Multiparameter critical quantum metrology with impurity probes

George Mihailescu [email protected] School of Physics, University College Dublin, Belfield, Dublin 4, Ireland Centre for Quantum Engineering, Science, and Technology, University College Dublin, Dublin 4, Ireland    Abolfazl Bayat [email protected] Institute of Fundamental and Frontier Sciences, University of Electronic Science and Technology of China, Chengdu 611731, China Key Laboratory of Quantum Physics and Photonic Quantum Information, Ministry of Education, University of Electronic Science and Technology of China, Chengdu 611731, China    Steve Campbell [email protected] School of Physics, University College Dublin, Belfield, Dublin 4, Ireland Centre for Quantum Engineering, Science, and Technology, University College Dublin, Dublin 4, Ireland Dahlem Center for Complex Quantum Systems, Freie Universität Berlin, Arnimallee 14, 14195 Berlin, Germany    Andrew K. Mitchell [email protected] School of Physics, University College Dublin, Belfield, Dublin 4, Ireland Centre for Quantum Engineering, Science, and Technology, University College Dublin, Dublin 4, Ireland
Abstract

Quantum systems can be used as probes in the context of metrology for enhanced parameter estimation. In particular, the delicacy of critical systems to perturbations can make them ideal sensors. Arguably the simplest realistic probe system is a spin-1212\tfrac{1}{2}divide start_ARG 1 end_ARG start_ARG 2 end_ARG impurity, which can be manipulated and measured in-situ when embedded in a fermionic environment. Although entanglement between a single impurity probe and its environment produces nontrivial many-body effects, criticality cannot be leveraged for sensing. Here we introduce instead the two-impurity Kondo (2IK) model as a novel paradigm for critical quantum metrology, and examine the multiparameter estimation scenario at finite temperature. We explore the full metrological phase diagram numerically and obtain exact analytic results near criticality. Enhanced sensitivity to the inter-impurity coupling driving a second-order phase transition is evidenced by diverging quantum Fisher information (QFI) and quantum signal-to-noise ratio (QSNR). However, with uncertainty in both coupling strength and temperature, the multiparameter QFI matrix becomes singular – even though the parameters to be estimated are independent – resulting in vanishing QSNRs. We demonstrate that by applying a known control field, the singularity can be removed and measurement sensitivity restored. For general systems, we show that the degradation in the QSNR due to uncertainties in another parameter is controlled by the degree of correlation between the unknown parameters.

I Introduction

Quantum probes are known for their advantage over classical sensors in achieving higher precision using the same resources Degen et al. (2017). Originally, such advantage has been accomplished through exploiting special forms of quantum superposition, known as GHZ states Greenberger et al. (1989), to reach quadratic improvement in the precision of detecting external signals Giovannetti et al. (2004); Leibfried et al. (2004). However, GHZ-based quantum sensors are hard to scale up and are prone to decoherence Demkowicz-Dobrzański et al. (2012); Zhou (2024) and perturbation De Pasquale et al. (2013). Alternatively, strongly correlated quantum many-body systems near their phase transitions have been identified as a resource for quantum enhanced sensitivity Campos Venuti and Zanardi (2007); Schwandt et al. (2009); Albuquerque et al. (2010); Gritsev and Polkovnikov (2010); Gu et al. (2008); Greschner et al. (2013); Frérot and Roscilde (2018); Zhou et al. (2020); Rams et al. (2018a); Chu et al. (2021); Di Fresco et al. (2022); Chu et al. (2023); Salvia et al. (2023); Rodríguez et al. (2023); Bressanini et al. (2024); Cavazzoni et al. (2024). Several kinds of phase transitions have been proposed for quantum sensing, including second-order Zanardi and Paunković (2006); Abasto et al. (2008); Sun et al. (2010); Zanardi et al. (2008); Damski and Rams (2013); Salvatori et al. (2014); Yang et al. (2022); Fernández-Lorenzo et al. (2018); Montenegro et al. (2022); Ozaydin and Altintas (2015); Garbe et al. (2022a); Mirkhalaf et al. (2021); Wald et al. (2020); Hotter et al. (2024); Ostermann and Gietka (2024); Lü et al. (2024); Alushi et al. (2024), topological Budich and Bergholtz (2020); Koch and Budich (2022); Sarkar et al. (2022); Zhang et al. (2023); Yang et al. (2024a), superradiant and Rabi type  Bin et al. (2019); Di Candia et al. (2023); Garbe et al. (2020); Heugel et al. (2019); Fernández-Lorenzo and Porras (2017); Wu and Shi (2021); Ying et al. (2022); Tang et al. (2023); Zhu et al. (2023); Garbe et al. (2022b), dynamical Tsang (2013); Macieszczak et al. (2016); Carollo et al. (2018), Floquet Lang et al. (2015); Mishra and Bayat (2021, 2022), continuous environmental monitoring Ilias et al. (2022); Boeyens et al. (2023), Stark localization  He et al. (2023); Yousefjani et al. (2023), disorder-induced  Bhattacharyya et al. (2023); Sahoo et al. (2023), and boundary time crystals Montenegro et al. (2023); Cabot et al. (2023). Furthermore, certain criticality-based sensing mechanisms have been experimentally realized in NV-centers in diamond Yu et al. (2022), NMR Liu et al. (2021), trapped ions Ilias et al. (2023), and Rydberg atoms Ding et al. (2022). Although quantum phase transitions strictly occur at zero temperature, in practice any physical realization will be performed at finite temperatures where thermal fluctuations become important. To further complicate matters, the temperature of the system itself might not be known precisely. As a consequence, there has been a growing interest in develo** thermometric schemes that exploit quantum systems Correa et al. (2015); Paris (2015); Stace (2010); Mitchison et al. (2020); Campbell et al. (2018); Brattegard and Mitchison (2023); Brenes and Segal (2023); Yu et al. (2023); Yang et al. (2024b); Srivastava et al. (2023); Verma et al. (2024). These issues give rise to several questions, including: (i) How does nonzero temperature affect the sensitivity of a criticality-based quantum sensor? (ii) Can quantum criticality also boost the sensitivity of temperature estimation?

Perhaps the most well-known theorem in metrology is the Cramér-Rao inequality Paris (2009) which puts a fundamental bound on the uncertainty of inferring one or more unknown parameters. Let us consider a quantum probe which encodes n𝑛nitalic_n parameters λ=(λ1,λ2,,λn)T𝜆superscriptsubscript𝜆1subscript𝜆2subscript𝜆𝑛𝑇\vec{\lambda}=(\lambda_{1},~{}\lambda_{2},~{}...,~{}\lambda_{n})^{T}over→ start_ARG italic_λ end_ARG = ( italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , … , italic_λ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT in its density matrix ϱ(λ)italic-ϱ𝜆\varrho(\vec{\lambda})italic_ϱ ( over→ start_ARG italic_λ end_ARG ). The uncertainty in estimating the parameters λ𝜆\vec{\lambda}over→ start_ARG italic_λ end_ARG, through performing a suitable measurement on the probe, can be quantified by the covariance matrix Cov[λ,λ]Cov𝜆𝜆\textbf{Cov}[\vec{\lambda},\vec{\lambda}]Cov [ over→ start_ARG italic_λ end_ARG , over→ start_ARG italic_λ end_ARG ] with components given by Cov(λi,λj)=(λiλi)(λjλj)Covsubscript𝜆𝑖subscript𝜆𝑗delimited-⟨⟩subscript𝜆𝑖delimited-⟨⟩subscript𝜆𝑖subscript𝜆𝑗delimited-⟨⟩subscript𝜆𝑗\text{Cov}(\lambda_{i},\lambda_{j})=\langle(\lambda_{i}-\langle\lambda_{i}% \rangle)(\lambda_{j}-\langle\lambda_{j}\rangle)\rangleCov ( italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) = ⟨ ( italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - ⟨ italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ⟩ ) ( italic_λ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT - ⟨ italic_λ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ⟩ ) ⟩. The individual variances are then simply the diagonal elements, Var(λi)Cov(λi,λi)Varsubscript𝜆𝑖Covsubscript𝜆𝑖subscript𝜆𝑖\text{Var}(\lambda_{i})\equiv\text{Cov}(\lambda_{i},\lambda_{i})Var ( italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) ≡ Cov ( italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ). The Cramér-Rao inequality is given by,

Cov[λ]1NF1Covdelimited-[]𝜆1𝑁superscriptF1\textbf{Cov}\left[\vec{\lambda}\right]\geq\frac{1}{N}\textbf{F}^{-1}Cov [ over→ start_ARG italic_λ end_ARG ] ≥ divide start_ARG 1 end_ARG start_ARG italic_N end_ARG F start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT (1)

where N𝑁Nitalic_N is the number of samples and F is the Fisher information matrix Helstrom (1969); Holevo (1982). A crucial point is that the Cramér-Rao bound is only meaningful when the Fisher information matrix is invertible. For example, in a two parameter setting, if the parameters are not independent from each other – that is, they can can be rescaled to a single effective parameter, one can show that the Fisher information matrix becomes singular and thus non-invertible. However, this is only a necessary condition and there might be other situations which result in a non-invertible Fisher information matrix. A potential scheme for such a problem is in criticality-based quantum sensing where the probe operates at a nonzero temperature that may not be known with certainty. In this case, one has to address the problem through a multiparameter sensing analysis, where temperature is also treated as an unknown parameter. In principle, the unknown microscopic parameter which drives the phase transition is independent from the temperature of the system. However, the way that these parameters are encoded in the quantum state of the probe may nonetheless result in a singular Fisher information matrix – especially in systems constrained by high symmetries or near quantum critical points where the low-energy physics exhibits emergent single-parameter scaling Sachdev (1999). In such cases, can one make the Fisher information matrix invertible again, thereby allowing for an effective sensing protocol?

In this work we consider the two-impurity Kondo (2IK) model Jayaprakash et al. (1981); Jones et al. (1988); Affleck and Ludwig (1992); Affleck et al. (1995); Mitchell et al. (2012); Sela et al. (2011); Mitchell and Sela (2012), a famous paradigm in condensed matter physics for quantum criticality in a strongly-correlated many-body system. Originally conceived to describe the through-lattice (RKKY) coupling between two magnetic impurities (such as iron atoms) embedded in a host metal (such as gold), the 2IK quantum phase transition also captures the essence of the competition between magnetic ordering and heavy-fermion physics in real correlated materials Hewson (1993). The model can be realized in semiconductor quantum dot devices van der Wiel et al. (2002); Izumida and Sakai (2000); Zaránd et al. (2006); Jeong et al. (2001), and indeed a closely related variant of the 2IK critical point was observed experimentally very recently in Ref. Pouse et al. (2023); *karki2023z. The 2IK model features two exchange-coupled spin-1212\tfrac{1}{2}divide start_ARG 1 end_ARG start_ARG 2 end_ARG “impurity” qubits, each coupled to its own fermionic environment (taken to be metallic continua of conduction electrons). Although such impurities are non-invasive in the sense that impurity effects in bulk systems are always non-extensive, they can still induce local criticality (also known as boundary critical phenomena) Affleck and Ludwig (1992). In the 2IK model, a nontrivial critical point, obtained by tuning the inter-impurity coupling strength, separates a phase in which the impurities bind together into a local spin-singlet state, from a phase in which each impurity is separately screened by conduction electrons through the Kondo effect Affleck and Ludwig (1992); Mitchell et al. (2012). The underlying physics is controlled by the development of strong many-body entanglement between the probe impurities and the electronic environment Bayat et al. (2012, 2014). The presence of a quantum phase transition whose critical properties survive thermal fluctuations (up to the so-called Kondo temperature) makes the 2IK model an excellent testbed for studying the performance of a critical sensor at nonzero temperature. Furthermore, we note that such interacting quantum condensed matter systems in the thermodynamic limit can in practice be straightforwardly tuned into their critical regimes at finite temperatures, without the need for meticulous and costly critical ground state preparation techniques Gietka et al. (2021); Rams et al. (2018b), as demonstrated experimentally in nanoelectronics device realizations Potok et al. (2007); Iftikhar et al. (2018); Pouse et al. (2023).

We regard the two coupled impurities in the 2IK model as a metrological probe, whose reduced state is characterized by the impurity singlet fraction (spin-spin correlator) – an experimentally relevant physical observable. The additional internal structure of the coupled impurity probe relative to a single impurity probe Mihailescu et al. (2023) endows a far richer metrological phase diagram. We consider estimation of either the environment temperature T𝑇Titalic_T or the inter-probe coupling K𝐾Kitalic_K, as well as the arguably more realistic multiparameter estimation scenario where neither T𝑇Titalic_T nor K𝐾Kitalic_K are known with absolute certainty. In the latter scenario, we find that the quantum Fisher information matrix (QFIM) becomes singular – despite temperature and impurity coupling being independent parameters. Such a situation prohibits inference of either parameter, and the corresponding quantum signal to noise ratio (QSNR) for both parameters vanishes. We propose a strategy to remedy this: by applying a known control field, the QFIM singularity is removed and the ability to perform multiparameter estimation is restored.

The paper is organised as follows: In Sec. II we briefly recapitulate the fundamentals of multiparameter estimation theory, introducing the QFIM and Cramér-Rao bound. We identify the QSNR as the key figure of merit, and introduce a novel generalization of this quantity in the multiparametric setting. In particular, we show that the QSNR for a given parameter is always reduced by uncertainties in another parameter, with the degree of degradation controlled by the degree of correlation between the unknown parameters. In Sec. III we introduce the 2IK model and contextualize the physical regimes in which we expect to observe quantum critical features. We take the two spin-1212\tfrac{1}{2}divide start_ARG 1 end_ARG start_ARG 2 end_ARG exchange coupled impurity qubits as our metrological probe, and show how parameter estimation sensitivity can be extracted from the spin-spin correlation function, an experimentally-motivated physical observable. In Sec. IV we provide analytical results for single parameter estimation in the simple but instructive limit of large inter-probe coupling strength. In Sec. V we investigate single-parameter metrology in the full many-body system, obtaining numerically exact results using the Numerical Renormalization Group (NRG) technique Wilson (1975); *bulla2008numerical, showing how strong correlations and quantum criticality affect sensing capabilities in the 2IK system. In Sec. VI we derive closed-form exact analytic results for the QSNRs in the vicinity of the quantum critical point. Our solution is obtained by relating the QFI to changes in the probe entropy, and constitutes a rare example in which exact results can be obtained for an interacting quantum many-body system at finite temperatures, near a nontrivial second-order quantum phase transition. In Sec. VII we consider explicitly the multiparameter scenario. Here we explore the QFIM singularity that arises when we have uncertainty in both system temperature T𝑇Titalic_T and probe coupling strength K𝐾Kitalic_K. The singularity in the QFIM that prevents multiparameter estimation is shown to be connected to the SU(2) spin symmetry of the probe reduced state. We further demonstrate that by applying a known control field that breaks this symmetry, the singularity is removed and multiparameter estimation sensitivity is restored. Our results indicate a dramatic difference in the effectiveness of critical quantum sensing when uncertainty in more than one parameter is taken into account. This is an essential practical consideration since any experiment must be performed at finite temperature, and there is typically some experimental uncertainty in determining this temperature.

II Parameter Estimation

We begin by introducing the tools necessary for multiparameter estimation, in particular, the quantum Fisher information and the Cramér-Rao bound (CRB) Paris (2009); Tóth and Apellaniz (2014); Liu et al. (2019). In what follows, we present the formalism for the multiparameter setting for generality. However we emphasize that the single parameter estimation scenario corresponds to the special case where all parameters except the one to be estimated are assumed to be known with certainty. We will extensively discuss single parameter estimation in Secs. III-VI and consider explicitly the multiparameter case in Sec. VII.

The Fisher information matrix appearing in Eq. (1) is an n×n𝑛𝑛n\times nitalic_n × italic_n positive semi-definite matrix for a system with n𝑛nitalic_n unknown parameters λ=(λ1,λ2,,λn)T𝜆superscriptsubscript𝜆1subscript𝜆2subscript𝜆𝑛𝑇\vec{\lambda}=(\lambda_{1},~{}\lambda_{2},~{}...,~{}\lambda_{n})^{T}over→ start_ARG italic_λ end_ARG = ( italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , … , italic_λ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT. Its elements are given by,

Fi,j=𝔼[(λiln(p(xkλi)))(λjln(p(xkλj)))]subscriptF𝑖𝑗𝔼delimited-[]subscriptsubscript𝜆𝑖𝑝conditionalsubscript𝑥𝑘subscript𝜆𝑖subscriptsubscript𝜆𝑗𝑝conditionalsubscript𝑥𝑘subscript𝜆𝑗\textbf{F}_{i,j}=\mathbb{E}\left[\left(\partial_{\lambda_{i}}\ln{p\left(x_{k}% \mid\lambda_{i}\right)}\right)\left(\partial_{\lambda_{j}}\ln{p\left(x_{k}\mid% \lambda_{j}\right)}\right)\right]F start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT = blackboard_E [ ( ∂ start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT roman_ln ( start_ARG italic_p ( italic_x start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ∣ italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) end_ARG ) ) ( ∂ start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_POSTSUBSCRIPT roman_ln ( start_ARG italic_p ( italic_x start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ∣ italic_λ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) end_ARG ) ) ] (2)

where p(xk|λi)𝑝conditionalsubscript𝑥𝑘subscript𝜆𝑖p(x_{k}|\lambda_{i})italic_p ( italic_x start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT | italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) denotes the conditional probability of obtaining outcome xksubscript𝑥𝑘x_{k}italic_x start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT given the parameter has value λisubscript𝜆𝑖\lambda_{i}italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT 111We remark that, in the limiting case, when we have a single parameter, λ𝜆\lambdaitalic_λ, that we wish to infer, the Fisher information in Eq. (2) reduces to F(λ)=Eλ[λln(p(xkλ))]2𝐹𝜆subscript𝐸𝜆superscriptdelimited-[]subscript𝜆𝑝conditionalsubscript𝑥𝑘𝜆2F\left(\lambda\right)=E_{\lambda}\left[\partial_{\lambda}\ln{p\left(x_{k}\mid% \lambda\right)}\right]^{2}italic_F ( italic_λ ) = italic_E start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT [ ∂ start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT roman_ln ( start_ARG italic_p ( italic_x start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ∣ italic_λ ) end_ARG ) ] start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT and the CRB becomes Var(λ)(NF)1Var𝜆superscript𝑁𝐹1\text{Var}\left(\lambda\right)\geq{\left(NF\right)}^{-1}Var ( italic_λ ) ≥ ( italic_N italic_F ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, establishing a lower-bound on the mean square error, Var(λ)=Eλ[(λ^({x})λ)2]Var𝜆subscript𝐸𝜆delimited-[]superscript^𝜆𝑥𝜆2\text{Var}\left(\lambda\right)=E_{\lambda}\left[\left(\hat{\lambda}\left(\{x\}% \right)-\lambda\right)^{2}\right]Var ( italic_λ ) = italic_E start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT [ ( over^ start_ARG italic_λ end_ARG ( { italic_x } ) - italic_λ ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ], of any estimator of the parameter λ𝜆\lambdaitalic_λ..

In the quantum setting Paris (2009), we consider parameter-encoded quantum states, ϱ^(λ)^italic-ϱ𝜆\hat{\varrho}(\vec{\lambda})over^ start_ARG italic_ϱ end_ARG ( over→ start_ARG italic_λ end_ARG ), whose measured outcome value, xksubscript𝑥𝑘x_{k}italic_x start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT, can be obtained through a set of positive operator-valued measurements (POVMs) denoted {Πi}subscriptΠ𝑖\{\Pi_{i}\}{ roman_Π start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT }. The parameter dependent conditional probability of these outcomes, defined through the Born rule p(xkλi)=Tr(Πiϱ^(λ))𝑝conditionalsubscript𝑥𝑘subscript𝜆𝑖tracesubscriptΠ𝑖^italic-ϱ𝜆p\left(x_{k}\mid\lambda_{i}\right)=\Tr{\Pi_{i}~{}\hat{\varrho}(\vec{\lambda})}italic_p ( italic_x start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ∣ italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) = roman_Tr ( start_ARG roman_Π start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT over^ start_ARG italic_ϱ end_ARG ( over→ start_ARG italic_λ end_ARG ) end_ARG ), allows for the construction of unbiased estimators. The quantum Fisher information matrix (QFIM) 𝓗𝓗\boldsymbol{\mathcal{H}}bold_caligraphic_H, is obtained through an optimization over all possible measurements. Its elements are defined in terms of the symmetric logarithmic derivative (SLD) operators as,

𝓗λi,λj=12Tr(ϱ^{L^i,L^j})subscript𝓗subscript𝜆𝑖subscript𝜆𝑗12Tr^italic-ϱanticommutatorsubscript^𝐿𝑖subscript^𝐿𝑗\boldsymbol{\mathcal{H}}_{\lambda_{i},\lambda_{j}}=\frac{1}{2}\text{Tr}\left(% \hat{\varrho}\anticommutator{\hat{L}_{i}}{\hat{L}_{j}}\right)bold_caligraphic_H start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 2 end_ARG Tr ( over^ start_ARG italic_ϱ end_ARG { start_ARG over^ start_ARG italic_L end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG , start_ARG over^ start_ARG italic_L end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG } ) (3)

where L^isubscript^𝐿𝑖\hat{L}_{i}over^ start_ARG italic_L end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT correspond to the ideal measurement for parameter λisubscript𝜆𝑖\lambda_{i}italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. Formally, the SLD is defined by the solution of the self-adjoint operator equation λiϱ^(λ)=12(Liϱ^(λ)+ϱ^(λ)Li)subscriptsubscript𝜆𝑖^italic-ϱ𝜆12subscript𝐿𝑖^italic-ϱ𝜆^italic-ϱ𝜆subscript𝐿𝑖\partial_{\lambda_{i}}\hat{\varrho}(\vec{\lambda})=\frac{1}{2}\left(L_{i}\hat{% \varrho}(\vec{\lambda})+\hat{\varrho}(\vec{\lambda})L_{i}\right)∂ start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT over^ start_ARG italic_ϱ end_ARG ( over→ start_ARG italic_λ end_ARG ) = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( italic_L start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT over^ start_ARG italic_ϱ end_ARG ( over→ start_ARG italic_λ end_ARG ) + over^ start_ARG italic_ϱ end_ARG ( over→ start_ARG italic_λ end_ARG ) italic_L start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ). As such, we note that the diagonal elements of the QFIM are identical to the single-parameter QFI, that is 𝓗λi,λi=(λi)subscript𝓗subscript𝜆𝑖subscript𝜆𝑖subscript𝜆𝑖\boldsymbol{\mathcal{H}}_{\lambda_{i},\lambda_{i}}=\mathcal{H}(\lambda_{i})bold_caligraphic_H start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT = caligraphic_H ( italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ).

The quantum multiparmater CRB reads Holevo (1982),

Cov[λ]1N𝓗1,Covdelimited-[]𝜆1𝑁superscript𝓗1\textbf{Cov}\left[\vec{\lambda}\right]\geq\frac{1}{N}\boldsymbol{\mathcal{H}}^% {-1}\;,Cov [ over→ start_ARG italic_λ end_ARG ] ≥ divide start_ARG 1 end_ARG start_ARG italic_N end_ARG bold_caligraphic_H start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT , (4)

where we set N=1𝑁1N\!=\!1italic_N = 1 for the single-shot measurement case considered hereafter, and we remark that the bound holds element-wise in this matrix inequality. While the diagonal elements of the QFIM, 𝓗λi,λisubscript𝓗subscript𝜆𝑖subscript𝜆𝑖\boldsymbol{\mathcal{H}}_{\lambda_{i},\lambda_{i}}bold_caligraphic_H start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT, on their own provide information only about the measurement precision for parameter λisubscript𝜆𝑖\lambda_{i}italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT in the single-parameter estimation scenario, the matrix inverse operation in Eq. (4) means that the other elements of the QFIM affect measurement precision when we have uncertainty in any of the other parameters. Indeed, as shown in Appendix A and discussed further below, the precision of estimating a given parameter is always reduced by uncertainties in other parameters. In addition, the bounds for all parameters may not be simultaneously saturable using a single measurement since the SLD operators L^isubscript^𝐿𝑖\hat{L}_{i}over^ start_ARG italic_L end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT for different parameters λisubscript𝜆𝑖\lambda_{i}italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT may be incompatible Zhu (2015); Heinosaari et al. (2016); Candeloro et al. (2024). Thus, a single optimal measurement basis shared by all the parameters might not exist, in which case precision trade-offs in the multiparameter estimation problem are unavoidable at the fundamental level.

It is important to establish how well one can distinguish the inferred parameter signal from the measurement noise. For example, in situations where the QFI indicates a region of high precision but the signal itself is extremely small in this region, accurate parameter estimation remains challenging in practice. For this reason, we focus on the QSNR

𝒬SP(λ)λ2(λ)subscript𝒬𝑆𝑃𝜆superscript𝜆2𝜆\mathcal{Q}_{SP}\left(\lambda\right)\equiv\lambda^{2}\mathcal{H}\left(\lambda\right)caligraphic_Q start_POSTSUBSCRIPT italic_S italic_P end_POSTSUBSCRIPT ( italic_λ ) ≡ italic_λ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT caligraphic_H ( italic_λ ) (5)

where we have here emphasized that the QSNR in question is the one corresponding to single-parameter estimation through the ‘SP’ subscript. For the multiparameter setting we consider a generalization of Eq. (5),

𝒬MP(λi,λj)|λiλj|[𝓗1]λi,λj|λiλj|Cov(λi,λj)subscript𝒬𝑀𝑃subscript𝜆𝑖subscript𝜆𝑗subscript𝜆𝑖subscript𝜆𝑗subscriptdelimited-[]superscript𝓗1subscript𝜆𝑖subscript𝜆𝑗subscript𝜆𝑖subscript𝜆𝑗Covsubscript𝜆𝑖subscript𝜆𝑗\mathcal{Q}_{MP}(\lambda_{i},\lambda_{j})\equiv\frac{|\lambda_{i}\lambda_{j}|}% {[\boldsymbol{\mathcal{H}}^{-1}]_{\lambda_{i},\lambda_{j}}}\geq\frac{|\lambda_% {i}\lambda_{j}|}{\text{Cov}(\lambda_{i},\lambda_{j})}caligraphic_Q start_POSTSUBSCRIPT italic_M italic_P end_POSTSUBSCRIPT ( italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) ≡ divide start_ARG | italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT | end_ARG start_ARG [ bold_caligraphic_H start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ] start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_POSTSUBSCRIPT end_ARG ≥ divide start_ARG | italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT | end_ARG start_ARG Cov ( italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) end_ARG (6)

where the bounds follow from the element-wise manipulation of the quantum CRB in Eq. (4). The maximum possible quantum signal to noise ratio λi2/Var(λi)superscriptsubscript𝜆𝑖2Varsubscript𝜆𝑖\lambda_{i}^{2}/\text{Var}(\lambda_{i})italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / Var ( italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) for the estimation of parameter λisubscript𝜆𝑖\lambda_{i}italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT in a system with multiple unknown quantities is given by 𝒬MP(λi,λi)subscript𝒬𝑀𝑃subscript𝜆𝑖subscript𝜆𝑖\mathcal{Q}_{MP}(\lambda_{i},\lambda_{i})caligraphic_Q start_POSTSUBSCRIPT italic_M italic_P end_POSTSUBSCRIPT ( italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ), which is a strictly non-negative quantity. The off-diagonal components 𝒬MP(λi,λj)subscript𝒬𝑀𝑃subscript𝜆𝑖subscript𝜆𝑗\mathcal{Q}_{MP}(\lambda_{i},\lambda_{j})caligraphic_Q start_POSTSUBSCRIPT italic_M italic_P end_POSTSUBSCRIPT ( italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) with λiλjsubscript𝜆𝑖subscript𝜆𝑗\lambda_{i}\neq\lambda_{j}italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ≠ italic_λ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT relate to covariances and therefore can be negative when the measurement outcomes of λisubscript𝜆𝑖\lambda_{i}italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and λjsubscript𝜆𝑗\lambda_{j}italic_λ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT are negatively correlated. Note that each element of the multiparameter QSNR defined in Eq. (6) is proportional to the determinant of the QFIM, det(𝓗)det𝓗{\rm det}(\boldsymbol{\mathcal{H}})roman_det ( bold_caligraphic_H ), due to the matrix inverse operation appearing in Eq. (4). As such the QSNR vanishes when the QFIM is singular since then det(𝓗)=0det𝓗0{\rm det}(\boldsymbol{\mathcal{H}})=0roman_det ( bold_caligraphic_H ) = 0. We discuss how to interpret and deal with a singular QFIM in Sec. VII.1.

Let us explicitly consider the estimation of two arbitrary parameters, λ=(λA,λB)T𝜆superscriptsubscript𝜆𝐴subscript𝜆𝐵𝑇\vec{\lambda}=\left(\lambda_{A},\lambda_{B}\right)^{T}over→ start_ARG italic_λ end_ARG = ( italic_λ start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT. Equation (4) requires the inverse of the 2×2222\times 22 × 2 QFIM,

𝓗1=1det(𝓗)(𝓗λB,λB𝓗λA,λB𝓗λA,λB𝓗λA,λA)superscript𝓗11det𝓗matrixsubscript𝓗subscript𝜆𝐵subscript𝜆𝐵subscript𝓗subscript𝜆𝐴subscript𝜆𝐵subscript𝓗subscript𝜆𝐴subscript𝜆𝐵subscript𝓗subscript𝜆𝐴subscript𝜆𝐴\boldsymbol{\mathcal{H}}^{-1}=\frac{1}{{\rm det}(\boldsymbol{\mathcal{H}})}% \begin{pmatrix}\boldsymbol{\mathcal{H}}_{\lambda_{B},\lambda_{B}}&-\boldsymbol% {\mathcal{H}}_{\lambda_{A},\lambda_{B}}\\ -\boldsymbol{\mathcal{H}}_{\lambda_{A},\lambda_{B}}&\boldsymbol{\mathcal{H}}_{% \lambda_{A},\lambda_{A}}\end{pmatrix}bold_caligraphic_H start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT = divide start_ARG 1 end_ARG start_ARG roman_det ( bold_caligraphic_H ) end_ARG ( start_ARG start_ROW start_CELL bold_caligraphic_H start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT end_POSTSUBSCRIPT end_CELL start_CELL - bold_caligraphic_H start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL - bold_caligraphic_H start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT end_POSTSUBSCRIPT end_CELL start_CELL bold_caligraphic_H start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT end_POSTSUBSCRIPT end_CELL end_ROW end_ARG ) (7)

where det(𝓗)=𝓗λA,λA𝓗λB,λB𝓗λA,λB2det𝓗subscript𝓗subscript𝜆𝐴subscript𝜆𝐴subscript𝓗subscript𝜆𝐵subscript𝜆𝐵superscriptsubscript𝓗subscript𝜆𝐴subscript𝜆𝐵2{\rm det}(\boldsymbol{\mathcal{H}})=\boldsymbol{\mathcal{H}}_{\lambda_{A},% \lambda_{A}}\boldsymbol{\mathcal{H}}_{\lambda_{B},\lambda_{B}}-\boldsymbol{% \mathcal{H}}_{\lambda_{A},\lambda_{B}}^{2}roman_det ( bold_caligraphic_H ) = bold_caligraphic_H start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT end_POSTSUBSCRIPT bold_caligraphic_H start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT end_POSTSUBSCRIPT - bold_caligraphic_H start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. Thus we find,

𝒬MP(λA,λA)=λA2det(𝓗)𝓗λB,λB=𝒬SP(λA)(λAλB𝓗λA,λB)2𝒬SP(λB)subscript𝒬𝑀𝑃subscript𝜆𝐴subscript𝜆𝐴superscriptsubscript𝜆𝐴2det𝓗subscript𝓗subscript𝜆𝐵subscript𝜆𝐵subscript𝒬𝑆𝑃subscript𝜆𝐴superscriptsubscript𝜆𝐴subscript𝜆𝐵subscript𝓗subscript𝜆𝐴subscript𝜆𝐵2subscript𝒬𝑆𝑃subscript𝜆𝐵\mathcal{Q}_{MP}\left(\lambda_{A},\lambda_{A}\right)=\lambda_{A}^{2}~{}\frac{{% \rm det}(\boldsymbol{\mathcal{H}})}{\boldsymbol{\mathcal{H}}_{\lambda_{B},% \lambda_{B}}}=\mathcal{Q}_{SP}(\lambda_{A})-\frac{(\lambda_{A}\lambda_{B}% \boldsymbol{\mathcal{H}}_{\lambda_{A},\lambda_{B}})^{2}}{\mathcal{Q}_{SP}(% \lambda_{B})}~{}~{}caligraphic_Q start_POSTSUBSCRIPT italic_M italic_P end_POSTSUBSCRIPT ( italic_λ start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ) = italic_λ start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT divide start_ARG roman_det ( bold_caligraphic_H ) end_ARG start_ARG bold_caligraphic_H start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT end_POSTSUBSCRIPT end_ARG = caligraphic_Q start_POSTSUBSCRIPT italic_S italic_P end_POSTSUBSCRIPT ( italic_λ start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ) - divide start_ARG ( italic_λ start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT bold_caligraphic_H start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG caligraphic_Q start_POSTSUBSCRIPT italic_S italic_P end_POSTSUBSCRIPT ( italic_λ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT ) end_ARG (8)

and similarly for 𝒬MP(λB,λB)subscript𝒬𝑀𝑃subscript𝜆𝐵subscript𝜆𝐵\mathcal{Q}_{MP}\left(\lambda_{B},\lambda_{B}\right)caligraphic_Q start_POSTSUBSCRIPT italic_M italic_P end_POSTSUBSCRIPT ( italic_λ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT ). For the off-diagonal terms,

𝒬MP(λA,λB)=|λA×λB|det(𝓗)𝓗λA,λBsubscript𝒬𝑀𝑃subscript𝜆𝐴subscript𝜆𝐵subscript𝜆𝐴subscript𝜆𝐵det𝓗subscript𝓗subscript𝜆𝐴subscript𝜆𝐵\mathcal{Q}_{MP}\left(\lambda_{A},\lambda_{B}\right)=-\absolutevalue{\lambda_{% A}\times\lambda_{B}}~{}\frac{{\rm det}(\boldsymbol{\mathcal{H}})}{\boldsymbol{% \mathcal{H}}_{\lambda_{A},\lambda_{B}}}caligraphic_Q start_POSTSUBSCRIPT italic_M italic_P end_POSTSUBSCRIPT ( italic_λ start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT ) = - | start_ARG italic_λ start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT × italic_λ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT end_ARG | divide start_ARG roman_det ( bold_caligraphic_H ) end_ARG start_ARG bold_caligraphic_H start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT end_POSTSUBSCRIPT end_ARG (9)

These expressions immediately provide insight into the multiparameter estimation problem. First, we see that the multiparameter QSNRs 𝒬MP(λ,λ)subscript𝒬𝑀𝑃𝜆𝜆\mathcal{Q}_{MP}(\lambda,\lambda)caligraphic_Q start_POSTSUBSCRIPT italic_M italic_P end_POSTSUBSCRIPT ( italic_λ , italic_λ ) can be decomposed into a piece corresponding to the single-parameter estimation of the same parameter 𝒬SP(λ)subscript𝒬𝑆𝑃𝜆\mathcal{Q}_{SP}(\lambda)caligraphic_Q start_POSTSUBSCRIPT italic_S italic_P end_POSTSUBSCRIPT ( italic_λ ) and a correction. This correction always lowers the multiparameter QSNR relative to its single-parameter counterpart. The multiparameter QSNR is therefore upper-bounded by the corresponding single-parameter QSNR – no additional precision in the estimation of parameter λ𝜆\lambdaitalic_λ may be obtained by uncertainties in other parameters, see also Appendix A. The magnitude of this correction is increased by the cross-correlation part of the QFIM 𝓗λA,λBsubscript𝓗subscript𝜆𝐴subscript𝜆𝐵\boldsymbol{\mathcal{H}}_{\lambda_{A},\lambda_{B}}bold_caligraphic_H start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT end_POSTSUBSCRIPT but it is decreased when the single-parameter QFI of the other parameter is larger. This makes physical sense, because the multiparameter QSNR for parameter λ𝜆\lambdaitalic_λ should approach its single-parameter QSNR value when λλsuperscript𝜆𝜆\lambda^{\prime}\neq\lambdaitalic_λ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ≠ italic_λ is known with certainty (whereby the QFI for λsuperscript𝜆\lambda^{\prime}italic_λ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT diverges and the deleterious correction vanishes).

This can be seen clearly by rearranging the equations to obtain the following identity:

𝒬MP(λA,λA)𝒬SP(λA)=𝒬MP(λB,λB)𝒬SP(λB)=1Cor(λA,λB)2subscript𝒬𝑀𝑃subscript𝜆𝐴subscript𝜆𝐴subscript𝒬𝑆𝑃subscript𝜆𝐴subscript𝒬𝑀𝑃subscript𝜆𝐵subscript𝜆𝐵subscript𝒬𝑆𝑃subscript𝜆𝐵1Corsuperscriptsubscript𝜆𝐴subscript𝜆𝐵2\frac{\mathcal{Q}_{MP}\left(\lambda_{A},\lambda_{A}\right)}{\mathcal{Q}_{SP}(% \lambda_{A})}=\frac{\mathcal{Q}_{MP}\left(\lambda_{B},\lambda_{B}\right)}{% \mathcal{Q}_{SP}(\lambda_{B})}=1-{\rm Cor}(\lambda_{A},\lambda_{B})^{2}divide start_ARG caligraphic_Q start_POSTSUBSCRIPT italic_M italic_P end_POSTSUBSCRIPT ( italic_λ start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ) end_ARG start_ARG caligraphic_Q start_POSTSUBSCRIPT italic_S italic_P end_POSTSUBSCRIPT ( italic_λ start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ) end_ARG = divide start_ARG caligraphic_Q start_POSTSUBSCRIPT italic_M italic_P end_POSTSUBSCRIPT ( italic_λ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT ) end_ARG start_ARG caligraphic_Q start_POSTSUBSCRIPT italic_S italic_P end_POSTSUBSCRIPT ( italic_λ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT ) end_ARG = 1 - roman_Cor ( italic_λ start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (10)

where we have defined Cor(λA,λB)=𝓗λA,λB/𝓗λA,λA𝓗λB,λBCorsubscript𝜆𝐴subscript𝜆𝐵subscript𝓗subscript𝜆𝐴subscript𝜆𝐵subscript𝓗subscript𝜆𝐴subscript𝜆𝐴subscript𝓗subscript𝜆𝐵subscript𝜆𝐵{\rm Cor}(\lambda_{A},\lambda_{B})\!=\!\boldsymbol{\mathcal{H}}_{\lambda_{A},% \lambda_{B}}/\sqrt{\boldsymbol{\mathcal{H}}_{\lambda_{A},\lambda_{A}}% \boldsymbol{\mathcal{H}}_{\lambda_{B},\lambda_{B}}}roman_Cor ( italic_λ start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT ) = bold_caligraphic_H start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT end_POSTSUBSCRIPT / square-root start_ARG bold_caligraphic_H start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT end_POSTSUBSCRIPT bold_caligraphic_H start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT end_POSTSUBSCRIPT end_ARG which is equal to the correlation Cov(λA,λB)/Var(λA)Var(λB)Covsubscript𝜆𝐴subscript𝜆𝐵Varsubscript𝜆𝐴Varsubscript𝜆𝐵{\rm Cov}(\lambda_{A},\lambda_{B})/\sqrt{{\rm Var}(\lambda_{A}){\rm Var}(% \lambda_{B})}roman_Cov ( italic_λ start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT ) / square-root start_ARG roman_Var ( italic_λ start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ) roman_Var ( italic_λ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT ) end_ARG between measurements of λAsubscript𝜆𝐴\lambda_{A}italic_λ start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT and λBsubscript𝜆𝐵\lambda_{B}italic_λ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT in the ‘best case scenario’ when the quantum multiparameter CRB in Eq. (4) is saturated. These relations embody the fact that the relative degradation in measurement precision when there is uncertainty in both parameters applies equally to both parameters. The degradation is controlled by the correlation between the parameters. As a consequence, in scenarios where the QFIM is singular we have Cor(λA,λB)1Corsubscript𝜆𝐴subscript𝜆𝐵1{\rm Cor}(\lambda_{A},\lambda_{B})\to 1roman_Cor ( italic_λ start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT ) → 1 and 𝒬MP0subscript𝒬𝑀𝑃0\mathcal{Q}_{MP}\to 0caligraphic_Q start_POSTSUBSCRIPT italic_M italic_P end_POSTSUBSCRIPT → 0.

These results have implications for critical quantum metrology, which typically assumes that only a single parameter is to be estimated, and we have perfect knowledge of all other system parameters. However, Eq. (8) for the two-parameter estimation scenario shows that uncertainty in one parameter can dramatically affect the sensing capability for another parameter. Therefore, quantum critical systems are not inherently good for parameter estimation unless the single-parameter QSNRs for all unknown parameters are large.

In general, elements of the QFIM may depend explicitly on both the state’s eigenvalues and eigenvectors Paris (2009). We now comment on a special but important case where the eigenvectors of the probe reduced density matrix do not explicitly depend on the parameters λ𝜆\vec{\lambda}over→ start_ARG italic_λ end_ARG to be estimated. Then,

𝓗λi,λj=kλiρk(λ)×λjρk(λ)ρk(λ),subscript𝓗subscript𝜆𝑖subscript𝜆𝑗subscript𝑘subscriptsubscript𝜆𝑖subscript𝜌𝑘𝜆subscriptsubscript𝜆𝑗subscript𝜌𝑘𝜆subscript𝜌𝑘𝜆\boldsymbol{\mathcal{H}}_{\lambda_{i},\lambda_{j}}=\sum_{k}\frac{\partial_{% \lambda_{i}}\rho_{k}(\vec{\lambda})\times\partial_{\lambda_{j}}\rho_{k}(\vec{% \lambda})}{\rho_{k}(\vec{\lambda})}\;,bold_caligraphic_H start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT divide start_ARG ∂ start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( over→ start_ARG italic_λ end_ARG ) × ∂ start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( over→ start_ARG italic_λ end_ARG ) end_ARG start_ARG italic_ρ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( over→ start_ARG italic_λ end_ARG ) end_ARG , (11)

where ρk(λ)subscript𝜌𝑘𝜆\rho_{k}(\vec{\lambda})italic_ρ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( over→ start_ARG italic_λ end_ARG ) are the parameter-imprinted eigenvalues of the probe reduced density matrix.

Furthermore, we note that if the probe’s populations are determined solely by a single observable ΩΩ\Omegaroman_Ω, i.e. ϱkϱk(Ω)subscriptitalic-ϱ𝑘subscriptitalic-ϱ𝑘Ω\varrho_{k}\equiv\varrho_{k}(\Omega)italic_ϱ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ≡ italic_ϱ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( roman_Ω ), then the QFIM elements follow as

𝓗λi,λj=(Ω)×(λiΩλjΩ)subscript𝓗subscript𝜆𝑖subscript𝜆𝑗Ωsubscriptsubscript𝜆𝑖Ωsubscriptsubscript𝜆𝑗Ω\boldsymbol{\mathcal{H}}_{\lambda_{i},\lambda_{j}}=\mathcal{H}(\Omega)\times(% \partial_{\lambda_{i}}\Omega~{}\partial_{\lambda_{j}}\Omega)bold_caligraphic_H start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_POSTSUBSCRIPT = caligraphic_H ( roman_Ω ) × ( ∂ start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT roman_Ω ∂ start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_POSTSUBSCRIPT roman_Ω ) (12)

where (Ω)=k(Ωϱk)2/ϱkΩsubscript𝑘superscriptsubscriptΩsubscriptitalic-ϱ𝑘2subscriptitalic-ϱ𝑘\mathcal{H}(\Omega)=\sum_{k}(\partial_{\Omega}\varrho_{k})^{2}/\varrho_{k}caligraphic_H ( roman_Ω ) = ∑ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( ∂ start_POSTSUBSCRIPT roman_Ω end_POSTSUBSCRIPT italic_ϱ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / italic_ϱ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT is an effective single-parameter estimation QFI. The factorized form of Eq. (12) immediately implies that the QFIM is singular, with det(𝓗)=0det𝓗0\text{det}(\boldsymbol{\mathcal{H}})=0det ( bold_caligraphic_H ) = 0. Thus, the multiparameter QSNRs in Eq. (6) identically vanish, 𝒬MP(λi,λj)=0subscript𝒬𝑀𝑃subscript𝜆𝑖subscript𝜆𝑗0\mathcal{Q}_{MP}(\lambda_{i},\lambda_{j})=0caligraphic_Q start_POSTSUBSCRIPT italic_M italic_P end_POSTSUBSCRIPT ( italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) = 0. This tells us that under such a scenario, absolutely no information can be extracted about multiple unknown parameters from the measurements of the single observable ΩΩ\Omegaroman_Ω. By contrast, if only λisubscript𝜆𝑖\lambda_{i}italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is to be estimated and all other parameters are known, then the corresponding single-parameter QSNR 𝒬SP(λi)subscript𝒬𝑆𝑃subscript𝜆𝑖\mathcal{Q}_{SP}(\lambda_{i})caligraphic_Q start_POSTSUBSCRIPT italic_S italic_P end_POSTSUBSCRIPT ( italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) is finite because the diagonal element of the QFIM 𝓗λi,λisubscript𝓗subscript𝜆𝑖subscript𝜆𝑖\boldsymbol{\mathcal{H}}_{\lambda_{i},\lambda_{i}}bold_caligraphic_H start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT is finite. As soon as we have two unknown parameters, λisubscript𝜆𝑖\lambda_{i}italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and λjsubscript𝜆𝑗\lambda_{j}italic_λ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT, nothing can be said about either of them from measurements on the probe because the other QFIM elements come into play.

Refer to caption
Figure 1: Illustration of the 2IK model studied in this work. Two spin-1212\tfrac{1}{2}divide start_ARG 1 end_ARG start_ARG 2 end_ARG ‘impurity’ qubits comprising the probe are exchanged-coupled together, and each is also coupled to its own fermionic environment. We treat the fermionic environments explicitly as metallic leads involving a continuum of electronic states, appropriate to a realization of the model in a quantum nanoelectronics device. The full system allows nontrivial correlations and quantum entanglement to build up between the impurity probes and the fermionic environment of the leads. In particular, many-body physics associated with the probe-lead coupling J𝐽Jitalic_J favouring the Kondo effect, competes with the intra-probe coupling K𝐾Kitalic_K which favours local spin-singlet formation. This frustration produces a quantum critical point with macroscopic probe-environment entanglement and fractionalized excitations.

III Physical System and Model

We consider a simple probe system H^probe=K𝐒^IL𝐒^IRsubscript^𝐻probe𝐾subscript^𝐒𝐼𝐿subscript^𝐒𝐼𝑅\hat{H}_{\rm probe}=K~{}\hat{\vec{\mathbf{S}}}_{IL}\cdot\hat{\vec{\mathbf{S}}}% _{IR}over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT roman_probe end_POSTSUBSCRIPT = italic_K over^ start_ARG over→ start_ARG bold_S end_ARG end_ARG start_POSTSUBSCRIPT italic_I italic_L end_POSTSUBSCRIPT ⋅ over^ start_ARG over→ start_ARG bold_S end_ARG end_ARG start_POSTSUBSCRIPT italic_I italic_R end_POSTSUBSCRIPT involving two coupled quantum spin-1212\tfrac{1}{2}divide start_ARG 1 end_ARG start_ARG 2 end_ARG ‘impurity’ qubits, where here and throughout we assume units such that =1Planck-constant-over-2-pi1\hbar=1roman_ℏ = 1. The unique spin-singlet state of the isolated probe is the ground state when the exchange coupling between the impurities is antiferromagnetic K>0𝐾0K>0italic_K > 0, whereas the degenerate spin-triplet state is the ground state for ferromagnetic coupling K<0𝐾0K<0italic_K < 0.

Importantly, in our setup we model explicitly the environment to be probed by the impurities. We take the environment to be a fermionic bath comprising a continuum of electronic states, in the thermodynamic limit, which we divide into α=L𝛼𝐿\alpha\!=\!Litalic_α = italic_L and α=R𝛼𝑅\alpha\!=\!Ritalic_α = italic_R regions (leads). Each probe impurity α𝛼\alphaitalic_α is then coupled to its own bath α𝛼\alphaitalic_α by an exchange coupling J𝐽Jitalic_J, see Fig. 1. The full ‘two impurity Kondo’ (2IK) model Jayaprakash et al. (1981); Jones et al. (1988); Affleck and Ludwig (1992); Affleck et al. (1995); Mitchell et al. (2012); Sela et al. (2011); Mitchell and Sela (2012) reads,

H^=H^E+H^probe+J(𝐒^IL𝐬^EL+𝐒^IR𝐬^ER)^𝐻subscript^𝐻𝐸subscript^𝐻probe𝐽subscript^𝐒𝐼𝐿subscript^𝐬𝐸𝐿subscript^𝐒𝐼𝑅subscript^𝐬𝐸𝑅\hat{H}=\hat{H}_{E}+\hat{H}_{\rm probe}+J~{}\left(\hat{\vec{\mathbf{S}}}_{IL}% \cdot\hat{\vec{\mathbf{s}}}_{EL}+\hat{\vec{\mathbf{S}}}_{IR}\cdot\hat{\vec{% \mathbf{s}}}_{ER}\right)over^ start_ARG italic_H end_ARG = over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT + over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT roman_probe end_POSTSUBSCRIPT + italic_J ( over^ start_ARG over→ start_ARG bold_S end_ARG end_ARG start_POSTSUBSCRIPT italic_I italic_L end_POSTSUBSCRIPT ⋅ over^ start_ARG over→ start_ARG bold_s end_ARG end_ARG start_POSTSUBSCRIPT italic_E italic_L end_POSTSUBSCRIPT + over^ start_ARG over→ start_ARG bold_S end_ARG end_ARG start_POSTSUBSCRIPT italic_I italic_R end_POSTSUBSCRIPT ⋅ over^ start_ARG over→ start_ARG bold_s end_ARG end_ARG start_POSTSUBSCRIPT italic_E italic_R end_POSTSUBSCRIPT ) (13)

where the fermionic environment is described by,

H^E=α=L,RH^E;α=α=L,Rk,σϵkc^αkσc^αkσsubscript^𝐻𝐸subscript𝛼𝐿𝑅subscript^𝐻𝐸𝛼subscript𝛼𝐿𝑅subscript𝑘𝜎subscriptitalic-ϵ𝑘subscriptsuperscript^𝑐𝛼𝑘𝜎subscriptsuperscript^𝑐absent𝛼𝑘𝜎\hat{H}_{E}=\sum_{\alpha=L,R}\hat{H}_{E;\alpha}=\sum_{\alpha=L,R}\sum_{k,% \sigma}\epsilon_{k}\hat{c}^{\dagger}_{\alpha k\sigma}\hat{c}^{\phantom{\dagger% }}_{\alpha k\sigma}over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_α = italic_L , italic_R end_POSTSUBSCRIPT over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT italic_E ; italic_α end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_α = italic_L , italic_R end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_k , italic_σ end_POSTSUBSCRIPT italic_ϵ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT over^ start_ARG italic_c end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α italic_k italic_σ end_POSTSUBSCRIPT over^ start_ARG italic_c end_ARG start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α italic_k italic_σ end_POSTSUBSCRIPT (14)

where c^αkσ()superscriptsubscript^𝑐𝛼𝑘𝜎\hat{c}_{\alpha k\sigma}^{(\dagger)}over^ start_ARG italic_c end_ARG start_POSTSUBSCRIPT italic_α italic_k italic_σ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( † ) end_POSTSUPERSCRIPT annihilates (creates) an electron in the single-particle momentum state k𝑘kitalic_k with spin σ=𝜎\sigma=\uparrowitalic_σ = ↑ or \downarrow in bath α𝛼\alphaitalic_α. Here ϵksubscriptitalic-ϵ𝑘\epsilon_{k}italic_ϵ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT is the dispersion, which for simplicity we take to be linear at low energies, giving a standard flat (metallic) electronic density of states within a band of halfwidth D𝐷Ditalic_D. The operator 𝐬^Eα=12ννc^α0σ𝝈σσc^α0σsubscript^𝐬𝐸𝛼12subscript𝜈superscript𝜈superscriptsubscript^𝑐𝛼0𝜎subscript𝝈𝜎superscript𝜎subscriptsuperscript^𝑐absent𝛼0superscript𝜎\hat{\vec{\mathbf{s}}}_{E\alpha}=\tfrac{1}{2}\sum_{\nu\nu^{\prime}}\hat{c}_{% \alpha 0\sigma}^{\dagger}\vec{\boldsymbol{\sigma}}_{\sigma\sigma^{\prime}}\hat% {c}^{\phantom{\dagger}}_{\alpha 0\sigma^{\prime}}over^ start_ARG over→ start_ARG bold_s end_ARG end_ARG start_POSTSUBSCRIPT italic_E italic_α end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ∑ start_POSTSUBSCRIPT italic_ν italic_ν start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT over^ start_ARG italic_c end_ARG start_POSTSUBSCRIPT italic_α 0 italic_σ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT over→ start_ARG bold_italic_σ end_ARG start_POSTSUBSCRIPT italic_σ italic_σ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT over^ start_ARG italic_c end_ARG start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α 0 italic_σ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT describes the spin density of bath α𝛼\alphaitalic_α at the probe position, where c^α0σsubscript^𝑐𝛼0𝜎\hat{c}_{\alpha 0\sigma}over^ start_ARG italic_c end_ARG start_POSTSUBSCRIPT italic_α 0 italic_σ end_POSTSUBSCRIPT is the corresponding local bath orbital to which impurity probe α𝛼\alphaitalic_α couples.

We note that the 2IK model is gapless when the leads are in the thermodynamic limit.

III.1 Phases and critical point

We consider the thermalized probe-lead system at equilibrium, treating the full many-body system exactly: non-perturbative renormalization effects at low temperatures produce macroscopic probe-lead entanglement through the Kondo effect Hewson (1993); Jones et al. (1988); Bayat et al. (2012, 2014), and the backaction effect of the probe on the fermionic environment cannot neglected when considering the probe response. The 2IK model embodies a nontrivial competition between the frustrated magnetic interactions K𝐾Kitalic_K and J𝐽Jitalic_J Jones et al. (1988). For K/J0𝐾𝐽0K/J\to 0italic_K / italic_J → 0 we have two decoupled single-impurity Kondo models. For antiferromagnetic J>0𝐽0J>0italic_J > 0 the Kondo effect produces strong-coupling physics at low temperatures TTKmuch-less-than𝑇subscript𝑇𝐾T\ll T_{K}italic_T ≪ italic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT, with TKDe2D/Jsimilar-tosubscript𝑇𝐾𝐷superscript𝑒2𝐷𝐽T_{K}\sim De^{-2D/J}italic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT ∼ italic_D italic_e start_POSTSUPERSCRIPT - 2 italic_D / italic_J end_POSTSUPERSCRIPT a low-energy scale called the Kondo temperature. The impurity spin is dynamically screened below TKsubscript𝑇𝐾T_{K}italic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT by surrounding conduction electrons in the environment by the formation of a many-body entanglement cloud Bayat et al. (2010). On the other hand, for J/K0𝐽𝐾0J/K\to 0italic_J / italic_K → 0, the leads are effectively decoupled and the full model reduces to just H^probesubscript^𝐻probe\hat{H}_{\rm probe}over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT roman_probe end_POSTSUBSCRIPT. The singlet and triplet states of the coupled impurities are essentially unaffected by the leads in this limit.

Frustration between incompatible singlet-formation mechanisms produces a second-order quantum phase transition in the thermodynamic limit of the 2IK model. A nontrivial critical point Affleck et al. (1995); Mitchell et al. (2012); Mitchell and Sela (2012) arises when the binding energy of the Kondo effect TKsubscript𝑇𝐾T_{K}italic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT driven by J𝐽Jitalic_J matches the magnetic binding energy between the impurities K𝐾Kitalic_K. The critical point at K=KcTK𝐾subscript𝐾𝑐similar-tosubscript𝑇𝐾K=K_{c}\sim T_{K}italic_K = italic_K start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ∼ italic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT separates a Kondo phase for K<Kc𝐾subscript𝐾𝑐K<K_{c}italic_K < italic_K start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT from a magnetic (RKKY) phase for K>Kc𝐾subscript𝐾𝑐K>K_{c}italic_K > italic_K start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. At zero temperature, tuning K𝐾Kitalic_K through the critical point at Kcsubscript𝐾𝑐K_{c}italic_K start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT results in a change in the many-body ground state of the system. However, signatures of criticality are observed over a window of K𝐾Kitalic_K around Kcsubscript𝐾𝑐K_{c}italic_K start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT at finite temperatures.

III.2 Impurity Quantum Metrology

We now explore the different regimes of the 2IK model in the context of single- and multiparameter metrology, focusing on the estimation of T𝑇Titalic_T and K𝐾Kitalic_K. The 2IK model captures a special case where the eigenvectors of the probe reduced density matrix do not explicitly depend on either T𝑇Titalic_T or K𝐾Kitalic_K, and therefore we may use Eq. 11 with λ=(T,K)T𝜆superscript𝑇𝐾𝑇\vec{\lambda}=\left(T,K\right)^{T}over→ start_ARG italic_λ end_ARG = ( italic_T , italic_K ) start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT to calculate the QFIM. As shown below, we find that the QFIM is singular for the 2IK model, and therefore estimation of the parameters T𝑇Titalic_T and/or K𝐾Kitalic_K by measurements on the probe is impossible when there is uncertainty in both T𝑇Titalic_T and K𝐾Kitalic_K. We therefore first consider the unproblematic single parameter estimation scenario in the following sections. Then in Sec. VII we show that the QFIM singularity can be removed by applying a known control field, thereby allowing us to recover true multiparameter estimation sensitivity in this system.

The global SU(2)2(2)( 2 ) spin symmetry of the full system is preserved on the level of the probe reduced density matrix (RDM) obtained by tracing out the electronic leads. This allows us to construct the probe RDM in the spin eigenbasis, which for the two-impurity system is diagonal:

ϱ^probe=diag(ϱS,ϱT;+1,ϱT;0,ϱT;1),subscript^italic-ϱprobediagsubscriptitalic-ϱ𝑆subscriptitalic-ϱ𝑇1subscriptitalic-ϱ𝑇0subscriptitalic-ϱ𝑇1\hat{\varrho}_{\rm probe}=\text{diag}\left(\varrho_{S},\varrho_{{T;+1}},% \varrho_{{T;0}},\varrho_{{T;-1}}\right),over^ start_ARG italic_ϱ end_ARG start_POSTSUBSCRIPT roman_probe end_POSTSUBSCRIPT = diag ( italic_ϱ start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT , italic_ϱ start_POSTSUBSCRIPT italic_T ; + 1 end_POSTSUBSCRIPT , italic_ϱ start_POSTSUBSCRIPT italic_T ; 0 end_POSTSUBSCRIPT , italic_ϱ start_POSTSUBSCRIPT italic_T ; - 1 end_POSTSUBSCRIPT ) , (15)

where ϱSsubscriptitalic-ϱ𝑆\varrho_{S}italic_ϱ start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT is the population of the reduced state spin-singlet, and ϱT;1=ϱT;0=ϱT;+1ϱTsubscriptitalic-ϱ𝑇1subscriptitalic-ϱ𝑇0subscriptitalic-ϱ𝑇1subscriptitalic-ϱ𝑇\varrho_{T;-1}=\varrho_{T;0}=\varrho_{T;+1}\equiv\varrho_{T}italic_ϱ start_POSTSUBSCRIPT italic_T ; - 1 end_POSTSUBSCRIPT = italic_ϱ start_POSTSUBSCRIPT italic_T ; 0 end_POSTSUBSCRIPT = italic_ϱ start_POSTSUBSCRIPT italic_T ; + 1 end_POSTSUBSCRIPT ≡ italic_ϱ start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT are the populations of the reduced state components of the spin-triplet (which are equal by symmetry when no external field acts).

Interestingly, even for the lead-coupled, many-body system, we can fully obtain ϱ^probesubscript^italic-ϱprobe\hat{\varrho}_{\rm probe}over^ start_ARG italic_ϱ end_ARG start_POSTSUBSCRIPT roman_probe end_POSTSUBSCRIPT from a single physical observable that can be measured on the probe system. We define the impurity spin-spin correlator (equivalent to the probe singlet fraction) as,

𝒞=𝐒^IL𝐒^IRTr((S^ILS^IR)ϱ^probe),𝒞delimited-⟨⟩subscript^𝐒𝐼𝐿subscript^𝐒𝐼𝑅tracesubscript^𝑆𝐼𝐿subscript^𝑆𝐼𝑅subscript^italic-ϱprobe\mathcal{C}=\langle\hat{\vec{\mathbf{S}}}_{IL}\cdot\hat{\vec{\mathbf{S}}}_{IR}% \rangle\equiv\Tr{\left(\hat{\vec{S}}_{IL}\cdot\hat{\vec{S}}_{IR}\right)~{}\hat% {\varrho}_{\rm probe}}\;,caligraphic_C = ⟨ over^ start_ARG over→ start_ARG bold_S end_ARG end_ARG start_POSTSUBSCRIPT italic_I italic_L end_POSTSUBSCRIPT ⋅ over^ start_ARG over→ start_ARG bold_S end_ARG end_ARG start_POSTSUBSCRIPT italic_I italic_R end_POSTSUBSCRIPT ⟩ ≡ roman_Tr ( start_ARG ( over^ start_ARG over→ start_ARG italic_S end_ARG end_ARG start_POSTSUBSCRIPT italic_I italic_L end_POSTSUBSCRIPT ⋅ over^ start_ARG over→ start_ARG italic_S end_ARG end_ARG start_POSTSUBSCRIPT italic_I italic_R end_POSTSUBSCRIPT ) over^ start_ARG italic_ϱ end_ARG start_POSTSUBSCRIPT roman_probe end_POSTSUBSCRIPT end_ARG ) , (16)

from which it follows that 𝒞=34(ϱTϱS)𝒞34subscriptitalic-ϱ𝑇subscriptitalic-ϱ𝑆\mathcal{C}=\tfrac{3}{4}\left(\varrho_{T}-\varrho_{S}\right)caligraphic_C = divide start_ARG 3 end_ARG start_ARG 4 end_ARG ( italic_ϱ start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT - italic_ϱ start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT ). Together with the normalization condition ϱS+3ϱT=1subscriptitalic-ϱ𝑆3subscriptitalic-ϱ𝑇1\varrho_{S}+3\varrho_{T}=1italic_ϱ start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT + 3 italic_ϱ start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT = 1 we can determine all of the RDM elements in terms of the observable correlator 𝒞𝒞\mathcal{C}caligraphic_C as:

ϱSsubscriptitalic-ϱ𝑆\displaystyle\varrho_{S}italic_ϱ start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT =\displaystyle== 14𝒞,14𝒞\displaystyle\frac{1}{4}-\mathcal{C}\;,divide start_ARG 1 end_ARG start_ARG 4 end_ARG - caligraphic_C , (17)
ϱTsubscriptitalic-ϱ𝑇\displaystyle\varrho_{T}italic_ϱ start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT =\displaystyle== 14+13𝒞.1413𝒞\displaystyle\frac{1}{4}+\frac{1}{3}\mathcal{C}\;.divide start_ARG 1 end_ARG start_ARG 4 end_ARG + divide start_ARG 1 end_ARG start_ARG 3 end_ARG caligraphic_C . (18)

We are now in a position to calculate the sensitivity in terms of the QFI via the correlator 𝒞𝒞\mathcal{C}caligraphic_C. Since the probe RDM is diagonal in the spin basis, the RDM eigenvectors are independent of the model parameters. For single-parameter estimation, where we assume there is only a single unknown parameter, Eq. (11) reduces from a matrix to a scalar given by

(λ)=i|λϱi|2ϱi=|λ𝒞|2(14𝒞)×(34+𝒞)𝜆subscript𝑖superscriptsubscript𝜆subscriptitalic-ϱ𝑖2subscriptitalic-ϱ𝑖superscriptsubscript𝜆𝒞214𝒞34𝒞\mathcal{H}\left(\lambda\right)=\sum_{i}\frac{|\partial_{\lambda}\varrho_{i}|^% {2}}{\varrho_{i}}=\frac{|\partial_{\lambda}\mathcal{C}|^{2}}{\left(\tfrac{1}{4% }-\mathcal{C}\right)\times\left(\tfrac{3}{4}+\mathcal{C}\right)}caligraphic_H ( italic_λ ) = ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT divide start_ARG | ∂ start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT italic_ϱ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_ϱ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG = divide start_ARG | ∂ start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT caligraphic_C | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG ( divide start_ARG 1 end_ARG start_ARG 4 end_ARG - caligraphic_C ) × ( divide start_ARG 3 end_ARG start_ARG 4 end_ARG + caligraphic_C ) end_ARG (19)

for λ=T𝜆𝑇\lambda=Titalic_λ = italic_T or K𝐾Kitalic_K. Since the correlator 𝒞𝒞\mathcal{C}caligraphic_C determines completely the probe RDM, it is equivalent to the SLD for this system and therefore corresponds to the ideal measurement for metrological purposes, saturating the quantum CRB.

IV Large K𝐾Kitalic_K-limit

We will first examine the single parameter estimation scenario for both temperature T𝑇Titalic_T and coupling K𝐾Kitalic_K. For simplicity and to provide physical insight into the behaviour of the full system, we focus here on an analytically tractable limiting case: the large-K𝐾Kitalic_K limit (K/J1)K/J\gg 1)italic_K / italic_J ≫ 1 ), where the fermionic leads play essentially no role in determining the reduced states of the probe. We may therefore approximate the full model as just that of the probe,

H^KL=H^probeK𝐒^IL𝐒^IR.subscript^𝐻𝐾𝐿subscript^𝐻probe𝐾subscript^𝐒𝐼𝐿subscript^𝐒𝐼𝑅\hat{H}_{KL}=\hat{H}_{\rm probe}\equiv K~{}\hat{\vec{\mathbf{S}}}_{IL}\cdot% \hat{\vec{\mathbf{S}}}_{IR}\;.over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT italic_K italic_L end_POSTSUBSCRIPT = over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT roman_probe end_POSTSUBSCRIPT ≡ italic_K over^ start_ARG over→ start_ARG bold_S end_ARG end_ARG start_POSTSUBSCRIPT italic_I italic_L end_POSTSUBSCRIPT ⋅ over^ start_ARG over→ start_ARG bold_S end_ARG end_ARG start_POSTSUBSCRIPT italic_I italic_R end_POSTSUBSCRIPT . (20)

The Hamiltonian is readily diagonalized and the probe density matrix follows immediately: ϱi=eEi/T/𝒵subscriptitalic-ϱ𝑖superscript𝑒subscript𝐸𝑖𝑇𝒵\varrho_{{i}}=e^{-E_{i}/T}/\mathcal{Z}italic_ϱ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_e start_POSTSUPERSCRIPT - italic_E start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT / italic_T end_POSTSUPERSCRIPT / caligraphic_Z are the thermal populations, properly normalised by the partition function 𝒵=ieEi/T𝒵subscript𝑖superscript𝑒subscript𝐸𝑖𝑇\mathcal{Z}=\sum_{i}e^{-E_{i}/T}caligraphic_Z = ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - italic_E start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT / italic_T end_POSTSUPERSCRIPT. Here ES=3K/4subscript𝐸𝑆3𝐾4E_{S}=-3K/4italic_E start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT = - 3 italic_K / 4 is the energy of the two-impurity probe spin-singlet state, and ET;Sz=+K/4subscript𝐸𝑇superscript𝑆𝑧𝐾4E_{T;S^{z}}=+K/4italic_E start_POSTSUBSCRIPT italic_T ; italic_S start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT end_POSTSUBSCRIPT = + italic_K / 4 is the energy of the three degenerate components of the spin-triplet state. Importantly, all populations are a function of the single rescaled parameter y=K/T𝑦𝐾𝑇y\!=\!K/Titalic_y = italic_K / italic_T. They can also be obtained in terms of the correlator 𝒞𝒞\mathcal{C}caligraphic_C.

Refer to caption
Figure 2: Single-parameter estimation of T𝑇Titalic_T and K𝐾Kitalic_K in the 2IK model. Top row (panels a-d): Results in the large-K𝐾Kitalic_K limit extracted from analytical solution. Bottom row (panels e-h): Numerical results for the full 2IK model from NRG. Columns 1 and 2 show the QFI for thermometry (T)𝑇\mathcal{H}(T)caligraphic_H ( italic_T ) and for estimation of the probe coupling strength (K)𝐾\mathcal{H}(K)caligraphic_H ( italic_K ) as a function of T𝑇Titalic_T and K𝐾Kitalic_K, whereas columns 3 and 4 show the corresponding single-parameter QSNR responses 𝒬SP(T)subscript𝒬𝑆𝑃𝑇\mathcal{Q}_{SP}\left(T\right)caligraphic_Q start_POSTSUBSCRIPT italic_S italic_P end_POSTSUBSCRIPT ( italic_T ) and 𝒬SP(K)subscript𝒬𝑆𝑃𝐾\mathcal{Q}_{SP}\left(K\right)caligraphic_Q start_POSTSUBSCRIPT italic_S italic_P end_POSTSUBSCRIPT ( italic_K ).

Although we consider only the probe Hamiltonian explicitly here, we take the fermionic leads implicitly and assume thermalization has occurred. The reduced states of the probe are therefore equivalent to the isolated probe states, and we can use the exact populations ϱisubscriptitalic-ϱ𝑖\varrho_{i}italic_ϱ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT obtained in Eqs. (17) and (18) to calculate the single-parameter QFI for estimation for T𝑇Titalic_T or K𝐾Kitalic_K by using Eq. (19). In this limit we find for thermometry,

(T)=3eK/TK2(3+eK/T)2T4,𝑇3superscript𝑒𝐾𝑇superscript𝐾2superscript3superscript𝑒𝐾𝑇2superscript𝑇4\mathcal{H}\left(T\right)=\frac{3~{}e^{K/T}~{}K^{2}}{\left(3+e^{K/T}\right)^{2% }T^{4}}\;,caligraphic_H ( italic_T ) = divide start_ARG 3 italic_e start_POSTSUPERSCRIPT italic_K / italic_T end_POSTSUPERSCRIPT italic_K start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG ( 3 + italic_e start_POSTSUPERSCRIPT italic_K / italic_T end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_T start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT end_ARG , (21)

and for estimation of K𝐾Kitalic_K,

(K)=3eK/T(3+eK/T)2T2.𝐾3superscript𝑒𝐾𝑇superscript3superscript𝑒𝐾𝑇2superscript𝑇2\mathcal{H}\left(K\right)=\frac{3~{}e^{K/T}}{\left(3+e^{K/T}\right)^{2}T^{2}}\;.caligraphic_H ( italic_K ) = divide start_ARG 3 italic_e start_POSTSUPERSCRIPT italic_K / italic_T end_POSTSUPERSCRIPT end_ARG start_ARG ( 3 + italic_e start_POSTSUPERSCRIPT italic_K / italic_T end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_T start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG . (22)

We show the behavior of these quantities in Fig. 2(a,b).

Evidently, the precision of parameter estimation is highly dependent on the singlet-triplet transitions in this limit. The best performance for estimating a given parameter by making local measurements on the probe is obtained when the population transfer between singlet and triplet states upon changing that parameter is maximal. Since the probe singlet and triplet states are separated by a single energy gap |K|𝐾|K|| italic_K |, this naturally happens when T|K|similar-to𝑇𝐾T\sim|K|italic_T ∼ | italic_K |.

A second interesting feature of the single-parameter QFIs obtained in the large-K𝐾Kitalic_K limit is the role of the sign of K𝐾Kitalic_K. This is because the degeneracy of the ground and excited probe states gets swapped when the sign of K𝐾Kitalic_K is flipped. Specifically, for positive (antiferromagnetic) coupling K>0𝐾0K>0italic_K > 0, the ground state is the unique spin-singlet state |S=12(||)|S\rangle=\tfrac{1}{\sqrt{2}}\left(|\uparrow\downarrow\rangle-|\downarrow% \uparrow\rangle\right)| italic_S ⟩ = divide start_ARG 1 end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG ( | ↑ ↓ ⟩ - | ↓ ↑ ⟩ ), while the excited states are the three degenerate components of the spin-triplet, |T;+1=||T;+1\rangle=|\uparrow\uparrow\rangle| italic_T ; + 1 ⟩ = | ↑ ↑ ⟩, |T;0=12(|+|)|T;0\rangle=\tfrac{1}{\sqrt{2}}\left(|\uparrow\downarrow\rangle+|\downarrow% \uparrow\rangle\right)| italic_T ; 0 ⟩ = divide start_ARG 1 end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG ( | ↑ ↓ ⟩ + | ↓ ↑ ⟩ ) and |T;1=||T;-1\rangle=|\downarrow\downarrow\rangle| italic_T ; - 1 ⟩ = | ↓ ↓ ⟩. For negative (ferromagnetic) coupling K<0𝐾0K<0italic_K < 0, it is the ground state that is degenerate and the excited state that is unique. As explored in previous works Correa et al. (2015); Campbell et al. (2018), systems with excited state degeneracies are known to give higher thermometric precision. We observe the same phenomenon here: better performance is obtained for K>0𝐾0K>0italic_K > 0 than K<0𝐾0K<0italic_K < 0. This gives rise to the asymmetric structure of the QFI plots in both Fig. 2(a) and (b) around K=0𝐾0K=0italic_K = 0. We note that the largest QFI arises at low temperatures and coupling strengths. In particular, it might seem counterintuitive that the best sensitivity to coupling strength K𝐾Kitalic_K is obtained when the probe impurities are actually decoupled, K=0𝐾0K=0italic_K = 0. This illustrates the need for the QSNR rather than the QFI itself when interpreting metrological capability. We further note that (T)𝑇\mathcal{H}(T)caligraphic_H ( italic_T ) and (K)𝐾\mathcal{H}(K)caligraphic_H ( italic_K ) look very different, even though the underlying probe populations in the large-K𝐾Kitalic_K limit depend only on the single rescaled parameter y=K/T𝑦𝐾𝑇y\!=\!K/Titalic_y = italic_K / italic_T, and so we expect the corresponding QFIs for K𝐾Kitalic_K and T𝑇Titalic_T to be simply related. Again, the QSNR helps to uncover these similarities.

First, we note that from Eq. (19) we may write (λ)=(y)/|λy|2𝜆𝑦superscriptsubscript𝜆𝑦2\mathcal{H}(\lambda)=\mathcal{H}(y)/|\partial_{\lambda}y|^{2}caligraphic_H ( italic_λ ) = caligraphic_H ( italic_y ) / | ∂ start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT italic_y | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. Since Ky=1/Tsubscript𝐾𝑦1𝑇\partial_{K}y=1/T∂ start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT italic_y = 1 / italic_T and Ty=K/T2subscript𝑇𝑦𝐾superscript𝑇2\partial_{T}y=-K/T^{2}∂ start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT italic_y = - italic_K / italic_T start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, it follows immediately that (T)×T2=(K)×K2𝑇superscript𝑇2𝐾superscript𝐾2\mathcal{H}(T)\times T^{2}=\mathcal{H}(K)\times K^{2}caligraphic_H ( italic_T ) × italic_T start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = caligraphic_H ( italic_K ) × italic_K start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. This is precisely the definition of the QSNR in Eq. (5), such that

𝒬SP(T)=𝒬SP(K)=3eK/TK2(3+eK/T)2T2,subscript𝒬𝑆𝑃𝑇subscript𝒬𝑆𝑃𝐾3superscript𝑒𝐾𝑇superscript𝐾2superscript3superscript𝑒𝐾𝑇2superscript𝑇2\mathcal{Q}_{SP}\left(T\right)=\mathcal{Q}_{SP}\left(K\right)=\frac{3~{}e^{K/T% }~{}K^{2}}{\left(3+e^{K/T}\right)^{2}T^{2}}\;,caligraphic_Q start_POSTSUBSCRIPT italic_S italic_P end_POSTSUBSCRIPT ( italic_T ) = caligraphic_Q start_POSTSUBSCRIPT italic_S italic_P end_POSTSUBSCRIPT ( italic_K ) = divide start_ARG 3 italic_e start_POSTSUPERSCRIPT italic_K / italic_T end_POSTSUPERSCRIPT italic_K start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG ( 3 + italic_e start_POSTSUPERSCRIPT italic_K / italic_T end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_T start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG , (23)

which is a universal function of the single parameter y=K/T𝑦𝐾𝑇y=K/Titalic_y = italic_K / italic_T. The QSNR results for the 2IK system in the large-K𝐾Kitalic_K limit are presented in Fig. 2(c,d) and demonstrate clearly the physical features discussed above. In particular, we see that the QSNR is indeed identical for single-parameter T𝑇Titalic_T and K𝐾Kitalic_K estimation, with the maximum sensitivity attained along the line K2.85Tsimilar-to-or-equals𝐾2.85𝑇K\simeq 2.85Titalic_K ≃ 2.85 italic_T for K>0𝐾0K>0italic_K > 0 with 𝒬Max1similar-to-or-equalssubscript𝒬𝑀𝑎𝑥1\mathcal{Q}_{Max}\simeq 1caligraphic_Q start_POSTSUBSCRIPT italic_M italic_a italic_x end_POSTSUBSCRIPT ≃ 1, whereas for K<0𝐾0K<0italic_K < 0 the maximum sensitivity is lower with 𝒬Max1/6similar-to-or-equalssubscript𝒬𝑀𝑎𝑥16\mathcal{Q}_{Max}\simeq 1/6caligraphic_Q start_POSTSUBSCRIPT italic_M italic_a italic_x end_POSTSUBSCRIPT ≃ 1 / 6 along the line K2.16Tsimilar-to-or-equals𝐾2.16𝑇K\simeq-2.16Titalic_K ≃ - 2.16 italic_T. We attribute the boosted robustness to measurement noise in the antiferromagnetic regime to the enhanced probe degeneracy for the excited state in this case.

Although the large-K𝐾Kitalic_K limit is over-simplified, we expect certain qualitative features (such as the difference between positive and negative K𝐾Kitalic_K) to carry over to the full solution. In particular, we remind that the QFI and QSNR signatures of the 2IK probe are fully determined by a single measureable observable, the spin-spin correlator 𝒞𝒞\mathcal{C}caligraphic_C, not only in the large-K𝐾Kitalic_K limit but also in the full lead-coupled model. Although many-body effects conspire to produce richer physics in the full lead-coupled system that are naturally reflected in a more complex structure for 𝒞𝒞\mathcal{C}caligraphic_C, we may still use Eq. (19) to extract metrological properties. On the other hand, we do not generally expect 𝒬SP(T)=𝒬SP(K)subscript𝒬𝑆𝑃𝑇subscript𝒬𝑆𝑃𝐾\mathcal{Q}_{SP}(T)=\mathcal{Q}_{SP}(K)caligraphic_Q start_POSTSUBSCRIPT italic_S italic_P end_POSTSUBSCRIPT ( italic_T ) = caligraphic_Q start_POSTSUBSCRIPT italic_S italic_P end_POSTSUBSCRIPT ( italic_K ) when the leads are attached because the probe RDM eigenvalues are then no longer simple functions of the single rescaled parameter K/T𝐾𝑇K/Titalic_K / italic_T and competition with other scales (J,D,TK𝐽𝐷subscript𝑇𝐾J,D,T_{K}italic_J , italic_D , italic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT) will become important. We refer to Appendix B where the nontrivial competition that arises when we have competing energy scales can already been seen in the narrow band limit (i.e. when J/D1much-greater-than𝐽𝐷1J/D\gg 1italic_J / italic_D ≫ 1 and wherein the electronic states in the leads can be approximated by a single local orbital in real space).

V 2IK model: NRG results

When a full continuum of electronic states is included in the metallic leads (which constitute the environment), the 2IK model, Eq. (13), is a famous strongly-correlated many-body problem Jones et al. (1988); Affleck et al. (1995); Mitchell and Sela (2012) whose solution requires the use of sophisticated methods. Here we use Wilson’s NRG method Bulla et al. (2008); Weichselbaum and von Delft (2007); Mitchell et al. (2014); *stadler2016interleaved to obtain the spin-spin correlation function 𝒞=𝐒^IL𝐒^IR𝒞delimited-⟨⟩subscript^𝐒𝐼𝐿subscript^𝐒𝐼𝑅\mathcal{C}=\langle\hat{\vec{\mathbf{S}}}_{IL}\cdot\hat{\vec{\mathbf{S}}}_{IR}\ranglecaligraphic_C = ⟨ over^ start_ARG over→ start_ARG bold_S end_ARG end_ARG start_POSTSUBSCRIPT italic_I italic_L end_POSTSUBSCRIPT ⋅ over^ start_ARG over→ start_ARG bold_S end_ARG end_ARG start_POSTSUBSCRIPT italic_I italic_R end_POSTSUBSCRIPT ⟩ numerically as a function of temperature T𝑇Titalic_T and couplings K𝐾Kitalic_K and J𝐽Jitalic_J. We set the conduction electron bandwidth D=1𝐷1D=1italic_D = 1, use NRG discretization parameter Λ=2.5Λ2.5\Lambda=2.5roman_Λ = 2.5 and keep Ns=8000subscript𝑁𝑠8000N_{s}=8000italic_N start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = 8000 states at each step of the iterative diagonalization procedure. As previously emphasized, a knowledge of 𝒞𝒞\mathcal{C}caligraphic_C (and its derivatives) is sufficient to determine fully the single parameter metrological capability of the 2IK probe – and so NRG, which provides numerically-exact access to this quantity, is an ideal tool. The evolution of 𝒞𝒞\mathcal{C}caligraphic_C as a function of T𝑇Titalic_T and K𝐾Kitalic_K is smooth in the full model, although at T=0𝑇0T=0italic_T = 0 we see a discontinuity in K𝒞subscript𝐾𝒞\partial_{K}\mathcal{C}∂ start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT caligraphic_C at K=Kc𝐾subscript𝐾𝑐K=K_{c}italic_K = italic_K start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, indicating the existence of a second-order quantum phase transition Affleck et al. (1995).

V.1 Overview of metrological phase diagram

Fig. 2(e-h) show the NRG results for the full lead-coupled 2IK model, as a function of K𝐾Kitalic_K and T𝑇Titalic_T. We set J=1𝐽1J=1italic_J = 1 here, and find that the critical point is located at K=Kc0.62𝐾subscript𝐾𝑐0.62K=K_{c}\approx 0.62italic_K = italic_K start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ≈ 0.62. Several qualitative features are reminiscent of the large-K𝐾Kitalic_K limit results in panels (a-d). In particular, the thermometric QFI (T)𝑇\mathcal{H}(T)caligraphic_H ( italic_T ) shows a split two-lobe structure; but in the full model this behavior is pushed to low temperature TJmuch-less-than𝑇𝐽T\ll Jitalic_T ≪ italic_J and is centred at the critical point K=Kc𝐾subscript𝐾𝑐K=K_{c}italic_K = italic_K start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT rather than K=0𝐾0K=0italic_K = 0. Likewise for (K)𝐾\mathcal{H}(K)caligraphic_H ( italic_K ) we see a single intense flair, but again it is now located at K=Kc𝐾subscript𝐾𝑐K=K_{c}italic_K = italic_K start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT.

The QSNR results in Fig. 2(g,h) tell the clearest story, however. Despite the large thermometric QFI near the critical point, this enhanced precision arises only at low temperatures where the thermometric signal is also small. Overall the thermometric QSNR is surprisingly poor at low temperatures. Only at larger T𝑇Titalic_T and K𝐾Kitalic_K do we see good temperature estimation capability due to the finite J𝐽Jitalic_J (or TKsubscript𝑇𝐾T_{K}italic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT) scale. This is because the measured probe observable 𝒞𝒞\mathcal{C}caligraphic_C does not change appreciably with temperature when TTKmuch-less-than𝑇subscript𝑇𝐾T\ll T_{K}italic_T ≪ italic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT, even at the critical point K=Kc𝐾subscript𝐾𝑐K=K_{c}italic_K = italic_K start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. On the other hand, the QSNR sensitivity to single-parameter estimation of the coupling constant K𝐾Kitalic_K is strongly enhanced near the critical point, especially at low temperatures. This is due to the the sharp crossover in 𝒞𝒞\mathcal{C}caligraphic_C as K𝐾Kitalic_K is tuned in the vicinity of the critical point at Kcsubscript𝐾𝑐K_{c}italic_K start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT.

We argue that these are generic features near a second-order quantum phase transition, expected on general grounds from a renormalization group (RG) perspective. In the critical regime, the physics of the system is controlled only by the critical fixed point at low temperatures, with the temperature-dependence of physical observables scaling weakly with RG irrelevant, or at best marginal, operators Sachdev (1999). By contrast, the dependence on a parameter driving the transition will typically be strong, since by definition its scaling is controlled by RG relevant operators.

V.2 Universal Kondo regime

Refer to caption
Figure 3: QSNR of the full 2IK model in the universal Kondo regime. Top panel (a): QSNR for thermometry 𝒬SP(T)subscript𝒬𝑆𝑃𝑇\mathcal{Q}_{SP}\left(T\right)caligraphic_Q start_POSTSUBSCRIPT italic_S italic_P end_POSTSUBSCRIPT ( italic_T ). Bottom panel (b): QSNR for coupling constant 𝒬SP(K)subscript𝒬𝑆𝑃𝐾\mathcal{Q}_{SP}\left(K\right)caligraphic_Q start_POSTSUBSCRIPT italic_S italic_P end_POSTSUBSCRIPT ( italic_K ). NRG calculations performed using J=0.15𝐽0.15J=0.15italic_J = 0.15, for which the Kondo temperature is found to be TK107Dsubscript𝑇𝐾superscript107𝐷T_{K}\approx 10^{-7}Ditalic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT ≈ 10 start_POSTSUPERSCRIPT - 7 end_POSTSUPERSCRIPT italic_D and the critical point is at K=Kc6TK𝐾subscript𝐾𝑐6subscript𝑇𝐾K=K_{c}\approx 6T_{K}italic_K = italic_K start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ≈ 6 italic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT. Note: axes are rescaled in terms of TKsubscript𝑇𝐾T_{K}italic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT and on a log scale.

The full 2IK model supports a nontrivial quantum phase transition that separates two distinct regimes Jones et al. (1988); Affleck et al. (1995); Mitchell et al. (2012). For K<Kc𝐾subscript𝐾𝑐K<K_{c}italic_K < italic_K start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT we have essentially two separated single-impurity Kondo models, whereas for K>Kc𝐾subscript𝐾𝑐K>K_{c}italic_K > italic_K start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT the two impurity probes lock up into a local spin-singlet and effectively decouple from the leads. Close to the critical point KKcsimilar-to𝐾subscript𝐾𝑐K\sim K_{c}italic_K ∼ italic_K start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT quantum fluctuations develop on long length and time scales and both impurities and leads become strongly entangled in one composite Bayat et al. (2012, 2014). The critical value of Kcsubscript𝐾𝑐K_{c}italic_K start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT depends on J𝐽Jitalic_J and is on the order of the single-channel TKsubscript𝑇𝐾T_{K}italic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT, which therefore gets small very quickly. In this section we use NRG to investigate the universal Kondo regime around TTKsimilar-to𝑇subscript𝑇𝐾T\sim T_{K}italic_T ∼ italic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT and KTKsimilar-to𝐾subscript𝑇𝐾K\sim T_{K}italic_K ∼ italic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT for J=0.15𝐽0.15J=0.15italic_J = 0.15, for which we find TK107subscript𝑇𝐾superscript107T_{K}\approx 10^{-7}italic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT ≈ 10 start_POSTSUPERSCRIPT - 7 end_POSTSUPERSCRIPT and Kc6TKsubscript𝐾𝑐6subscript𝑇𝐾K_{c}\approx 6T_{K}italic_K start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ≈ 6 italic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT. The reason for this choice is that the nonuniversal physics associated with the bare scales D𝐷Ditalic_D and J𝐽Jitalic_J is then unimportant. The critical point at Kcsubscript𝐾𝑐K_{c}italic_K start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT is determined from NRG calculations by tuning K𝐾Kitalic_K to achieve a vanishing Fermi liquid scale.

In Fig. 3 we consider the QSNR for single-parameter estimation of T𝑇Titalic_T and K𝐾Kitalic_K in this regime, with results obtained by NRG. We see very clearly from the numerical results the onset of critical physics and the demarcation of the two phases of the model, when using the impurity probes for metrology – especially so for 𝒬SP(K)subscript𝒬𝑆𝑃𝐾\mathcal{Q}_{SP}(K)caligraphic_Q start_POSTSUBSCRIPT italic_S italic_P end_POSTSUBSCRIPT ( italic_K ) in the lower panel.

For thermometry, Fig. 3(a), the measurement sensitivity at low temperatures TTKsimilar-to𝑇subscript𝑇𝐾T\sim T_{K}italic_T ∼ italic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT is almost entirely lost in the Kondo phase K<Kc𝐾subscript𝐾𝑐K<K_{c}italic_K < italic_K start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. This is expected from the results for the single-impurity Kondo probe explored in Ref. Mihailescu et al. (2023), because only a small amount of information about the state is imprinted locally on the impurity probe due to the macroscopic Kondo entanglement with the leads. For K>Kc𝐾subscript𝐾𝑐K>K_{c}italic_K > italic_K start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT however, we see that singlet-triplet transitions on the probe impurities give good measurement sensitivity when the temperature T𝑇Titalic_T is on the order of the renormalized probe gap, which scales as Ksimilar-toabsent𝐾\sim K∼ italic_K.

In Fig. 3(b) we investigate 𝒬SP(K)subscript𝒬𝑆𝑃𝐾\mathcal{Q}_{SP}(K)caligraphic_Q start_POSTSUBSCRIPT italic_S italic_P end_POSTSUBSCRIPT ( italic_K ) in the same system. In the Kondo phase K<Kc𝐾subscript𝐾𝑐K<K_{c}italic_K < italic_K start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT we again expect low measurement sensitivity because the probe populations do not change much with K𝐾Kitalic_K when both probes are separately being Kondo screened by their attached leads. But for K>Kc𝐾subscript𝐾𝑐K>K_{c}italic_K > italic_K start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT we again see enhanced sensitivity for KTsimilar-to𝐾𝑇K\sim Titalic_K ∼ italic_T due to the internal singlet-triplet transitions on the probe. The major difference is that 𝒬SP(K)subscript𝒬𝑆𝑃𝐾\mathcal{Q}_{SP}(K)caligraphic_Q start_POSTSUBSCRIPT italic_S italic_P end_POSTSUBSCRIPT ( italic_K ) is also very sensitive to the critical point in the model, with boosted precision for measurement of the coupling constant around Kcsubscript𝐾𝑐K_{c}italic_K start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT at low temperatures. This makes intuitive sense since at low-T𝑇Titalic_T, changing K𝐾Kitalic_K away from its critical value of Kcsubscript𝐾𝑐K_{c}italic_K start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT causes dramatic changes, driving the system into one or other of the stable phases of the model.

VI 2IK model: exact results near criticality

We now examine carefully the full 2IK model in the close vicinity of the critical point. For a given J𝐽Jitalic_J, the critical point at K=Kc𝐾subscript𝐾𝑐K=K_{c}italic_K = italic_K start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT is characterized by a single energy scale TKsubscript𝑇𝐾T_{K}italic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT, equivalent to the usual single-impurity Kondo scale Jones et al. (1988). When TKJ,Dmuch-less-thansubscript𝑇𝐾𝐽𝐷T_{K}\ll J,Ditalic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT ≪ italic_J , italic_D, physical properties are universal scaling functions of the rescaled parameter T/TK𝑇subscript𝑇𝐾T/T_{K}italic_T / italic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT, controlled by the 2IK quantum critical fixed point Affleck et al. (1995). However, detuning the impurity coupling introduces a finite relevant perturbation δK=KKc𝛿𝐾𝐾subscript𝐾𝑐\delta K=K-K_{c}italic_δ italic_K = italic_K - italic_K start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT which destabilizes the critical point and generates a nontrivial RG flow towards a Fermi liquid fixed point on the new scale of Tsuperscript𝑇T^{*}italic_T start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT Sela et al. (2011); Mitchell and Sela (2012). For TTmuch-less-than𝑇superscript𝑇T\ll T^{*}italic_T ≪ italic_T start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT the system flows towards one of the two stable phases of the model, depending on the sign of the perturbation. For δK<0𝛿𝐾0\delta K<0italic_δ italic_K < 0 the system flows to the Kondo phase, whereas for δK>0𝛿𝐾0\delta K>0italic_δ italic_K > 0 the system flows to the local inter-impurity singlet phase. However, for small perturbations |δK|TKmuch-less-than𝛿𝐾subscript𝑇𝐾|\delta K|\ll T_{K}| italic_δ italic_K | ≪ italic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT, we have good scale separation TTKmuch-less-thansuperscript𝑇subscript𝑇𝐾T^{*}\ll T_{K}italic_T start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ≪ italic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT, with Tsuperscript𝑇T^{*}italic_T start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT given by Affleck et al. (1995)

TTK=c(KKCTK)2,superscript𝑇subscript𝑇𝐾𝑐superscript𝐾subscript𝐾𝐶subscript𝑇𝐾2\frac{T^{*}}{T_{K}}=c~{}\left(\frac{K-K_{C}}{T_{K}}\right)^{2}\;,divide start_ARG italic_T start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT end_ARG start_ARG italic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT end_ARG = italic_c ( divide start_ARG italic_K - italic_K start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT end_ARG start_ARG italic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , (24)

where c𝑐citalic_c is a constant. For TTKmuch-less-than𝑇subscript𝑇𝐾T\ll T_{K}italic_T ≪ italic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT physical quantities are universal functions of the rescaled parameter T/T𝑇superscript𝑇T/T^{*}italic_T / italic_T start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT and are entirely characteristic of the 2IK quantum critical point, independent of microscopic details. Remarkably, in this universal critical regime the 2IK admits an exact analytical solution Sela et al. (2011); Mitchell and Sela (2012) – despite it being a strongly-correlated many-body problem. In particular, the entropy flow along the crossover on the scale of Tsuperscript𝑇T^{*}italic_T start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT is known analytically in closed form Mitchell and Sela (2012),

S(T)=12log(2)+S¯(TT):TTKS\left(T\right)=\frac{1}{2}\log{2}+\bar{S}\left(\frac{T}{T^{*}}\right)\qquad% \qquad:~{}T\ll T_{K}italic_S ( italic_T ) = divide start_ARG 1 end_ARG start_ARG 2 end_ARG roman_log ( start_ARG 2 end_ARG ) + over¯ start_ARG italic_S end_ARG ( divide start_ARG italic_T end_ARG start_ARG italic_T start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT end_ARG ) : italic_T ≪ italic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT (25)

where S¯¯𝑆\bar{S}over¯ start_ARG italic_S end_ARG is defined as,

S¯(t)=1t[ψ(12+1t)1]log[1πΓ(12+1t)]¯𝑆𝑡1𝑡delimited-[]𝜓121𝑡11𝜋Γ121𝑡\bar{S}\left(t\right)=\frac{1}{t}\left[\psi\left(\frac{1}{2}+\frac{1}{t}\right% )-1\right]-\log\left[\frac{1}{\sqrt{\pi}}\Gamma\left(\frac{1}{2}+\frac{1}{t}% \right)\right]over¯ start_ARG italic_S end_ARG ( italic_t ) = divide start_ARG 1 end_ARG start_ARG italic_t end_ARG [ italic_ψ ( divide start_ARG 1 end_ARG start_ARG 2 end_ARG + divide start_ARG 1 end_ARG start_ARG italic_t end_ARG ) - 1 ] - roman_log [ divide start_ARG 1 end_ARG start_ARG square-root start_ARG italic_π end_ARG end_ARG roman_Γ ( divide start_ARG 1 end_ARG start_ARG 2 end_ARG + divide start_ARG 1 end_ARG start_ARG italic_t end_ARG ) ] (26)

with Γ()Γ\Gamma(\cdot)roman_Γ ( ⋅ ) and ψ()𝜓\psi(\cdot)italic_ψ ( ⋅ ) being the gamma and the digamma functions, respectively. Full NRG calculations for the 2IK entropy performed in the critical regime TTKmuch-less-thansuperscript𝑇subscript𝑇𝐾T^{*}\ll T_{K}italic_T start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ≪ italic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT confirm Eqs. (24) and (25) precisely. Taking the standard operational definition of the Kondo temperature through S(T=TK)=log(2)𝑆𝑇subscript𝑇𝐾2S(T=T_{K})=\log(2)italic_S ( italic_T = italic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT ) = roman_log ( start_ARG 2 end_ARG ) we extract the constant c0.035similar-to-or-equals𝑐0.035c\simeq 0.035italic_c ≃ 0.035 from the NRG thermodynamics.

For the purposes of single-parameter estimation, we need access to the correlator 𝒞=𝐒^IL𝐒^IR𝒞delimited-⟨⟩subscript^𝐒𝐼𝐿subscript^𝐒𝐼𝑅\mathcal{C}=\langle\hat{\vec{\mathbf{S}}}_{IL}\cdot\hat{\vec{\mathbf{S}}}_{IR}\ranglecaligraphic_C = ⟨ over^ start_ARG over→ start_ARG bold_S end_ARG end_ARG start_POSTSUBSCRIPT italic_I italic_L end_POSTSUBSCRIPT ⋅ over^ start_ARG over→ start_ARG bold_S end_ARG end_ARG start_POSTSUBSCRIPT italic_I italic_R end_POSTSUBSCRIPT ⟩ so that we may compute the QFI via Eq. (19). The behavior of this correlator has not previously been discussed in the 2IK critical region. However, here we note that 𝒞=K𝒞subscript𝐾\mathcal{C}=\partial_{K}~{}\mathcal{F}caligraphic_C = ∂ start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT caligraphic_F is an exact identity, where =Tln(𝒵)𝑇𝒵\mathcal{F}=-T\ln{\mathcal{Z}}caligraphic_F = - italic_T roman_ln ( start_ARG caligraphic_Z end_ARG ) is the equilibrium thermodynamic free energy (grand potential), and 𝒵𝒵\mathcal{Z}caligraphic_Z is the partition function of the full system. Meanwhile, the thermodynamic entropy is also related to the free energy, S=T𝑆subscript𝑇S=-\partial_{T}\mathcal{F}italic_S = - ∂ start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT caligraphic_F. Therefore at thermal equilibrium we may apply a Maxwell relation to connect the entropy to the correlator,

TK=KTsubscript𝑇subscript𝐾subscript𝐾subscript𝑇\partial_{T}\partial_{K}\mathcal{F}=\partial_{K}\partial_{T}\mathcal{F}∂ start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT ∂ start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT caligraphic_F = ∂ start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT ∂ start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT caligraphic_F (27)

Thus it follows that,

T𝒞=KSsubscript𝑇𝒞subscript𝐾𝑆\partial_{T}\mathcal{C}=-\partial_{K}S∂ start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT caligraphic_C = - ∂ start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT italic_S (28)

which holds as an exact identity for any T𝑇Titalic_T and K𝐾Kitalic_K (not just in the critical region). However, the derivative on the right hand side can be evaluated explicitly using Eqs. (25) and (26) if we confine our attention now to the critical region. This yields,

T𝒞(T,K)=2cδK[TTKcδK2ψ(Φ)]T2TK2subscript𝑇𝒞𝑇𝐾2𝑐𝛿𝐾delimited-[]𝑇subscript𝑇𝐾𝑐𝛿superscript𝐾2superscript𝜓Φsuperscript𝑇2superscriptsubscript𝑇𝐾2\displaystyle\partial_{T}\mathcal{C}(T,K)=\frac{2c~{}\delta K\left[T~{}T_{K}-c% ~{}\delta K^{2}~{}\psi^{\prime}\left(\Phi\right)\right]}{T^{2}~{}T_{K}^{2}}∂ start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT caligraphic_C ( italic_T , italic_K ) = divide start_ARG 2 italic_c italic_δ italic_K [ italic_T italic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT - italic_c italic_δ italic_K start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ψ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( roman_Φ ) ] end_ARG start_ARG italic_T start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG (29)

where we have defined ΦΦ(T,K)=12+TT=12+cδK2TTKΦΦ𝑇𝐾12superscript𝑇𝑇12𝑐𝛿superscript𝐾2𝑇subscript𝑇𝐾\Phi\equiv\Phi(T,K)=\tfrac{1}{2}+\tfrac{T^{*}}{T}=\tfrac{1}{2}+\tfrac{c~{}% \delta K^{2}}{T~{}T_{K}}roman_Φ ≡ roman_Φ ( italic_T , italic_K ) = divide start_ARG 1 end_ARG start_ARG 2 end_ARG + divide start_ARG italic_T start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT end_ARG start_ARG italic_T end_ARG = divide start_ARG 1 end_ARG start_ARG 2 end_ARG + divide start_ARG italic_c italic_δ italic_K start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_T italic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT end_ARG, and with ψsuperscript𝜓\psi^{\prime}italic_ψ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT the trigamma function. An exact expression for the impurity spin-spin correlation function itself can now be obtained by integrating

𝒞(T,K)𝒞𝑇𝐾\displaystyle\mathcal{C}\left(T,K\right)caligraphic_C ( italic_T , italic_K ) =𝑑TKS(T,K)absentdifferential-d𝑇subscript𝐾𝑆𝑇𝐾\displaystyle=-\int dT~{}\partial_{K}S\left(T,K\right)= - ∫ italic_d italic_T ∂ start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT italic_S ( italic_T , italic_K )
=2cδK[log(T)+ψ(Φ)]TK+𝒞absent2𝑐𝛿𝐾delimited-[]𝑇𝜓Φsubscript𝑇𝐾superscript𝒞\displaystyle=\frac{2c~{}\delta K\left[\log{T}+\psi\left(\Phi\right)\right]}{T% _{K}}+\mathcal{C}^{*}= divide start_ARG 2 italic_c italic_δ italic_K [ roman_log ( start_ARG italic_T end_ARG ) + italic_ψ ( roman_Φ ) ] end_ARG start_ARG italic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT end_ARG + caligraphic_C start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT (30)

where the constant of integration 𝒞=𝒞(T=0,K=Kc)superscript𝒞𝒞formulae-sequence𝑇0𝐾subscript𝐾𝑐\mathcal{C}^{*}=\mathcal{C}(T=0,K=K_{c})caligraphic_C start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = caligraphic_C ( italic_T = 0 , italic_K = italic_K start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ) is found to be the value of the spin-spin correlator at the critical point. This is determined by noting that the indefinite integral is defined up to a function of K𝐾Kitalic_K, and that 𝒞(TTK,KC)=𝒞𝒞much-less-than𝑇subscript𝑇𝐾subscript𝐾𝐶superscript𝒞\mathcal{C}(T\ll T_{K},K_{C})=\mathcal{C}^{*}caligraphic_C ( italic_T ≪ italic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT , italic_K start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ) = caligraphic_C start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT and 𝒞(T,K)=0𝒞𝑇𝐾0\mathcal{C}(T\to\infty,K)=0caligraphic_C ( italic_T → ∞ , italic_K ) = 0. Although 𝒞superscript𝒞\mathcal{C}^{*}caligraphic_C start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT depends in general on J𝐽Jitalic_J, it can be calculated numerically with NRG. With this information, we have the full temperature and coupling dependence of the correlator in the critical region via Eq. (30). We note that similar Maxwell relations have been used recently in reverse, to determine the entropy changes for a process from measureable observables in experiments on quantum devices Han et al. (2022); Child et al. (2022).

We can now take the derivative with respect to K𝐾Kitalic_K to obtain,

K𝒞(T,K)=2c[TTK(log(T)+ψ(Φ))+2cδK2ψ(Φ)]TTK2subscript𝐾𝒞𝑇𝐾2𝑐delimited-[]𝑇subscript𝑇𝐾𝑇𝜓Φ2𝑐𝛿superscript𝐾2superscript𝜓Φ𝑇superscriptsubscript𝑇𝐾2\partial_{K}\mathcal{C}\left(T,K\right)=\frac{2c~{}\left[T~{}T_{K}\left(\log{T% }+\psi\left(\Phi\right)\right)+2c~{}\delta K^{2}~{}\psi^{\prime}\left(\Phi% \right)\right]}{T~{}T_{K}^{2}}∂ start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT caligraphic_C ( italic_T , italic_K ) = divide start_ARG 2 italic_c [ italic_T italic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT ( roman_log ( start_ARG italic_T end_ARG ) + italic_ψ ( roman_Φ ) ) + 2 italic_c italic_δ italic_K start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ψ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( roman_Φ ) ] end_ARG start_ARG italic_T italic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG (31)

We validate these analytic results in Appendix C by explicit comparison to NRG results in the universal regime.

We now have everything we need to compute the QFIs through Eq. (19). Our exact analytic expressions for the corresponding QSNRs in the critical region follow as,

𝒬SP(T)=4c2δK2[TTKcδK2ψ(Φ)]2T2TK4(14𝒞)×(34+𝒞)subscript𝒬𝑆𝑃𝑇4superscript𝑐2𝛿superscript𝐾2superscriptdelimited-[]𝑇subscript𝑇𝐾𝑐𝛿superscript𝐾2superscript𝜓Φ2superscript𝑇2superscriptsubscript𝑇𝐾414𝒞34𝒞\mathcal{Q}_{SP}\left(T\right)=\frac{4c^{2}~{}\delta K^{2}\left[T~{}T_{K}-c~{}% \delta K^{2}~{}\psi^{\prime}\left(\Phi\right)\right]^{2}}{T^{2}~{}T_{K}^{4}~{}% \left(\tfrac{1}{4}-\mathcal{C}\right)\times\left(\tfrac{3}{4}+\mathcal{C}% \right)}caligraphic_Q start_POSTSUBSCRIPT italic_S italic_P end_POSTSUBSCRIPT ( italic_T ) = divide start_ARG 4 italic_c start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_δ italic_K start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT [ italic_T italic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT - italic_c italic_δ italic_K start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ψ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( roman_Φ ) ] start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_T start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT ( divide start_ARG 1 end_ARG start_ARG 4 end_ARG - caligraphic_C ) × ( divide start_ARG 3 end_ARG start_ARG 4 end_ARG + caligraphic_C ) end_ARG (32)

and

𝒬SP(K)=4c2K2[TTK(log(T)+ψ(Φ))+2cδK2ψ(Φ)]2T2TK4(14𝒞)×(34+𝒞)subscript𝒬𝑆𝑃𝐾4superscript𝑐2superscript𝐾2superscriptdelimited-[]𝑇subscript𝑇𝐾𝑇𝜓Φ2𝑐𝛿superscript𝐾2superscript𝜓Φ2superscript𝑇2superscriptsubscript𝑇𝐾414𝒞34𝒞\mathcal{Q}_{SP}\left(K\right)=\frac{4c^{2}K^{2}\left[T~{}T_{K}\left(\log{T}+% \psi\left(\Phi\right)\right)+2c~{}\delta K^{2}~{}\psi^{\prime}\left(\Phi\right% )\right]^{2}}{T^{2}~{}T_{K}^{4}~{}\left(\tfrac{1}{4}-\mathcal{C}\right)\times% \left(\tfrac{3}{4}+\mathcal{C}\right)}caligraphic_Q start_POSTSUBSCRIPT italic_S italic_P end_POSTSUBSCRIPT ( italic_K ) = divide start_ARG 4 italic_c start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_K start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT [ italic_T italic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT ( roman_log ( start_ARG italic_T end_ARG ) + italic_ψ ( roman_Φ ) ) + 2 italic_c italic_δ italic_K start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ψ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( roman_Φ ) ] start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_T start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT ( divide start_ARG 1 end_ARG start_ARG 4 end_ARG - caligraphic_C ) × ( divide start_ARG 3 end_ARG start_ARG 4 end_ARG + caligraphic_C ) end_ARG (33)

We present the corresponding exact QSNR results for the critical region in Fig. 4 using J=1𝐽1J=1italic_J = 1, for which we find from NRG that Kc0.618similar-to-or-equalssubscript𝐾𝑐0.618K_{c}\simeq 0.618italic_K start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ≃ 0.618, TK0.362similar-to-or-equalssubscript𝑇𝐾0.362T_{K}\simeq 0.362italic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT ≃ 0.362 and 𝒞0.385similar-to-or-equalssuperscript𝒞0.385\mathcal{C}^{*}\simeq-0.385caligraphic_C start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ≃ - 0.385. We note that although Eqs. (32) and (33) involve the exact correlator 𝒞𝒞\mathcal{C}caligraphic_C from Eq. (30), in the universal regime of interest where δK/TK102less-than-or-similar-to𝛿𝐾subscript𝑇𝐾superscript102\delta K/T_{K}\lesssim 10^{-2}italic_δ italic_K / italic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT ≲ 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT deviations of 𝒞𝒞\mathcal{C}caligraphic_C away from its critical value 𝒞superscript𝒞\mathcal{C}^{*}caligraphic_C start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT are tiny. Therefore our results are essentially indistinguishable if one replaces the functions 𝒞𝒞\mathcal{C}caligraphic_C in the denominator of the expressions for the QSNR with the constant 𝒞superscript𝒞\mathcal{C}^{*}caligraphic_C start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT.

Refer to caption
Figure 4: Single parameter estimation of T𝑇Titalic_T and K𝐾Kitalic_K in the universal quantum critical regime of the full 2IK model. (a) Exact analytic result for the critical thermometric QSNR from Eq. (32) as a function of rescaled temperature T/TK𝑇subscript𝑇𝐾T/T_{K}italic_T / italic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT and coupling detuning δK/TK𝛿𝐾subscript𝑇𝐾\delta K/T_{K}italic_δ italic_K / italic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT. (b) Corresponding QSNR for measuring the coupling constant K𝐾Kitalic_K, obtained from Eq. (33). (c,d) Maximum QSNR in the critical regime for each value of the detuning δK/TK𝛿𝐾subscript𝑇𝐾\delta K/T_{K}italic_δ italic_K / italic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT. Results obtained for J=1𝐽1J=1italic_J = 1.

Our results show a highly nontrivial evolution of measurement sensitivity in the quantum critical regime of this model. For thermometry, perhaps surprisingly, strong quantum critical correlations do not help with equilibrium measurement sensitivity at low temperatures. Physically, this is because in the region of parameter space described by the critical fixed point (TTTKmuch-less-thansuperscript𝑇𝑇much-less-thansubscript𝑇𝐾T^{*}\ll T\ll T_{K}italic_T start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ≪ italic_T ≪ italic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT with TδK2similar-tosuperscript𝑇𝛿superscript𝐾2T^{*}\sim\delta K^{2}italic_T start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ∼ italic_δ italic_K start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT), the value of the probe spin-spin correlation function is 𝒞𝒞similar-to-or-equals𝒞superscript𝒞\mathcal{C}\simeq\mathcal{C}^{*}caligraphic_C ≃ caligraphic_C start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT and has very little temperature dependence. Therefore T𝒞subscript𝑇𝒞\partial_{T}\mathcal{C}∂ start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT caligraphic_C is small and the corresponding QSNR is small. In particular, for δK=0𝛿𝐾0\delta K=0italic_δ italic_K = 0 at the critical point, there is no RG flow for TTKmuch-less-than𝑇subscript𝑇𝐾T\ll T_{K}italic_T ≪ italic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT and δT𝒞0similar-to-or-equalssubscript𝛿𝑇𝒞0\delta_{T}\mathcal{C}\simeq 0italic_δ start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT caligraphic_C ≃ 0, meaning that we have zero measurement sensitivity to temperature at the critical point below the Kondo temperature. This is confirmed in Fig. 4(a,c). This behavior is expected from the single-probe results of Ref. Mihailescu et al. (2023), where strong probe-environment entanglement hinders thermometric precision when measurements are only made on the probe. Likewise here, only a small amount of information about the environment temperature is imprinted on the probes when the probe-environment composite is in a strongly multipartite many-body entangled state. Indeed, in the small T𝑇Titalic_T and small δK𝛿𝐾\delta Kitalic_δ italic_K limit we extract the asymptotic behavior from our exact solution 𝒬SP(T)T4δK2/(aδK8+T4)similar-tosubscript𝒬𝑆𝑃𝑇superscript𝑇4𝛿superscript𝐾2𝑎𝛿superscript𝐾8superscript𝑇4\mathcal{Q}_{SP}(T)\sim T^{4}\delta K^{2}/(a\delta K^{8}+T^{4})caligraphic_Q start_POSTSUBSCRIPT italic_S italic_P end_POSTSUBSCRIPT ( italic_T ) ∼ italic_T start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT italic_δ italic_K start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / ( italic_a italic_δ italic_K start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT + italic_T start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT ) with a0.04similar-to-or-equals𝑎0.04a\simeq 0.04italic_a ≃ 0.04, which is strongly suppressed near the critical point. Good thermometric precision is instead obtained at much higher temperatures outside the critical window, where the probes act essentially as coupled thermal qubits.

On the other hand, measuring other model parameters such as the internal probe coupling constant K𝐾Kitalic_K is a very different story. This is because changing δK𝛿𝐾\delta Kitalic_δ italic_K induces a change in the scale Tsuperscript𝑇T^{*}italic_T start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT which pushes the system out of the critical window. The correlator 𝒞𝒞\mathcal{C}caligraphic_C then changes significantly over a narrow range of δK𝛿𝐾\delta Kitalic_δ italic_K near the critical point. As such, the derivative K𝒞subscript𝐾𝒞\partial_{K}\mathcal{C}∂ start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT caligraphic_C is strongly increased at lower temperatures and smaller δK𝛿𝐾\delta Kitalic_δ italic_K, where the critical window is narrower and small parameter changes have the largest effect. Indeed, as shown in Fig. 4(b,d) measurement sensitivity is in fact diverging rather than vanishing at the critical point, with 𝒬SP(K)log2(bT+δK2)similar-tosubscript𝒬𝑆𝑃𝐾superscript2𝑏𝑇𝛿superscript𝐾2\mathcal{Q}_{SP}(K)\sim\log^{2}(bT+\delta K^{2})caligraphic_Q start_POSTSUBSCRIPT italic_S italic_P end_POSTSUBSCRIPT ( italic_K ) ∼ roman_log start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_b italic_T + italic_δ italic_K start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) with b0.13similar-to-or-equals𝑏0.13b\simeq 0.13italic_b ≃ 0.13 at small T𝑇Titalic_T and δK𝛿𝐾\delta Kitalic_δ italic_K. Our exact results in the close vicinity of the critical point are consistent with the broader NRG results for the QSNR presented in Figs. 2 and 3. We note that the behavior discussed above should arise in any system in the Ising universality class Affleck et al. (1995) of boundary critical phenomena.

Finally we comment on the role of fractionalization phenomena for metrology. Away from the critical point of the 2IK model, all of the probe degrees of freedom are quenched at low enough temperatures – either by formation of a decoupled probe spin-singlet state for large KKcmuch-greater-than𝐾subscript𝐾𝑐K\gg K_{c}italic_K ≫ italic_K start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT or through the Kondo effect by environment-probe entanglement for KKcmuch-less-than𝐾subscript𝐾𝑐K\ll K_{c}italic_K ≪ italic_K start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. Good measurement precision is then afforded by overcoming the excitation gap so that the reduced probe state populations become sensitive to changes in the parameter of interest. In the critical regime however, we have a somewhat different story. Here, the probe degrees of freedom are only partially screened: at the critical point, a degree of freedom remains unscreened even down to zero temperature, due to the frustration driving the phase transition Affleck et al. (1995). In the critical regime, it is the fate of this residual critical degree of freedom that is responsible for metrological performance when one manipulates the system by changing temperature or model parameters. Remarkably, this degree of freedom in the 2IK model is not just a qubit spin or electron, but a fractionalized Majorana fermion Affleck et al. (1995); Gan (1995); Mitchell et al. (2012); Sela et al. (2011); Mitchell and Sela (2012), with an exotic 12ln(2)122\tfrac{1}{2}\ln(2)divide start_ARG 1 end_ARG start_ARG 2 end_ARG roman_ln ( start_ARG 2 end_ARG ) residual entropy – see Eq. (25). This Majorana is localized on the probe, and we measure it when we make measurements on the probe near the critical point. Instead of using qubits or electronic degrees of freedom for sensing, here we effectively leverage the unusual properties of the Majorana fermion when we do metrology near the 2IK critical point.

VII Multiparameter estimation

We now consider multiparameter estimation in the 2IK model for T𝑇Titalic_T and K𝐾Kitalic_K. The QFIM follows in this case from Eq. (11) with λ=(T,K)T𝜆superscript𝑇𝐾𝑇\vec{\lambda}=(T,K)^{T}over→ start_ARG italic_λ end_ARG = ( italic_T , italic_K ) start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT. However, since the probe reduced state populations ϱkϱk(𝒞)subscriptitalic-ϱ𝑘subscriptitalic-ϱ𝑘𝒞\varrho_{k}\equiv\varrho_{k}(\mathcal{C})italic_ϱ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ≡ italic_ϱ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( caligraphic_C ) in Eq. (15) are entirely determined by the probe spin-spin correlator 𝒞𝒞\mathcal{C}caligraphic_C, then Eq. (12) holds and the QFIM is singular. The 2IK model is therefore a prime example where the single- and multiparameter estimation schemes yield totally different results. For example, for the estimation of the temperature T𝑇Titalic_T, any uncertainty in K𝐾Kitalic_K completely collapses the QSNR, and vice versa. Multiparametric QFIM singularities are therefore crucial to identify in any practical setup.

VII.1 Control field

Here we demonstrate that when the QFIM is singular, metrological sensitivity can be recovered by applying a known control field. In the 2IK model, the singularity in the QFIM hindering multiparameter estimation is a consequence of the large SU(2) spin symmetry of the coupled spin-1212\tfrac{1}{2}divide start_ARG 1 end_ARG start_ARG 2 end_ARG qubit probes, which tightly constrains the properties of the probe reduced states. One might expect similar metrological problems in other systems with many conserved quantities (especially so for the class of integrable systems). For the 2IK model, breaking the SU(2) spin symmetry provides more flexibility and a route to precision parameter estimation. Here we focus on adding a known control field B𝐵Bitalic_B to our system, which we take to be just a simple Zeeman magnetic field oriented along the z𝑧zitalic_z-direction. The model therefore becomes,

H^=H^E+H^probe+J(𝐒^IL𝐬^EL+𝐒^IR𝐬^ER)+H^field^𝐻subscript^𝐻𝐸subscript^𝐻probe𝐽subscript^𝐒𝐼𝐿subscript^𝐬𝐸𝐿subscript^𝐒𝐼𝑅subscript^𝐬𝐸𝑅subscript^𝐻field\hat{H}=\hat{H}_{E}+\hat{H}_{\rm probe}+J~{}\left(\hat{\vec{\mathbf{S}}}_{IL}% \cdot\hat{\vec{\mathbf{s}}}_{EL}+\hat{\vec{\mathbf{S}}}_{IR}\cdot\hat{\vec{% \mathbf{s}}}_{ER}\right)+\hat{H}_{\rm field}over^ start_ARG italic_H end_ARG = over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT + over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT roman_probe end_POSTSUBSCRIPT + italic_J ( over^ start_ARG over→ start_ARG bold_S end_ARG end_ARG start_POSTSUBSCRIPT italic_I italic_L end_POSTSUBSCRIPT ⋅ over^ start_ARG over→ start_ARG bold_s end_ARG end_ARG start_POSTSUBSCRIPT italic_E italic_L end_POSTSUBSCRIPT + over^ start_ARG over→ start_ARG bold_S end_ARG end_ARG start_POSTSUBSCRIPT italic_I italic_R end_POSTSUBSCRIPT ⋅ over^ start_ARG over→ start_ARG bold_s end_ARG end_ARG start_POSTSUBSCRIPT italic_E italic_R end_POSTSUBSCRIPT ) + over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT roman_field end_POSTSUBSCRIPT (34)

with control field Hamiltonian H^field=BS^totzsubscript^𝐻field𝐵subscriptsuperscript^𝑆𝑧tot\hat{H}_{\rm field}=B\hat{S}^{z}_{\rm tot}over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT roman_field end_POSTSUBSCRIPT = italic_B over^ start_ARG italic_S end_ARG start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_tot end_POSTSUBSCRIPT, where the total spin projection is decomposed into α=L,R𝛼𝐿𝑅\alpha=L,Ritalic_α = italic_L , italic_R probe and environment parts S^totz=α(S^Iαz+s^Eα;totz)subscriptsuperscript^𝑆𝑧totsubscript𝛼superscriptsubscript^𝑆𝐼𝛼𝑧superscriptsubscript^𝑠𝐸𝛼tot𝑧\hat{S}^{z}_{\rm tot}=\sum_{\alpha}\left(\hat{S}_{I\alpha}^{z}+\hat{s}_{E% \alpha;{\rm tot}}^{z}\right)over^ start_ARG italic_S end_ARG start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_tot end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( over^ start_ARG italic_S end_ARG start_POSTSUBSCRIPT italic_I italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT + over^ start_ARG italic_s end_ARG start_POSTSUBSCRIPT italic_E italic_α ; roman_tot end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT ) with S^Iαzsubscriptsuperscript^𝑆𝑧𝐼𝛼\hat{S}^{z}_{I\alpha}over^ start_ARG italic_S end_ARG start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_I italic_α end_POSTSUBSCRIPT the spin projection for impurity probe α𝛼\alphaitalic_α and with s^Eα;totz=12k(cαkcαkcαkcαk)superscriptsubscript^𝑠𝐸𝛼tot𝑧12subscript𝑘superscriptsubscript𝑐𝛼𝑘absentsuperscriptsubscript𝑐𝛼𝑘absentabsentsuperscriptsubscript𝑐𝛼𝑘absentsuperscriptsubscript𝑐𝛼𝑘absentabsent\hat{s}_{E\alpha;{\rm tot}}^{z}=\tfrac{1}{2}\sum_{k}\left(c_{\alpha k\uparrow}% ^{\dagger}c_{\alpha k\uparrow}^{\phantom{\dagger}}-c_{\alpha k\downarrow}^{% \dagger}c_{\alpha k\downarrow}^{\phantom{\dagger}}\right)over^ start_ARG italic_s end_ARG start_POSTSUBSCRIPT italic_E italic_α ; roman_tot end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ∑ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_c start_POSTSUBSCRIPT italic_α italic_k ↑ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_c start_POSTSUBSCRIPT italic_α italic_k ↑ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT - italic_c start_POSTSUBSCRIPT italic_α italic_k ↓ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_c start_POSTSUBSCRIPT italic_α italic_k ↓ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT ) an operator for the total lead-α𝛼\alphaitalic_α (environment) spin projection. We note that the results for a field applied locally to the probe rather than globally to the full probe-environment system are qualitatively and quantitatively very similar.

For B0𝐵0B\neq 0italic_B ≠ 0 the full SU(2) spin symmetry is lifted, but a U(1) symmetry corresponding to conserved total Stotzsubscriptsuperscript𝑆𝑧totS^{z}_{\rm tot}italic_S start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_tot end_POSTSUBSCRIPT remains. The probe reduced density matrix retains its diagonal structure in the spin basis Eq. (15), but now the triplet populations are not equal, ϱT;1ϱT;0ϱT;+1subscriptitalic-ϱ𝑇1subscriptitalic-ϱ𝑇0subscriptitalic-ϱ𝑇1\varrho_{T;-1}\neq\varrho_{T;0}\neq\varrho_{T;+1}italic_ϱ start_POSTSUBSCRIPT italic_T ; - 1 end_POSTSUBSCRIPT ≠ italic_ϱ start_POSTSUBSCRIPT italic_T ; 0 end_POSTSUBSCRIPT ≠ italic_ϱ start_POSTSUBSCRIPT italic_T ; + 1 end_POSTSUBSCRIPT. In the full lead-coupled 2IK model, there is no simple exact relation between these reduced state populations. The probe spin-spin correlator 𝒞=𝐒^IL𝐒^IRTr(𝐒^IL𝐒^IRϱ^probe)𝒞delimited-⟨⟩subscript^𝐒𝐼𝐿subscript^𝐒𝐼𝑅tracesubscript^𝐒𝐼𝐿subscript^𝐒𝐼𝑅subscript^italic-ϱprobe\mathcal{C}\!=\!\langle\hat{\vec{\mathbf{S}}}_{IL}\cdot\hat{\vec{\mathbf{S}}}_% {IR}\rangle\!\equiv\!\Tr{\hat{\vec{\mathbf{S}}}_{IL}\cdot\hat{\vec{\mathbf{S}}% }_{IR}~{}\hat{\varrho}_{\rm probe}}caligraphic_C = ⟨ over^ start_ARG over→ start_ARG bold_S end_ARG end_ARG start_POSTSUBSCRIPT italic_I italic_L end_POSTSUBSCRIPT ⋅ over^ start_ARG over→ start_ARG bold_S end_ARG end_ARG start_POSTSUBSCRIPT italic_I italic_R end_POSTSUBSCRIPT ⟩ ≡ roman_Tr ( start_ARG over^ start_ARG over→ start_ARG bold_S end_ARG end_ARG start_POSTSUBSCRIPT italic_I italic_L end_POSTSUBSCRIPT ⋅ over^ start_ARG over→ start_ARG bold_S end_ARG end_ARG start_POSTSUBSCRIPT italic_I italic_R end_POSTSUBSCRIPT over^ start_ARG italic_ϱ end_ARG start_POSTSUBSCRIPT roman_probe end_POSTSUBSCRIPT end_ARG ) is no longer sufficient to fully determine the probe reduced state. We therefore introduce two other physically-motivated and experimentally-feasible observables to fully characterize the probe: the probe magnetization =S^ILz+S^IRzTr((S^ILz+S^IRz)ϱ^probe)delimited-⟨⟩subscriptsuperscript^𝑆𝑧𝐼𝐿subscriptsuperscript^𝑆𝑧𝐼𝑅tracesuperscriptsubscript^𝑆𝐼𝐿𝑧superscriptsubscript^𝑆𝐼𝑅𝑧subscript^italic-ϱprobe\mathcal{M}=\langle\hat{S}^{z}_{IL}+\hat{S}^{z}_{IR}\rangle\equiv\Tr{\left(% \hat{S}_{IL}^{z}+\hat{S}_{IR}^{z}\right)~{}\hat{\varrho}_{\rm probe}}caligraphic_M = ⟨ over^ start_ARG italic_S end_ARG start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_I italic_L end_POSTSUBSCRIPT + over^ start_ARG italic_S end_ARG start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_I italic_R end_POSTSUBSCRIPT ⟩ ≡ roman_Tr ( start_ARG ( over^ start_ARG italic_S end_ARG start_POSTSUBSCRIPT italic_I italic_L end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT + over^ start_ARG italic_S end_ARG start_POSTSUBSCRIPT italic_I italic_R end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT ) over^ start_ARG italic_ϱ end_ARG start_POSTSUBSCRIPT roman_probe end_POSTSUBSCRIPT end_ARG ) and χ=(S^ILz+S^IRz)2Tr((S^ILz+S^IRz)2ϱ^probe)𝜒delimited-⟨⟩superscriptsubscriptsuperscript^𝑆𝑧𝐼𝐿subscriptsuperscript^𝑆𝑧𝐼𝑅2tracesuperscriptsuperscriptsubscript^𝑆𝐼𝐿𝑧superscriptsubscript^𝑆𝐼𝑅𝑧2subscript^italic-ϱprobe\mathcal{\chi}=\langle(\hat{S}^{z}_{IL}+\hat{S}^{z}_{IR})^{2}\rangle\equiv\Tr{% \left(\hat{S}_{IL}^{z}+\hat{S}_{IR}^{z}\right)^{2}~{}\hat{\varrho}_{\rm probe}}italic_χ = ⟨ ( over^ start_ARG italic_S end_ARG start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_I italic_L end_POSTSUBSCRIPT + over^ start_ARG italic_S end_ARG start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_I italic_R end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ ≡ roman_Tr ( start_ARG ( over^ start_ARG italic_S end_ARG start_POSTSUBSCRIPT italic_I italic_L end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT + over^ start_ARG italic_S end_ARG start_POSTSUBSCRIPT italic_I italic_R end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT over^ start_ARG italic_ϱ end_ARG start_POSTSUBSCRIPT roman_probe end_POSTSUBSCRIPT end_ARG ) which is related to the probe magnetic susceptibility.

Since the probe reduced state eigenvalues ϱkϱk(𝒞,,χ)subscriptitalic-ϱ𝑘subscriptitalic-ϱ𝑘𝒞𝜒\varrho_{k}\equiv\varrho_{k}(\mathcal{C},\mathcal{M},\mathcal{\chi})italic_ϱ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ≡ italic_ϱ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( caligraphic_C , caligraphic_M , italic_χ ) are fully determined by these observables, the QFIM is clearly also a function of these quantities. Importantly, the factorized form of Eq. (12) no longer applies, and the QFIM singularity is removed due to the known control field. We find that,

ϱS=14𝒞,ϱT;0=34+𝒞χ,ϱT;1=12[χ],ϱT;+1=12[χ+].\begin{split}&\varrho_{S}={}\frac{1}{4}-\mathcal{C}\;,~{}~{}~{}~{}~{}~{}~{}~{}% ~{}~{}~{}~{}~{}~{}~{}\varrho_{T;0}={}\frac{3}{4}+\mathcal{C}-\mathcal{\chi}\;,% \\ &\varrho_{T;-1}={}\frac{1}{2}\left[\mathcal{\chi}-\mathcal{M}\right]\;,~{}~{}~% {}\varrho_{T;+1}={}\frac{1}{2}\left[\mathcal{\chi}+\mathcal{M}\right]\;.\end{split}start_ROW start_CELL end_CELL start_CELL italic_ϱ start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 4 end_ARG - caligraphic_C , italic_ϱ start_POSTSUBSCRIPT italic_T ; 0 end_POSTSUBSCRIPT = divide start_ARG 3 end_ARG start_ARG 4 end_ARG + caligraphic_C - italic_χ , end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL italic_ϱ start_POSTSUBSCRIPT italic_T ; - 1 end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 2 end_ARG [ italic_χ - caligraphic_M ] , italic_ϱ start_POSTSUBSCRIPT italic_T ; + 1 end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 2 end_ARG [ italic_χ + caligraphic_M ] . end_CELL end_ROW (35)

We may now use Eq. (11) to obtain the QFIM in terms of this set of observables,

𝓗λ,λ=λ(χ+)λ(χ+)2(χ+)+λ(χ)λ(χ)2(χ)+λ(𝒞χ)λ(𝒞χ)3/4+𝒞χ+λ𝒞λ𝒞1/4𝒞subscript𝓗𝜆superscript𝜆subscript𝜆𝜒subscriptsuperscript𝜆𝜒2𝜒subscript𝜆𝜒subscriptsuperscript𝜆𝜒2𝜒subscript𝜆𝒞𝜒subscriptsuperscript𝜆𝒞𝜒34𝒞𝜒subscript𝜆𝒞subscriptsuperscript𝜆𝒞14𝒞\begin{split}\boldsymbol{\mathcal{H}}_{\lambda,\lambda^{\prime}}=&\frac{% \partial_{\lambda}(\mathcal{\chi}+\mathcal{M})\partial_{\lambda^{\prime}}(% \mathcal{\chi}+\mathcal{M})}{2(\mathcal{\chi}+\mathcal{M})}+\frac{\partial_{% \lambda}(\mathcal{\chi}-\mathcal{M})\partial_{\lambda^{\prime}}(\mathcal{\chi}% -\mathcal{M})}{2(\mathcal{\chi}-\mathcal{M})}\\ &+\frac{\partial_{\lambda}(\mathcal{C}-\mathcal{\chi})\partial_{\lambda^{% \prime}}(\mathcal{C}-\mathcal{\chi})}{3/4+\mathcal{C}-\mathcal{\chi}}+\frac{% \partial_{\lambda}\mathcal{C}\partial_{\lambda^{\prime}}\mathcal{C}}{1/4-% \mathcal{C}}\end{split}start_ROW start_CELL bold_caligraphic_H start_POSTSUBSCRIPT italic_λ , italic_λ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT = end_CELL start_CELL divide start_ARG ∂ start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ( italic_χ + caligraphic_M ) ∂ start_POSTSUBSCRIPT italic_λ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ( italic_χ + caligraphic_M ) end_ARG start_ARG 2 ( italic_χ + caligraphic_M ) end_ARG + divide start_ARG ∂ start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ( italic_χ - caligraphic_M ) ∂ start_POSTSUBSCRIPT italic_λ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ( italic_χ - caligraphic_M ) end_ARG start_ARG 2 ( italic_χ - caligraphic_M ) end_ARG end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL + divide start_ARG ∂ start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ( caligraphic_C - italic_χ ) ∂ start_POSTSUBSCRIPT italic_λ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ( caligraphic_C - italic_χ ) end_ARG start_ARG 3 / 4 + caligraphic_C - italic_χ end_ARG + divide start_ARG ∂ start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT caligraphic_C ∂ start_POSTSUBSCRIPT italic_λ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT caligraphic_C end_ARG start_ARG 1 / 4 - caligraphic_C end_ARG end_CELL end_ROW (36)

In general, a finite magnetization 00\mathcal{M}\neq 0caligraphic_M ≠ 0 is enough to remove the QFIM singularity. However, we note in passing that if the field B𝐵Bitalic_B is also unknown, then the 3×3333\times 33 × 3 QFIM for T,K,B𝑇𝐾𝐵T,K,Bitalic_T , italic_K , italic_B can once again become singular. The added information about the known field B𝐵Bitalic_B is crucial to this multiparameter estimation protocol.

In Appendix D we provide an approximate ansatz for the probe reduced state populations, which holds exactly in the large-K𝐾Kitalic_K and small B𝐵Bitalic_B limits, and is generally found to be rather accurate throughout the phase diagram. The simplification allows us to express the QFIM in terms of only 𝒞𝒞\mathcal{C}caligraphic_C and \mathcal{M}caligraphic_M. This might be advantageous in an experimental setting since simultaneous measurements of different observables may be challenging in practice. In principle, since the SLD operators for T𝑇Titalic_T and K𝐾Kitalic_K commute, there exists a single optimal measurement basis (effectively the Bell basis on the probe states). However, this is not a physically meaningful or feasibly measurable observable.

VII.2 Multiparameter estimation in the large-K𝐾Kitalic_K limit

As a simple but nontrivial demonstration of multiparameter estimation, we consider now the large-K𝐾Kitalic_K limit of the 2IK model with an applied control field B𝐵Bitalic_B, focusing our discussion on the multiparameter QSNR 𝒬MP(λ,λ)subscript𝒬𝑀𝑃𝜆𝜆\mathcal{Q}_{MP}\left(\lambda,\lambda\right)caligraphic_Q start_POSTSUBSCRIPT italic_M italic_P end_POSTSUBSCRIPT ( italic_λ , italic_λ ) for λ=T𝜆𝑇\lambda=Titalic_λ = italic_T or K𝐾Kitalic_K.

By way of comparison, we consider first the analogous single parameter QSNRs 𝒬SP(λ)subscript𝒬𝑆𝑃𝜆\mathcal{Q}_{SP}\left(\lambda\right)caligraphic_Q start_POSTSUBSCRIPT italic_S italic_P end_POSTSUBSCRIPT ( italic_λ ), shown in Fig. 5(a,c). See Fig. 2(c,d) for the corresponding zero field case discussed already. The impact of the field itself on single-parameter estimation is discussed in depth in Appendix E and F.

Refer to caption
Figure 5: Multiparameter estimation in the large-K𝐾Kitalic_K limit of the 2IK model. (a) Single-parameter and (b) multiparameter QSNRs for thermometery. (c) Single-parameter and (d) multiparameter QSNRs for estimation of the coupling K𝐾Kitalic_K. The multiparameter QSNR 𝒬MP(λ,λ)subscript𝒬𝑀𝑃𝜆𝜆\mathcal{Q}_{MP}(\lambda,\lambda)caligraphic_Q start_POSTSUBSCRIPT italic_M italic_P end_POSTSUBSCRIPT ( italic_λ , italic_λ ) is a universal scaling function of T/B𝑇𝐵T/Bitalic_T / italic_B and K/B𝐾𝐵K/Bitalic_K / italic_B, such that the regime of good sensitivity shrinks as the field strength is reduced, and vanishes in the limit B0𝐵0B\to 0italic_B → 0.

From Eq. (8) we see that the QSNR for either T𝑇Titalic_T or K𝐾Kitalic_K, in the case where neither T𝑇Titalic_T or K𝐾Kitalic_K are known with certainty, involves information from the full QFIM and not just its diagonal components. The behavior of the QFIM elements themselves are discussed in Appendix G. We note that 𝒬MP(λ,λ)subscript𝒬𝑀𝑃𝜆𝜆\mathcal{Q}_{MP}(\lambda,\lambda)caligraphic_Q start_POSTSUBSCRIPT italic_M italic_P end_POSTSUBSCRIPT ( italic_λ , italic_λ ) and the quantity B2𝓗λ,λsuperscript𝐵2subscript𝓗𝜆𝜆B^{2}\boldsymbol{\mathcal{H}}_{\lambda,\lambda}italic_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT bold_caligraphic_H start_POSTSUBSCRIPT italic_λ , italic_λ end_POSTSUBSCRIPT are universal scaling functions of the rescaled parameters t=T/B𝑡𝑇𝐵t=T/Bitalic_t = italic_T / italic_B and k=K/B𝑘𝐾𝐵k=K/Bitalic_k = italic_K / italic_B in the large-K𝐾Kitalic_K limit of the 2IK model. For example,

𝒬MP(T,T)=2e2/t(2+cosh(1/t))(1+e1/t+e2/t)(1+e1/t(1+e1/t+ek/t))t2subscript𝒬𝑀𝑃𝑇𝑇2superscript𝑒2𝑡21𝑡1superscript𝑒1𝑡superscript𝑒2𝑡1superscript𝑒1𝑡1superscript𝑒1𝑡superscript𝑒𝑘𝑡superscript𝑡2\mathcal{Q}_{MP}\left(T,T\right)=\frac{2e^{2/t}\left(2+\cosh{1/t}\right)}{% \left(1+e^{1/t}+e^{2/t}\right)\left(1+e^{1/t}\left(1+e^{1/t}+e^{k/t}\right)% \right)t^{2}}caligraphic_Q start_POSTSUBSCRIPT italic_M italic_P end_POSTSUBSCRIPT ( italic_T , italic_T ) = divide start_ARG 2 italic_e start_POSTSUPERSCRIPT 2 / italic_t end_POSTSUPERSCRIPT ( 2 + roman_cosh ( start_ARG 1 / italic_t end_ARG ) ) end_ARG start_ARG ( 1 + italic_e start_POSTSUPERSCRIPT 1 / italic_t end_POSTSUPERSCRIPT + italic_e start_POSTSUPERSCRIPT 2 / italic_t end_POSTSUPERSCRIPT ) ( 1 + italic_e start_POSTSUPERSCRIPT 1 / italic_t end_POSTSUPERSCRIPT ( 1 + italic_e start_POSTSUPERSCRIPT 1 / italic_t end_POSTSUPERSCRIPT + italic_e start_POSTSUPERSCRIPT italic_k / italic_t end_POSTSUPERSCRIPT ) ) italic_t start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG (37)

and similarly for 𝒬MP(K,K)subscript𝒬𝑀𝑃𝐾𝐾\mathcal{Q}_{MP}\left(K,K\right)caligraphic_Q start_POSTSUBSCRIPT italic_M italic_P end_POSTSUBSCRIPT ( italic_K , italic_K ). We present these universal results as a function of t𝑡titalic_t and k𝑘kitalic_k in Fig. 5. Interestingly, 𝒬MP(λ,λ)subscript𝒬𝑀𝑃𝜆𝜆\mathcal{Q}_{MP}(\lambda,\lambda)caligraphic_Q start_POSTSUBSCRIPT italic_M italic_P end_POSTSUBSCRIPT ( italic_λ , italic_λ ) and B2𝓗λ,λsuperscript𝐵2subscript𝓗𝜆𝜆B^{2}\boldsymbol{\mathcal{H}}_{\lambda,\lambda}italic_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT bold_caligraphic_H start_POSTSUBSCRIPT italic_λ , italic_λ end_POSTSUBSCRIPT are also universal functions of t𝑡titalic_t and k𝑘kitalic_k in the full 2IK model provided all model parameters (in this case j=J/B𝑗𝐽𝐵j=J/Bitalic_j = italic_J / italic_B and d=D/B𝑑𝐷𝐵d=D/Bitalic_d = italic_D / italic_B) are similarly scaled, although of course many-body effects change the functional form.

Turning now to our numerical results, we see a stark difference between the multiparametric QSNR for thermometry 𝒬MP(T,T)subscript𝒬𝑀𝑃𝑇𝑇\mathcal{Q}_{MP}\left(T,T\right)caligraphic_Q start_POSTSUBSCRIPT italic_M italic_P end_POSTSUBSCRIPT ( italic_T , italic_T ) shown in Fig. 5(b) compared with its single-parameter estimation counterpart 𝒬SP(T)subscript𝒬𝑆𝑃𝑇\mathcal{Q}_{SP}\left(T\right)caligraphic_Q start_POSTSUBSCRIPT italic_S italic_P end_POSTSUBSCRIPT ( italic_T ) shown in Fig. 5(a). Uncertainty in K𝐾Kitalic_K dramatically affects our ability to estimate T𝑇Titalic_T. In particular, 𝒬MP(T,T)subscript𝒬𝑀𝑃𝑇𝑇\mathcal{Q}_{MP}\left(T,T\right)caligraphic_Q start_POSTSUBSCRIPT italic_M italic_P end_POSTSUBSCRIPT ( italic_T , italic_T ) is small for all K>B𝐾𝐵K>Bitalic_K > italic_B and rapidly attenuates to zero at large K𝐾Kitalic_K. Appreciable sensitivity to T𝑇Titalic_T is instead afforded for K<B𝐾𝐵K<Bitalic_K < italic_B because here the triplet state |T;1ket𝑇1|T;-1\rangle| italic_T ; - 1 ⟩ is the ground state, and excited states |T;0ket𝑇0|T;0\rangle| italic_T ; 0 ⟩ and |T;+1ket𝑇1|T;+1\rangle| italic_T ; + 1 ⟩ are respectively B𝐵Bitalic_B and 2B2𝐵2B2 italic_B higher in energy. Since the control field strength B𝐵Bitalic_B is known, thermometric sensitivity is boosted at temperatures on the order of these excitation energies, since then the rate of population transfer is maximal. There is effectively no sensitivity at temperatures below the minimum gap TBmuch-less-than𝑇𝐵T\ll Bitalic_T ≪ italic_B. The singlet state |Sket𝑆|S\rangle| italic_S ⟩ is of order |BK|𝐵𝐾|B-K|| italic_B - italic_K | higher in energy, but K𝐾Kitalic_K is unknown. Therefore it is the triplet excitations which dominate thermometry in this system. This behavior is not evident from simply looking at the QFI for T𝑇Titalic_T, i.e. the element 𝓗T,Tsubscript𝓗𝑇𝑇\boldsymbol{\mathcal{H}}_{T,T}bold_caligraphic_H start_POSTSUBSCRIPT italic_T , italic_T end_POSTSUBSCRIPT in the QFIM, as discussed in Appendix G.

We show the complementary behavior when estimating the coupling K𝐾Kitalic_K in the presence of uncertainty in the temperature T𝑇Titalic_T in Fig. 5(d) where we see the QSNR 𝒬MP(K,K)subscript𝒬𝑀𝑃𝐾𝐾\mathcal{Q}_{MP}\left(K,K\right)caligraphic_Q start_POSTSUBSCRIPT italic_M italic_P end_POSTSUBSCRIPT ( italic_K , italic_K ) flares sharply for KBsimilar-to𝐾𝐵K\sim Bitalic_K ∼ italic_B at the level crossing point between ground states, thus showing a more consistent behavior with the single-parameter estimation shown in panel (c). The lack of low-temperature thermometric sensitivity at K=B𝐾𝐵K=Bitalic_K = italic_B means that estimation of K𝐾Kitalic_K near KBsimilar-to𝐾𝐵K\sim Bitalic_K ∼ italic_B at low temperatures is also poor.

VII.3 Multiparameter estimation near criticality

Richer behavior is naturally expected in the full 2IK model near the quantum critical point. The magnetic field B𝐵Bitalic_B is a relevant perturbation that also destabilizes the critical point. For small B𝐵Bitalic_B in the close vicinity of the critical point, the physics is again controlled by a single emergent scale Tsuperscript𝑇T^{*}italic_T start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT, to which all relevant perturbations contribute additively Sela et al. (2011); Mitchell and Sela (2012), TcKδK2+cBB2similar-tosuperscript𝑇subscript𝑐𝐾𝛿superscript𝐾2subscript𝑐𝐵superscript𝐵2T^{*}\sim c_{K}\delta K^{2}+c_{B}B^{2}italic_T start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ∼ italic_c start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT italic_δ italic_K start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_c start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT with constants cKsubscript𝑐𝐾c_{K}italic_c start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT and cBsubscript𝑐𝐵c_{B}italic_c start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT of order TK1superscriptsubscript𝑇𝐾1T_{K}^{-1}italic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT. This expression replaces Eq. (24), but Eqs. (25), (26) still hold. Therefore similar methods leading to Eqs. (32), (33) may be applied to obtain exact results for the multiparameter QSNRs. The generalization requires expressions for \mathcal{M}caligraphic_M and χ𝜒\mathcal{\chi}italic_χ, but these can again be obtained from the entropy via the appropriate Maxwell relations by noting that =Bsubscript𝐵\mathcal{M}=\partial_{B}\mathcal{F}caligraphic_M = ∂ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT caligraphic_F and χ=B2𝜒superscriptsubscript𝐵2\mathcal{\chi}=\partial_{B}^{2}\mathcal{F}italic_χ = ∂ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT caligraphic_F. We leave detailed calculations for future work.

In the full 2IK model our basic conclusion remains unchanged: multiparameter estimation is much more complex and subtle than its single-parameter counterpart. In particular, it is important to understand when a system might have a singular QFIM and how to remove that singularity. It is also crucial in assessing metrological performance to examine the QSNR rather than simply elements of the QFIM, which can paint a misleading picture in practice.

VIII Conclusions and Discussions

Quantum criticality is a well-known resource for quantum metrology at zero temperature. In practice, however, thermal fluctuations are inevitable and even the temperature might itself be unknown. In this work, we considered the 2IK model as a rich playground to explore finite-temperature multiparameter estimation in an interacting quantum many-body system that supports a nontrivial second-order quantum phase transition. We regard the two spin-1212\tfrac{1}{2}divide start_ARG 1 end_ARG start_ARG 2 end_ARG impurities as in-situ probes for the continuum electronic environment in which they are embedded. Quantum sensing is performed by measuring physically-motivated observables on the accessible impurity qubit probes. The 2IK model and its variants can be realized in nanoelectronics quantum dot devices. Experimental signatures of finite-temperature quantum criticality in related Kondo systems have been recently reported Potok et al. (2007); Iftikhar et al. (2018); Pouse et al. (2023), but their metrological capacity was not addressed.

Through the NRG method, we have computed the finite-temperature reduced density matrix of the two impurities as our probe state, and extracted from it elements of the QFIM for estimation of the system temperature T𝑇Titalic_T and the probe coupling strength K𝐾Kitalic_K. Interestingly, information on T𝑇Titalic_T and K𝐾Kitalic_K appear only in the eigenvalues of the density matrix, which makes the optimal measurement basis independent of these parameters. As a novel figure of merit for quantifying the performance of the probe, we introduced the multiparameter QSNR.

Remarkably, we were able to find exact analytic results for the QSNRs near the 2IK critical point at finite temperatures by relating them to physical observables on the probe, which were in turn obtained via Maxwell relations from an exact solution for the probe entropy.

In the context of single parameter estimation, our analysis shows that one can get strongly enhanced sensitivity (i.e. a large QSNR which diverges as T0𝑇0T\to 0italic_T → 0) for the coupling K𝐾Kitalic_K around the critical point at Kcsubscript𝐾𝑐K_{c}italic_K start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT and up to the Kondo temperature TKsubscript𝑇𝐾T_{K}italic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT. High quality sensitivity to K𝐾Kitalic_K is also achieved outside the critical regime for TTKmuch-greater-than𝑇subscript𝑇𝐾T\gg T_{K}italic_T ≫ italic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT – albeit comparatively diminished with respect to achievable sensitivity near the critical point. This shows that criticality-based quantum sensors can indeed be beneficial even at finite temperatures. We also demonstrated that the probe can act as an effective thermometer, however only when the temperature T𝑇Titalic_T is larger than the Kondo temperature TKsubscript𝑇𝐾T_{K}italic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT.

In the context of two-parameter estimation, where both the impurity coupling and the temperature are unknown, we find that the QFI matrix becomes singular, thus preventing estimation of either parameter. This is a counterintuitive result since QFI matrix singularities are typically associated with the interdependence of parameters to be estimated. Here by contrast, the parameters T𝑇Titalic_T and K𝐾Kitalic_K are strictly independent, yet the QFI matrix is still singular. The reason for this is the partial accessibility to the system, which restricts the probe state to only the reduced density matrix of the two impurities. In order to restore metrological sensitivity, we show that simply applying a known control field is sufficient to remove the QFI matrix singularity. Thanks to this simple recipe, the QFI matrix becomes invertible and simultaneous two parameter estimation becomes possible. Our analysis shows that the precision of estimating any parameter is enhanced if the other parameters are known. If this is not possible, the precision can still be strongly enhanced even if another parameter is unknown, but its QFI is large. Since uncertainties are unavoidable in any practical situation, this provides a conceptual route to the design of optimized and robust quantum sensors: one should seek to maximize the QFI of all model parameters – not just the one directly of interest for metrology.

Our results provide important insights for probe-based parameter estimation of complex quantum systems, in particular highlighting the importance of the QSNR as a key figure of merit to accurately assess a system’s metrological utility. The 2IK model provides a versatile setting to demonstrate our results; however we expect that similar effects will be relevant to other candidate systems for critical quantum metrology.

Refer to caption
Figure 6: Single-parameter estimation of T𝑇Titalic_T and K𝐾Kitalic_K in the narrow band limit of the 2IK model. Panels a-b: QFI for thermometry (T)𝑇\mathcal{H}\left(T\right)caligraphic_H ( italic_T ) and for the estimation of probe coupling strength (K)𝐾\mathcal{H}\left(K\right)caligraphic_H ( italic_K ) as a function of T𝑇Titalic_T and K𝐾Kitalic_K. Panels c-d: Corresponding single parameter QSNR responses 𝒬SP(T)subscript𝒬𝑆𝑃𝑇\mathcal{Q}_{SP}\left(T\right)caligraphic_Q start_POSTSUBSCRIPT italic_S italic_P end_POSTSUBSCRIPT ( italic_T ) and 𝒬SP(K)subscript𝒬𝑆𝑃𝐾\mathcal{Q}_{SP}\left(K\right)caligraphic_Q start_POSTSUBSCRIPT italic_S italic_P end_POSTSUBSCRIPT ( italic_K ) as a function of T𝑇Titalic_T and K𝐾Kitalic_K. Results obtained by exact diagonalization.
Acknowledgements.
We acknowledge fruitful exchanges with Mark Mitchison and Gabriel Landi. GM acknowledges support from Equal1 Laboratories Ireland Limited. SC is supported by the Science Foundation Ireland Starting Investigator Research Grant “SpeedDemon” No. 18/SIRG/5508, the John Templeton Foundation Grant ID 62422, and the Alexander von Humboldt Foundation. AKM acknowledges funding from the Irish Research Council through grant EPSPG/2022/59 and from Science Foundation Ireland through grant 21/RP-2TF/10019. AB acknowledges support from the National Natural Science Foundation of China (Grants Nos. 12050410253, 92065115, and 12274059) and the Ministry of Science and Technology of China (Grant No. QNJ2021167001L).

Appendix A Uncertainties in multiparameter estimation

One can multiply both sides of the multiparameter Cramér-Rao inequality inequality, given in Eq. (4), by a positive weight matrix W0𝑊0W\geq 0italic_W ≥ 0 and take the trace of both sides to get an inequality for scalar quantities, namely

Tr[WCov[λ]]1NTr[W𝓗1].Trdelimited-[]𝑊Covdelimited-[]𝜆1𝑁Trdelimited-[]𝑊superscript𝓗1\textrm{Tr}\left[W\textbf{Cov}\left[\vec{\lambda}\right]\right]\geq\frac{1}{N}% \textrm{Tr}\left[W\boldsymbol{\mathcal{H}}^{-1}\right].Tr [ italic_W Cov [ over→ start_ARG italic_λ end_ARG ] ] ≥ divide start_ARG 1 end_ARG start_ARG italic_N end_ARG Tr [ italic_W bold_caligraphic_H start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ] . (38)

If one is only interested in estimating one of the parameters, i.e. λisubscript𝜆𝑖\lambda_{i}italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, then we can choose W=diag(0,,0,1,0,,0)𝑊diag00100W=\textrm{diag}(0,\cdots,0,1,0,\cdots,0)italic_W = diag ( 0 , ⋯ , 0 , 1 , 0 , ⋯ , 0 ), where only Wii=1subscript𝑊𝑖𝑖1W_{ii}=1italic_W start_POSTSUBSCRIPT italic_i italic_i end_POSTSUBSCRIPT = 1. In this case the Eq. (38) becomes Var(λi)1N(𝓗1)iiVarsubscript𝜆𝑖1𝑁subscriptsuperscript𝓗1𝑖𝑖\textrm{Var}(\lambda_{i})\geq\frac{1}{N}(\boldsymbol{\mathcal{H}}^{-1})_{ii}Var ( italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) ≥ divide start_ARG 1 end_ARG start_ARG italic_N end_ARG ( bold_caligraphic_H start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ) start_POSTSUBSCRIPT italic_i italic_i end_POSTSUBSCRIPT. Note that here the other parameters are assumed to be unknown. In the case that all other parameters are known, the problem reduces to a single parameter estimation in which Var(λi)1N(𝓗ii)1Varsubscript𝜆𝑖1𝑁superscriptsubscript𝓗𝑖𝑖1\textrm{Var}(\lambda_{i})\geq\frac{1}{N}(\boldsymbol{\mathcal{H}}_{ii})^{-1}Var ( italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) ≥ divide start_ARG 1 end_ARG start_ARG italic_N end_ARG ( bold_caligraphic_H start_POSTSUBSCRIPT italic_i italic_i end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT. The QFI matrix is positive semi-definite, i.e. 𝓗0𝓗0\boldsymbol{\mathcal{H}}\geq 0bold_caligraphic_H ≥ 0, which implies that (𝓗1)ii(𝓗ii)1subscriptsuperscript𝓗1𝑖𝑖superscriptsubscript𝓗𝑖𝑖1(\boldsymbol{\mathcal{H}}^{-1})_{ii}\geq(\boldsymbol{\mathcal{H}}_{ii})^{-1}( bold_caligraphic_H start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ) start_POSTSUBSCRIPT italic_i italic_i end_POSTSUBSCRIPT ≥ ( bold_caligraphic_H start_POSTSUBSCRIPT italic_i italic_i end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT. This clearly shows that the uncertainty in estimating λisubscript𝜆𝑖\lambda_{i}italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT reduces if one knows other parameters of the system.

Refer to caption
Figure 7: Exact analytic results for the spin-spin correlation function 𝒞𝒞\mathcal{C}caligraphic_C and its derivatives T𝒞subscript𝑇𝒞\partial_{T}\mathcal{C}∂ start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT caligraphic_C and K𝒞subscript𝐾𝒞\partial_{K}\mathcal{C}∂ start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT caligraphic_C (colored lines), compared with numerically-exact NRG results (black dashed lines) in the universal quantum critical regime of the full 2IK model. Plotted for J=1𝐽1J=1italic_J = 1.

Appendix B Narrow band limit

In Sec. IV of the main text, we consider an analytically tractable limit of the full 2IK system where the fermionic leads play no role in determining the reduced state of the probe. Another important limit is that arising when J/D1much-greater-than𝐽𝐷1J/D\gg 1italic_J / italic_D ≫ 1 (with D𝐷Ditalic_D the conduction electron bandwidth), wherein the continuum of electronic states in the leads can be well-approximated by just the single orbital in real-space to which each impurity couples (previously denoted c^α0νsubscript^𝑐𝛼0𝜈\hat{c}_{\alpha 0\nu}over^ start_ARG italic_c end_ARG start_POSTSUBSCRIPT italic_α 0 italic_ν end_POSTSUBSCRIPT). This is known as the ‘narrow band limit’ (NBL), and is described by the Hamiltonian,

H^NBL=K𝐒^IL𝐒^IR+J(𝐒^IL𝐬^EL+𝐒^IR𝐬^ER)subscript^𝐻𝑁𝐵𝐿𝐾subscript^𝐒𝐼𝐿subscript^𝐒𝐼𝑅𝐽subscript^𝐒𝐼𝐿subscript^𝐬𝐸𝐿subscript^𝐒𝐼𝑅subscript^𝐬𝐸𝑅\hat{H}_{NBL}=K~{}\hat{\vec{\mathbf{S}}}_{IL}\cdot\hat{\vec{\mathbf{S}}}_{IR}+% J~{}\left(\hat{\vec{\mathbf{S}}}_{IL}\cdot\hat{\vec{\mathbf{s}}}_{EL}+\hat{% \vec{\mathbf{S}}}_{IR}\cdot\hat{\vec{\mathbf{s}}}_{ER}\right)over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT italic_N italic_B italic_L end_POSTSUBSCRIPT = italic_K over^ start_ARG over→ start_ARG bold_S end_ARG end_ARG start_POSTSUBSCRIPT italic_I italic_L end_POSTSUBSCRIPT ⋅ over^ start_ARG over→ start_ARG bold_S end_ARG end_ARG start_POSTSUBSCRIPT italic_I italic_R end_POSTSUBSCRIPT + italic_J ( over^ start_ARG over→ start_ARG bold_S end_ARG end_ARG start_POSTSUBSCRIPT italic_I italic_L end_POSTSUBSCRIPT ⋅ over^ start_ARG over→ start_ARG bold_s end_ARG end_ARG start_POSTSUBSCRIPT italic_E italic_L end_POSTSUBSCRIPT + over^ start_ARG over→ start_ARG bold_S end_ARG end_ARG start_POSTSUBSCRIPT italic_I italic_R end_POSTSUBSCRIPT ⋅ over^ start_ARG over→ start_ARG bold_s end_ARG end_ARG start_POSTSUBSCRIPT italic_E italic_R end_POSTSUBSCRIPT ) (39)

The competition between J𝐽Jitalic_J and K𝐾Kitalic_K scales is captured in the NBL, although many-body effects (such as Kondo and the quantum phase transition) are missing. The two-impurity probe RDM is still a function of the single observable 𝒞𝒞\mathcal{C}caligraphic_C, but it is no longer just a function of the single reduced parameter K/T𝐾𝑇K/Titalic_K / italic_T since we also now have a nontrivial dependence on J/T𝐽𝑇J/Titalic_J / italic_T. These properties have implications for metrology, as shown below.

The NBL, which features just one electronic mode in each bath as per Eq. (39), is nonetheless a valid description of the system when JDmuch-greater-than𝐽𝐷J\gg Ditalic_J ≫ italic_D. This limit has the advantage that it is simple enough to be analytically tractable, like the large-K𝐾Kitalic_K limit (although in practice we do not present closed-form formulae here since they are extremely cumbersome). For concreteness in this section we set J=1𝐽1J=1italic_J = 1 and explore the (T,K)𝑇𝐾(T,K)( italic_T , italic_K ) parameter space as before.

We stress that, although the physical setup includes the explicit addition of environment degrees of freedom, the two spin-1212\tfrac{1}{2}divide start_ARG 1 end_ARG start_ARG 2 end_ARG impurity composite probe is still fully characterised by the spin-spin correlation function 𝒞𝒞\mathcal{C}caligraphic_C, and metrological properties can be extracted from Eq. (19). We note that explicit inclusion of the impurity-environment coupling can significantly alter the physics of the system, as manifest in the behavior of 𝒞𝒞\mathcal{C}caligraphic_C. This is because the impurity probes can now become entangled with the environment degrees of freedom. Unlike the large-K𝐾Kitalic_K limit, in the NBL the overall ground state of the system is always a net spin-singlet state and there is no ground state transition as a function of intra-probe coupling strength K𝐾Kitalic_K. However, the character of the reduced state on the two-impurity probe changes continuously when J𝐽Jitalic_J is finite from being predominantly singlet type on the impurity probes for K>0𝐾0K>0italic_K > 0 to predominantly triplet type for K<0𝐾0K<0italic_K < 0. This is clearly demonstrated through the smooth evolution of the impurity probe spin-spin correlator 𝒞𝒞\mathcal{C}caligraphic_C with coupling strength K𝐾Kitalic_K (not shown), which in turn controls metrological performance in this system. Therefore we do again expect qualitative differences between positive and negative K𝐾Kitalic_K regimes, but a smoother crossover in behavior around K=0𝐾0K=0italic_K = 0. We also note however that the NBL does support transitions (level crossings) in the excited states, which play a role in finite-temperature metrology. We expect the finite J𝐽Jitalic_J scale to shift the maximum measurement precision point to finite K𝐾Kitalic_K and T𝑇Titalic_T (by contrast, in the large-K𝐾Kitalic_K limit the maximum QFI was somewhat artificially pinned to K=T=0𝐾𝑇0K=T=0italic_K = italic_T = 0).

We now turn to our numerical results for single-parameter estimation in the NBL of the 2IK model in panels (a-d) of Fig. 6. A first glance at the QFI plots shows a richer structure compared with the large-K𝐾Kitalic_K limit shown in the main text in panels (a-d) of Fig. 2. We see that the maximum in both (T)𝑇\mathcal{H}(T)caligraphic_H ( italic_T ) and (K)𝐾\mathcal{H}(K)caligraphic_H ( italic_K ) occurs around KJsimilar-to𝐾𝐽K\sim Jitalic_K ∼ italic_J at finite temperatures. More instructive are the QSNR responses in Figs. 6(c,d). First, we observe that the QSNR for estimation of T𝑇Titalic_T and K𝐾Kitalic_K are now different from each other. We note that for large KJmuch-greater-than𝐾𝐽K\gg Jitalic_K ≫ italic_J we start to recover the large-K𝐾Kitalic_K limit behavior in panels (c,d) of Fig. 2, especially at higher temperatures. However, the low-T𝑇Titalic_T behavior around KJsimilar-to𝐾𝐽K\sim Jitalic_K ∼ italic_J is dramatically altered, with a fragmentation of the thermometric response 𝒬SP(T)subscript𝒬𝑆𝑃𝑇\mathcal{Q}_{SP}(T)caligraphic_Q start_POSTSUBSCRIPT italic_S italic_P end_POSTSUBSCRIPT ( italic_T ) to a double-peak structure, and a pronounced knee in 𝒬SP(K)subscript𝒬𝑆𝑃𝐾\mathcal{Q}_{SP}(K)caligraphic_Q start_POSTSUBSCRIPT italic_S italic_P end_POSTSUBSCRIPT ( italic_K ). These features anticipate what will become an actual quantum phase transition in the full many-body model. The results seem to show that with coupling to the leads, good thermometric sensitivity is not possible at low temperatures TJmuch-less-than𝑇𝐽T\ll Jitalic_T ≪ italic_J. By contrast, the peak sensitivity in estimating K𝐾Kitalic_K is at around KJsimilar-to𝐾𝐽K\sim Jitalic_K ∼ italic_J. For thermometry we explicitly note that only at larger T𝑇Titalic_T and K𝐾Kitalic_K do we see good temperature estimation capability due to finite J𝐽Jitalic_J scale - this is rather similar to the full 2IK model shown in panel (g) of Fig. 2. Similarly, the single-parameter estimation of the coupling constant K𝐾Kitalic_K becomes relatively enhanced when K𝐾Kitalic_K is of order J𝐽Jitalic_J which anticipates the onset of criticality.

Appendix C Validation of exact result near criticality

In Sec. VI of main text we present exact analytical results for the spin-spin correlation function 𝒞(T,K)𝒞𝑇𝐾\mathcal{C}\left(T,K\right)caligraphic_C ( italic_T , italic_K ) in Eq. (30), as well as its derivatives T𝒞(T,K)subscript𝑇𝒞𝑇𝐾\partial_{T}\mathcal{C}\left(T,K\right)∂ start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT caligraphic_C ( italic_T , italic_K ) in Eq. (29) and K𝒞(T,K)subscript𝐾𝒞𝑇𝐾\partial_{K}\mathcal{C}\left(T,K\right)∂ start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT caligraphic_C ( italic_T , italic_K ) in Eq. (31), for the full 2IK model in the close vicinity of the critical point. In turn, this allows us to find closed form solutions for both the thermometric precision 𝒬SP2(T)superscriptsubscript𝒬𝑆𝑃2𝑇\mathcal{Q}_{SP}^{2}\left(T\right)caligraphic_Q start_POSTSUBSCRIPT italic_S italic_P end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_T ) in Eq. (32) and sensitivity to coupling strength 𝒬SP(K)subscript𝒬𝑆𝑃𝐾\mathcal{Q}_{SP}\left(K\right)caligraphic_Q start_POSTSUBSCRIPT italic_S italic_P end_POSTSUBSCRIPT ( italic_K ) in Eq. (33). In this section we validate these analytic expressions by comparison with full NRG calculations. For concreteness we carry out our NRG calculations with a coupling J=1𝐽1J=1italic_J = 1, for which we find that KC0.618similar-to-or-equalssubscript𝐾𝐶0.618K_{C}\simeq 0.618italic_K start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ≃ 0.618, TK=0.362subscript𝑇𝐾0.362T_{K}=0.362italic_T start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT = 0.362, 𝒞0.385similar-to-or-equalssuperscript𝒞0.385\mathcal{C}^{*}\simeq-0.385caligraphic_C start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ≃ - 0.385 and we extract the constant c0.035similar-to-or-equals𝑐0.035c\simeq 0.035italic_c ≃ 0.035 from NRG thermodynamics.

In panel (a) of Fig. 7 the NRG computed spin-spin correlation function is plotted for different values of impurity detuning, δK=KKC𝛿𝐾𝐾subscript𝐾𝐶\delta K=K-K_{C}italic_δ italic_K = italic_K - italic_K start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT, where blue and red lines correspond to positive and negative detunings δK𝛿𝐾\delta Kitalic_δ italic_K respectively. The corresponding exact results of Eq. (30) are denoted by the black dashed line. We see excellent agreement between the NRG and exact solutions, and note that the spin-spin correlation function remains largely unchanged as a function of T𝑇Titalic_T at low temperatures, but changes sharply as a function of the impurity detuning δK𝛿𝐾\delta Kitalic_δ italic_K. Similarly, in panels (b) and (c) of Fig. 7 we show the spin-spin correlation function derivatives, T𝒞(T,K)subscript𝑇𝒞𝑇𝐾\partial_{T}\mathcal{C}\left(T,K\right)∂ start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT caligraphic_C ( italic_T , italic_K ) and K𝒞(T,K)subscript𝐾𝒞𝑇𝐾\partial_{K}\mathcal{C}\left(T,K\right)∂ start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT caligraphic_C ( italic_T , italic_K ) respectively in the colored red and blue lines for the NRG solution, and the black dashed lines shows the derived exact results Eq. (29) and Eq. (31). We again note the excellent quantitative agreement.

Refer to caption
Figure 8: Sub-optimal thermometry in the large-K𝐾Kitalic_K limit of the 2IK model with an applied field B𝐵Bitalic_B. (a) Single-parameter quantum signal to noise ratio, 𝒬SP(T)subscript𝒬𝑆𝑃𝑇\mathcal{Q}_{SP}\left(T\right)caligraphic_Q start_POSTSUBSCRIPT italic_S italic_P end_POSTSUBSCRIPT ( italic_T ) as a function of rescaled temperature T/B𝑇𝐵T/Bitalic_T / italic_B and coupling strength K/B𝐾𝐵K/Bitalic_K / italic_B; compared with (b) the (sub-optimal) single-parameter signal to noise ratio SP;𝒞(T)subscript𝑆𝑃𝒞𝑇\mathcal{R}_{SP;\mathcal{C}}\left(T\right)caligraphic_R start_POSTSUBSCRIPT italic_S italic_P ; caligraphic_C end_POSTSUBSCRIPT ( italic_T ) using the Fisher information obtained from measurement of only the spin-spin correlation function 𝒞𝒞\mathcal{C}caligraphic_C; and (c) the (sub-optimal) signal to noise ratio SP;(T)subscript𝑆𝑃𝑇\mathcal{R}_{SP;\mathcal{M}}\left(T\right)caligraphic_R start_POSTSUBSCRIPT italic_S italic_P ; caligraphic_M end_POSTSUBSCRIPT ( italic_T ) corresponding to a magnetization measurement \mathcal{M}caligraphic_M. The maximum QSNR and SNRs are shown in (d) for a given K/B𝐾𝐵K/Bitalic_K / italic_B. Panel (a) for the QSNR with finite field should be compared with the B=0𝐵0B=0italic_B = 0 result shown in Fig. 2(c).

Appendix D Approximate ansatz for the probe reduced
state in a B𝐵Bitalic_B field

The situation described in Sec. VII.1, which is formally exact, can be simplified somewhat by making an approximate ansatz for the probe reduced state populations. We motivate this by looking at the large-K𝐾Kitalic_K limit of the 2IK model, where we take the probe reduced density matrix to be that of the isolated probe, thermalized to the same temperature as the environment. In that case, we may write ϱT;0ϱTsubscriptitalic-ϱ𝑇0subscriptitalic-ϱ𝑇\varrho_{T;0}\equiv\varrho_{T}italic_ϱ start_POSTSUBSCRIPT italic_T ; 0 end_POSTSUBSCRIPT ≡ italic_ϱ start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT and ϱT;+1=ϱT×eB/Tsubscriptitalic-ϱ𝑇1subscriptitalic-ϱ𝑇superscript𝑒𝐵𝑇\varrho_{T;+1}=\varrho_{T}\times e^{-B/T}italic_ϱ start_POSTSUBSCRIPT italic_T ; + 1 end_POSTSUBSCRIPT = italic_ϱ start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT × italic_e start_POSTSUPERSCRIPT - italic_B / italic_T end_POSTSUPERSCRIPT, ϱT;1=ϱT×e+B/Tsubscriptitalic-ϱ𝑇1subscriptitalic-ϱ𝑇superscript𝑒𝐵𝑇\varrho_{T;-1}=\varrho_{T}\times e^{+B/T}italic_ϱ start_POSTSUBSCRIPT italic_T ; - 1 end_POSTSUBSCRIPT = italic_ϱ start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT × italic_e start_POSTSUPERSCRIPT + italic_B / italic_T end_POSTSUPERSCRIPT due to the energy splitting of the spin-triplet states in a field. With explicit coupling to the leads (finite J𝐽Jitalic_J), this is no longer the case because the probe and leads become entangled (the full states are not probe-lead separable). However, one might expect that the Sz=±1superscript𝑆𝑧plus-or-minus1S^{z}=\pm 1italic_S start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT = ± 1 reduced states of the probe are still split approximately symmetrically around the Sz=0superscript𝑆𝑧0S^{z}=0italic_S start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT = 0 triplet state, with some renormalized effective field. Taking the ansatz ϱT;+1/ϱT;0=ϱT;0/ϱT;1subscriptitalic-ϱ𝑇1subscriptitalic-ϱ𝑇0subscriptitalic-ϱ𝑇0subscriptitalic-ϱ𝑇1\varrho_{T;+1}/\varrho_{T;0}=\varrho_{T;0}/\varrho_{T;-1}italic_ϱ start_POSTSUBSCRIPT italic_T ; + 1 end_POSTSUBSCRIPT / italic_ϱ start_POSTSUBSCRIPT italic_T ; 0 end_POSTSUBSCRIPT = italic_ϱ start_POSTSUBSCRIPT italic_T ; 0 end_POSTSUBSCRIPT / italic_ϱ start_POSTSUBSCRIPT italic_T ; - 1 end_POSTSUBSCRIPT, which holds exactly in the large-K𝐾Kitalic_K limit and also in the small-B𝐵Bitalic_B limit, we can derive an expression for χ𝜒\mathcal{\chi}italic_χ in terms of 𝒞𝒞\mathcal{C}caligraphic_C and \mathcal{M}caligraphic_M,

χ1+43𝒞(12+23𝒞)2132similar-to-or-equals𝜒143𝒞superscript1223𝒞213superscript2\mathcal{\chi}\simeq 1+\tfrac{4}{3}\mathcal{C}-\sqrt{(\tfrac{1}{2}+\tfrac{2}{3% }\mathcal{C})^{2}-\tfrac{1}{3}\mathcal{M}^{2}}italic_χ ≃ 1 + divide start_ARG 4 end_ARG start_ARG 3 end_ARG caligraphic_C - square-root start_ARG ( divide start_ARG 1 end_ARG start_ARG 2 end_ARG + divide start_ARG 2 end_ARG start_ARG 3 end_ARG caligraphic_C ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - divide start_ARG 1 end_ARG start_ARG 3 end_ARG caligraphic_M start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG (40)

This allows us to eliminate the dependence of χ𝜒\mathcal{\chi}italic_χ and reduce the parametric complexity of the problem to just two observables – the probe spin-spin correlator 𝒞𝒞\mathcal{C}caligraphic_C and the magnetization \mathcal{M}caligraphic_M. Even in the full 2IK model with finite K,J,B,D𝐾𝐽𝐵𝐷K,J,B,Ditalic_K , italic_J , italic_B , italic_D we find from our NRG calculations that Eq. (40) holds rather accurately across the full parameter space investigated.

Appendix E Single-parameter thermometry with a control field

In Sec. VII we apply a known control field B𝐵Bitalic_B to our system which removes the singularity of the QFIM. Here we briefly discuss how a finite field B𝐵Bitalic_B alters the physics of the system by summarizing results for the single-parameter estimation case, focusing on the QSNR for thermometry in the large-K𝐾Kitalic_K limit of the 2IK model. The situation here is somewhat different from that of Fig. 2(c) due to the inclusion of the field. In Fig. 5(a) of the main text we plot 𝒬SP(T)subscript𝒬𝑆𝑃𝑇\mathcal{Q}_{SP}(T)caligraphic_Q start_POSTSUBSCRIPT italic_S italic_P end_POSTSUBSCRIPT ( italic_T ) as a function of the rescaled parameters t=T/B𝑡𝑇𝐵t=T/Bitalic_t = italic_T / italic_B and k=K/B𝑘𝐾𝐵k=K/Bitalic_k = italic_K / italic_B. The single-parameter QSNR is a universal function of t𝑡titalic_t and k𝑘kitalic_k. This holds exactly in the large-K𝐾Kitalic_K limit, but also holds generally in the full 2IK model provided all parameters are similarly scaled. This scaling implies that we get the same behavior as a function of t𝑡titalic_t and k𝑘kitalic_k independent of the field strength B𝐵Bitalic_B (but over an ever-narrowing window of the bare parameters T𝑇Titalic_T and K𝐾Kitalic_K as B0𝐵0B\to 0italic_B → 0).

The single-parameter QSNR in Fig. 5(a) of the main text shows level-crossing transition behavior at K=B𝐾𝐵K\!=\!Bitalic_K = italic_B corresponding to a change in ground state from the collective probe spin-singlet |Sket𝑆|S\rangle| italic_S ⟩ for K>B𝐾𝐵K>Bitalic_K > italic_B to a component of the spin-triplet |T;1ket𝑇1|T;-1\rangle| italic_T ; - 1 ⟩ for K<B𝐾𝐵K<Bitalic_K < italic_B. The transition between singlet and triplet ground states occurs at K=0𝐾0K=0italic_K = 0 when B=0𝐵0B=0italic_B = 0, as shown in Fig. 2(c). However, at B=0𝐵0B=0italic_B = 0 we saw that the transition is also associated with a change in the degeneracy of the ground and excited states, and this produced a marked difference in the maximum attainable QSNR for K>0𝐾0K>0italic_K > 0 vs K<0𝐾0K<0italic_K < 0. At finite B𝐵Bitalic_B and modest K𝐾Kitalic_K we see no such difference between the phases K>B𝐾𝐵K>Bitalic_K > italic_B and K<B𝐾𝐵K<Bitalic_K < italic_B because all states are unique (non-degenerate) due to splittings in the field. At very large |K/B|𝐾𝐵|K/B|| italic_K / italic_B | and K>0𝐾0K>0italic_K > 0 we expect to recover the QSNR profile obtained at B=0𝐵0B=0italic_B = 0, with boosted sensitivity around TKsimilar-to𝑇𝐾T\sim Kitalic_T ∼ italic_K as the excited triplet states become quasi-degenerate. For large |K/B|𝐾𝐵|K/B|| italic_K / italic_B | but K<0𝐾0K<0italic_K < 0, we have relatively poor sensitivity at TKsimilar-to𝑇𝐾T\sim Kitalic_T ∼ italic_K because the excited singlet state is non-degenerate, but transitions between the triplet states now provide additional sensitivity for TBsimilar-to𝑇𝐵T\sim Bitalic_T ∼ italic_B.

Refer to caption
Figure 9: Elements of the multiparameter QFI matrix in the large-K𝐾Kitalic_K limit of the 2IK model. We plot B2𝓗T,Tsuperscript𝐵2subscript𝓗𝑇𝑇B^{2}\boldsymbol{\mathcal{H}}_{T,T}italic_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT bold_caligraphic_H start_POSTSUBSCRIPT italic_T , italic_T end_POSTSUBSCRIPT, B2𝓗K,Ksuperscript𝐵2subscript𝓗𝐾𝐾B^{2}\boldsymbol{\mathcal{H}}_{K,K}italic_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT bold_caligraphic_H start_POSTSUBSCRIPT italic_K , italic_K end_POSTSUBSCRIPT, and B2𝓗T,Ksuperscript𝐵2subscript𝓗𝑇𝐾B^{2}\boldsymbol{\mathcal{H}}_{T,K}italic_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT bold_caligraphic_H start_POSTSUBSCRIPT italic_T , italic_K end_POSTSUBSCRIPT (B2𝓗K,Tabsentsuperscript𝐵2subscript𝓗𝐾𝑇\equiv B^{2}\boldsymbol{\mathcal{H}}_{K,T}≡ italic_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT bold_caligraphic_H start_POSTSUBSCRIPT italic_K , italic_T end_POSTSUBSCRIPT) in panels (a), (b), and (c) respectively. The quantities B2𝓗λ,λsuperscript𝐵2subscript𝓗𝜆superscript𝜆B^{2}\boldsymbol{\mathcal{H}}_{\lambda,\lambda^{\prime}}italic_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT bold_caligraphic_H start_POSTSUBSCRIPT italic_λ , italic_λ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT are universal scaling functions of T/B𝑇𝐵T/Bitalic_T / italic_B and K/B𝐾𝐵K/Bitalic_K / italic_B.

Appendix F Sub-optimal measurements

The QSNR plotted in Fig. 5(a) of the main text and in Fig. 8(a) is the ‘best case scenario’ obtained from the (single parameter) thermometric QFI – it corresponds to the optimal measurement on the probe. This optimal measurement (the thermometric SLD) always exists and is an observable in principle. For the 2IK model, the optimal measurement basis is effectively the Bell basis, but such measurements are not practical in an experimental setting. Previously, we have assumed that we have simultaneous access to the set of physical observables 𝒞𝒞\mathcal{C}caligraphic_C, \mathcal{M}caligraphic_M and χ𝜒\mathcal{\chi}italic_χ, in terms of which the SLD can be decomposed. Here we ask about the thermometric sensitivity obtainable for a sub-optimal measurement – for example just the spin-spin correlator 𝒞𝒞\mathcal{C}caligraphic_C or the magnetization \mathcal{M}caligraphic_M. With only partial access to the probe state, one naturally expects lower parameter estimation precision.

This can be understood quantitatively from the Fisher information FΩ(T)subscript𝐹Ω𝑇F_{\Omega}(T)italic_F start_POSTSUBSCRIPT roman_Ω end_POSTSUBSCRIPT ( italic_T ) for a given measurement ΩΩ\Omegaroman_Ω. The corresponding (sub-optimal, single-parameter) signal-to-noise ratio (SNR) is defined analogously to the QSNR,

SP:Ω(T)=T2FΩ(T).subscript:𝑆𝑃Ω𝑇superscript𝑇2subscript𝐹Ω𝑇\mathcal{R}_{SP:\Omega}\left(T\right)=T^{2}~{}F_{\Omega}\left(T\right)\;.caligraphic_R start_POSTSUBSCRIPT italic_S italic_P : roman_Ω end_POSTSUBSCRIPT ( italic_T ) = italic_T start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_F start_POSTSUBSCRIPT roman_Ω end_POSTSUBSCRIPT ( italic_T ) . (41)

with FΩ(λ)=|λΩ|2/Var[Ω]subscript𝐹Ω𝜆superscriptsubscript𝜆expectation-valueΩ2Vardelimited-[]ΩF_{\Omega}\left(\lambda\right)=\absolutevalue{\partial_{\lambda}% \expectationvalue{\Omega}}^{2}/\text{Var}\left[\Omega\right]italic_F start_POSTSUBSCRIPT roman_Ω end_POSTSUBSCRIPT ( italic_λ ) = | start_ARG ∂ start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ⟨ start_ARG roman_Ω end_ARG ⟩ end_ARG | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / Var [ roman_Ω ] and λ=T𝜆𝑇\lambda=Titalic_λ = italic_T for thermometry, defined from the error propagation formula Tóth and Apellaniz (2014). In Fig. 8(b,c) we plot the thermometric SNR corresponding to measurement of just 𝒞𝒞\mathcal{C}caligraphic_C or \mathcal{M}caligraphic_M on the probe (which are the dominant contributions). Fig. 8(d) shows the maximum QSNR or SNR at a given K/B𝐾𝐵K/Bitalic_K / italic_B. We see immediately that the full QSNR can be decomposed into contributions from the different measurements (although not in a simple additive fashion, as Eq. (36) attests). In particular, note that the measurement of only 𝒞𝒞\mathcal{C}caligraphic_C may yield better results than measuring just \mathcal{M}caligraphic_M in certain parameter regimes (in this case, for K>B𝐾𝐵K>Bitalic_K > italic_B) – with the SNR SP:𝒞subscript:𝑆𝑃𝒞\mathcal{R}_{SP:\mathcal{C}}caligraphic_R start_POSTSUBSCRIPT italic_S italic_P : caligraphic_C end_POSTSUBSCRIPT even approaching the QSNR 𝒬SPsubscript𝒬𝑆𝑃\mathcal{Q}_{SP}caligraphic_Q start_POSTSUBSCRIPT italic_S italic_P end_POSTSUBSCRIPT for certain parameters. However measurement of \mathcal{M}caligraphic_M might be preferable in other cases (for example K<B𝐾𝐵K<Bitalic_K < italic_B). The differences can be traced to the different types of transitions controlling the SNR in the two regimes: for K>B𝐾𝐵K>Bitalic_K > italic_B the dominant contributions come from spin singlet-triplet transitions picked up by the spin-spin correlator 𝒞𝒞\mathcal{C}caligraphic_C, whereas for K<B𝐾𝐵K<Bitalic_K < italic_B magnetic transitions between the triplet components dominate, and these sensitively affect the magnetization \mathcal{M}caligraphic_M. Therefore intimate details about the spectrum can be inferred by examining the SNRs for different but complementary observables in this way. Broadly similar results are obtained for the single-parameter estimation of the coupling constant K𝐾Kitalic_K (not shown).

Appendix G QFIM for multiparameter estimation

In the main text Sec. VII.2 we present the multiparameter QSNRs in the large K𝐾Kitalic_K-limit in the presence of a magnetic field. In Fig. 9 we provide the elements of the QFIM for reference. In particular, we note that the off-diagonal element, 𝓗T,Ksubscript𝓗𝑇𝐾\boldsymbol{\mathcal{H}}_{T,K}bold_caligraphic_H start_POSTSUBSCRIPT italic_T , italic_K end_POSTSUBSCRIPT presented in panel (c) can be negative if there is a negative correlation between the parameters; the variance or covariance does not correspond simply to any one of the QFIM elements, but requires information from all of them.

Formally, the diagonal terms of the QFIM are related to variances in the single-parameter estimation case, and so these quantities are strictly positive. On the other hand, the off-diagonal terms can be either positive or negative – the sign telling us in what way the parameters are correlated.

Our results illustrate the importance of using the QSNR rather than the QFIM itself to quantify metrological performance. For example, the QFIM element 𝓗T,Tsubscript𝓗𝑇𝑇\boldsymbol{\mathcal{H}}_{T,T}bold_caligraphic_H start_POSTSUBSCRIPT italic_T , italic_T end_POSTSUBSCRIPT shown in Fig. 9(a) only provides us with information about how temperature sensitivity is imprinted on our probe in the context of single parameter estimation. The true multiparameter QSNR 𝒬MP(T,T)subscript𝒬𝑀𝑃𝑇𝑇\mathcal{Q}_{MP}\left(T,T\right)caligraphic_Q start_POSTSUBSCRIPT italic_M italic_P end_POSTSUBSCRIPT ( italic_T , italic_T ) shown in Fig. 5(b) tells a very different story about the actual metrological utility of the probe.

References

  • Degen et al. (2017) C. L. Degen, F. Reinhard,  and P. Cappellaro, “Quantum sensing,” Rev. Mod. Phys. 89, 035002 (2017).
  • Greenberger et al. (1989) D. M. Greenberger, M. A. Horne,  and A. Zeilinger, “Going beyond bell’s theorem,” in Bell’s theorem, quantum theory and conceptions of the universe (Springer, 1989) pp. 69–72.
  • Giovannetti et al. (2004) V. Giovannetti, S. Lloyd,  and L. Maccone, “Quantum-enhanced measurements: beating the standard quantum limit,” Science 306, 1330–1336 (2004).
  • Leibfried et al. (2004) D. Leibfried, M. D. Barrett, T. Schaetz, J. Britton, J. Chiaverini, W. M. Itano, J. D. Jost, C. Langer,  and D. J. Wineland, “Toward Heisenberg-limited spectroscopy with multiparticle entangled states,” Science 304, 1476–1478 (2004).
  • Demkowicz-Dobrzański et al. (2012) R. Demkowicz-Dobrzański, J. Kołodyński,  and M. Guţă, “The elusive Heisenberg limit in quantum-enhanced metrology,” Nat. Commun.  (2012), 10.1038/ncomms2067.
  • Zhou (2024) S. Zhou, “Limits of noisy quantum metrology with restricted quantum controls,”  (2024), arXiv:2402.18765 [quant-ph] .
  • De Pasquale et al. (2013) A. De Pasquale, D. Rossini, P. Facchi,  and V. Giovannetti, “Quantum parameter estimation affected by unitary disturbance,” Phys. Rev. A 88, 052117 (2013).
  • Campos Venuti and Zanardi (2007) L. Campos Venuti and P. Zanardi, “Quantum critical scaling of the geometric tensors,” Phys. Rev. Lett. 99, 095701 (2007).
  • Schwandt et al. (2009) D. Schwandt, F. Alet,  and S. Capponi, “Quantum monte carlo simulations of fidelity at magnetic quantum phase transitions,” Phys. Rev. Lett. 103, 170501 (2009).
  • Albuquerque et al. (2010) A. F. Albuquerque, F. Alet, C. Sire,  and S. Capponi, “Quantum critical scaling of fidelity susceptibility,” Phys. Rev. B 81, 064418 (2010).
  • Gritsev and Polkovnikov (2010) V. Gritsev and A. Polkovnikov, “Universal dynamics near quantum critical points,”  (2010), arXiv:0910.3692 [cond-mat.stat-mech] .
  • Gu et al. (2008) S.J. Gu, H.M. Kwok, W.Q. Ning,  and H.Q. Lin, “Fidelity susceptibility, scaling, and universality in quantum critical phenomena,” Phys. Rev. B 77, 245109 (2008).
  • Greschner et al. (2013) S. Greschner, A. K. Kolezhuk,  and T. Vekua, “Fidelity susceptibility and conductivity of the current in one-dimensional lattice models with open or periodic boundary conditions,” Phys. Rev. B 88, 195101 (2013).
  • Frérot and Roscilde (2018) I. Frérot and T. Roscilde, “Quantum critical metrology,” Phys. Rev. Lett. 121, 020402 (2018).
  • Zhou et al. (2020) H. Zhou, J. Choi, S. Choi, R. Landig, A. M. Douglas, J. Isoya, F. Jelezko, S. Onoda, H. Sumiya, P. Cappellaro, H. S. Knowles, H. Park,  and M. D. Lukin, “Quantum metrology with strongly interacting spin systems,” Phys. Rev. X 10, 031003 (2020).
  • Rams et al. (2018a) M. M. Rams, P. Sierant, O. Dutta, P. Horodecki,  and J. Zakrzewski, “At the limits of criticality-based quantum metrology: Apparent super-Heisenberg scaling revisited,” Phys. Rev. X 8, 021022 (2018a).
  • Chu et al. (2021) Y. Chu, S. Zhang, B. Yu,  and J. Cai, “Dynamic framework for criticality-enhanced quantum sensing,” Phys. Rev. Lett. 126, 010502 (2021).
  • Di Fresco et al. (2022) G. Di Fresco, B. Spagnolo, D. Valenti,  and A. Carollo, “Multiparameter quantum critical metrology,”  (2022), arXiv:2203.12676 [quant-ph] .
  • Chu et al. (2023) Y. Chu, X. Li,  and J. Cai, “Strong quantum metrological limit from many-body physics,” Phys. Rev. Lett. 130, 170801 (2023).
  • Salvia et al. (2023) R. Salvia, M. Mehboudi,  and M. Perarnau-Llobet, “Critical quantum metrology assisted by real-time feedback control,” Phys. Rev. Lett. 130, 240803 (2023).
  • Rodríguez et al. (2023) R. R. Rodríguez, M. Mehboudi, M. Horodecki,  and M. Perarnau-Llobet, “Strongly coupled fermionic probe for nonequilibrium thermometry,”  (2023), arXiv:2310.14655 [quant-ph] .
  • Bressanini et al. (2024) G. Bressanini, M.G Genoni, M. S. Kim,  and M.G.A Paris, “Multi-parameter quantum estimation of single- and two-mode pure gaussian states,”  (2024), arXiv:2403.03919 [quant-ph] .
  • Cavazzoni et al. (2024) S. Cavazzoni, M. Adani, P. Bordone,  and M.G.A Paris, “Characterization of partially accessible anisotropic spin chains in the presence of anti-symmetric exchange,”  (2024), arXiv:2401.14479 [quant-ph] .
  • Zanardi and Paunković (2006) P. Zanardi and N. Paunković, “Ground state overlap and quantum phase transitions,” Phys. Rev. E 74, 031123 (2006).
  • Abasto et al. (2008) D. F. Abasto, A. Hamma,  and P. Zanardi, “Fidelity analysis of topological quantum phase transitions,” Phys. Rev. A 78, 010301 (2008).
  • Sun et al. (2010) Z. Sun, J. Ma, X.M. Lu,  and X. Wang, “Fisher information in a quantum-critical environment,” Phys. Rev. A 82, 022306 (2010).
  • Zanardi et al. (2008) P. Zanardi, M. G. A. Paris,  and L. Campos Venuti, “Quantum criticality as a resource for quantum estimation,” Phys. Rev. A 78, 042105 (2008).
  • Damski and Rams (2013) B. Damski and M. M. Rams, “Exact results for fidelity susceptibility of the quantum ising model: the interplay between parity, system size, and magnetic field,” J. Phys. A Math. Theor. 47, 025303 (2013).
  • Salvatori et al. (2014) G. Salvatori, A. Mandarino,  and M. G. A. Paris, “Quantum metrology in lipkin-meshkov-glick critical systems,” Phys. Rev. A 90, 022111 (2014).
  • Yang et al. (2022) J. Yang, S. Pang, A. del Campo,  and A. N. Jordan, “Super-Heisenberg scaling in hamiltonian parameter estimation in the long-range kitaev chain,” Phys. Rev. Res. 4, 013133 (2022).
  • Fernández-Lorenzo et al. (2018) S. Fernández-Lorenzo, J. A. Dunningham,  and D. Porras, “Heisenberg scaling with classical long-range correlations,” Phys. Rev. A 97, 023843 (2018).
  • Montenegro et al. (2022) V. Montenegro, G. S. Jones, S. Bose,  and A. Bayat, “Sequential measurements for quantum-enhanced magnetometry in spin chain probes,” Phys. Rev. Lett. 129, 120503 (2022).
  • Ozaydin and Altintas (2015) F. Ozaydin and A. A. Altintas, “Quantum metrology: Surpassing the shot-noise limit with dzyaloshinskii-moriya interaction,” Sci. Rep. 5, 1–6 (2015).
  • Garbe et al. (2022a) L. Garbe, O. Abah, S. Felicetti,  and R. Puebla, “Critical quantum metrology with fully-connected models: from Heisenberg to kibble–zurek scaling,” Quantum Sci. Technol. 7, 035010 (2022a).
  • Mirkhalaf et al. (2021) S. S. Mirkhalaf, D. Benedicto Orenes, M. W. Mitchell,  and E. Witkowska, “Criticality-enhanced quantum sensing in ferromagnetic bose-einstein condensates: Role of readout measurement and detection noise,” Phys. Rev. A 103, 023317 (2021).
  • Wald et al. (2020) S. Wald, S. V. Moreira,  and F. L. Semião, “In- and out-of-equilibrium quantum metrology with mean-field quantum criticality,” Phys. Rev. E 101, 052107 (2020).
  • Hotter et al. (2024) C. Hotter, H. Ritsch,  and K. Gietka, “Combining critical and quantum metrology,” Phys. Rev. Lett. 132, 060801 (2024).
  • Ostermann and Gietka (2024) L. Ostermann and K. Gietka, “Temperature-enhanced critical quantum metrology,”  (2024), arXiv:2312.04176 [quant-ph] .
  • Lü et al. (2024) J. Lü, W. Ning, F. Wu, R.H Zheng, K. Chen, X. Zhu, Z.B Yang, H.Z Wu,  and S.B Zheng, “Critical quantum metrology robust against dissipation and non-adiabaticity,”  (2024), arXiv:2403.04475 [quant-ph] .
  • Alushi et al. (2024) U. Alushi, W. Górecki, S. Felicetti,  and R.D Candia, “Optimality and noise-resilience of critical quantum sensing,”  (2024), arXiv:2402.15559 [quant-ph] .
  • Budich and Bergholtz (2020) J. C. Budich and E. J. Bergholtz, “Non-hermitian topological sensors,” Phys. Rev. Lett. 125, 180403 (2020).
  • Koch and Budich (2022) F. Koch and J. C. Budich, “Quantum non-hermitian topological sensors,” Phys. Rev. Res. 4, 013113 (2022).
  • Sarkar et al. (2022) S. Sarkar, C. Mukhopadhyay, A. Alase,  and A. Bayat, “Free-fermionic topological quantum sensors,” Phys. Rev. Lett. 129, 090503 (2022).
  • Zhang et al. (2023) T. Zhang, J. Hu,  and X. Qiu, “Topological waveguide quantum sensors,”  (2023), arXiv:2311.01370 [quant-ph] .
  • Yang et al. (2024a) Y. Yang, A. Yuan,  and F. Li, “Quantum multiparameter estimation enhanced by a topological phase transition,” Physical Review A 109 (2024a), 10.1103/physreva.109.022604.
  • Bin et al. (2019) S.W. Bin, X.Y. Lü, T.S. Yin, G.L. Zhu, Q. Bin,  and Y. Wu, “Mass sensing by quantum criticality,” Optics Letters 44, 630–633 (2019).
  • Di Candia et al. (2023) R. Di Candia, F. Minganti, K. V. Petrovnin, G. S. Paraoanu,  and S. Felicetti, “Critical parametric quantum sensing,” Npj Quantum Inf. 9 (2023), 10.1038/s41534-023-00690-z.
  • Garbe et al. (2020) L. Garbe, M. Bina, A. Keller, M. G. A. Paris,  and S. Felicetti, “Critical quantum metrology with a finite-component quantum phase transition,” Phys. Rev. Lett. 124, 120504 (2020).
  • Heugel et al. (2019) T. L. Heugel, M. Biondi, O. Zilberberg,  and R. Chitra, “Quantum transducer using a parametric driven-dissipative phase transition,” Phys. Rev. Lett. 123, 173601 (2019).
  • Fernández-Lorenzo and Porras (2017) S. Fernández-Lorenzo and D. Porras, “Quantum sensing close to a dissipative phase transition: Symmetry breaking and criticality as metrological resources,” Phys. Rev. A 96, 013817 (2017).
  • Wu and Shi (2021) W. Wu and C. Shi, “Criticality-enhanced quantum sensor at finite temperature,” Phys. Rev. A 104, 022612 (2021).
  • Ying et al. (2022) Z.J. Ying, S. Felicetti, G. Liu,  and D. Braak, “Critical quantum metrology in the non-linear quantum rabi model,” Entropy 24 (2022), 10.3390/e24081015.
  • Tang et al. (2023) S.B Tang, H. Qin, D. Y. Wang, K. Cui, S. L. Su, L. L. Yan,  and G. Chen, “Enhancement of quantum sensing in a cavity optomechanical system around quantum critical point,”  (2023), arXiv:2303.16486 [quant-ph] .
  • Zhu et al. (2023) X. Zhu, J. J. Lü, W. Ning, F. Wu, L. T. Shen, z. B. Yang,  and S. B. Zheng, “Criticality-enhanced quantum sensing in the anisotropic quantum rabi model,” SCI CHINA PHYS MECH 66 (2023), 10.1007/s11433-022-2073-9.
  • Garbe et al. (2022b) L. Garbe, O. Abah, S. Felicetti,  and R. Puebla, “Exponential time-scaling of estimation precision by reaching a quantum critical point,” Phys. Rev. Res. 4, 043061 (2022b).
  • Tsang (2013) M. Tsang, “Quantum transition-edge detectors,” Phys. Rev. A 88, 021801 (2013).
  • Macieszczak et al. (2016) K. Macieszczak, M. Guţă, I. Lesanovsky,  and J. P. Garrahan, “Dynamical phase transitions as a resource for quantum enhanced metrology,” Phys. Rev. A 93, 022103 (2016).
  • Carollo et al. (2018) A. Carollo, B. Spagnolo,  and D. Valenti, “Uhlmann curvature in dissipative phase transitions,” Sci. Rep. 8, 9852 (2018).
  • Lang et al. (2015) J. E. Lang, R. B. Liu,  and T. S. Monteiro, “Dynamical-decoupling-based quantum sensing: Floquet spectroscopy,” Phys. Rev. X 5, 041016 (2015).
  • Mishra and Bayat (2021) U. Mishra and A. Bayat, “Driving enhanced quantum sensing in partially accessible many-body systems,” Phys. Rev. Lett. 127, 080504 (2021).
  • Mishra and Bayat (2022) U. Mishra and A. Bayat, “Integrable quantum many-body sensors for AC field sensing,” Sci. Rep. 12 (2022), 10.1038/s41598-022-17381-y.
  • Ilias et al. (2022) Theodoros Ilias, Dayou Yang, Susana F. Huelga,  and Martin B. Plenio, “Criticality-enhanced quantum sensing via continuous measurement,” PRX Quantum 3, 010354 (2022).
  • Boeyens et al. (2023) J. Boeyens, B. Annby-Andersson, P. Bakhshinezhad, G. Haack, M. Perarnau-Llobet, S Nimmrichter, P. P. Potts,  and M. Mehboudi, “Probe thermometry with continuous measurements,”  (2023), arXiv:2307.13407 [quant-ph] .
  • He et al. (2023) X. He, R. Yousefjani,  and A. Bayat, “Stark localization as a resource for weak-field sensing with super-Heisenberg precision,” Phys. Rev. Lett. 131, 010801 (2023).
  • Yousefjani et al. (2023) R. Yousefjani, X. He,  and A. Bayat, “Long-range interacting stark many-body probes with super-Heisenberg precision,” Chinese Physics B 32, 100313 (2023).
  • Bhattacharyya et al. (2023) A. Bhattacharyya, A. Ghoshal,  and U. Sen, “Disorder-induced enhancement of precision in quantum metrology,”  (2023), arXiv:2212.08523 [quant-ph] .
  • Sahoo et al. (2023) A. Sahoo, U. Mishra,  and D. Rakshit, “Localization driven quantum sensing,”  (2023), arXiv:2305.02315 [quant-ph] .
  • Montenegro et al. (2023) V. Montenegro, M. G. Genoni, A. Bayat,  and M. G. A. Paris, “Quantum metrology with boundary time crystals,” Communications Physics 6 (2023), 10.1038/s42005-023-01423-6.
  • Cabot et al. (2023) A. Cabot, F. Carollo,  and I. Lesanovsky, “Continuous sensing and parameter estimation with the boundary time-crystal,”  (2023), arXiv:2307.13277 [quant-ph] .
  • Yu et al. (2022) M. Yu, X. Li, y. Chu, B. Mera, F. N. Ünal, P. Yang, Y. Liu, N. Goldman,  and J. Cai, “Experimental demonstration of topological bounds in quantum metrology,”  (2022), arXiv:2206.00546 [quant-ph] .
  • Liu et al. (2021) R. Liu, Y. Chen, M. Jiang, X. Yang, Z. Wu, Y. Li, H. Yuan, X. Peng,  and J. Du, “Experimental critical quantum metrology with the Heisenberg scaling,” Npj Quantum Inf. 7, 1–7 (2021).
  • Ilias et al. (2023) Y. Ilias, D. Yang, S. F. Huelga,  and B. B. Plenio, “Criticality-enhanced electromagnetic field sensor with single trapped ions,”  (2023), arXiv:2304.02050 [quant-ph] .
  • Ding et al. (2022) D. S. Ding, Z. K. Liu, B. S. Shi, G. C. Guo, K. Mølmer,  and C. S. Adams, “Enhanced metrology at the critical point of a many-body Rydberg atomic system,” Nature Physics 18, 1447–1452 (2022).
  • Correa et al. (2015) L. A. Correa, M. Mehboudi, G. Adesso,  and A. Sanpera, “Individual quantum probes for optimal thermometry,” Phys. Rev. Lett. 114, 220405 (2015).
  • Paris (2015) M. G. A. Paris, “Achieving the Landau bound to precision of quantum thermometry in systems with vanishing gap,” J. Phys. A: Math. Theor. 49, 03LT02 (2015).
  • Stace (2010) T. M. Stace, “Quantum limits of thermometry,” Phys. Rev. A 82, 011611 (2010).
  • Mitchison et al. (2020) M. T. Mitchison, T. Fogarty, G. Guarnieri, S. Campbell, T. Busch,  and J. Goold, “In situ thermometry of a cold fermi gas via dephasing impurities,” Phys. Rev. Lett. 125, 080402 (2020).
  • Campbell et al. (2018) S. Campbell, M. G. Genoni,  and Deffner S., “Precision thermometry and the quantum speed limit,” Quantum Sci. Technol 3, 025002 (2018).
  • Brattegard and Mitchison (2023) S. Brattegard and M. T. Mitchison, “Thermometry by correlated dephasing of impurities in a 1d fermi gas,”  (2023), arXiv:2307.10132v2 [cond-mat.quant-gas] .
  • Brenes and Segal (2023) M. Brenes and D. Segal, “Multispin probes for thermometry in the strong-coupling regime,” Phys. Rev. A 108, 032220 (2023).
  • Yu et al. (2023) M. Yu, H. C. Nguyen,  and S. Nimmrichter, “Criticality-enhanced precision in phase thermometry,”  (2023), arXiv:2311.14578 [quant-ph] .
  • Yang et al. (2024b) Y. Yang, V. Montenegro,  and A. Bayat, “Sequential measurements candeloetry with quantum many-body probes,”  (2024b), arXiv:2403.10084 [quant-ph] .
  • Srivastava et al. (2023) A.K Srivastava, U. Bhattacharya, M. Lewenstein,  and M. Płodzień, “Topological quantum thermometry,”  (2023), arXiv:2311.14524 [quant-ph] .
  • Verma et al. (2024) H. Verma, M. Zych,  and F. Costa, “Measuring two temperatures using a single thermometer,”  (2024), arXiv:2403.15186 [quant-ph] .
  • Paris (2009) M. G. A. Paris, “Quantum estimation for quantum technology,” Int. J. Quantum Inf. 07, 125–137 (2009).
  • Helstrom (1969) C. W. Helstrom, “Quantum detection and estimation theory,” J Stat Phy 1, 231–252 (1969).
  • Holevo (1982) A. Holevo, Probabilistic and Statistical Aspects of Quantum Theory (Edizioni della Normale Pisa, 1982) p. 324.
  • Sachdev (1999) Subir Sachdev, “Quantum phase transitions,” Physics world 12, 33 (1999).
  • Jayaprakash et al. (1981) C. Jayaprakash, H. R. Krishna-murthy,  and J. W. Wilkins, “Two-impurity Kondo problem,” Phys. Rev. Lett. 47, 737–740 (1981).
  • Jones et al. (1988) B. A. Jones, C. M. Varma,  and J. W. Wilkins, “Low-temperature properties of the two-impurity Kondo hamiltonian,” Phys. Rev. Lett. 61, 125–128 (1988).
  • Affleck and Ludwig (1992) I. Affleck and A. W. W. Ludwig, “Exact critical theory of the two-impurity Kondo model,” Phys. Rev. Lett. 68, 1046–1049 (1992).
  • Affleck et al. (1995) I. Affleck, A. W. W. Ludwig,  and B. A. Jones, “Conformal-field-theory approach to the two-impurity Kondo problem: Comparison with numerical renormalization-group results,” Phys. Rev. B 52, 9528–9546 (1995).
  • Mitchell et al. (2012) A. K. Mitchell, E. Sela,  and D. E. Logan, “Two-channel Kondo physics in two-impurity Kondo models,” Phys. Rev. Lett. 108, 086405 (2012).
  • Sela et al. (2011) E. Sela, A. K. Mitchell,  and L. Fritz, “Exact crossover green function in the two-channel and two-impurity Kondo models,” Phys. Rev. Lett. 106, 147202 (2011).
  • Mitchell and Sela (2012) A. K. Mitchell and E. Sela, “Universal low-temperature crossover in two-channel Kondo models,” Phys. Rev. B 85, 235127 (2012).
  • Hewson (1993) A. Hewson, The Kondo problem to heavy fermions (Cambridge Studies in Magnetism, CUP, 1993).
  • van der Wiel et al. (2002) W. G. van der Wiel, S. De Franceschi, J. M. Elzerman, T. Fujisawa, S. Tarucha,  and L. P. Kouwenhoven, “Electron transport through double quantum dots,” Rev. Mod. Phys. 75, 1–22 (2002).
  • Izumida and Sakai (2000) W. Izumida and O. Sakai, “Two-impurity Kondo effect in double-quantum-dot systems: Effect of interdot kinetic exchange coupling,” Phys. Rev. B 62, 10260–10267 (2000).
  • Zaránd et al. (2006) G. Zaránd, C. H. Chung, P. Simon,  and M. Vojta, “Quantum criticality in a double-quantum-dot system,” Phys. Rev. Lett. 97, 166802 (2006).
  • Jeong et al. (2001) H. Jeong, A. M. Chang,  and M. R. Melloch, “The Kondo effect in an artificial quantum dot molecule,” Science 293, 2221–2223 (2001).
  • Pouse et al. (2023) W. Pouse, L. Peeters, C. L. Hsueh, U. Gennser, A. Cavanna, M. A. Kastner, A. K. Mitchell,  and D. Goldhaber-Gordon, “Quantum simulation of an exotic quantum critical point in a two-site charge Kondo circuit,” Nat. Phys. 19, 492–499 (2023).
  • Karki et al. (2023) D. B. Karki, E. Boulat, W. Pouse, D. Goldhaber-Gordon, A. K. Mitchell,  and C. Mora, “Z3 parafermion in the double charge Kondo model,” Phys. Rev. Lett. 130, 146201 (2023).
  • Bayat et al. (2012) A. Bayat, S. Bose, P. Sodano,  and H. Johannesson, “Entanglement probe of two-impurity Kondo physics in a spin chain,” Phys. Rev. Lett. 109, 066403 (2012).
  • Bayat et al. (2014) A. Bayat, H. Johannesson, S. Bose,  and P. Sodano, “An order parameter for impurity systems at quantum criticality,” Nat. Commun. 5, 3784 (2014).
  • Gietka et al. (2021) K. Gietka, F. Metz, T. Keller,  and J. Li, “Adiabatic critical quantum metrology cannot reach the Heisenberg limit even when shortcuts to adiabaticity are applied,” Quantum 5, 489 (2021).
  • Rams et al. (2018b) M. M. Rams, P. Sierant, O. Dutta, P/ Horodecki,  and J. Zakrzewski, “At the limits of criticality-based quantum metrology: Apparent super-Heisenberg scaling revisited,” Phys. Rev. X 8, 021022 (2018b).
  • Potok et al. (2007) R. M Potok, I. G. Rau, H. Shtrikman, Y. Oreg,  and D. Goldhaber-Gordon, “Observation of the two-channel Kondo effect,” Nature 446, 167–171 (2007).
  • Iftikhar et al. (2018) Z. Iftikhar, A. Anthore, A. K. Mitchell, F. D. Parmentier, U. Gennser, A. Ouerghi, A. Cavanna, C. Mora, P. Simon,  and F. Pierre, “Tunable quantum criticality and super-ballistic transport in a “charge” Kondo circuit,” Science 360, 1315–1320 (2018).
  • Mihailescu et al. (2023) G. Mihailescu, S. Campbell,  and A. K. Mitchell, “Thermometry of strongly correlated fermionic quantum systems using impurity probes,” Phys. Rev. A 107, 042614 (2023).
  • Wilson (1975) K. G. Wilson, “The renormalization group: Critical phenomena and the Kondo problem,” Rev. Mod. Phys. 47, 773–840 (1975).
  • Bulla et al. (2008) R. Bulla, T. A. Costi,  and T. Pruschke, “Numerical renormalization group method for quantum impurity systems,” Rev. Mod. Phys. 80, 395–450 (2008).
  • Tóth and Apellaniz (2014) G. Tóth and I. Apellaniz, “Quantum metrology from a quantum information science perspective,” J. Phys. A Math. Theor. 47, 424006 (2014).
  • Liu et al. (2019) J. Liu, H. Yuan, X.M. Lu,  and X. Wang, “Quantum fisher information matrix and multiparameter estimation,” J. Phys. A Math. Theor. 53, 023001 (2019).
  • Note (1) We remark that, in the limiting case, when we have a single parameter, λ𝜆\lambdaitalic_λ, that we wish to infer, the Fisher information in Eq. (2) reduces to F(λ)=Eλ[λln(p(xkλ))]2𝐹𝜆subscript𝐸𝜆superscriptdelimited-[]subscript𝜆𝑝conditionalsubscript𝑥𝑘𝜆2F\left(\lambda\right)=E_{\lambda}\left[\partial_{\lambda}\ln{p\left(x_{k}\mid% \lambda\right)}\right]^{2}italic_F ( italic_λ ) = italic_E start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT [ ∂ start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT roman_ln ( start_ARG italic_p ( italic_x start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ∣ italic_λ ) end_ARG ) ] start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT and the CRB becomes Var(λ)(NF)1Var𝜆superscript𝑁𝐹1\text{Var}\left(\lambda\right)\geq{\left(NF\right)}^{-1}Var ( italic_λ ) ≥ ( italic_N italic_F ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, establishing a lower-bound on the mean square error, Var(λ)=Eλ[(λ^({x})λ)2]Var𝜆subscript𝐸𝜆delimited-[]superscript^𝜆𝑥𝜆2\text{Var}\left(\lambda\right)=E_{\lambda}\left[\left(\hat{\lambda}\left(\{x\}% \right)-\lambda\right)^{2}\right]Var ( italic_λ ) = italic_E start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT [ ( over^ start_ARG italic_λ end_ARG ( { italic_x } ) - italic_λ ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ], of any estimator of the parameter λ𝜆\lambdaitalic_λ.
  • Zhu (2015) H. Zhu, “Information complementarity: A new paradigm for decoding quantum incompatibility,” Sci. Rep. 5, 14317 (2015).
  • Heinosaari et al. (2016) T. Heinosaari, T. Miyadera,  and M. Ziman, “An invitation to quantum incompatibility,” J. Phys. A Math. Theor. 49, 123001 (2016).
  • Candeloro et al. (2024) A. Candeloro, Z. Pazhotan,  and M.G.A Paris, “Dimension matters: precision and incompatibility in multi-parameter quantum estimation models,”  (2024), arXiv:2403.07106 [quant-ph] .
  • Bayat et al. (2010) A. Bayat, P. Sodano,  and S. Bose, “Negativity as the entanglement measure to probe the Kondo regime in the spin-chain Kondo model,” Phys. Rev. B 81, 064429 (2010).
  • Weichselbaum and von Delft (2007) A. Weichselbaum and J. von Delft, “Sum-rule conserving spectral functions from the numerical renormalization group,” Phys. Rev. Lett. 99, 076402 (2007).
  • Mitchell et al. (2014) A. K. Mitchell, M. R. Galpin, S. Wilson-Fletcher, D. E. Logan,  and R. Bulla, “Generalized wilson chain for solving multichannel quantum impurity problems,” Phys. Rev. B 89, 121105 (2014).
  • Stadler et al. (2016) K. M. Stadler, A. K. Mitchell, J. von Delft,  and A. Weichselbaum, “Interleaved numerical renormalization group as an efficient multiband impurity solver,” Phys. Rev. B 93, 235101 (2016).
  • Han et al. (2022) C. Han, Z. Iftikhar, Y. Kleeorin, A. Anthore, F. Pierre, Y. Meir, A. K. Mitchell,  and E. Sela, “Fractional entropy of multichannel Kondo systems from conductance-charge relations,” Phys. Rev. Lett. 128, 146803 (2022).
  • Child et al. (2022) T. Child, O. Sheekey, S. Lüscher, S. Fallahi, G. C. Gardner, M. Manfra, A. K. Mitchell, E. Sela, Y. Kleeorin, Y. Meir,  and J. Folk, “Entropy measurement of a strongly coupled quantum dot,” Phys. Rev. Lett. 129, 227702 (2022).
  • Gan (1995) J. Gan, “Map** the critical point of the two-impurity Kondo model to a two-channel problem,” Phys. Rev. Lett. 74, 5287–5287 (1995).