions of the channel with reasonably good precision and high success probability. On the contrary, we present a simple scheme utilizing entanglement with ancillary quantum memory (i.e., entanglement-assisted) that can complete the same learning task with a sample complexity independent of n𝑛nitalic_n, given access to two-mode squeezed vacuum (TMSV) states with sufficiently large squeezing parameter and Bell measurements (BM). This establishes an exponential separation between learning with and without entanglement in the bosonic system. The two learning scenarios are illustrated in Fig. 1. Note that our hardness results hold for arbitrarily high-energy input states and arbitrary measurements, while the presented entanglement-assisted scheme only requires a finite-energy TMSV and BM.

Furthermore, we analyze the robustness of this entanglement-enabled advantage under realistic experimental conditions. Specifically, we study the photon-loss effect, the most common noise source in optical platforms. Our results suggest that for squeezing parameters and loss rates achievable in a state-of-the-art bosonic experiment platform, the separation in sample complexity remains significant. Therefore, we anticipate that an experimental demonstration of entanglement-enabled advantage in CV quantum systems can be achieved in the near future.

Problem Setup.— We consider the task of learning an n𝑛nitalic_n-mode random displacement channel characterized by a probability distribution p(α)𝑝𝛼p(\alpha)italic_p ( italic_α ) with αn𝛼superscript𝑛\alpha\in\mathbb{C}^{n}italic_α ∈ blackboard_C start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT, which transforms an input state ρ^^𝜌\hat{\rho}over^ start_ARG italic_ρ end_ARG as

Λ(ρ^)=d2nαp(α)D^(α)ρ^D^(α),Λ^𝜌superscript𝑑2𝑛𝛼𝑝𝛼^𝐷𝛼^𝜌superscript^𝐷𝛼\displaystyle\Lambda(\hat{\rho})=\int d^{2n}\alpha~{}p(\alpha)\hat{D}(\alpha)% \hat{\rho}\hat{D}^{\dagger}(\alpha),roman_Λ ( over^ start_ARG italic_ρ end_ARG ) = ∫ italic_d start_POSTSUPERSCRIPT 2 italic_n end_POSTSUPERSCRIPT italic_α italic_p ( italic_α ) over^ start_ARG italic_D end_ARG ( italic_α ) over^ start_ARG italic_ρ end_ARG over^ start_ARG italic_D end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ( italic_α ) , (1)

where D^(α):=i=1nD^(αi)\hat{D}(\alpha)\mathrel{\mathop{:}}\nobreak\mkern-1.2mu=\otimes_{i=1}^{n}\hat{% D}(\alpha_{i})over^ start_ARG italic_D end_ARG ( italic_α ) : = ⊗ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT over^ start_ARG italic_D end_ARG ( italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) and D^(αi):=exp(αia^iαi*a^i)\hat{D}(\alpha_{i})\mathrel{\mathop{:}}\nobreak\mkern-1.2mu=\exp(\alpha_{i}% \hat{a}_{i}^{\dagger}-\alpha_{i}^{*}\hat{a}_{i})over^ start_ARG italic_D end_ARG ( italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) : = roman_exp ( italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT over^ start_ARG italic_a end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT - italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT over^ start_ARG italic_a end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) is the displacement operator for the i𝑖iitalic_ith mode. The random displacement channel can also be equivalently described by the characteristic function of p(α)𝑝𝛼p(\alpha)italic_p ( italic_α ), i.e., its Fourier transform, as (see SM S1 [29] for the derivation)

Λ(ρ^)=1πnd2nβλ(β)Tr[ρ^D^(β)]D^(β),Λ^𝜌1superscript𝜋𝑛superscript𝑑2𝑛𝛽𝜆𝛽Tr^𝜌^𝐷𝛽superscript^𝐷𝛽\displaystyle\Lambda(\hat{\rho})=\frac{1}{\pi^{n}}\int d^{2n}\beta~{}\lambda(% \beta)\operatorname{Tr}[\hat{\rho}\hat{D}(\beta)]\hat{D}^{\dagger}(\beta),roman_Λ ( over^ start_ARG italic_ρ end_ARG ) = divide start_ARG 1 end_ARG start_ARG italic_π start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_ARG ∫ italic_d start_POSTSUPERSCRIPT 2 italic_n end_POSTSUPERSCRIPT italic_β italic_λ ( italic_β ) roman_Tr [ over^ start_ARG italic_ρ end_ARG over^ start_ARG italic_D end_ARG ( italic_β ) ] over^ start_ARG italic_D end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ( italic_β ) , (2)
λ(β):=d2nαp(α)eαββα.\displaystyle\lambda(\beta)\mathrel{\mathop{:}}\nobreak\mkern-1.2mu=\int d^{2n% }\alpha~{}p(\alpha)e^{\alpha^{\dagger}\beta-\beta^{\dagger}\alpha}.italic_λ ( italic_β ) : = ∫ italic_d start_POSTSUPERSCRIPT 2 italic_n end_POSTSUPERSCRIPT italic_α italic_p ( italic_α ) italic_e start_POSTSUPERSCRIPT italic_α start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_β - italic_β start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT . (3)

Here, because of the Fourier relation, λ(β)𝜆𝛽\lambda(\beta)italic_λ ( italic_β ) with a large β𝛽\betaitalic_β contributes to rapidly oscillating p(α)𝑝𝛼p(\alpha)italic_p ( italic_α ). Since the domain of β𝛽\betaitalic_β is infinite in principle, we will focus on a restricted finite domain specified later. The goal is to learn the channel by estimating the characteristic function λ(β)𝜆𝛽\lambda(\beta)italic_λ ( italic_β ). We emphasize that the goal is to characterize the channel, as opposed to identifying a particular displacement that is drawn from the distribution p(α)𝑝𝛼p(\alpha)italic_p ( italic_α ). The value of β𝛽\betaitalic_β for which λ(β)𝜆𝛽\lambda(\beta)italic_λ ( italic_β ) is to be estimated is revealed only after all measurements are completed.

We focus on the separation between two types of learning schemes for the random displacement channel distinguished by whether or not the scheme uses entanglement between the system and an ancilla, as illustrated in Fig. 1. Throughout this work, we define an entanglement-free scheme to be both ancilla-free and concatenation-free, i.e., the output of the channel is measured destructively after each channel use. However, the entanglement-free scheme is allowed to be adaptive; for each channel use, the input to the channel and the measurement performed on the output may depend on measurement outcomes obtained in earlier rounds. This scenario is similar to Refs. [17, 22]. Several recent works have obtained lower bounds on learning DV channels that hold even with concatenation [23, 30, 31], but we will not analyze the consequences of concatenating CV channels in this work for simplicity.

Refer to caption
Figure 1: Schemes for learning an n𝑛nitalic_n-mode random displacement channel ΛΛ\Lambdaroman_Λ. (a) TMSV+BM, a specific entanglement-assisted (EA) scheme. (b) General entanglement-free (EF) scheme. Here we assume no concatenation is allowed, i.e., each copy of the channel acts on some input state ρ0subscript𝜌0\rho_{0}italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and is measured destructively by some POVM {E}𝐸\{E\}{ italic_E }. The input state and measurement are allowed to be adaptively chosen depending on previous outcomes. An example of EF scheme is Vacuum+Heterodyne (see the main text).

Schemes.— Now, we present an entanglement-assisted scheme (see Fig. 1) inspired by a similar scheme to which has been proposed in DV Pauli channel estimation in Ref. [23]. Consider an n𝑛nitalic_n-mode random displacement channel ΛBsubscriptΛ𝐵\Lambda_{B}roman_Λ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT acting on the n𝑛nitalic_n-mode system B𝐵Bitalic_B. To learn this channel, we prepare n𝑛nitalic_n CV Bell states with a finite squeezing parameter r𝑟ritalic_r, which is a two-mode squeezed vacuum (TMSV) state, and half of the states go through the channel while the other half stays in quantum memory. Finally, we measure the output state by CV Bell measurement (BM), which can be implemented by passing through a 50:50 beam splitter and performing homodyne measurement on output ports along different quadratures [27]. Formally, the BM POVM element labeled by {ζn}𝜁superscript𝑛\{\zeta\in\mathbb{C}^{n}\}{ italic_ζ ∈ blackboard_C start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT } has the following form: (ID^(ζ))|ΨΨ|(ID^(ζ))/πntensor-product𝐼^𝐷𝜁ketΨbraΨtensor-product𝐼superscript^𝐷𝜁superscript𝜋𝑛(I\otimes\hat{D}(\zeta))|\Psi\rangle\langle\Psi|(I\otimes\hat{D}^{\dagger}(% \zeta))/\pi^{n}( italic_I ⊗ over^ start_ARG italic_D end_ARG ( italic_ζ ) ) | roman_Ψ ⟩ ⟨ roman_Ψ | ( italic_I ⊗ over^ start_ARG italic_D end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ( italic_ζ ) ) / italic_π start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT; here |ΨketΨ|\Psi\rangle| roman_Ψ ⟩ denotes the tensor product of n𝑛nitalic_n infinitely squeezed TMSV states, each proportional to k=0|k|ksuperscriptsubscript𝑘0ket𝑘ket𝑘\sum_{k=0}^{\infty}|k\rangle|k\rangle∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT | italic_k ⟩ | italic_k ⟩ when expressed in the Fock basis. To see how to learn a random displacement channel using this TMSV+BM scheme, we invoke the probability of obtaining outcome ζ𝜁\zetaitalic_ζ from BM (see SM S2 A [29] for the derivation):

pEA(ζ)subscript𝑝𝐸𝐴𝜁\displaystyle p_{EA}(\zeta)italic_p start_POSTSUBSCRIPT italic_E italic_A end_POSTSUBSCRIPT ( italic_ζ ) =1π2nd2nαλ(α)ee2r|α|2eαζζα.absent1superscript𝜋2𝑛superscript𝑑2𝑛𝛼𝜆𝛼superscript𝑒superscript𝑒2𝑟superscript𝛼2superscript𝑒superscript𝛼𝜁superscript𝜁𝛼\displaystyle=\frac{1}{\pi^{2n}}\int d^{2n}\alpha~{}\lambda(\alpha)e^{-e^{-2r}% |\alpha|^{2}}e^{\alpha^{\dagger}\zeta-\zeta^{\dagger}\alpha}.= divide start_ARG 1 end_ARG start_ARG italic_π start_POSTSUPERSCRIPT 2 italic_n end_POSTSUPERSCRIPT end_ARG ∫ italic_d start_POSTSUPERSCRIPT 2 italic_n end_POSTSUPERSCRIPT italic_α italic_λ ( italic_α ) italic_e start_POSTSUPERSCRIPT - italic_e start_POSTSUPERSCRIPT - 2 italic_r end_POSTSUPERSCRIPT | italic_α | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT italic_α start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_ζ - italic_ζ start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT . (4)

Fourier transforming to invert this relation, we obtain

λ(β)𝜆𝛽\displaystyle\lambda(\beta)italic_λ ( italic_β ) =ee2r|β|2d2nζpEA(ζ)eζββζabsentsuperscript𝑒superscript𝑒2𝑟superscript𝛽2superscript𝑑2𝑛𝜁subscript𝑝𝐸𝐴𝜁superscript𝑒superscript𝜁𝛽superscript𝛽𝜁\displaystyle=e^{e^{-2r}|\beta|^{2}}\int d^{2n}\zeta~{}p_{EA}(\zeta)e^{\zeta^{% \dagger}\beta-\beta^{\dagger}\zeta}= italic_e start_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - 2 italic_r end_POSTSUPERSCRIPT | italic_β | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT ∫ italic_d start_POSTSUPERSCRIPT 2 italic_n end_POSTSUPERSCRIPT italic_ζ italic_p start_POSTSUBSCRIPT italic_E italic_A end_POSTSUBSCRIPT ( italic_ζ ) italic_e start_POSTSUPERSCRIPT italic_ζ start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_β - italic_β start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_ζ end_POSTSUPERSCRIPT (5)
:=ee2r|β|2λEA(β).\displaystyle\mathrel{\mathop{:}}\nobreak\mkern-1.2mu=e^{e^{-2r}|\beta|^{2}}% \lambda_{EA}(\beta).: = italic_e start_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - 2 italic_r end_POSTSUPERSCRIPT | italic_β | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT italic_λ start_POSTSUBSCRIPT italic_E italic_A end_POSTSUBSCRIPT ( italic_β ) .

This expression indicates that, by sampling N𝑁Nitalic_N measurement outcomes {ζ(i)}i=1Nsuperscriptsubscriptsuperscript𝜁𝑖𝑖1𝑁\{\zeta^{(i)}\}_{i=1}^{N}{ italic_ζ start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT } start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT from a TMSV+BM scheme, one can obtain an unbiased estimator λ~(β)~𝜆𝛽\tilde{\lambda}(\beta)over~ start_ARG italic_λ end_ARG ( italic_β ) of λ(β)𝜆𝛽\lambda(\beta)italic_λ ( italic_β ) by defining λ~(β):=1Nee2r|β|2i=1Neζ(i)ββζ(i)\tilde{\lambda}(\beta)\mathrel{\mathop{:}}\nobreak\mkern-1.2mu=\frac{1}{N}e^{e% ^{-2r}|\beta|^{2}}\sum_{i=1}^{N}e^{\zeta^{(i)\dagger}\beta-\beta^{\dagger}% \zeta^{(i)}}over~ start_ARG italic_λ end_ARG ( italic_β ) : = divide start_ARG 1 end_ARG start_ARG italic_N end_ARG italic_e start_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - 2 italic_r end_POSTSUPERSCRIPT | italic_β | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT italic_ζ start_POSTSUPERSCRIPT ( italic_i ) † end_POSTSUPERSCRIPT italic_β - italic_β start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_ζ start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT. Note that the same set of samples can be used to estimate λ(β)𝜆𝛽\lambda(\beta)italic_λ ( italic_β ) for different values of β𝛽\betaitalic_β just by modifying the estimator. Using the Hoeffding’s bound, we prove the following theorem (see SM S2 A for the proof):

Theorem 1.

For any n𝑛nitalic_n-mode random displacement channel Λnormal-Λ\Lambdaroman_Λ, after the TMSV+BM scheme with squeezing parameter r𝑟ritalic_r has learned from N𝑁Nitalic_N copies of Λnormal-Λ\Lambdaroman_Λ, and then received a query βn𝛽superscript𝑛\beta\in\mathbb{C}^{n}italic_β ∈ blackboard_C start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT, it can provide an estimator λ~(β)normal-~𝜆𝛽\tilde{\lambda}(\beta)over~ start_ARG italic_λ end_ARG ( italic_β ) of Λnormal-Λ\Lambdaroman_Λ’s characteristic function λ(β)𝜆𝛽\lambda(\beta)italic_λ ( italic_β ) such that |λ~(β)λ(β)|ϵnormal-~𝜆𝛽𝜆𝛽italic-ϵ|\tilde{\lambda}(\beta)-\lambda(\beta)|\leq\epsilon| over~ start_ARG italic_λ end_ARG ( italic_β ) - italic_λ ( italic_β ) | ≤ italic_ϵ with probability at least 1δ1𝛿1-\delta1 - italic_δ, with the number of samples N=8e2e2r|β|2ϵ2log4δ1𝑁8superscript𝑒2superscript𝑒2𝑟superscript𝛽2superscriptitalic-ϵ24superscript𝛿1N=8e^{2e^{-2r}|\beta|^{2}}\epsilon^{-2}\log 4\delta^{-1}italic_N = 8 italic_e start_POSTSUPERSCRIPT 2 italic_e start_POSTSUPERSCRIPT - 2 italic_r end_POSTSUPERSCRIPT | italic_β | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT italic_ϵ start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT roman_log 4 italic_δ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT.

Let us compare the TMSV+BM scheme with a particular entanglement-free scheme that uses the vacuum state as input and heterodyne detection (Vacuum+Heterodyne). Here, heterodyne detection is defined as a projection onto the (overcomplete) basis of coherent states, i.e., |ζζ|/πnket𝜁bra𝜁superscript𝜋𝑛|\zeta\rangle\langle\zeta|/\pi^{n}| italic_ζ ⟩ ⟨ italic_ζ | / italic_π start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT with ζn𝜁superscript𝑛\zeta\in\mathbb{C}^{n}italic_ζ ∈ blackboard_C start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT. Though it may not be the optimal entanglement-free scheme, this specific scheme helps us understand the limitations of entanglement-free schemes, which we capture more generally in Theorem 2 below.

In this scheme, the probability of obtaining POVM outcome ζ𝜁\zetaitalic_ζ is (see SM S2 B [29])

pVH(ζ)=1π2nd2nαλ(α)eαζζαe|α|2.subscript𝑝𝑉𝐻𝜁1superscript𝜋2𝑛superscript𝑑2𝑛𝛼𝜆𝛼superscript𝑒superscript𝛼𝜁superscript𝜁𝛼superscript𝑒superscript𝛼2\displaystyle p_{{VH}}(\zeta)=\frac{1}{\pi^{2n}}\int d^{2n}\alpha~{}\lambda(% \alpha)e^{\alpha^{\dagger}\zeta-\zeta^{\dagger}\alpha}e^{-|\alpha|^{2}}.italic_p start_POSTSUBSCRIPT italic_V italic_H end_POSTSUBSCRIPT ( italic_ζ ) = divide start_ARG 1 end_ARG start_ARG italic_π start_POSTSUPERSCRIPT 2 italic_n end_POSTSUPERSCRIPT end_ARG ∫ italic_d start_POSTSUPERSCRIPT 2 italic_n end_POSTSUPERSCRIPT italic_α italic_λ ( italic_α ) italic_e start_POSTSUPERSCRIPT italic_α start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_ζ - italic_ζ start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - | italic_α | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT . (6)

In fact, the Vacuum+Heterodyne scheme can be understood as the TMSV+BM scheme with r=0𝑟0r=0italic_r = 0. Inverting this relation by Fourier transforming, we may express the channel’s characteristic function in terms of the measurement probability distribution:

λ(β)=e|β|2d2nζpVH(ζ)eζββζ:=e|β|2λVH(β),\displaystyle\lambda(\beta)=e^{|\beta|^{2}}\int d^{2n}\zeta~{}p_{{VH}}(\zeta)e% ^{\zeta^{\dagger}\beta-\beta^{\dagger}\zeta}\mathrel{\mathop{:}}\nobreak\mkern% -1.2mu=e^{|\beta|^{2}}\lambda_{VH}(\beta),italic_λ ( italic_β ) = italic_e start_POSTSUPERSCRIPT | italic_β | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT ∫ italic_d start_POSTSUPERSCRIPT 2 italic_n end_POSTSUPERSCRIPT italic_ζ italic_p start_POSTSUBSCRIPT italic_V italic_H end_POSTSUBSCRIPT ( italic_ζ ) italic_e start_POSTSUPERSCRIPT italic_ζ start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_β - italic_β start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_ζ end_POSTSUPERSCRIPT : = italic_e start_POSTSUPERSCRIPT | italic_β | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT italic_λ start_POSTSUBSCRIPT italic_V italic_H end_POSTSUBSCRIPT ( italic_β ) , (7)

which yields another unbiased estimator λ~(β):=1Ne|β|2i=1Neζ(i)ββζ(i)\tilde{\lambda}(\beta)\mathrel{\mathop{:}}\nobreak\mkern-1.2mu=\frac{1}{N}e^{|% \beta|^{2}}\sum_{i=1}^{N}e^{\zeta^{(i)\dagger}\beta-\beta^{\dagger}\zeta^{(i)}}over~ start_ARG italic_λ end_ARG ( italic_β ) : = divide start_ARG 1 end_ARG start_ARG italic_N end_ARG italic_e start_POSTSUPERSCRIPT | italic_β | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT italic_ζ start_POSTSUPERSCRIPT ( italic_i ) † end_POSTSUPERSCRIPT italic_β - italic_β start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_ζ start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT given N𝑁Nitalic_N samples {ζ(i)}i=1Nsuperscriptsubscriptsuperscript𝜁𝑖𝑖1𝑁\{\zeta^{(i)}\}_{i=1}^{N}{ italic_ζ start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT } start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT. Comparing to (5), we see that the r𝑟ritalic_r-dependent prefactor is missing from (7). Specifically, if we confine β𝛽\betaitalic_β to |β|2κnsuperscript𝛽2𝜅𝑛|\beta|^{2}\leq\kappa n| italic_β | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ≤ italic_κ italic_n with a constant κ>0𝜅0\kappa>0italic_κ > 0, we obtain upper bounds on the sample complexity for achieving an error ϵitalic-ϵ\epsilonitalic_ϵ with success probability 1δ1𝛿1{-}\delta1 - italic_δ using each scheme:

NEAsubscript𝑁𝐸𝐴\displaystyle N_{EA}italic_N start_POSTSUBSCRIPT italic_E italic_A end_POSTSUBSCRIPT =O(e2e2rκnϵ2logδ1),absent𝑂superscript𝑒2superscript𝑒2𝑟𝜅𝑛superscriptitalic-ϵ2superscript𝛿1\displaystyle=O(e^{2e^{-2r}\kappa n}\epsilon^{-2}\log\delta^{-1}),= italic_O ( italic_e start_POSTSUPERSCRIPT 2 italic_e start_POSTSUPERSCRIPT - 2 italic_r end_POSTSUPERSCRIPT italic_κ italic_n end_POSTSUPERSCRIPT italic_ϵ start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT roman_log italic_δ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ) , (8)
NVHsubscript𝑁𝑉𝐻\displaystyle N_{VH}italic_N start_POSTSUBSCRIPT italic_V italic_H end_POSTSUBSCRIPT =O(e2κnϵ2logδ1).absent𝑂superscript𝑒2𝜅𝑛superscriptitalic-ϵ2superscript𝛿1\displaystyle=O(e^{2\kappa n}\epsilon^{-2}\log\delta^{-1}).= italic_O ( italic_e start_POSTSUPERSCRIPT 2 italic_κ italic_n end_POSTSUPERSCRIPT italic_ϵ start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT roman_log italic_δ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ) . (9)

In particular, if we choose the squeezing parameter as r=Ω(logn)𝑟Ω𝑛r=\Omega(\log n)italic_r = roman_Ω ( roman_log italic_n ), the sample complexity NEA=O(ϵ2logδ1)subscript𝑁𝐸𝐴𝑂superscriptitalic-ϵ2superscript𝛿1N_{EA}=O(\epsilon^{-2}\log\delta^{-1})italic_N start_POSTSUBSCRIPT italic_E italic_A end_POSTSUBSCRIPT = italic_O ( italic_ϵ start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT roman_log italic_δ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ) of the entanglement-assisted scheme becomes independent of the number of modes n𝑛nitalic_n, while our upper bound on sample complexity of the entanglement-free scheme increases exponentially with n𝑛nitalic_n. Since the accessible squeezing parameter is bounded in practice, though, we will compare the sample complexities of the two schemes when r𝑟ritalic_r is an n𝑛nitalic_n-independent constant below.

To illustrate the difference, we compare TMSV+BM and Vacuum+Heterodyne strategies with an example in Fig. 2. We consider a single-mode channel for ease of visualization, characterized by

p(α)𝑝𝛼\displaystyle p(\alpha)italic_p ( italic_α ) =2σ2πe2σ2|α|2[cos2(αrγiαiγr)\displaystyle=\frac{2\sigma^{2}}{\pi}e^{-2\sigma^{2}|\alpha|^{2}}[\cos^{2}(% \alpha_{r}\gamma_{i}-\alpha_{i}\gamma_{r})= divide start_ARG 2 italic_σ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_π end_ARG italic_e start_POSTSUPERSCRIPT - 2 italic_σ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT | italic_α | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT [ roman_cos start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_α start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT italic_γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_γ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT )
+sin2(αrγr+αiγi)],\displaystyle~{}~{}~{}~{}~{}+\sin^{2}(\alpha_{r}\gamma_{r}+\alpha_{i}\gamma_{i% })],+ roman_sin start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_α start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT italic_γ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT + italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) ] , (10)
λ(β)𝜆𝛽\displaystyle\lambda(\beta)italic_λ ( italic_β ) =e|β|22σ2+14e|βγ|22σ2+14e|β+γ|22σ2absentsuperscript𝑒superscript𝛽22superscript𝜎214superscript𝑒superscript𝛽𝛾22superscript𝜎214superscript𝑒superscript𝛽𝛾22superscript𝜎2\displaystyle=e^{-\frac{|\beta|^{2}}{2\sigma^{2}}}+\frac{1}{4}e^{-\frac{|\beta% -\gamma|^{2}}{2\sigma^{2}}}+\frac{1}{4}e^{-\frac{|\beta+\gamma|^{2}}{2\sigma^{% 2}}}= italic_e start_POSTSUPERSCRIPT - divide start_ARG | italic_β | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_σ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_POSTSUPERSCRIPT + divide start_ARG 1 end_ARG start_ARG 4 end_ARG italic_e start_POSTSUPERSCRIPT - divide start_ARG | italic_β - italic_γ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_σ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_POSTSUPERSCRIPT + divide start_ARG 1 end_ARG start_ARG 4 end_ARG italic_e start_POSTSUPERSCRIPT - divide start_ARG | italic_β + italic_γ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_σ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_POSTSUPERSCRIPT
14e|βiγ|22σ214e|β+iγ|22σ2,14superscript𝑒superscript𝛽𝑖𝛾22superscript𝜎214superscript𝑒superscript𝛽𝑖𝛾22superscript𝜎2\displaystyle~{}~{}~{}~{}~{}-\frac{1}{4}e^{-\frac{|\beta-i\gamma|^{2}}{2\sigma% ^{2}}}-\frac{1}{4}e^{-\frac{|\beta+i\gamma|^{2}}{2\sigma^{2}}},- divide start_ARG 1 end_ARG start_ARG 4 end_ARG italic_e start_POSTSUPERSCRIPT - divide start_ARG | italic_β - italic_i italic_γ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_σ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_POSTSUPERSCRIPT - divide start_ARG 1 end_ARG start_ARG 4 end_ARG italic_e start_POSTSUPERSCRIPT - divide start_ARG | italic_β + italic_i italic_γ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_σ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_POSTSUPERSCRIPT , (11)

with σ=0.3𝜎0.3\sigma=0.3italic_σ = 0.3, γr=1.6subscript𝛾𝑟1.6\gamma_{r}=1.6italic_γ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT = 1.6, γi=0subscript𝛾𝑖0\gamma_{i}=0italic_γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 0 (γ:=γr+iγi\gamma\mathrel{\mathop{:}}\nobreak\mkern-1.2mu=\gamma_{r}+i\gamma_{i}italic_γ : = italic_γ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT + italic_i italic_γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT), and r=2𝑟2r=2italic_r = 2 for the TMSV+BM scheme. The figure, where we present the underlying output probability distributions and their characteristic functions from Eqs. (4),(5),(6), and (7), clearly shows that in the TMSV+BM scheme with a sufficiently large squeezing parameter, the resultant probability distribution and characteristic function are almost identical to the ideal case. However, for the Vacuum+Heterodyne scheme, the vacuum noise distorts the initial probability distribution so significantly that we cannot see the signal clearly, which thus makes it harder to estimate the original characteristic function.

Refer to caption
Figure 2: Comparison between (a) the true distribution, (b) TMSV+BM, and (c) Vacuum+Heterodyne strategies. The left panel represents the probability distribution of the true distribution and measurement probability distributions for each scheme. The right panel represents the characteristic function of probability distributions.

Lower bound.— Our upper bound on the sample complexity of the Vacuum+Heterodyne scheme scales exponentially with n𝑛nitalic_n. Can this scaling be improved using more advanced entanglement-free schemes, such as homodyne or general-dyne detection [32, 27], or by non-Gaussian resources like GKP states [33] or photon-number resolving measurements [34]? Here, using information-theoretic methods, we prove an exponential sample complexity lower bound for any entanglement-free scheme. This highlights the indispensable role of entanglement for efficiently learning bosonic random displacement channels. Our result is as follows:

Theorem 2.

Let Λnormal-Λ\Lambdaroman_Λ be an arbitrary n𝑛nitalic_n-mode random displacement channel (n8𝑛8n\geq 8italic_n ≥ 8) and consider an entanglement-free scheme that uses N𝑁Nitalic_N copies of Λnormal-Λ\Lambdaroman_Λ. After all measurements are completed, the scheme receives the query βn𝛽superscript𝑛\beta\in\mathbb{C}^{n}italic_β ∈ blackboard_C start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT and returns an estimate λ~(β)normal-~𝜆𝛽\tilde{\lambda}(\beta)over~ start_ARG italic_λ end_ARG ( italic_β ) of Λnormal-Λ\Lambdaroman_Λ’s characteristic function λ(β)𝜆𝛽\lambda(\beta)italic_λ ( italic_β ). Suppose that, with success probability at least 2/3232/32 / 3, |λ~(β)λ(β)|ϵ0.24normal-~𝜆𝛽𝜆𝛽italic-ϵ0.24|\tilde{\lambda}(\beta)-\lambda(\beta)|\leq\epsilon\leq 0.24| over~ start_ARG italic_λ end_ARG ( italic_β ) - italic_λ ( italic_β ) | ≤ italic_ϵ ≤ 0.24 for all β𝛽\betaitalic_β such that |β|2nκsuperscript𝛽2𝑛𝜅|\beta|^{2}\leq n\kappa| italic_β | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ≤ italic_n italic_κ. Then N0.01ϵ2(1+1.98κ)n𝑁0.01superscriptitalic-ϵ2superscript11.98𝜅𝑛N\geq 0.01\epsilon^{-2}(1+1.98\kappa)^{n}italic_N ≥ 0.01 italic_ϵ start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT ( 1 + 1.98 italic_κ ) start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT.

Here, the choice of success probability 2/3232/32 / 3 is arbitrary and can be easily amplified. Comparing with the entanglement-assisted sample complexity given in Eq. (8), Theorem 2 establishes a separation exponential in n𝑛nitalic_n for cutoff coefficient κ=O(1)𝜅𝑂1\kappa=O(1)italic_κ = italic_O ( 1 ) and squeezing parameter r=Ω(logn)𝑟Ω𝑛r=\Omega(\log n)italic_r = roman_Ω ( roman_log italic_n ). The intuition underlying this theorem is that displacement operators D^(β)^𝐷𝛽\hat{D}(\beta)over^ start_ARG italic_D end_ARG ( italic_β ) do not generally commute with each other. Consequently, entanglement-free measurements can resolve λ(β)𝜆𝛽\lambda(\beta)italic_λ ( italic_β ) for only a small portion of β𝛽\betaitalic_β space. We sketch the proof below and leave the full details to SM S3 [29].

Proof Sketch.

Our proof extends the techniques of Refs. [23, 18] to the CV case. We begin by defining the following family of “3333-peak” random displacement channels 𝚲3peakϵ,σ={Λγ}γnsubscriptsuperscript𝚲italic-ϵ𝜎3peaksubscriptsubscriptΛ𝛾𝛾superscript𝑛\bm{\Lambda}^{\epsilon,\sigma}_{\mathrm{3\mhyphen peak}}=\{\Lambda_{\gamma}\}_% {\gamma\in\mathbb{C}^{n}}bold_Λ start_POSTSUPERSCRIPT italic_ϵ , italic_σ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 3 roman_p roman_e roman_a roman_k end_POSTSUBSCRIPT = { roman_Λ start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT } start_POSTSUBSCRIPT italic_γ ∈ blackboard_C start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_POSTSUBSCRIPT whose characteristic functions and distributions of displacement are, respectively,

λγ(β)subscript𝜆𝛾𝛽\displaystyle\lambda_{\gamma}(\beta)italic_λ start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT ( italic_β ) =e|β|22σ2+2iϵ0e|βγ|22σ22iϵ0e|β+γ|22σ2,absentsuperscript𝑒superscript𝛽22superscript𝜎22𝑖subscriptitalic-ϵ0superscript𝑒superscript𝛽𝛾22superscript𝜎22𝑖subscriptitalic-ϵ0superscript𝑒superscript𝛽𝛾22superscript𝜎2\displaystyle=e^{-\frac{|\beta|^{2}}{2\sigma^{2}}}+2i\epsilon_{0}e^{-\frac{|% \beta-\gamma|^{2}}{2\sigma^{2}}}-2i\epsilon_{0}e^{-\frac{|\beta+\gamma|^{2}}{2% \sigma^{2}}},= italic_e start_POSTSUPERSCRIPT - divide start_ARG | italic_β | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_σ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_POSTSUPERSCRIPT + 2 italic_i italic_ϵ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - divide start_ARG | italic_β - italic_γ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_σ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_POSTSUPERSCRIPT - 2 italic_i italic_ϵ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - divide start_ARG | italic_β + italic_γ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_σ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_POSTSUPERSCRIPT , (12)
pγ(α)subscript𝑝𝛾𝛼\displaystyle p_{\gamma}(\alpha)italic_p start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT ( italic_α ) e2σ2|α|2(1+4ϵ0sin(2(γiαrγrαi))).proportional-toabsentsuperscript𝑒2superscript𝜎2superscript𝛼214subscriptitalic-ϵ02subscript𝛾𝑖subscript𝛼𝑟subscript𝛾𝑟subscript𝛼𝑖\displaystyle\propto e^{-2\sigma^{2}|\alpha|^{2}}\left(1+4\epsilon_{0}\sin(2(% \gamma_{i}\alpha_{r}-\gamma_{r}\alpha_{i}))\right).∝ italic_e start_POSTSUPERSCRIPT - 2 italic_σ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT | italic_α | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT ( 1 + 4 italic_ϵ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT roman_sin ( 2 ( italic_γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_α start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT - italic_γ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) ) ) . (13)

with positive parameters σ𝜎\sigmaitalic_σ and ϵ:=0.98ϵ0\epsilon\mathrel{\mathop{:}}\nobreak\mkern-1.2mu=0.98\epsilon_{0}italic_ϵ : = 0.98 italic_ϵ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. We will show that, even with the prior knowledge that the channel is from this family, it is still hard for entanglement-free schemes to complete the learning tasks.

The key idea is to reduce learning to binary hypothesis testing. Consider the following game between Alice and Bob: Alice chooses one of two hypotheses with equal probability: (1) Set Λ=Λ0ΛsubscriptΛ0\Lambda=\Lambda_{0}roman_Λ = roman_Λ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT; (2) Set Λ=ΛγΛsubscriptΛ𝛾\Lambda=\Lambda_{\gamma}roman_Λ = roman_Λ start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT, for γ𝛾\gammaitalic_γ sampled from a zero-mean homogeneous Gaussian distribution whose variance is determined by κ𝜅\kappaitalic_κ. Next, Alice allows Bob to use the channel ΛΛ\Lambdaroman_Λ N𝑁Nitalic_N times, and Bob uses his favorite entanglement-free scheme to learn from these channel uses. After Bob has finished all quantum measurements and keeps only classical data, Alice reveals some auxiliary information to Bob, who is then asked to decide whether Alice has chosen (1) or (2).

Given a learning scheme satisfying the assumptions of Theorem 2, Bob can guess correctly with high probability. This means that the outcome distributions of Bob’s scheme under hypotheses (1) and (2) must have a sufficiently large total variation distance (TVD). On the other hand, we can upper bound the contribution from each use of ΛΛ\Lambdaroman_Λ to the TVD to be exponentially small, where we use a technique inspired by Ref. [35] which derived the maximum fidelity of Gaussian random displacement channels. Therefore, the number of channel uses N𝑁Nitalic_N must be exponentially large to ensure a large enough TVD, which gives us the desired lower bound. ∎

Refer to caption
Figure 3: (a) Comparison of TMSV+BM (with different loss rates), Vacuum+Heterodyne, and the entanglement-free lower bound at κ=1𝜅1\kappa=1italic_κ = 1. The task is to estimate any λ(β)𝜆𝛽\lambda(\beta)italic_λ ( italic_β ) such that |β|2κnsuperscript𝛽2𝜅𝑛|\beta|^{2}\leq\kappa n| italic_β | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ≤ italic_κ italic_n with precision ε=0.2𝜀0.2\varepsilon=0.2italic_ε = 0.2 and success probability 1δ=2/31𝛿231-\delta=2/31 - italic_δ = 2 / 3. The orange region represents a rigorous advantage over all entanglement-free schemes. The blue region represents an advantage over noiseless Vacuum+Heterodyne. (b) Comparison of the TMSV+BM scheme with squeezing parameter r=1.0𝑟1.0r=1.0italic_r = 1.0 and loss rate 1T=0.11𝑇0.11-T=0.11 - italic_T = 0.1 with the entanglement-free lower bound of Theorem 2. (See SM S3 A for further practical considerations.) The task is the same as (a). The brown solid contour lines represent the sample complexity of TMSV+BM given by Theorem 3. The blue dashed contour lines represent the ratio of sample complexity between the entanglement-free lower bound and TMSV+BM, indicating the entanglement-enabled advantage.

Effect of loss.— Now, for practical applications, we study how the entanglement-assisted scheme is affected by photon loss, a dominant noise source in optical platforms (see SM S2 D for a discussion of more general noise models, such as phase diffusion). Photon loss transforms the relevant bosonic operator a^^𝑎\hat{a}over^ start_ARG italic_a end_ARG to Ta^+1Te^𝑇^𝑎1𝑇^𝑒\sqrt{T}\hat{a}+\sqrt{1-T}\hat{e}square-root start_ARG italic_T end_ARG over^ start_ARG italic_a end_ARG + square-root start_ARG 1 - italic_T end_ARG over^ start_ARG italic_e end_ARG, where T𝑇Titalic_T is the transmission rate and e^^𝑒\hat{e}over^ start_ARG italic_e end_ARG is the environmental mode, i.e., 1T1𝑇1-T1 - italic_T is the loss rate. We consider two different places where the loss occurs: one is before applying the channel with loss rate 1Tb1subscript𝑇𝑏1-T_{b}1 - italic_T start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT to model the preparation imperfection, and the other is after applying the channel and before the perfect BM with loss rate 1Ta1subscript𝑇𝑎1-T_{a}1 - italic_T start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT, which models the finite efficiency of detection, i.e., an imperfect BM [27]. As before, we derive the relation between the measurement probability distribution and the characteristic function of the channel (with appropriate rescaling of the phase):

λ(β)𝜆𝛽\displaystyle\lambda(\beta)italic_λ ( italic_β ) =ee2reff|β|2d2nζploss(ζ)e(ζββζ)/Ta,absentsuperscript𝑒superscript𝑒2subscript𝑟effsuperscript𝛽2superscript𝑑2𝑛𝜁subscript𝑝𝑙𝑜𝑠𝑠𝜁superscript𝑒superscript𝜁𝛽superscript𝛽𝜁subscript𝑇𝑎\displaystyle=e^{e^{-2r_{\mathrm{eff}}}|\beta|^{2}}\int d^{2n}\zeta~{}p_{loss}% (\zeta)e^{(\zeta^{\dagger}\beta-\beta^{\dagger}\zeta)/\sqrt{T_{a}}},= italic_e start_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - 2 italic_r start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT end_POSTSUPERSCRIPT | italic_β | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT ∫ italic_d start_POSTSUPERSCRIPT 2 italic_n end_POSTSUPERSCRIPT italic_ζ italic_p start_POSTSUBSCRIPT italic_l italic_o italic_s italic_s end_POSTSUBSCRIPT ( italic_ζ ) italic_e start_POSTSUPERSCRIPT ( italic_ζ start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_β - italic_β start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_ζ ) / square-root start_ARG italic_T start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT end_ARG end_POSTSUPERSCRIPT , (14)

where we define an effective squeezing parameter

reff:=12log(Tbe2r+(1Tb)+1TaTa),r_{\text{eff}}\mathrel{\mathop{:}}\nobreak\mkern-1.2mu=-\frac{1}{2}\log\left(T% _{b}e^{-2r}+(1-T_{b})+\frac{1-T_{a}}{T_{a}}\right),italic_r start_POSTSUBSCRIPT eff end_POSTSUBSCRIPT : = - divide start_ARG 1 end_ARG start_ARG 2 end_ARG roman_log ( italic_T start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - 2 italic_r end_POSTSUPERSCRIPT + ( 1 - italic_T start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ) + divide start_ARG 1 - italic_T start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT end_ARG start_ARG italic_T start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT end_ARG ) , (15)

which incorporates the loss rates. Because loss degrades the advantage from squeezing, the upper bound on sample complexity in Theorem 1 is modified in the presence of loss (see SM S2 C for the proof):

Theorem 3.

For the same task as in Theorem 1, a TMVS+BM scheme with squeezing parameter r𝑟ritalic_r and transmission rates before and after the channel to be Tbsubscript𝑇𝑏T_{b}italic_T start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT and Tasubscript𝑇𝑎T_{a}italic_T start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT, respectively, can estimate any λ(β)𝜆𝛽\lambda(\beta)italic_λ ( italic_β ) to error ϵitalic-ϵ\epsilonitalic_ϵ with success probability 1δ1𝛿1{-}\delta1 - italic_δ using the number of samples N=8e2e2r𝑒𝑓𝑓|β|2ϵ2log4δ1𝑁8superscript𝑒2superscript𝑒2subscript𝑟𝑒𝑓𝑓superscript𝛽2superscriptitalic-ϵ24superscript𝛿1N=8e^{2e^{-2r_{\text{eff}}|\beta|^{2}}}\epsilon^{-2}\log 4\delta^{-1}italic_N = 8 italic_e start_POSTSUPERSCRIPT 2 italic_e start_POSTSUPERSCRIPT - 2 italic_r start_POSTSUBSCRIPT eff end_POSTSUBSCRIPT | italic_β | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT italic_ϵ start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT roman_log 4 italic_δ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, where reffsubscript𝑟normal-effr_{\mathrm{eff}}italic_r start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT is defined according to Eq. (15).

Thus, when |β|2κnsuperscript𝛽2𝜅𝑛|\beta|^{2}\leq\kappa n| italic_β | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ≤ italic_κ italic_n with a constant κ>0𝜅0\kappa>0italic_κ > 0, Tb=1O(1/n)subscript𝑇𝑏1𝑂1𝑛T_{b}=1-O(1/n)italic_T start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT = 1 - italic_O ( 1 / italic_n ), Ta=1O(1/n)subscript𝑇𝑎1𝑂1𝑛T_{a}=1-O(1/n)italic_T start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT = 1 - italic_O ( 1 / italic_n ) and r=Ω(logn)𝑟Ω𝑛r=\Omega(\log n)italic_r = roman_Ω ( roman_log italic_n ), the sample complexity becomes N=O(ϵ2logδ1)𝑁𝑂superscriptitalic-ϵ2superscript𝛿1N=O(\epsilon^{-2}\log\delta^{-1})italic_N = italic_O ( italic_ϵ start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT roman_log italic_δ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ) as in the lossless case. For practically relevant squeezing and including loss prior to Bell measurement, we compare the sample complexity for the lossy TMSV+BM protocol and the lossless entanglement-free lower bound in Fig. 3, finding a significant entanglement-enabled advantage in realistic experimental settings. Specifically, for reasonable parameter choices such as squeezing parameter r=1𝑟1r=1italic_r = 1, loss rate 10%percent1010\%10 %, and κ=O(1)𝜅𝑂1\kappa=O(1)italic_κ = italic_O ( 1 ), we can achieve a factor of 104(108)superscript104superscript10810^{4}~{}(10^{8})10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT ( 10 start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT ) advantage for around n=30(60)𝑛3060n=30~{}(60)italic_n = 30 ( 60 ) modes. Although the 109superscript10910^{9}10 start_POSTSUPERSCRIPT 9 end_POSTSUPERSCRIPT number of samples required to achieve the advantage seems large, the state-of-the-art quantum optics experiments (e.g., Refs. [36, 37]) can attain such number of samples in a reasonable time with high sampling rate up to 160 GHz.

Discussion.— We proved that schemes that exploit entanglement with an ancillary quantum memory can learn n𝑛nitalic_n-mode bosonic random displacement channels with exponentially fewer samples compared to entanglement-free schemes. Our results show that the information-theoretic framework for learning studied in DV quantum systems [17, 19] can be generalized to the CV setting and have powerful implications. We anticipate that these techniques can be applied to other CV learning tasks as well. In addition, our analysis suggests that the separation in sample complexity between entanglement-assisted and entanglement-free protocols may be realized in the near future.

Apart from their theoretical interest, random displacement channels can also be practically relevant in, e.g., modeling noise in bosonic systems. As in the qubit case [38], we expect that noise tailoring methods can transform more general noise models into random displacement channels; therefore efficiently learning random displacement channels can be useful for benchmarking CV quantum systems [39, 40] and for error mitigation.

Displacement estimation is also studied in quantum metrology (see e.g. [41, 42, 43]). A task often considered in metrology is learning an unknown unitary displacement or phase transformation acting independently on each mode [44, 42, 36, 45, 46, 47] whereas the task analyzed in this word is learning an unknown mixture of multimode displacements. Furthermore, while the goal in metrology is typically to learn one or a few parameters, in our case, the parameter space is very large. Therefore, the methodology in the two settings is quite different. Connections between metrology and bosonic channel learning are worthy of further exploration.


Acknowledgements.
We thank Mankei Tsang, Yuxin Wang, Ronald de Wolf, Mingxing Yao, Ming Yuan for insightful discussions. C.O., S.C., Y.W., L.J. acknowledge support from the ARO(W911NF-23-1-0077), ARO MURI (W911NF-21-1-0325), AFOSR MURI (FA9550-19-1-0399, FA9550-21-1-0209), NSF (OMA-1936118, ERC-1941583, OMA-2137642), NTT Research, Packard Foundation (2020-71479). J.P. acknowledges support from the U.S. Department of Energy Office of Science, Office of Advanced Scientific Computing Research (DE-NA0003525, DE-SC0020290), the U.S. Department of Energy, Office of Science, National Quantum Information Science Research Centers, Quantum Systems Accelerator, and the National Science Foundation (PHY-1733907). The Institute for Quantum Information and Matter is an NSF Physics Frontiers Center. S.Z. acknowledges funding provided by the Institute for Quantum Information and Matter and Perimeter Institute for Theoretical Physics, a research institute supported in part by the Government of Canada through the Department of Innovation, Science and Economic Development Canada and by the Province of Ontario through the Ministry of Colleges and Universities. J.A.H.N, Z.L., J.S.N and U.L.A acknowledge support from DNRF (bigQ, DNRF142), IFD (photoQ) and EU (CLUSTEC, ClusterQ ERC-101055224, GTGBS MC-101106833).

References

  • Nielsen and Chuang [2002] M. A. Nielsen and I. Chuang, Quantum computation and quantum information (2002).
  • Gisin and Thew [2007] N. Gisin and R. Thew, Quantum communication, Nature photonics 1, 165 (2007).
  • Kimble [2008] H. J. Kimble, The quantum internet, Nature 453, 1023 (2008).
  • Giovannetti et al. [2006] V. Giovannetti, S. Lloyd, and L. Maccone, Quantum metrology, Physical review letters 96, 010401 (2006).
  • Giovannetti et al. [2011] V. Giovannetti, S. Lloyd, and L. Maccone, Advances in quantum metrology, Nature photonics 5, 222 (2011).
  • Polino et al. [2020] E. Polino, M. Valeri, N. Spagnolo, and F. Sciarrino, Photonic quantum metrology, AVS Quantum Science 2, 024703 (2020).
  • Preskill [2018] J. Preskill, Quantum computing in the NISQ era and beyond, Quantum 2, 79 (2018).
  • Aaronson and Arkhipov [2011] S. Aaronson and A. Arkhipov, The computational complexity of linear optics, in Proceedings of the forty-third annual ACM symposium on Theory of computing (2011) pp. 333–342.
  • Boixo et al. [2018] S. Boixo, S. V. Isakov, V. N. Smelyanskiy, R. Babbush, N. Ding, Z. Jiang, M. J. Bremner, J. M. Martinis, and H. Neven, Characterizing quantum supremacy in near-term devices, Nature Physics 14, 595 (2018).
  • Arute et al. [2019] F. Arute, K. Arya, R. Babbush, D. Bacon, J. C. Bardin, R. Barends, R. Biswas, S. Boixo, F. G. Brandao, D. A. Buell, et al., Quantum supremacy using a programmable superconducting processor, Nature 574, 505 (2019).
  • Wu et al. [2021] Y. Wu, W.-S. Bao, S. Cao, F. Chen, M.-C. Chen, X. Chen, T.-H. Chung, H. Deng, Y. Du, D. Fan, et al., Strong quantum computational advantage using a superconducting quantum processor, Physical review letters 127, 180501 (2021).
  • Zhong et al. [2020] H.-S. Zhong, H. Wang, Y.-H. Deng, M.-C. Chen, L.-C. Peng, Y.-H. Luo, J. Qin, D. Wu, X. Ding, Y. Hu, et al., Quantum computational advantage using photons, Science 370, 1460 (2020).
  • Zhong et al. [2021] H.-S. Zhong, Y.-H. Deng, J. Qin, H. Wang, M.-C. Chen, L.-C. Peng, Y.-H. Luo, D. Wu, S.-Q. Gong, H. Su, et al., Phase-programmable Gaussian boson sampling using stimulated squeezed light, Physical review letters 127, 180502 (2021).
  • 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).
  • Morvan et al. [2023] A. Morvan, B. Villalonga, X. Mi, S. Mandra, A. Bengtsson, P. Klimov, Z. Chen, S. Hong, C. Erickson, I. Drozdov, et al., Phase transition in random circuit sampling, arXiv preprint arXiv:2304.11119  (2023).
  • Deng et al. [2023] Y.-H. Deng, Y.-C. Gu, H.-L. Liu, S.-Q. Gong, H. Su, Z.-J. Zhang, H.-Y. Tang, M.-H. Jia, J.-M. Xu, M.-C. Chen, J. Qin, L.-C. Peng, J. Yan, Y. Hu, J. Huang, H. Li, Y. Li, Y. Chen, X. Jiang, L. Gan, G. Yang, L. You, L. Li, H.-S. Zhong, H. Wang, N.-L. Liu, J. J. Renema, C.-Y. Lu, and J.-W. Pan, Gaussian boson sampling with pseudo-photon-number-resolving detectors and quantum computational advantage, Phys. Rev. Lett. 131, 150601 (2023).
  • Huang et al. [2021] H.-Y. Huang, R. Kueng, and J. Preskill, Information-theoretic bounds on quantum advantage in machine learning, Physical Review Letters 126, 190505 (2021).
  • Huang et al. [2022] H.-Y. Huang, M. Broughton, J. Cotler, S. Chen, J. Li, M. Mohseni, H. Neven, R. Babbush, R. Kueng, J. Preskill, and J. R. McClean, Quantum advantage in learning from experiments, Science 376, 1182 (2022).
  • Chen et al. [2022a] S. Chen, J. Cotler, H.-Y. Huang, and J. Li, Exponential separations between learning with and without quantum memory, in 2021 IEEE 62nd Annual Symposium on Foundations of Computer Science (FOCS) (IEEE, 2022) pp. 574–585.
  • Caro [2022] M. C. Caro, Learning quantum processes and hamiltonians via the pauli transfer matrix, arXiv preprint arXiv:2212.04471  (2022).
  • Bubeck et al. [2020] S. Bubeck, S. Chen, and J. Li, Entanglement is necessary for optimal quantum property testing, in 2020 IEEE 61st Annual Symposium on Foundations of Computer Science (FOCS) (IEEE, 2020) pp. 692–703.
  • Aharonov et al. [2022] D. Aharonov, J. Cotler, and X.-L. Qi, Quantum algorithmic measurement, Nature communications 13, 1 (2022).
  • Chen et al. [2022b] S. Chen, S. Zhou, A. Seif, and L. Jiang, Quantum advantages for pauli channel estimation, Phys. Rev. A 105, 032435 (2022b).
  • Rossi et al. [2022] Z. M. Rossi, J. Yu, I. L. Chuang, and S. Sugiura, Quantum advantage for noisy channel discrimination, Physical Review A 105, 032401 (2022).
  • Braunstein and Van Loock [2005] S. L. Braunstein and P. Van Loock, Quantum information with continuous variables, Reviews of modern physics 77, 513 (2005).
  • Weedbrook et al. [2012] C. Weedbrook, S. Pirandola, R. García-Patrón, N. J. Cerf, T. C. Ralph, J. H. Shapiro, and S. Lloyd, Gaussian quantum information, Reviews of Modern Physics 84, 621 (2012).
  • Serafini [2017] A. Serafini, Quantum continuous variables: a primer of theoretical methods (CRC press, 2017).
  • Wu et al. [2023] Y.-D. Wu, G. Chiribella, and N. Liu, Quantum-enhanced learning of continuous-variable quantum states, arXiv preprint arXiv:2303.05097  (2023).
  • [29] Supplemental material.
  • Chen and Gong [2023] S. Chen and W. Gong, Futility and utility of a few ancillas for pauli channel learning, arXiv preprint arXiv:2309.14326  (2023).
  • Chen et al. [2023] S. Chen, C. Oh, S. Zhou, H.-Y. Huang, and L. Jiang, Tight bounds on pauli channel learning without entanglement, arXiv preprint arXiv:2309.13461  (2023).
  • Schleich [2011] W. P. Schleich, Quantum optics in phase space (John Wiley & Sons, 2011).
  • Gottesman et al. [2001] D. Gottesman, A. Kitaev, and J. Preskill, Encoding a qubit in an oscillator, Physical Review A 64, 012310 (2001).
  • Schuster et al. [2007] D. Schuster, A. A. Houck, J. Schreier, A. Wallraff, J. Gambetta, A. Blais, L. Frunzio, J. Majer, B. Johnson, M. Devoret, et al., Resolving photon number states in a superconducting circuit, Nature 445, 515 (2007).
  • Caves and Wódkiewicz [2004] C. M. Caves and K. Wódkiewicz, Fidelity of gaussian channels, Open Systems & Information Dynamics 11, 309 (2004).
  • Guo et al. [2020] X. Guo, C. R. Breum, J. Borregaard, S. Izumi, M. V. Larsen, T. Gehring, M. Christandl, J. S. Neergaard-Nielsen, and U. L. Andersen, Distributed quantum sensing in a continuous-variable entangled network, Nature Physics 16, 281 (2020).
  • Inoue et al. [2023] A. Inoue, T. Kashiwazaki, T. Yamashima, N. Takanashi, T. Kazama, K. Enbutsu, K. Watanabe, T. Umeki, M. Endo, and A. Furusawa, Toward a multi-core ultra-fast optical quantum processor: 43-ghz bandwidth real-time amplitude measurement of 5-db squeezed light using modularized optical parametric amplifier with 5g technology, Applied Physics Letters 122, 104001 (2023).
  • Wallman and Emerson [2016] J. J. Wallman and J. Emerson, Noise tailoring for scalable quantum computation via randomized compiling, Physical Review A 94, 052325 (2016).
  • Wu and Sanders [2019] Y.-D. Wu and B. C. Sanders, Efficient verification of bosonic quantum channels via benchmarking, New Journal of Physics 21, 073026 (2019).
  • Bai and Chiribella [2018] G. Bai and G. Chiribella, Test one to test many: a unified approach to quantum benchmarks, Physical Review Letters 120, 150502 (2018).
  • Shi and Zhuang [2023] H. Shi and Q. Zhuang, Ultimate precision limit of noise sensing and dark matter search, npj Quantum Information 9, 27 (2023).
  • Zhuang et al. [2018] Q. Zhuang, Z. Zhang, and J. H. Shapiro, Distributed quantum sensing using continuous-variable multipartite entanglement, Physical Review A 97, 032329 (2018).
  • Xia et al. [2020] Y. Xia, W. Li, W. Clark, D. Hart, Q. Zhuang, and Z. Zhang, Demonstration of a reconfigurable entangled radio-frequency photonic sensor network, Physical Review Letters 124, 150502 (2020).
  • Duivenvoorden et al. [2017] K. Duivenvoorden, B. M. Terhal, and D. Weigand, Single-mode displacement sensor, Physical Review A 95, 012305 (2017).
  • Oh et al. [2020] C. Oh, C. Lee, S. H. Lie, and H. Jeong, Optimal distributed quantum sensing using gaussian states, Physical Review Research 2, 023030 (2020).
  • Oh et al. [2022] C. Oh, L. Jiang, and C. Lee, Distributed quantum phase sensing for arbitrary positive and negative weights, Physical Review Research 4, 023164 (2022).
  • Kwon et al. [2022] H. Kwon, Y. Lim, L. Jiang, H. Jeong, and C. Oh, Quantum metrological power of continuous-variable quantum networks, Physical Review Letters 128, 180503 (2022).