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

Communication, Dynamical Resource Theory, and Thermodynamics

Chung-Yun Hsieh [email protected] ICFO - Institut de Ciències Fotòniques, The Barcelona Institute of Science and Technology, 08860 Castelldefels, Spain
Abstract

Recently, new insights have been obtained by jointly studying communication and resource theory. This interplay consequently serves as a potential platform for interdisciplinary studies. To continue this line, we analyze the role of dynamical resources in a communication setup, and further apply our analysis to thermodynamics. To start with, we study classical communication scenarios constrained by a given resource, in the sense that the information processing channel is unable to supply additional amounts of the resource. We show that the one-shot classical capacity is upper bounded by resource preservability, which is a measure of the ability to preserve the resource. A lower bound can be further obtained when the resource is asymmetry. As an application, unexpectedly, under a recently-studied thermalization model, we found that the smallest bath size needed to thermalize all outputs of a Gibbs-preserving coherence-annihilating channel upper bounds its one-shot classical capacity. When the channel is coherence non-generating, the upper bound is given by a sum of coherence preservability and the bath size of the channel’s incoherent version. In this sense, bath sizes can be interpreted as the thermodynamic cost of transmitting classical information. This finding provides a dynamical analogue of Landauer’s principle, and therefore bridges classical communication and thermodynamics. As another implication, we show that, in bipartite settings, classically correlated local baths can admit classical communication even when both local systems are completely thermalized. Hence, thermalizations can transmit information by accessing only classical correlation as a resource. Our results demonstrate interdisciplinary applications enabled by dynamical resource theory.

I Introduction

Resource is a concept widely used in the study of physics: It can be an effect or a phenomenon, helping us achieve advantages that can never occur in its absence. A quantitative understanding of different resources is thus vital for further applications. For this reason, an approach called resource theory comes, aiming to provide a general strategy to depict different resources RT-RMP .

A resource theory can be interpreted as a triplet (R,R,𝒪R)(R,\mathcal{F}_{R},\mathcal{O}_{R}), consisting of the resource itself RR (e.g., entanglement Ent-RMP ), the set of quantities without the resource R\mathcal{F}_{R} (called free quantities; e.g., separable states), and the set of physical processes that will not generate the resource 𝒪R\mathcal{O}_{R} (called free operations of RR; e.g., local operation and classical communication channels QCI-book ). Every 𝒪R\mathcal{E}\in\mathcal{O}_{R} must satisfy (η)RηR\mathcal{E}(\eta)\in\mathcal{F}_{R}\;\forall\eta\in\mathcal{F}_{R}, which is sometimes called the golden rule of resource theories RT-RMP . Operations satisfying this condition are called resource non-generating for RR, which form the largest possible set of free operations of RR. A resource theory allows ones to quantify the resource via a resource monotone, QRQ_{R}, which is a non-negative-valued function satisfying two conditions: (i) QR(q)=0Q_{R}(q)=0 if qRq\in\mathcal{F}_{R}; and (ii) QR[(q)]QR(q)q&𝒪RQ_{R}[\mathcal{E}(q)]\leq Q_{R}(q)\;\forall q\;\&\;\forall\mathcal{E}\in\mathcal{O}_{R}. This is a “ruler” attributing numbers to different resource contents.

Adopting this general approach, one can study specific resources such as (but not limited to) entanglement Ent-RMP ; Vedral1997 ; Vidal2002 , coherence  Coherence-RMP ; Baumgratz2014 , nonlocality Bell ; Bell-RMP ; Wolfe2019 , steering Wiseman2007 ; Jones2007 ; steering-review ; Skrzypczyk2014 ; Piani2015 ; Gallego2015 ; RMP-steering , asymmetry Gour2008 ; Marvian2016 ; Takagi2019-4 , and athermality Brandao2013 ; Brandao2015 ; Horodecki2013 ; Lostaglio2018 ; Serafini2019 ; Narasimhachar2019 . Together with various features of general resource theories Horodecki2013-2 ; Brandao2015-2 ; del_Rio2015 ; Coecke2016 ; Gour2017 ; Anshu2018 ; Regula2018 ; Bu2018 ; Liu2017 ; Lami2018 ; RT-RMP ; Takagi2019 ; Takagi2019-2 ; Liu2019 ; Fang2019 ; Korzekwa2019 ; Regula2020 , one is able to concretely picture the originally vague notion of resources for states – while the unique roles of dynamical resources have not been noticed until recently. Resource theories of channels footnote0 have therefore drawn much attention lately and been studied intensively Hsieh2017 ; Kuo2018 ; Pirandola2017 ; Dana2017 ; Bu2018 ; Wilde2018 ; Diaz2018 ; Zhuang2018 ; Gour2019-3 ; Theurer2019 ; Seddon2019 ; Rosset2018 ; LiuWinter2019 ; LiuYuan2019 ; Gour2019 ; Gour2019-2 ; Bauml2019 ; Wang2019 ; Berk2019 ; Takagi2019-3 ; Hsieh2020-1 ; Saxena2019 ; Zhang2020 . Unlike the state resources, which are static, channel resources are dynamical properties, thereby providing links to dynamical problems such as communication Takagi2019-3 and resource preservation Hsieh2020-1 ; Saxena2019 .

Very recently, the interplay between resource theories and classical communication has been investigated Korzekwa2019 ; Takagi2019-3 (see also Ref. Kristjansson2020 ), successfully providing new insights and widening our understanding. For instance, a neat proof of the strong converse property of non-signaling assisted classical capacity has been established Takagi2019-3 . Also, amounts of classical messages encodable into the resource content of states has been estimated, and different physical meanings can be concluded by considering specific resources Korzekwa2019 . Hence, the interplay between resource theory and classical communication is a potential platform for interdisciplinary studies. To continue this research line, it is thus necessary to understand communication setups constrained by different static resources. A general treatment on this can clarify the role of static resources in communication and provide potential applications in different physical settings. This motivates us to ask:

How do resource constraints affect classical communications?

In this work, we consider classical communication scenarios where the information processing channel is forbidden to supply additional resources, thereby being a free operation (there are some subtleties about this setting, and we refer the reader to Appendix A for a detailed discussion). The basic setup will be given in Sec. II. In Sec. III, we show that the corresponding one-shot classical capacity is upper bounded by the ability to preserve the resource, which is called resource preservability Hsieh2020-1 , plus a resourceless contribution term. Furthermore, when the underlying resource is asymmetry, a lower bound can be obtained. As an application, we use our approach to bridge classical communication and thermodynamics in Sec. IV: Under the thermalization model introduced in Ref. Sparaciari2019 , the one-shot classical capacity of a Gibbs-preserving coherence non-generating channel is upper bounded by its coherence preservability plus the smallest bath size needed to thermalize all outputs of its incoherent version Sparaciari2019 ; Hsieh2020-1 . A direct physical message is that, under this setting, transmitting classical information through a coherence-annihilating channel necessarily needs a large enough bath size as the thermodynamic cost. This provides a dynamical analogue of the famous Landauer’s principle Landauer1961 and illustrates how dynamical resource theory can connect seemingly different fields, inspiring us to further investigate the interplay of classical communication and thermalization. To this end, in Sec. V we first study a tool related to the ability to simultaneously maintain orthogonality and maximal entanglement. Using this concept, in Sec. VI we found that it is possible for locally performed thermalization channels to globally transmit classical information when the local baths are correlated classically through pre-shared randomness. This result reveals the huge difference between local and global dynamics, providing the recent discovery in Ref. Hsieh2020-2 a generalization and application to communication. Finally, we conclude in Sec. VII.

II Formulation

To process classical information depicted by a finite sequence of integers {m}m=0M1\{m\}_{m=0}^{M-1} through a quantum channel 𝒩\mathcal{N}, one needs to encode them into a set of quantum states {ρm}m=0M1\{\rho_{m}\}_{m=0}^{M-1}; likewise, a decoding is needed to extract the information from outputs of 𝒩\mathcal{N}, which can be done by a positive operator-valued measurement (POVM) {Em}m=0M1\{E_{m}\}_{m=0}^{M-1} QCI-book . They can be written jointly as ΘM=({ρ}m=0M1,{Em}m=0M1)\Theta_{M}=(\{\rho\}_{m=0}^{M-1},\{E_{m}\}_{m=0}^{M-1}), called an MM-code, which depicts the transformation ρmtr[Em𝒩(ρm)]\rho_{m}\mapsto{\rm tr}\left[E_{m}\mathcal{N}(\rho_{m})\right] for each mm. To see how faithfully one can extract the input messages {m}m=0M1\{m\}_{m=0}^{M-1}, the one-shot classical capacity with error ϵ\epsilon Wang2012 ; Datta2013 of 𝒩\mathcal{N} can be defined as a measure:

C(1)ϵ(𝒩)max{log2M|ΘM,ps(ΘM,𝒩)1ϵ},\displaystyle C_{\rm(1)}^{\epsilon}(\mathcal{N})\coloneqq\max\left\{\log_{2}{M}\;|\;\exists\Theta_{M},\;p_{s}(\Theta_{M},\mathcal{N})\geq 1-\epsilon\right\}, (1)

where the average success probability is given by

ps(ΘM,𝒩)1Mm=0M1tr[Em𝒩(ρm)].\displaystyle p_{s}(\Theta_{M},\mathcal{N})\coloneqq\frac{1}{M}\sum_{m=0}^{M-1}{\rm tr}\left[E_{m}\mathcal{N}(\rho_{m})\right]. (2)

Before introducing the main results, we briefly review relevant ingredients of resource preservability Hsieh2020-1 (or simply RR-preservability when the state resource RR is given), which is a dynamical resource depicting the ability to preserve RR. To start with, we impose basic assumptions on a given state resource theory (R,R,𝒪R)(R,\mathcal{F}_{R},\mathcal{O}_{R}):

  1. 1.

    Identity and partial trace are both free operations.

  2. 2.

    𝒪R\mathcal{O}_{R} is closed under tensor products, convex sums, and compositions  footnote:Markovianity .

These assumptions are strict enough for an analytically feasible study and still general enough to be shared by many known resource theories (see Appendix B for a further discussion). For a given state resource theory (R,R,𝒪R)(R,\mathcal{F}_{R},\mathcal{O}_{R}), the induced RR-preservability theory is a channel resource theory written as (RR-preservability, 𝒪RN\mathcal{O}_{R}^{N}, 𝔽R\mathbb{F}_{R}). It is defined on all channels in 𝒪R\mathcal{O}_{R}. In this channel resource theory, the free quantities are members of the set 𝒪RN\mathcal{O}_{R}^{N} called resource annihilating channels Hsieh2020-1 ; footnote:R-DestroyingMap :

𝒪RN{Λ𝒪R|Λ(ρ)Rρ}.\displaystyle\mathcal{O}_{R}^{N}\coloneqq\{\Lambda\in\mathcal{O}_{R}\;|\;\Lambda(\rho)\in\mathcal{F}_{R}\;\forall\rho\}. (3)

They are channels in 𝒪R\mathcal{O}_{R} which can only output free states. A special class of resource annihilating channels are those who cannot output any resourceful state even assisted by ancillary resource annihilating channels; specifically, no RR-preservability can be activated Palazuelos2012 ; Cavalcanti2013 ; Hsieh2016 ; Quintino2016 ; Hsieh2020-1 ; Zhang2020 . Such channels are called absolutely resource annihilating channels Hsieh2020-1 , which are elements of the set 𝒪~RN{Λ~𝒪RN|Λ~Λ𝒪RNΛ𝒪RN}.\widetilde{\mathcal{O}}_{R}^{N}\coloneqq\{\widetilde{\Lambda}\in\mathcal{O}_{R}^{N}\;|\;\widetilde{\Lambda}\otimes\Lambda\in\mathcal{O}_{R}^{N}\;\forall\Lambda\in\mathcal{O}_{R}^{N}\}. Free operations of RR-preservability, which are collectively denoted by the set 𝔽R\mathbb{F}_{R}, are super-channels Chiribella2008 ; Chiribella2008-2 given by Hsieh2020-1 Λ+(Λ~)Λ\mathcal{E}\mapsto\Lambda_{+}\circ(\mathcal{E}\otimes\widetilde{\Lambda})\circ\Lambda_{-} with Λ+,Λ𝒪R\Lambda_{+},\Lambda_{-}\in\mathcal{O}_{R} and Λ~𝒪~RN\widetilde{\Lambda}\in\widetilde{\mathcal{O}}_{R}^{N}. Finally, to quantify RR-preservability, consider a contractive generalized distance measure DD on quantum states; that is, it is a real-valued function satisfying (i) D(ρ,σ)0D(\rho,\sigma)\geq 0 and equality holds if and only if ρ=σ\rho=\sigma, and (ii) (data-processing inequality) D[(ρ),(σ)]D(ρ,σ)D[\mathcal{E}(\rho),\mathcal{E}(\sigma)]\leq D(\rho,\sigma) for all ρ,σ\rho,\sigma and channels \mathcal{E}. The following RR-preservability monotone induced by DD Hsieh2020-1 will be used in this work:

PD()infΛ𝒪RNDR(,Λ),\displaystyle P_{D}(\mathcal{E})\coloneqq\inf_{\Lambda\in\mathcal{O}_{R}^{N}}D^{R}(\mathcal{E},\Lambda), (4)

where DR(S,ΛS)supΛ~A,ρSAD[(SΛ~A)(ρSA),(ΛSΛ~A)(ρSA)]D^{R}(\mathcal{E}_{\rm S},\Lambda_{\rm S})\coloneqq\sup_{\widetilde{\Lambda}_{\rm A},\rho_{\rm SA}}D[(\mathcal{E}_{\rm S}\otimes\widetilde{\Lambda}_{\rm A})(\rho_{\rm SA}),(\Lambda_{\rm S}\otimes\widetilde{\Lambda}_{\rm A})(\rho_{\rm SA})] maximizes over every possible ancillary system A{\rm A}, joint input ρSA\rho_{\rm SA}, and absolutely resource annihilating channel Λ~SA\widetilde{\Lambda}_{\rm SA}. Geometrically, Eq. (4) can be understood as a distance between \mathcal{E} and the set 𝒪RN\mathcal{O}_{R}^{N} that is adjusted by absolutely resource annihilating channels. With Assumptions 1 and 2, in Appendix B we show that PDP_{D} is indeed a monotone Hsieh2020-1 , in the sense that it is a non-negative-valued function such that PD()=0P_{D}(\mathcal{E})=0 if 𝒪RN\mathcal{E}\in\mathcal{O}_{R}^{N} footnote:FaithfulnessRemark , and PD[𝔉()]PD()P_{D}[\mathfrak{F}(\mathcal{E})]\leq P_{D}(\mathcal{E}) for every channel \mathcal{E} and 𝔉𝔽R\mathfrak{F}\in\mathbb{F}_{R}. Note that, in this work, we only ask RR-preservability monotones to satisfy these two conditions, which are the core features of a monotone. Further properties, such as Eq. (7) in Ref. Hsieh2020-1 , will need additional assumptions, and we leave the details in Appendix B. Finally, the distance measure that will be mainly used in this work is the max-relative entropy Datta2009 for states defined as (and, conventionally, we adopt inf=\inf\emptyset=\infty)

Dmax(ρσ)log2inf{λ0|ρλσ}.\displaystyle D_{\rm max}(\rho\|\sigma)\coloneqq\log_{2}\inf\{\lambda\geq 0\,|\,\rho\leq\lambda\sigma\}. (5)

Hence, PDmax(𝒩)P_{D_{\rm max}}(\mathcal{N}) is the minimal amount of noise one needs to add to turn 𝒩\mathcal{N} into resource annihilating [see also Eq. (C)]. For a given error κ>0\kappa>0, we define 𝒪RN(κ;𝒩){Λ𝒪RN||DmaxR(𝒩Λ)PDmax(𝒩)|κ}\mathcal{O}_{R}^{N}(\kappa;\mathcal{N})\coloneqq\{\Lambda\in\mathcal{O}_{R}^{N}\,|\,|D_{\rm max}^{R}(\mathcal{N}\|\Lambda)-P_{D_{\rm max}}(\mathcal{N})|\leq\kappa\}. It contains resource annihilating channels achievable by adding the smallest amount of noise to 𝒩\mathcal{N}, up to the error κ\kappa. We call them resourceless versions of 𝒩\mathcal{N} up to the error κ\kappa, and use the notation Λ𝒩𝒪RN(κ;𝒩)\Lambda^{\mathcal{N}}\in\mathcal{O}_{R}^{N}(\kappa;\mathcal{N}) to emphasize their dependence on 𝒩\mathcal{N}.

III Bounds On Classical Capacity

To introduce the first result, define

Γκ(𝒩)log2infΛ𝒩𝒪RN(κ;𝒩)supΘMm=0M1tr[EmΛ𝒩(ρm)],\displaystyle\Gamma_{\kappa}(\mathcal{N})\coloneqq\log_{2}\inf_{\Lambda^{\mathcal{N}}\in\mathcal{O}_{R}^{N}(\kappa;\mathcal{N})}\sup_{\Theta_{M}}\sum_{m=0}^{M-1}{\rm tr}\left[E_{m}\Lambda^{\mathcal{N}}(\rho_{m})\right], (6)

where supΘM\sup_{\Theta_{M}} maximizes over every MM\in\mathbb{N} and MM-code ΘM\Theta_{M}. 2Γκ(𝒩)2^{\Gamma_{\kappa}(\mathcal{N})} tells us the highest number of discriminable states through every resourceless version of 𝒩\mathcal{N}, up to the error κ\kappa. We also consider PDδ()inf2δPD()P_{D}^{\delta}(\mathcal{E})\coloneqq\inf_{\left\|\mathcal{E}-\mathcal{E}^{\prime}\right\|_{\diamond}\leq 2\delta}P_{D}(\mathcal{E}^{\prime}) and Γκδ()sup2δΓκδ()\Gamma_{\kappa}^{\delta}(\mathcal{E})\coloneqq\sup_{\left\|\mathcal{E}-\mathcal{E}^{\prime}\right\|_{\diamond}\leq 2\delta}\Gamma_{\kappa}^{\delta}(\mathcal{E}^{\prime}), which smooth the optimizations over all channels \mathcal{E}^{\prime} closed to \mathcal{E}. Also, supA,ρSA(A)(ρSA)1\left\|\mathcal{E}\right\|_{\diamond}\coloneqq\sup_{{\rm A},\rho_{\rm SA}}\left\|(\mathcal{E}\otimes\mathcal{I}_{\rm A})(\rho_{\rm SA})\right\|_{1} is the diamond norm. Combining Refs. Takagi2019-3 ; Hsieh2020-1 , in Appendix C we prove the following upper bound:

Theorem 1.

Given ϵ,δ0& 0<κ<1\epsilon,\delta\geq 0\;\&\;0<\kappa<1 satisfying ϵ+δ<1\epsilon+\delta<1. Then for every 𝒩𝒪R\mathcal{N}\in\mathcal{O}_{R} we have

C(1)ϵ(𝒩)PDmaxδ(𝒩)+Γκδ(𝒩)+log22κ1ϵδ.\displaystyle C_{\rm(1)}^{\epsilon}(\mathcal{N})\leq P_{D_{\rm max}}^{\delta}(\mathcal{N})+{\Gamma_{\kappa}^{\delta}(\mathcal{N})}+\log_{2}\frac{2^{\kappa}}{1-\epsilon-\delta}. (7)

This upper bound Eq. (7) can be interpreted as follows: It is the highest amount of carriable classical information by every resourceless version of 𝒩\mathcal{N}, i.e., Γκδ(𝒩){\Gamma_{\kappa}^{\delta}}(\mathcal{N}), plus the contribution from the ability of 𝒩\mathcal{N} to preserve RR, i.e., PDmaxδ(𝒩)P_{D_{\rm max}}^{\delta}(\mathcal{N}). This also suggests that C(1)ϵ(𝒩)Γκδ(𝒩)C_{\rm(1)}^{\epsilon}(\mathcal{N})-{\Gamma_{\kappa}^{\delta}(\mathcal{N})} characterizes the resource advantage, since it estimates the amount of transmissible classical information via the ability of 𝒩\mathcal{N} to preserve RR. To see the tightness of Eq. (7) (up to error terms containing ϵ,δ\epsilon,\delta), the inequality is saturated by every state preparation channel outputting a fixed free state ()η(\cdot)\mapsto\eta with ηR\eta\in\mathcal{F}_{R}. There also exist examples attending the inequality with non-zero classical capacity. For instance, in a dd-dimensional system, when RR is coherence and 𝒩\mathcal{N} is the dephasing channel, i.e., ()i=1d|ii||ii|(\cdot)\mapsto\sum_{i=1}^{d}|i\rangle\langle i|\cdot|i\rangle\langle i|, we have both side as log2d\log_{2}d.

As the last remark on Eq. (7), when the optimal amount of classical information can be encoded into free states, the optimal capacity should intuitively be achievable by channels with zero RR-preservability, and any general result should respect this fact. Hence, resource preservability monotone cannot upper bound classical capacity solely, and Theorem 1 is consistent with this expectation due to the term Γκδ(𝒩)\Gamma_{\kappa}^{\delta}(\mathcal{N}). However, such resourceless advantages no longer exist when more constraints are made for specific purposes (e.g., Sec. V).

Theorem 1 can link different physical properties to classical communication, which is illustrated by the following result. Let PD|RP_{D|R} denote Eq. (4) with the state resource RR and write R=γR=\gamma when the state resource is the athermality induced by the thermal state γ\gamma (we postpone its formal definition to Sec. IV). Then in Appendix C we show that:

Corollary 2.

Given 0ϵ<1& 0<κ<10\leq\epsilon<1\;\&\;0<\kappa<1 and a full-rank thermal state γ\gamma. For 𝒩𝒪R\mathcal{N}\in\mathcal{O}_{R} that is also Gibbs-preserving and every Λ𝒩𝒪RN(κ;𝒩)\Lambda^{\mathcal{N}}\in\mathcal{O}_{R}^{N}(\kappa;\mathcal{N}), we have

C(1)ϵ(𝒩)PDmax|R(𝒩)+PDmax|γ(Λ𝒩)+log22κ1ϵ.\displaystyle C_{\rm(1)}^{\epsilon}(\mathcal{N})\leq P_{D_{\rm max}|R}(\mathcal{N})+P_{D_{\rm max}|\gamma}\left(\Lambda^{\mathcal{N}}\right)+\log_{2}\frac{2^{\kappa}}{1-\epsilon}. (8)

Corollary 2 implies that the ability of a Gibbs-preserving 𝒩𝒪R\mathcal{N}\in\mathcal{O}_{R} to transmit classical information is limited by its ability to preserve RR, plus the ability of its resourceless version (to RR) to preserve athermality. This observation will allow us to prove one of the main results of this work in Sec. IV.

III.1 Asymmetry and Lower Bounds

When the underlying state resource is the asymmetry of a given unitary group GG, a lower bound on the classical capacity can also be obtained. Formally, asymmetry of a given group GG, or simply GG-asymmetry, has free states as those invariant under group actions; that is, ρ=UρU\rho=U\rho U^{\dagger} for all UGU\in G. One option of free operations, which is adopted here, is GG-covariant channels, which are channels commuting with unitaries in GG: U()U=[U()U]U\mathcal{E}(\cdot)U^{\dagger}=\mathcal{E}[U(\cdot)U^{\dagger}] for all UGU\in G (see, e.g., Ref. MarvianThesis ; Marvian2013 ; QRF-RMP ; Takagi2019-4 ).

To introduce the result, we need to use the information spectrum relative entropy Hayashi2003 ; Tomamichel2013 (see also Ref. Korzekwa2019 ) given by Dsδ(ρσ)sup{ω|tr(ρΠρ2ωσ)δ},D_{s}^{\delta}(\rho\|\sigma)\coloneqq\sup\{\omega\,|\,{\rm tr}\left(\rho\Pi_{\rho\leq 2^{\omega}\sigma}\right)\leq\delta\}, where Πρ2ωσ\Pi_{\rho\leq 2^{\omega}\sigma} is the projection onto the union of eigenspaces of 2ωσρ2^{\omega}\sigma-\rho with non-negative eigenvalues Korzekwa2019 . Despite its name, the information spectrum relative entropy is not a proper contractive distance measure, since it will not satisfy data-processing inequality and can output negative values Leditzky-PhD . However, it allows us to obtain a lower bound on C(1)C_{\rm(1)}. In Appendix D, we apply results in Ref. Korzekwa2019 and show the following bound (now 𝒪RN\mathcal{O}_{R}^{N} denotes the set of GG-covariant channels that cannot preserve any GG-asymmetry):

Theorem 3.

Given R=R= GG-asymmetry, then for every GG-covariant channel 𝒩\mathcal{N} and 0δ<ϵ<10\leq\delta<\epsilon<1, we have

max{0;1ln2P~Dsϵδ(𝒩)+log2δ1}C(1)ϵ(𝒩),\displaystyle\max\left\{0;\frac{1}{\ln{2}}\widetilde{P}_{D_{s}^{\epsilon-\delta}}(\mathcal{N})+\log_{2}\delta-1\right\}\leq C_{\rm(1)}^{\epsilon}(\mathcal{N}), (9)

where P~Dsϵδ(𝒩)infΛ𝒪RNsupρDsϵδ[𝒩(ρ)Λ(ρ)]\widetilde{P}_{D_{s}^{\epsilon-\delta}}(\mathcal{N})\coloneqq\inf_{\Lambda\in\mathcal{O}_{R}^{N}}\sup_{\rho}D_{s}^{\epsilon-\delta}[\mathcal{N}(\rho)\,\|\,\Lambda(\rho)].

This provides an RR-preservability-like lower bound on the one-shot classical capacity for GG-covariant channels, which also shows a witness of resourceful advantages, i.e., C(1)ϵ(𝒩)Γκδ(𝒩)C_{\rm(1)}^{\epsilon}(\mathcal{N})-{\Gamma_{\kappa}^{\delta}(\mathcal{N})}. Using (UU)(U\otimes U^{*})-asymmetry as an example (see also Appendix C.2 for more details), the advantage from asymmetry can be witnessed when P~Dsϵδ(𝒩)>2ln2+lnd2d21lnδ\widetilde{P}_{D_{s}^{\epsilon-\delta}}(\mathcal{N})>2\ln{2}+\ln\frac{d^{2}}{d^{2}-1}-\ln\delta, which is approximately P~Dsϵδ(𝒩)>2ln2lnδ\widetilde{P}_{D_{s}^{\epsilon-\delta}}(\mathcal{N})>2\ln 2-\ln\delta when d1d\gg 1.

IV Application: Classical Communications And Thermodynamics

It is worth mentioning that our result bridges two seemingly different physical concepts: Classical capacity Takagi2019-3 and heat bath size needed for thermalization Sparaciari2019 ; Hsieh2020-1 . To introduce the result, we give a quick review of the resource theory of athermality and related ingredients for thermalization bath sizes Sparaciari2019 . Athermality is the status out of thermal equilibrium. With a fixed system dimension dd, the unique free state is the thermal equilibrium state, or the thermal state. With a given system Hamiltonian HSH_{\rm S} and temperature TT, the thermal state is uniquely given by γ=eβHStr(eβHS),\gamma=\frac{e^{-\beta H_{\rm S}}}{{\rm tr}(e^{-\beta H_{\rm S}})}, where β=1kBT\beta=\frac{1}{k_{B}T} is the inverse temperature and kBk_{B} is the Boltzmann constant. For multiple systems with tensor product, all free states in this resource theory are γk\gamma^{\otimes k} for some positive integer kk (i.e., all allowed dimensions are dkd^{k} with some kk). In this work, we adopt Gibbs-preserving channels as the free operations. They are channels \mathcal{E} keeping thermal states invariant: (γk)=γl\mathcal{E}(\gamma^{\otimes k})=\gamma^{\otimes l}, where dkd^{k} and dld^{l} are the input and output dimensions, respectively. Physically, these are dynamics that will not drive thermal equilibrium away from equilibrium.

To formally study thermalization, we follow Ref. Sparaciari2019 and define a channel (jointly acting on system S{\rm S} plus bath B{\rm B}) SB:SBSB\mathcal{E}_{\rm SB}:{\rm SB}\to{\rm SB} to ϵ\epsilon-thermalize a system state ρS\rho_{\rm S} if

SB(ρSγ(n1))γn1ϵ.\displaystyle\left\|\mathcal{E}_{\rm SB}\left(\rho_{\rm S}\otimes\gamma^{\otimes(n-1)}\right)-\gamma^{\otimes n}\right\|_{1}\leq\epsilon. (10)

That is, SB\mathcal{E}_{\rm SB} needs to globally thermalize SB{\rm SB}, where the thermal state γ\gamma is determined by the Hamiltonian and the temperature of S{\rm S}, and the initial state of B{\rm B} is the n1n-1 copies of γ\gamma. To depict such thermalization processes dynamically, we consider the collision model introduced in Ref. Sparaciari2019 . To avoid complexity, we refer the reader to Appendix E for a brief introduction of this model, and here we let 𝒞n\mathcal{C}_{n} be the set of all channels on SB{\rm SB} that can be realized by this model. Then the quantity nγϵ(ρS)inf{n|SB𝒞ns.t.Eq.(10)holds}n^{\epsilon}_{\gamma}(\rho_{\rm S})\coloneqq\inf\{n\in\mathbb{N}\,|\,\exists\,\mathcal{E}_{\rm SB}\in\mathcal{C}_{n}\;{\rm s.t.\;Eq.~{}\eqref{Eq:Thermalize-Def}\;holds}\} can be understood as the smallest bath size needed to ϵ\epsilon-thermalize ρS\rho_{\rm S} under this model Sparaciari2019 . This concept can be generalized to any channel 𝒩\mathcal{N} by defining Hsieh2020-1

γϵ(𝒩)supρnγϵ[𝒩(ρ)]1,\displaystyle\mathcal{B}^{\epsilon}_{\gamma}(\mathcal{N})\coloneqq\sup_{\rho}n^{\epsilon}_{\gamma}[\mathcal{N}(\rho)]-1, (11)

which maximizes over all the smallest bath sizes among all outputs of 𝒩\mathcal{N}. This is thus the smallest bath size needed to ϵ\epsilon-thermalize all outputs of 𝒩\mathcal{N} under the given collision model.

Now we mention a core assumption made in Ref. Sparaciari2019 used to regularize the analysis. A given Hamiltonian HH with energy levels {Ei}i=1d\{E_{i}\}_{i=1}^{d} is said to satisfy the energy subspace condition if for every positive integer MM and every pair of different vectors {𝐦𝐦}({0})d\{{\bf m}\neq{\bf m}^{\prime}\}\subset\left(\mathbb{N}\cup\{0\}\right)^{d} satisfying i=1dmi=i=1dmi=M\sum_{i=1}^{d}m_{i}=\sum_{i=1}^{d}m^{\prime}_{i}=M, we have i=1dmiEii=1dmiEi\sum_{i=1}^{d}m_{i}E_{i}\neq\sum_{i=1}^{d}m^{\prime}_{i}E_{i}. Hence, energy levels cannot be integer multiples of each other, and energy degeneracy is also forbidden. As an application of Theorem 1 (and Corollary 2), in Appendix F we show the following bound [we implicitly assume the system Hamiltonian is the one realizing the given thermal state γ\gamma with some temperature, and its energy eigenbasis defines the coherence, R=CohR={\rm Coh}; also, we use pmin(γ)p_{\rm min}(\gamma) to denote the smallest eigenvalue of γ\gamma]:

Theorem 4.

Given 0ϵ,δ<1& 0<κ<10\leq\epsilon,\delta<1\;\&\;0<\kappa<1 and a full-rank thermal state γ\gamma. Assume the system Hamiltonian satisfies the energy subspace condition. Then for a Gibbs-preserving map 𝒩\mathcal{N} of γ\gamma that is also coherence non-generating, we have

C(1)ϵ(𝒩)\displaystyle C_{\rm(1)}^{\epsilon}(\mathcal{N})\leq PDmax|Coh(𝒩)+log2(γδ(Λ𝒩)+2δpmin(γ)+1)\displaystyle P_{D_{\rm max}|{\rm Coh}}(\mathcal{N})+\log_{2}\left(\mathcal{B}^{\delta}_{\gamma}\left({\Lambda^{\mathcal{N}}}\right)+\frac{2\sqrt{\delta}}{p_{\rm min}(\gamma)}+1\right)
+log22κ1ϵ\displaystyle+\log_{2}\frac{2^{\kappa}}{1-\epsilon} (12)

for every Λ𝒩𝒪CohN(κ,𝒩)\Lambda^{\mathcal{N}}\in\mathcal{O}_{\rm Coh}^{N}(\kappa,\mathcal{N}).

Theorem 4 illustrates how a dynamical resource theory can bridge a pure thermodynamic quantity to a pure communication quantity. To illustrate this, let us first consider the special case when 𝒩\mathcal{N} is coherence-annihilating footnote:Coherence-Annihilating . In this case, one is able to choose 𝒩=Λ𝒩\mathcal{N}=\Lambda^{\mathcal{N}} and κ=0\kappa=0. Within this setup, if 𝒩\mathcal{N} can communicate a high amount of classical information, it necessarily requires a large bath to thermalize all its outputs. On the other hand, if this channel has a small thermalization bath size, it unavoidably has a weak ability to communicate classical information. Importantly, Theorem 4 provides a physical message that is in spirit similar to the Landauer’s principle Landauer1961 . Landauer’s principle says that preparing a pre-defined pure state, e.g., |0|0\rangle, from an initial state ρ\rho requires at least S(ρ)kBTln2S(\rho)k_{B}T\ln 2 amount of energy del_Lio2011 , where S(ρ)tr(ρlog2ρ)S(\rho)\coloneqq-{\rm tr}(\rho\log_{2}\rho) is the von Neumann entropy of ρ\rho. In this sense, energy can be regarded as the thermodynamic cost needed to create classical information carried by an orthonormal basis {|m}m=0M1\{|m\rangle\}_{m=0}^{M-1}. Theorem 4 provides a different, dynamical perspective: Under the given setting, transmitting nn bits of classical information, i.e., C(1)ϵ(𝒩)=nC_{\rm(1)}^{\epsilon}(\mathcal{N})=n, necessarily requires the bath size γϵ(𝒩)\mathcal{B}^{\epsilon}_{\gamma}(\mathcal{N}) to be at least 2n12^{n}-1, up to some small terms. In this case, the bath size can be interpreted as the thermodynamic cost needed to transmit classical information. When the ability of the channel 𝒩\mathcal{N} to preserve coherence is turned on, interestingly, the cost to transmit classical information becomes a hybrid term: It is the bath size of 𝒩\mathcal{N}’s incoherent version [i.e., Λ𝒩𝒪CohN(κ,𝒩)\Lambda^{\mathcal{N}}\in\mathcal{O}_{\rm Coh}^{N}(\kappa,\mathcal{N})] plus the ability of 𝒩\mathcal{N} to preserve coherence. In other words, this is the sum of the thermodynamic cost of 𝒩\mathcal{N}’s classical counterpart, and the quantum effect maintained by 𝒩\mathcal{N}. By treating Landauer’s principle as a bridge between thermodynamics and a static property of classical information, Theorem 4 brings a connection between thermodynamics and a dynamical feature of classical information.

Note that, as expected, a full thermalization process (i.e., a state preparation channel of the given thermal state) cannot transmit any amount of classical information, since the bath size needed for thermalization is precisely zero. This also implies that the inequality in Theorem 4 is tight. However, locally-performed thermalization processes can actually allow global transmission of classical information, which only needs shared randomness as a resource. In this sense, thermalization together with a classical resource can still achieve nontrivial classical communication. We will introduce this result more in-dept in Sec. VI, and the following section is for a tool needed for its proof.

V Application: Maintaining Orthogonal Maximal Entanglement

Once a question can be formulated into a classical communication problem, our approach can be used to study connections between the given question and different resource constraints. To illustrate this, we study the following question: How robust is the structure of orthogonal maximal entanglement under dynamics? As the motivation, a maximally entangled basis is a well-known tool in quantum information science, promising applications such as quantum teleportation Bennett1993 and superdense coding QCI-book . The key is the simultaneous existence of maximal entanglement and orthogonality, and maintaining both of them through a physical evolution is vital for applications afterward. To model this question, we impose two restrictions in the classical communication scenarios used in this work: (i) The encoding {ρm}m=0M1\{\rho_{m}\}_{m=0}^{M-1} are mutually orthonormal maximally entangled states {|Φm}m=0M1\{|\Phi_{m}\rangle\}_{m=0}^{M-1}. (ii) The decoding {Em}m=0M1\{E_{m}\}_{m=0}^{M-1} are projective measurements done by a (sub-)basis of orthogonal maximally entangled states {|ΦmΦm|}m=0M1\{|\Phi^{\prime}_{m}\rangle\langle\Phi^{\prime}_{m}|\}_{m=0}^{M-1}. The corresponding one-shot classical capacity characterizes the ability of a given channel 𝒩\mathcal{N} footnote:Local-Dimension to simultaneously maintain orthogonality and maximal entanglement:

CME,(1)ϵ(𝒩)log2max{M|ps|ME(M,𝒩)1ϵ},\displaystyle C_{{\rm ME},(1)}^{\epsilon}(\mathcal{N})\coloneqq\log_{2}\max\{M\,|\,p_{s|{\rm ME}}(M,\mathcal{N})\geq 1-\epsilon\}, (13)

where the success probability reads ps|ME(M,𝒩)sup{|Φm},{|Φm}1Mm=0M1Φm|𝒩(|ΦmΦm|)|Φm,p_{s|{\rm ME}}(M,\mathcal{N})\coloneqq\sup_{\{|\Phi_{m}\rangle\},\{|\Phi^{\prime}_{m}\rangle\}}\frac{1}{M}\sum_{m=0}^{M-1}\langle\Phi^{\prime}_{m}|\mathcal{N}(|\Phi_{m}\rangle\langle\Phi_{m}|)|\Phi^{\prime}_{m}\rangle, and the maximization is taken over all sets of orthogonal maximally entangled states of size MM, denoted by {|Φm},{|Φm}\{|\Phi_{m}\rangle\},\{|\Phi^{\prime}_{m}\rangle\}. Thus, CME,(1)ϵ(𝒩)C_{{\rm ME},(1)}^{\epsilon}(\mathcal{N}) is the highest maintainable pairs of mutually orthonormal maximally entangled states under the dynamics 𝒩\mathcal{N}, up to an error smaller than ϵ\epsilon. To introduce the result, we say a state ρ\rho is multi-copy nonlocal/steerable Palazuelos2012 ; Cavalcanti2013 ; Hsieh2016 ; Quintino2016 if there exists an integer kk such that ρk\rho^{\otimes k} is nonlocal/steerable. Also, recall that 𝔽R\mathbb{F}_{R} is the set of free operation of RR-preservability defined in Sec. II. Then in Appendix G we show that

Theorem 5.

For a given 𝒩𝒪R\mathcal{N}\in\mathcal{O}_{R} and 0ϵ,δ<10\leq\epsilon,\delta<1 satisfying ϵ+δ<1\epsilon+\delta<1, we have

α×sup𝔉𝔽RCME,(1)ϵ[𝔉(𝒩)]PDmaxδ(𝒩)+log211ϵδ\displaystyle\alpha\times\sup_{\mathfrak{F}\in\mathbb{F}_{R}}C^{\epsilon}_{{\rm ME},(1)}\left[\mathfrak{F}(\mathcal{N})\right]\leq P^{\delta}_{D_{\rm max}}(\mathcal{N})+\log_{2}\frac{1}{1-\epsilon-\delta} (14)

with α=1\alpha=1 when R=R= athermality; α=12\alpha=\frac{1}{2} when R=R= entanglement, free entanglement Horodecki1998 , multi-copy nonlocality, and multi-copy steerability.

Theorem 5 provides upper bounds on CME,(1)(𝒩)C_{\rm ME,(1)}(\mathcal{N}) when 𝒩\mathcal{N} is a free operation of specific resources, which holds even when the channel is assisted by additional structures constrained by the resource (which is given by 𝔽R\mathbb{F}_{R}). Theorem 5 also brings an alternative operational interpretation of RR-preservability: For the resources RR mentioned above, RR-preservability bounds the channel’s simultaneous maintainability of orthogonality and maximal entanglement in resource-constrained scenarios.

We remark that one can also interpret Eq. (13) as a measure of the ability to admit superdense coding, and Theorem 5 therefore serves as an upper bound on this ability. Furthermore, Theorem 5 brings a connection between fully entangled fraction Horodecki1999-2 ; Albeverio2002 and RR-preservability. We leave the details in Appendix H.

VI Application: Transmitting Information Through Thermalization

The connection between Communication and Thermodynamics given by Theorem 4 implies that it is impossible for a full thermalization, i.e., Φγ:()γtr()\Phi_{\gamma}:(\cdot)\mapsto\gamma{\rm tr}(\cdot) for a given thermal state γ\gamma, to transmit any amount of classical information. This is consistent with our physical intuition since a full thermalization means a complete destroy of the input information, leaving no freedom for the output. However, this impossibility can be flipped with the help of pre-shared randomness, which is a pure classical resource. More precisely, when local baths in a bipartite setting are allowed to be correlated through shared randomness, it is possible for locally performed full thermalizations to globally transmit classical information.

To make the statement precise, let us recall the following notions from Ref. Hsieh2020-2 . A bipartite channel AB\mathcal{E}_{\rm AB} on AB{\rm AB} is called a local thermalization to a given pair of thermal states (γA,γB)(\gamma_{\rm A},\gamma_{\rm B}) if (1) there exist an ancillary system AB{\rm A^{\prime}B^{\prime}}, unitary channels 𝒰AA,𝒰BB\mathcal{U}_{\rm AA^{\prime}},\mathcal{U}_{\rm BB^{\prime}}, and a thermal state γ~AB\widetilde{\gamma}_{\rm A^{\prime}B^{\prime}} separable in AB{\rm AB} bipartition such that AB(ρ)=trAB[𝒰AA𝒰BB(ργ~AB)],\mathcal{E}_{\rm AB}(\rho)={\rm tr}_{\rm A^{\prime}B^{\prime}}\left[\mathcal{U}_{\rm AA^{\prime}}\otimes\mathcal{U}_{\rm BB^{\prime}}(\rho\otimes\widetilde{\gamma}_{\rm A^{\prime}B^{\prime}})\right], and (2) trBAB(ρ)=γA{\rm tr}_{\rm B}\circ\mathcal{E}_{\rm AB}(\rho)=\gamma_{\rm A} and trAAB(ρ)=γB{\rm tr}_{\rm A}\circ\mathcal{E}_{\rm AB}(\rho)=\gamma_{\rm B} for all ρ\rho. As mentioned in Ref. Hsieh2020-2 , local thermalization is local in two senses: It is a local operation plus pre-shared randomness channel, and it is locally equivalent to a full thermalization process. The main message from Ref. Hsieh2020-2 is that entanglement preserving local thermalization exists; that is, for every non-pure full-rank γA,γB\gamma_{\rm A},\gamma_{\rm B}, there exists a local thermalization AB\mathcal{E}_{\rm AB} such that AB(ρ)\mathcal{E}_{\rm AB}(\rho) is entangled for some ρ\rho. It turns out that such channels can also transmit classical information. To state the result, consider a tripartite system ABC{\rm ABC} with local dimension d,d,d2+1d,d,d^{2}+1, respectively. We focus on the bipartition A|BC{\rm A|BC}; namely, the subsystem C{\rm C} can be understood as an ancillary system possessed by the agent in B{\rm B}. Let γX\gamma_{{\rm X}} be the thermal state of the subsystem X=A,B,C{\rm X=A,B,C} with temperature TXT_{\rm X} and Hamiltonian HXH_{\rm X}. Following Ref. Hsieh2020-2 , we assume each HXH_{\rm X} is non-degenerate and finite-energy. Then, Combining Ref. Hsieh2020-2 and results in Sec. V, in Appendix I we prove that locally performed thermalization is able to transmit classical information with the help of pre-shared randomness [let pmin|ABmin{pmin(γA);pmin(γB)}p_{\rm min|AB}\coloneqq\min\{p_{\rm min}(\gamma_{{\rm A}});p_{\rm min}(\gamma_{{\rm B}})\} and γBCγBγC\gamma_{\rm BC}\coloneqq\gamma_{\rm B}\otimes\gamma_{\rm C}; see also Fig. 1]:

Theorem 6.

When d<d<\infty and TX>0T_{{\rm X}}>0, there exists an entanglement preserving local thermalization to (γA,γBC)(\gamma_{\rm A},\gamma_{\rm BC}), denoted by A|BC\mathcal{E}_{\rm A|BC}, such that

C(1)ϵ(A|BC)log2d2\displaystyle C_{\rm(1)}^{\epsilon}(\mathcal{E}_{\rm A|BC})\geq\log_{2}d^{2} (15)

for every ϵ(11d2)(1dpmin|AB)\epsilon\geq\left(1-\frac{1}{d^{2}}\right)\left(1-dp_{\rm min|AB}\right).

Under the global dynamics given by Theorem 6, although the local agents will observe a full thermalization process, transmitting classical information is still possible via the joint bipartite dynamics. This is because classical information can be locally encoded in the global quantum correlation. Note that the amount of transmissible classical information largely depends on local temperatures: When the local systems are too cold (which corresponds to the limit pmin|AB0p_{\rm min|AB}\to 0), a low temperature will force the local thermal state be closed to a pure state (here we assume no energy degeneracy), resulting in no possibility to maintain global correlation. Also note that although A|BC\mathcal{E}_{\rm A|BC} is locally identical to a full thermalization process, its global behavior is far from a thermalization. This gap between local and global dynamics leads to the possibility for a classical communication through local thermalizations when shared randomness is accessed to. Notably, this illustrate that, although being a classical resource, shared randomness provides advantages to maintain quantum correlation and assist classical communication through thermalization processes.

Refer to caption
Figure 1: Schematically, Theorem 6 implies that when local baths are classically correlated (the dashed line), it is possible to transmit classical information through locally performed full thermalization processes.

VII Conclusion

We study classical communication scenarios with free operations of a given resource as the information processor. The one-shot classical capacity is upper bounded by resource preservability Hsieh2020-1 plus a term of resourceless contribution. This upper bound provides an alternative interpretation of resource preservability. Furthermore, when asymmetry is the resource, a lower bound can also be obtained.

As an application, we use our result to bridge two seemingly different concepts: Under the thermalization model given by Ref. Sparaciari2019 , for every Gibbs-preserving coherence non-generating channel, its smallest channel bath size (i.e., a bath size needed to thermalize all outputs of a given channel) plus its coherence preservability will upper bound its one-shot classical capacity. Thus, under this setting, transmitting nn bits of classical information through a coherence-annihilating channel requires the channel bath size to be at least 2n12^{n}-1, up to some small terms. This reveals an implicit thermodynamic cost of transmitting classical information, providing a dynamical analogue of the Landauer’s principle Landauer1961 and illustrating how a dynamical resource theory allows applications to connect different dynamical phenomena.

We further apply our approach to study channel’s simultaneous maintainability of orthogonality and maximal entanglements. Formulating the question into a communication form, a capacity-like measure can be introduced and upper bounded by resource preservability.

Finally, applying the thermalization channel introduced in Ref. Hsieh2020-2 to a bipartite setting, we show that classically correlated local baths allow a decent amount of one-shot classical capacity even when both local systems are completely thermalized. Hence, classical information processing and thermalization processes can coexist, which only requires shared randomness as a resource. This result also means that when a many-body system is in contact with a global bath having classical correlations within, it is possible to maintain classical information even after it has been thermalized locally. Share randomness as a resource is enough to guarantee that certain amounts of classical information can be extracted after local thermalizations, provided that local temperatures are not too low.

Several open questions remain. First, currently all the results related to resource preservability (e.g., Ref. Hsieh2020-1 ) are focusing on the one-shot regime, and a future direction is to understand its asymptotic behavior. Furthermore, whether one can derive a lower bound similar to Theorem 3 in terms of resource preservability and even extend the result to other state resources are still unknown. These questions could be difficult and largely depend on the choice of resources, since, e.g., the lower bound on the capacity used in Ref. Korzekwa2019 is given by the information spectrum relative entropy, which is not a contractive generalized distance measure Leditzky-PhD and hence cannot induce legal resource preservability monotone. Also, whether one can obtain any result similar to Theorem 5 in the context of coherence is still unknown. Finally, as a direct consequence of Theorem 4, we have the following conjecture: Transmitting nn bits of classical information through a Gibbs-preserving coherence-annihilating channel requires the corresponding channel bath size to be at least 2n12^{n}-1. This conjecture could largely depend on the underlying communication setup and the thermalization model.

We hope the physical messages provided by this work can offer alternative interpretations in the interplay of dynamical resource theory, classical communication, thermodynamics, and different forms of inseparability.

Acknowledgements

We thank Antonio Acín, Stefan Ba¨\ddot{\rm a}uml, Dario De Santis, Yeong-Cherng Liang, Matteo Lostaglio, Mohammad Mehboudi, Gabriel Senno, and Ryuji Takagi for fruitful discussions and comments. This project is part of the ICFOstepstone - PhD Programme for Early-Stage Researchers in Photonics, funded by the Marie Skłodowska-Curie Co-funding of regional, national and international programmes (GA665884) of the European Commission, as well as by the ‘Severo Ochoa 2016-2019’ program at ICFO (SEV-2015-0522), funded by the Spanish Ministry of Economy, Industry, and Competitiveness (MINECO). We also acknowledge support from the Spanish MINECO (Severo Ochoa SEV-2015-0522), Fundació Cellex and Mir-Puig, Generalitat de Catalunya (SGR1381 and CERCA Programme).

Appendix A Being Realizable Without Consuming Resource and Being Resource Non-Generating Are Not Equivalent

In order to study channels constrained by a given static resource RR, it is straightforward to expect these channels to be free from RR. In this work, we depict this by requiring those channels to be “unable to generate RR.” This coincides with the notion of being RR non-generating RT-RMP , and hence being free operations of RR. Meanwhile, there is another possible approach, which is requiring those channels to be “realizable without consuming RR.” It turns out that this latter option is less generic and accessible compared with the former. This Appendix aims to briefly make this difference clear.

In a given resource theory, either static, dynamical, or a more general one, requiring an operation to be free from the resource sometimes includes two seemingly equivalent concepts implicitly; namely, being realizable without consuming the resource, and being unable to generate the resource. While these two concepts match for some settings, in general they are not equivalent. For instance, the resource theories of entanglement equipped with local operation and classical communication (LOCC) channels or local operation plus pre-shared randomness (LOSR) channels allow this property, and so does the resource theory of nonlocality with LOSR channels. This is because LOCC (so do LOSR) channels can be realized without touching and using any entangled state. Nevertheless, the resource theory of athermality demonstrates a counterexample. In this case, the only free state is the state in thermal equilibrium, i.e., the thermal state γ\gamma. Physically, it is impossible to realize any non-trivial channel only within thermal equilibrium (the only realizable one is the state preparation channel of γ\gamma, since one can artificially switch γ\gamma with the input and discard the original system). On the other hand, a commonly used free operation is the thermal operation, which takes the form ()AtrB[UAB(()AγB)UAB](\cdot)_{\rm A}\mapsto{\rm tr}_{\rm B}[U_{\rm AB}((\cdot)_{\rm A}\otimes\gamma_{\rm B})U_{\rm AB}^{\dagger}], where UABU_{\rm AB} conserves the total energy (i.e., it commutes with the total Hamiltonian). One can see that any non-trivial thermal operation (which is athermality non-generating) needs a non-trivial unitary UABU_{\rm AB}, and hence includes effects out of thermal equilibrium; i.e., it is not realizable without consuming athermality.

Apart from resource theories of states, there are also instances in dynamical resource theories illustrating this difference. In the resource theory for non-signaling assisted classical communication Takagi2019-3 , the free quantities are state preparation channels, and it is impossible to output channels useful for classical communication if one only uses state preparation channels to implement free super-channels. Similarly, in the theory of resource preservability Hsieh2020-1 , it is again impossible to output resourceful channels when one only uses resource annihilating channels to implement free super-channels.

If one upgrades the discussion to general and abstract considerations, it can be tough to access the detailed physical structures of operations. Consequently, the best one can do is to analyze an operation by comparing its inputs and outputs. This is also the only way to check whether an operation is free from the given resource. Hence, “zero ability to generate the resource” ends up to be the most feasible and well-defined way to depict “begin free” in the most general extend when further structures and contexts are not available. Being realizable without consuming the resource is an additional property that can be satisfied in certain cases, but this notion could be generally ill-defined.

Note that this is also why in a general, model-independent level, the definition of being resource non-generating only requires no generation of the resource for free inputs: Before introducing free operations, we cannot compare and order different resourceful states, and the only existing concept before defining free operations is “whether the quantity is resourceful or not.” This gives us the most extensive range to clarify the notion of “being free from the resource.” It also briefly summarizes the features of central ingredients in a resource theory: Free states give us detection, free operations give us comparison, and monotones give us quantification.

Due to the above discussion, in this work, we depict a channel as constrained by a resource if it is a free operation.

Appendix B Assumptions on State Resource Theories for Resource Preservability Theories

To have a general study that is also analytically feasible, we need to impose certain assumptions on the state resource theories considered in this work. Let (R,R,𝒪R)(R,\mathcal{F}_{R},\mathcal{O}_{R}) be a given state resource theory. Then we consider

  1. 1.

    Identity channel and partial trace are both free operations; namely, they are both in 𝒪R\mathcal{O}_{R}.

  2. 2.

    Free operations are closed under tensor products, convex sums, and compositions: For every ,𝒪R\mathcal{E},\mathcal{E}^{\prime}\in\mathcal{O}_{R} and p[0,1]p\in[0,1], we have 𝒪R\mathcal{E}\otimes\mathcal{E}^{\prime}\in\mathcal{O}_{R}, p+(1p)𝒪Rp\mathcal{E}+(1-p)\mathcal{E}^{\prime}\in\mathcal{O}_{R}, and 𝒪R\mathcal{E}\circ\mathcal{E}^{\prime}\in\mathcal{O}_{R}.

  3. 3.

    For every system S{\rm S^{\prime}} there exists a state σS\sigma_{\rm S^{\prime}} such that ()S()SσS(\cdot)_{\rm S}\mapsto(\cdot)_{\rm S}\otimes\sigma_{\rm S^{\prime}} is a free operation.

Assumptions 1 and 2 are always assumed in this work in order to capture the necessary properties of a monotone, and we leave Assumption 3 optional. This is slightly different from Ref. Hsieh2020-1 , and our motivation is to relax the assumptions made by Ref. Hsieh2020-1 to achieve a general consideration admitting more applicable cases. We briefly explain each assumptions. Assumption 1 follows from our conceptual expectation; that is, “doing nothing” and “ignoring part of the system” are both unable to generate RR. Assumption 2 implies that if two channels are unable to generate RR, then neither can their simultaneous application (tensor product), classical mixture (convex sum), and sequential application (composition). Finally, Assumption 3 ensures that there always exists a “free extension,” which automatically implies the state σS\sigma_{\rm S^{\prime}} is free an hence R\mathcal{F}_{R}\neq\emptyset (to see this, consider trS{\rm tr}_{\rm S} and use Assumptions 1 and 2). Note that Assumption 3 is only imposed on systems with proper system sizes. For example, in the resource theory of entanglement, steering, and nonlocality, S{\rm S^{\prime}} must be bipartite (and we always assume equal local dimension); in the resource theory of athermality, S{\rm S^{\prime}} can only have dimension dkd^{k} with some positive integer kk, where dd is the dimension of the given thermal state.

Many known resource theories share these assumptions. For instance, Assumptions 12, and 3 are satisfied by the sets of LOCC channels, LOSR channels, Gibbs-preserving maps, and GG-covariant channels (in multipartite cases, we consider the group Gk{i=1kUi|UiGi}G^{\otimes k}\coloneqq\{\bigotimes_{i=1}^{k}U_{i}\,|\,U_{i}\in G\;\forall i\}). Note that Assumption 3 holds since for every system S{\rm S^{\prime}} with dimension dSd_{\rm S^{\prime}} the mapping ()()𝕀SdS(\cdot)\mapsto(\cdot)\otimes\frac{\mathbb{I}_{\rm S^{\prime}}}{d_{\rm S^{\prime}}} is an LOSR and GG-covariant channel. The case of Gibbs-preserving maps (with the thermal state γ\gamma) follows from the fact that ()()γk(\cdot)\mapsto(\cdot)\otimes\gamma^{\otimes k} is Gibbs-preserving for all kk. This implies the validity of Assumptions 12, and 3 in the following state resource theories, which covers most of the cases studied in this work: (i) entanglement and free entanglement Horodecki1998 equipped with LOCC or LOSR channels, (ii) nonlocality, steering, multi-copy nonlocality, and multi-copy steering equipped with LOSR channels (see Appendix B.2 for a discussion), (iii) GG-asymmetry equipped with GG-covariant channels, and (iv) athermality equipped with Gibbs-preserving maps.

It turns out that, by using Assumptions 12, and 3, we can prove a generalized version of Theorem 2 in Ref. Hsieh2020-1 , which is summarized as follows:

Theorem B.1.

Hsieh2020-1 (R,R,𝒪R)(R,\mathcal{F}_{R},\mathcal{O}_{R}) is a state resource theory satisfying Assumptions 1 and 2. DD is a contractive generalized distance measure of states. Then PDP_{D} satisfies

  1. 1.

    PD(𝒩)0P_{D}(\mathcal{N})\geq 0 and PD(𝒩)=0P_{D}(\mathcal{N})=0 if 𝒩𝒪RN\mathcal{N}\in\mathcal{O}_{R}^{N}.

  2. 2.

    PD[𝔉()]PD()P_{D}[\mathfrak{F}(\mathcal{E})]\leq P_{D}(\mathcal{E}) for every channel \mathcal{E} and free super-channel 𝔉𝔽R\mathfrak{F}\in\mathbb{F}_{R}.

If Assumption 3 holds, then we have

PD(𝒩𝒩)PD(𝒩)\displaystyle P_{D}(\mathcal{N}\otimes\mathcal{N}^{\prime})\geq P_{D}(\mathcal{N}) (16)

for every 𝒩,𝒩𝒪R\mathcal{N},\mathcal{N}^{\prime}\in\mathcal{O}_{R}, and the equality holds if 𝒩𝒪~RN\mathcal{N}^{\prime}\in\widetilde{\mathcal{O}}_{R}^{N}.

Proof.

Apart from Eq. (49) in Ref. Hsieh2020-1 , the proof is the same with the one of Theorem 2 in Ref. Hsieh2020-1 (see Eqs. (48) and (50) in Ref. Hsieh2020-1 ). Note that Eq. (48) in Ref. Hsieh2020-1 works for every channel, which explains the validity of statement 2 in this theorem. It remains to show Eq. (7) in Ref. Hsieh2020-1 , which can be seen by the following alternative proof:

PD(SS)\displaystyle P_{D}(\mathcal{E}_{\rm S}\otimes\mathcal{E}_{\rm S^{\prime}}) =infΛSS𝒪RNsupAD[(SSΛ~A)(ρSSA),(ΛSSΛ~A)(ρSSA)]\displaystyle=\inf_{\Lambda_{\rm SS^{\prime}}\in\mathcal{O}_{R}^{N}}\sup_{\rm A}D\left[(\mathcal{E}_{\rm S}\otimes\mathcal{E}_{\rm S^{\prime}}\otimes\widetilde{\Lambda}_{\rm A})(\rho_{\rm SS^{\prime}A}),(\Lambda_{\rm SS^{\prime}}\otimes\widetilde{\Lambda}_{\rm A})(\rho_{\rm SS^{\prime}A})\right]
infΛSS𝒪RNsupAD[(SΛ~A)(ρSA),[(trSΛSS)Λ~A](ρSSA)]\displaystyle\geq\inf_{\Lambda_{\rm SS^{\prime}}\in\mathcal{O}_{R}^{N}}\sup_{\rm A}D\left[(\mathcal{E}_{\rm S}\otimes\widetilde{\Lambda}_{\rm A})(\rho_{\rm SA}),[({\rm tr}_{\rm S^{\prime}}\circ\Lambda_{\rm SS^{\prime}})\otimes\widetilde{\Lambda}_{\rm A}](\rho_{\rm SS^{\prime}A})\right]
infΛSS𝒪RNsupAD[(SΛ~A)(ρSA),[(trSΛSS)Λ~A](ρSAσS)]\displaystyle\geq\inf_{\Lambda_{\rm SS^{\prime}}\in\mathcal{O}_{R}^{N}}\sup_{\rm A}D\left[(\mathcal{E}_{\rm S}\otimes\widetilde{\Lambda}_{\rm A})(\rho_{\rm SA}),[({\rm tr}_{\rm S^{\prime}}\circ\Lambda_{\rm SS^{\prime}})\otimes\widetilde{\Lambda}_{\rm A}](\rho_{\rm SA}\otimes\sigma_{\rm S^{\prime}})\right]
infΛS𝒪RNsupAD[(SΛ~A)(ρSA),(ΛSΛ~A)(ρSA)]\displaystyle\geq\inf_{\Lambda_{\rm S}\in\mathcal{O}_{R}^{N}}\sup_{\rm A}D\left[(\mathcal{E}_{\rm S}\otimes\widetilde{\Lambda}_{\rm A})(\rho_{\rm SA}),(\Lambda_{\rm S}\otimes\widetilde{\Lambda}_{\rm A})(\rho_{\rm SA})\right]
=PD(S).\displaystyle=P_{D}(\mathcal{E}_{\rm S}). (17)

The second line follows from the data-processing inequality and the fact that SS\mathcal{E}_{\rm S}\otimes\mathcal{E}_{\rm S^{\prime}} has a well-defined marginal in S{\rm S} Acknowledgement . In the third line, ρSAσS\rho_{\rm SA}\otimes\sigma_{\rm S^{\prime}} forms a sub-optimal range of the maximization, where σS\sigma_{\rm S^{\prime}} is the state guaranteed by Assumption 3 that allows the map ()()σS(\cdot)\mapsto(\cdot)\otimes\sigma_{\rm S^{\prime}} to be a free operation of RR. Together with Assumptions 1 and 2, we learn that (trSΛSS)[()σS]𝒪RN({\rm tr}_{\rm S^{\prime}}\circ\Lambda_{\rm SS^{\prime}})[(\cdot)\otimes\sigma_{\rm S^{\prime}}]\in\mathcal{O}_{R}^{N} is a resource annihilating channel, which forms a sub-optimal range of the minimization and implies the fourth line. Hence, Assumptions 12, and 3 are enough to ensure the correcteness of Theorem 2 in Ref. Hsieh2020-1 . ∎

Theorem B.1 generalizes Theorem 2 in Ref. Hsieh2020-1 by relaxing the assumptions of absolutely free states (i.e., the assumptions (R1) and (R3) in Ref. Hsieh2020-1 ) into Assumption 3. Furthermore, there is no need to assume the convexity of R\mathcal{F}_{R}. Another remark is that non-increasing under free super-channel actually works for every channel, including channels that are not free operations. This is a useful observation when one needs to consider the smooth version of RR-preservability, e.g., in the next sub-section.

B.1 Properties of PDδP_{D}^{\delta}

We remark that PDδP_{D}^{\delta}, which can be interpreted as the smooth version of PDP_{D}, still possesses the expected properties that a monotone should have. First, if 𝒩𝒪RN\mathcal{N}\in\mathcal{O}_{R}^{N}, then we have PDδ(𝒩)inf𝒩𝒩2δPD(𝒩)PD(𝒩)=0P_{D}^{\delta}(\mathcal{N})\coloneqq\inf_{\left\|\mathcal{N}-\mathcal{N}^{\prime}\right\|_{\diamond}\leq 2\delta}P_{D}(\mathcal{N}^{\prime})\leq P_{D}(\mathcal{N})=0. The non-increasing property under free super-channels can be summarized in the following lemma:

Lemma B.2.

For every channel 𝒩\mathcal{N}, 0δ10\leq\delta\leq 1, and 𝔉𝔽R\mathfrak{F}\in\mathbb{F}_{R}, we have

PDδ[𝔉(𝒩)]PDδ(𝒩).\displaystyle P_{D}^{\delta}[\mathfrak{F}(\mathcal{N})]\leq P_{D}^{\delta}(\mathcal{N}). (18)
Proof.

We note the following estimate first:

(𝒩𝒩)Λ~\displaystyle\left\|(\mathcal{N}-\mathcal{N}^{\prime})\otimes\widetilde{\Lambda}\right\|_{\diamond} (𝒩𝒩)\displaystyle\leq\left\|(\mathcal{N}-\mathcal{N}^{\prime})\otimes\mathcal{I}\right\|_{\diamond}
𝒩𝒩.\displaystyle\leq\left\|\mathcal{N}-\mathcal{N}^{\prime}\right\|_{\diamond}. (19)

The first inequality follows from the data processing inequality, or equivalently, the contractivity of the trace norm; the second ineuqality follows from the definition of the diamond norm. Recall that 𝔉𝔽R\mathfrak{F}\in\mathbb{F}_{R} will take the form 𝔉(𝒩)=Λ+(𝒩Λ~)Λ\mathfrak{F}(\mathcal{N})=\Lambda_{+}\circ(\mathcal{N}\otimes\widetilde{\Lambda})\circ\Lambda_{-}. We conclude that

𝔉(𝒩)𝔉(𝒩)\displaystyle\left\|\mathfrak{F}(\mathcal{N})-\mathfrak{F}(\mathcal{N}^{\prime})\right\|_{\diamond} =Λ+[(𝒩𝒩)Λ~]Λ\displaystyle=\left\|\Lambda_{+}\circ\left[(\mathcal{N}-\mathcal{N}^{\prime})\otimes\widetilde{\Lambda}\right]\circ\Lambda_{-}\right\|_{\diamond}
[(𝒩𝒩)Λ~]Λ\displaystyle\leq\left\|\left[(\mathcal{N}-\mathcal{N}^{\prime})\otimes\widetilde{\Lambda}\right]\circ\Lambda_{-}\right\|_{\diamond}
supA;ρSA[(𝒩𝒩)Λ~A](ΛA)(ρSA)1\displaystyle\coloneqq\sup_{{\rm A};\rho_{\rm SA}}\left\|\left[(\mathcal{N}-\mathcal{N}^{\prime})\otimes\widetilde{\Lambda}\otimes\mathcal{I}_{\rm A}\right]\circ(\Lambda_{-}\otimes\mathcal{I}_{\rm A})(\rho_{\rm SA})\right\|_{1}
(𝒩𝒩)Λ~\displaystyle\leq\left\|(\mathcal{N}-\mathcal{N}^{\prime})\otimes\widetilde{\Lambda}\right\|_{\diamond}
𝒩𝒩,\displaystyle\leq\left\|\mathcal{N}-\mathcal{N}^{\prime}\right\|_{\diamond}, (20)

where the second line follows from data-processing inequality, the fourth line is because (ΛA)(ρSA)(\Lambda_{-}\otimes\mathcal{I}_{\rm A})(\rho_{\rm SA}) induces a sub-optimal range of the maximization in the definition of diamond norm. In the last line we use Eq. (B.1). Now, direct computation shows

PDδ(𝒩)\displaystyle P_{D}^{\delta}(\mathcal{N}) inf𝒩𝒩2δPD(𝒩)\displaystyle\coloneqq\inf_{\left\|\mathcal{N}-\mathcal{N}^{\prime}\right\|_{\diamond}\leq 2\delta}P_{D}(\mathcal{N}^{\prime})
inf𝔉(𝒩)𝔉(𝒩)2δPD(𝒩)\displaystyle\geq\inf_{\left\|\mathfrak{F}(\mathcal{N})-\mathfrak{F}(\mathcal{N}^{\prime})\right\|_{\diamond}\leq 2\delta}P_{D}(\mathcal{N}^{\prime})
inf𝔉(𝒩)𝔉(𝒩)2δPD[𝔉(𝒩)]\displaystyle\geq\inf_{\left\|\mathfrak{F}(\mathcal{N})-\mathfrak{F}(\mathcal{N}^{\prime})\right\|_{\diamond}\leq 2\delta}P_{D}[\mathfrak{F}(\mathcal{N}^{\prime})]
inf𝔉(𝒩)𝒩′′2δPD(𝒩′′)\displaystyle\geq\inf_{\left\|\mathfrak{F}(\mathcal{N})-\mathcal{N}^{\prime\prime}\right\|_{\diamond}\leq 2\delta}P_{D}(\mathcal{N}^{\prime\prime})
=PDδ[𝔉(𝒩)].\displaystyle=P_{D}^{\delta}[\mathfrak{F}(\mathcal{N})]. (21)

From Eq. (B.1) we learn that all channels 𝒩\mathcal{N}^{\prime} satisfying 𝒩𝒩2δ\left\|\mathcal{N}-\mathcal{N}^{\prime}\right\|_{\diamond}\leq 2\delta form a subset of all channels 𝒩\mathcal{N}^{\prime} satisfying 𝔉(𝒩)𝔉(𝒩)2δ\left\|\mathfrak{F}(\mathcal{N})-\mathfrak{F}(\mathcal{N}^{\prime})\right\|_{\diamond}\leq 2\delta. This explains the second line. The third line follows from Theorem B.1 (note that 𝒩\mathcal{N}^{\prime} could be outside 𝒪R\mathcal{O}_{R}, but it is still a channel). In the fourth line, we have the set of all channels of the form 𝔉(𝒩)\mathfrak{F}(\mathcal{N}^{\prime}) be a subset of the set of all channels. This shows the desired claim. ∎

B.2 LOSR Channels as Free Operations of Nonlocality, Steering, Multi-Copy Nonlocality, and Multi-Copy Steering

Appendix A.1 in Ref. Hsieh2020-1 explains that LOSR channels can be free operations of nonlocality. Here, we briefly show that LOSR channels can also be free operations of three other different forms of inseparabilities: Steering Wiseman2007 ; Jones2007 ; steering-review ; RMP-steering , multi-copy nonlocality Palazuelos2012 ; Cavalcanti2013 , and multi-copy steering Hsieh2016 ; Quintino2016 . Formally, a LOSR channel in a given bipartite system AB{\rm AB} is given by the following form:

(λAλB)pλ𝑑λ,\displaystyle\mathcal{E}\coloneqq\int(\mathcal{E}^{\rm A}_{\lambda}\otimes\mathcal{E}^{\rm B}_{\lambda})p_{\lambda}d\lambda, (22)

where the integration is taken over the parameter λ\lambda. Physically, it is a convex mixture of local dynamics. Now, in the given bipartite system AB{\rm AB}, a state is unsteerable from A{\rm A} to B{\rm B} Wiseman2007 ; Jones2007 ; steering-review ; RMP-steering , or simply AB{\rm A\to B} unsteerable, if for every local POVMs {Ea|xA}\{E_{a|x}^{\rm A}\} in A{\rm A} and {Eb|yB}\{E_{b|y}^{\rm B}\} in B{\rm B}, one can write

tr[(Ea|xAEb|yB)ρ]=λΛLHSP(a|x,λ)tr(Eb|yBσλ)pλ𝑑λ\displaystyle{\rm tr}\left[\left(E_{a|x}^{\rm A}\otimes E_{b|y}^{\rm B}\right)\rho\right]=\int_{\lambda\in\Lambda_{\rm LHS}}P(a|x,\lambda){\rm tr}\left(E_{b|y}^{\rm B}\sigma_{\lambda}\right)p_{\lambda}d\lambda (23)

for some variable λ\lambda in a set ΛLHS\Lambda_{\rm LHS}, some probability distributions P(a|x,λ),pλP(a|x,\lambda),p_{\lambda}, and some local states σλ\sigma_{\lambda} in B{\rm B}. In other words, a state is AB{\rm A\to B} unsteerable if every outcome of local measurements in B{\rm B} is indistinguishable from the outputs of pre-shared randomness combined with local quantum theory in B{\rm B}. Such models are called local hidden state models Wiseman2007 ; Jones2007 ; steering-review ; RMP-steering , as depicted by ΛLHS\Lambda_{\rm LHS}. States that are not AB{\rm A\to B} unsteerable are said to be AB{\rm A\to B} steerable.

To see why LOSR channels can be free operations for steering, consider an LOSR channel ν(νAνB)qν𝑑ν\mathcal{E}\coloneqq\int_{\nu}\left(\mathcal{E}_{\nu}^{\rm A}\otimes\mathcal{E}_{\nu}^{\rm B}\right)q_{\nu}d\nu and the following computation

tr[(Ea|xAEb|yB)(ρ)]\displaystyle{\rm tr}\left[\left(E_{a|x}^{\rm A}\otimes E_{b|y}^{\rm B}\right)\mathcal{E}(\rho)\right] =νtr[(Ea|xAEb|yB)(νAνB)(ρ)]qν𝑑ν\displaystyle=\int_{\nu}{\rm tr}\left[\left(E_{a|x}^{\rm A}\otimes E_{b|y}^{\rm B}\right)\left(\mathcal{E}_{\nu}^{\rm A}\otimes\mathcal{E}_{\nu}^{\rm B}\right)(\rho)\right]q_{\nu}d\nu
=νtr[(νA,(Ea|xA)νB,(Eb|yB))ρ]qν𝑑ν,\displaystyle=\int_{\nu}{\rm tr}\left[\left(\mathcal{E}_{\nu}^{\rm A,\dagger}\left(E_{a|x}^{\rm A}\right)\otimes\mathcal{E}_{\nu}^{\rm B,\dagger}\left(E_{b|y}^{\rm B}\right)\right)\rho\right]q_{\nu}d\nu, (24)

where νA,(Ea|xA)\mathcal{E}_{\nu}^{\rm A,\dagger}\left(E_{a|x}^{\rm A}\right) and νB,(Eb|yB)\mathcal{E}_{\nu}^{\rm B,\dagger}\left(E_{b|y}^{\rm B}\right) again form local POVMs since νA,,νB,\mathcal{E}_{\nu}^{\rm A,\dagger},\mathcal{E}_{\nu}^{\rm B,\dagger} are completely-positive unital maps. Hence, when ρ\rho is AB{\rm A\to B} unsteerable, it means, for every ν\nu, we can write tr[(νA,(Ea|xA)νB,(Eb|yB))ρ]{\rm tr}\left[\left(\mathcal{E}_{\nu}^{\rm A,\dagger}\left(E_{a|x}^{\rm A}\right)\otimes\mathcal{E}_{\nu}^{\rm B,\dagger}\left(E_{b|y}^{\rm B}\right)\right)\rho\right] as Eq. (23). This means the output of Eq. (B.2) is again described by Eq. (23).

With the notions of nonlocality and steering, we say a state ρ\rho is multi-copy nonlocal Palazuelos2012 (and, similarly, multi-copy AB{\rm A\to B} steerable Hsieh2016 ; Quintino2016 ) if ρk\rho^{\otimes k} is nonlocal (AB{\rm A\to B} steerable) for some positive integer kk. One can see that LOSR channels again act as free operations for these two resources. To see this, it suffices to observe that if \mathcal{E} is an LOSR channel in a given bipartition, then k\mathcal{E}^{\otimes k} will again be an LOSR channel in the same bipartition. More precisely, consider an LOSR channel AB\mathcal{E}_{\rm AB} in AB{\rm AB} bipartition. Suppose ρ\rho is multi-copy local (AB{\rm A\to B} unsteerable) in this bipartition; namely, ρk\rho^{\otimes k} is local (AB{\rm A\to B} unsteerable) for all kk. Then, for all kk, [AB(ρ)]k=ABk(ρk)\left[\mathcal{E}_{\rm AB}(\rho)\right]^{\otimes k}=\mathcal{E}_{\rm AB}^{\otimes k}\left(\rho^{\otimes k}\right) must be local (AB{\rm A\to B} unsteerable) since ρk\rho^{\otimes k} is local (AB{\rm A\to B} unsteerable) and ABk\mathcal{E}_{\rm AB}^{\otimes k} is again an LOSR channel in the AB{\rm AB} bipartition. This shows that LOSR channels can be free operations of multi-copy nonlocality and multi-copy steering.

Appendix C Proofs of Theorem 1 and Corollary 2

First, we note the following lemma similar to Fact E.2 in Ref. Hsieh2020-1 . This will enable us to obtain an equivalent representation of PDmaxP_{D_{\rm max}} defined in Eq. (4). In what follows, the maximization supΛ~A,ρSA\sup_{\widetilde{\Lambda}_{\rm A},\rho_{\rm SA}} is taken over all ancillary systems A{\rm A}, absolutely resource annihilating channels Λ~A\widetilde{\Lambda}_{\rm A}, and joint input states ρSA\rho_{\rm SA}. Note that the maximization includes the trivial ancillary system (i.e., the one with dimension 1), which means it also covers the case when there is no ancillary system.

Lemma C.1.

Given two channels 𝒩\mathcal{N} and \mathcal{E}, then we have

supΛ~A,ρSAinf{λ0| 0[(λ𝒩)Λ~A](ρSA)}=inf{λ0| 0[(λ𝒩)Λ~A](ρSA)A,Λ~A,ρSA}.\displaystyle\sup_{\widetilde{\Lambda}_{\rm A},\rho_{\rm SA}}\inf\left\{\lambda\geq 0\,|\,0\leq[(\lambda\mathcal{E}-\mathcal{N})\otimes\widetilde{\Lambda}_{\rm A}](\rho_{\rm SA})\right\}=\inf\left\{\lambda\geq 0\,|\,0\leq[(\lambda\mathcal{E}-\mathcal{N})\otimes\widetilde{\Lambda}_{\rm A}](\rho_{\rm SA})\;\forall{\rm A},\widetilde{\Lambda}_{\rm A},\rho_{\rm SA}\right\}. (25)
Proof.

Let 𝐀{λ| 0[(λ𝒩)Λ~A](ρSA)}\mathcal{L}_{\bf A}\coloneqq\left\{\lambda\,|\,0\leq[(\lambda\mathcal{E}-\mathcal{N})\otimes\widetilde{\Lambda}_{\rm A}](\rho_{\rm SA})\right\}, where 𝐀(A,Λ~A,ρSA){\bf A}\coloneqq({\rm A},\widetilde{\Lambda}_{\rm A},\rho_{\rm SA}) is a specific combination of A{\rm A}, Λ~A\widetilde{\Lambda}_{\rm A}, and ρSA\rho_{\rm SA}. Then the left-hand-side is sup𝐀inf{λ|λ𝐀}\sup_{\bf A}\inf\{\lambda\,|\,\lambda\in\mathcal{L}_{\bf A}\}, and the right-hand-side is inf{λ|λ𝐀𝐀}\inf\left\{\lambda\,|\,\lambda\in\bigcap_{\bf A}\mathcal{L}_{\bf A}\right\}. The inequality “\leq” follows since 𝐀𝐀𝐀\bigcap_{\bf A}\mathcal{L}_{\bf A}\subseteq\mathcal{L}_{\bf A^{\prime}} for all 𝐀{\bf A^{\prime}}. To show the opposite, consider an arbitrary positive integer kk. Then there exist 𝐀k{\bf A}_{k} and λk𝐀k\lambda_{k}\in\mathcal{L}_{{\bf A}_{k}} such that

inf{λ|λ𝐀k}sup𝐀inf{λ|λ𝐀}<inf{λ|λ𝐀k}+1k;\displaystyle\inf\left\{\lambda\,|\,\lambda\in\mathcal{L}_{{\bf A}_{k}}\right\}\leq\sup_{\bf A}\inf\{\lambda\,|\,\lambda\in\mathcal{L}_{\bf A}\}<\inf\left\{\lambda\,|\,\lambda\in\mathcal{L}_{{\bf A}_{k}}\right\}+\frac{1}{k}; (26)
λk1k<inf{λ|λ𝐀k}λk.\displaystyle\lambda_{k}-\frac{1}{k}<\inf\left\{\lambda\,|\,\lambda\in\mathcal{L}_{{\bf A}_{k}}\right\}\leq\lambda_{k}. (27)

This means inf{λ|λ𝐀}<λk+1k\inf\{\lambda\,|\,\lambda\in\mathcal{L}_{\bf A}\}<\lambda_{k}+\frac{1}{k} for all 𝐀{\bf A}, which further implies λk+1k𝐀𝐀\lambda_{k}+\frac{1}{k}\in\bigcap_{\bf A}\mathcal{L}_{\bf A}. We conclude that

inf{λ|λ𝐀𝐀}\displaystyle\inf\left\{\lambda\,|\,\lambda\in\bigcap_{\bf A}\mathcal{L}_{\bf A}\right\} λk+1k\displaystyle\leq\lambda_{k}+\frac{1}{k}
inf{λ|λ𝐀k}+2k\displaystyle\leq\inf\left\{\lambda\,|\,\lambda\in\mathcal{L}_{{\bf A}_{k}}\right\}+\frac{2}{k}
sup𝐀inf{λ|λ𝐀}+2k,\displaystyle\leq\sup_{\bf A}\inf\{\lambda\,|\,\lambda\in\mathcal{L}_{\bf A}\}+\frac{2}{k}, (28)

and the desired claim follows by considering all possible kk. ∎

Combining Eq. (5) and Lemma C.1, we note the following:

DmaxR(𝒩)\displaystyle D^{R}_{\rm max}(\mathcal{N}\|\mathcal{E}) supΛ~A,ρSADmax[(𝒩Λ~A)(ρSA)(Λ~A)(ρSA)]\displaystyle\coloneqq\sup_{\widetilde{\Lambda}_{\rm A},\rho_{\rm SA}}D_{\rm max}\left[(\mathcal{N}\otimes\widetilde{\Lambda}_{\rm A})(\rho_{\rm SA})\,\|\,(\mathcal{E}\otimes\widetilde{\Lambda}_{\rm A})(\rho_{\rm SA})\right]
log2supΛ~A,ρSAinf{λ0| 0[(λ𝒩)Λ~A](ρSA)}\displaystyle\coloneqq\log_{2}\sup_{\widetilde{\Lambda}_{\rm A},\rho_{\rm SA}}\inf\left\{\lambda\geq 0\,|\,0\leq[(\lambda\mathcal{E}-\mathcal{N})\otimes\widetilde{\Lambda}_{\rm A}](\rho_{\rm SA})\right\}
=log2inf{λ0|(λ𝒩)Λ~AisapositivemapA,Λ~A}.\displaystyle=\log_{2}\inf\left\{\lambda\geq 0\,|\,(\lambda\mathcal{E}-\mathcal{N})\otimes\widetilde{\Lambda}_{\rm A}\;{\rm is\;a\;positive\;map}\;\forall{\rm A},\widetilde{\Lambda}_{\rm A}\right\}. (29)

A direct observation from Eq. (C) is

Fact C.2.

(2DmaxR(𝒩)𝒩)Λ~A\left(2^{D^{R}_{\rm max}(\mathcal{N}\|\mathcal{E})}\mathcal{E}-\mathcal{N}\right)\otimes\widetilde{\Lambda}_{\rm A} is a positive map A,Λ~A\forall{\rm A},\widetilde{\Lambda}_{\rm A}.

Proof.

Suppose the opposite was correct. Then there exists an ancillary system A{\rm A}_{*}, an absolutely resource annihilating channel Λ~A\widetilde{\Lambda}_{\rm A_{*}}, two states ρSA\rho_{\rm SA_{*}}, |ϕ|\phi\rangle such that ϕ|(2DmaxR(𝒩)𝒩)Λ~A(ρSA)|ϕ<0\langle\phi|\left(2^{D^{R}_{\rm max}(\mathcal{N}\|\mathcal{E})}\mathcal{E}-\mathcal{N}\right)\otimes\widetilde{\Lambda}_{\rm A_{*}}(\rho_{\rm SA_{*}})|\phi\rangle<0. However, we have ϕ|(2[DmaxR(𝒩)+1k]𝒩)Λ~A(ρSA)|ϕ0k\langle\phi|\left(2^{\left[D^{R}_{\rm max}(\mathcal{N}\|\mathcal{E})+\frac{1}{k}\right]}\mathcal{E}-\mathcal{N}\right)\otimes\widetilde{\Lambda}_{\rm A_{*}}(\rho_{\rm SA_{*}})|\phi\rangle\geq 0\;\forall k\in\mathbb{N} due to Eq. (C). This leads to a contradiction when kk\to\infty. ∎

Finally, we note the following alternative form of Eq. (4):

PDmax(𝒩)=log2infΛ𝒪RNinf{λ0|(λΛ𝒩)Λ~AisapositivemapA,Λ~A}.\displaystyle P_{D_{\rm max}}(\mathcal{N})=\log_{2}\inf_{\Lambda\in\mathcal{O}_{R}^{N}}\inf\left\{\lambda\geq 0\,|\,(\lambda\Lambda-\mathcal{N})\otimes\widetilde{\Lambda}_{\rm A}\;{\rm is\;a\;positive\;map}\;\forall{\rm A},\widetilde{\Lambda}_{\rm A}\right\}. (30)

Now, we can present the proofs of Theorem 1 and Corollary 2.

C.1 Proof of Theorem 1

Proof.

Consider a channel 𝒩\mathcal{N}^{\prime} satisfying 𝒩𝒩2δ\left\|\mathcal{N}-\mathcal{N}^{\prime}\right\|_{\diamond}\leq 2\delta. For a given error κ>0\kappa>0, recall that 𝒪RN(κ;𝒩){Λ𝒪RN||DmaxR(𝒩Λ)PDmax(𝒩)|κ}\mathcal{O}_{R}^{N}(\kappa;\mathcal{N}^{\prime})\coloneqq\{\Lambda\in\mathcal{O}_{R}^{N}\,|\,|D_{\rm max}^{R}(\mathcal{N}^{\prime}\|\Lambda)-P_{D_{\rm max}}(\mathcal{N}^{\prime})|\leq\kappa\}, which is by definition non-empty. Then for every Λ𝒩𝒪RN(κ;𝒩)\Lambda^{\mathcal{N}^{\prime}}\in\mathcal{O}_{R}^{N}(\kappa;\mathcal{N}^{\prime}), Fact C.2 implies the existence of a positive map 𝒫\mathcal{P} such that

𝒫Λ~AisapositivemapA,Λ~A;\displaystyle\mathcal{P}\otimes\widetilde{\Lambda}_{\rm A}\;{\rm is\;a\;positive\;map}\;\forall{\rm A},\widetilde{\Lambda}_{\rm A}; (31)
𝒩+𝒫=2DmaxR(𝒩Λ𝒩)Λ𝒩.\displaystyle\mathcal{N}^{\prime}+\mathcal{P}=2^{D^{R}_{\rm max}\left(\mathcal{N}^{\prime}\|\Lambda^{\mathcal{N}^{\prime}}\right)}\Lambda^{\mathcal{N}^{\prime}}. (32)

Note that the positivity of 𝒫\mathcal{P} actually follows from Eq. (31) and the fact that one is allowed to consider the trivial ancillary system, i.e., the case when there is no ancillary system. Now, with a given MM-code ΘM=({ρm}m=0M1,{Em}m=0M1)\Theta_{M}=(\{\rho_{m}\}_{m=0}^{M-1},\{E_{m}\}_{m=0}^{M-1}), we have [recall the definition from Eq. (2)]:

ps(ΘM,𝒩)\displaystyle p_{s}\left(\Theta_{M},\mathcal{N}^{\prime}\right) 1Mm=0M1tr[Em𝒩(ρm)]\displaystyle\coloneqq\frac{1}{M}\sum_{m=0}^{M-1}{\rm tr}\left[E_{m}\mathcal{N}^{\prime}(\rho_{m})\right]
=2DmaxR(𝒩Λ𝒩)Mm=0M1tr[EmΛ𝒩(ρm)]1Mm=0M1tr[Em𝒫(ρm)]\displaystyle=\frac{2^{D^{R}_{\rm max}\left(\mathcal{N}^{\prime}\|\Lambda^{\mathcal{N}^{\prime}}\right)}}{M}\sum_{m=0}^{M-1}{\rm tr}\left[E_{m}\Lambda^{\mathcal{N}^{\prime}}(\rho_{m})\right]-\frac{1}{M}\sum_{m=0}^{M-1}{\rm tr}\left[E_{m}\mathcal{P}(\rho_{m})\right]
2[PDmax(𝒩)+κ]MsupΘMm=0M1tr[EmΛ𝒩(ρm)],\displaystyle\leq\frac{2^{\left[P_{D_{\rm max}}(\mathcal{N}^{\prime})+\kappa\right]}}{M}\sup_{\Theta_{M^{\prime}}}\sum_{m=0}^{M^{\prime}-1}{\rm tr}\left[E^{\prime}_{m}\Lambda^{\mathcal{N}^{\prime}}(\rho^{\prime}_{m})\right], (33)

where the facts that Λ𝒩𝒪RN(κ,𝒩)\Lambda^{\mathcal{N}^{\prime}}\in\mathcal{O}_{R}^{N}(\kappa,\mathcal{N}^{\prime}) and tr[Em𝒫(ρm)]0{\rm tr}\left[E_{m}\mathcal{P}(\rho_{m})\right]\geq 0 for all mm imply the third line, and the maximization supΘM\sup_{\Theta_{M^{\prime}}} is taken over every MM^{\prime}\in\mathbb{N} and MM^{\prime}-code ΘM=({ρm}m=0M1,{Em}m=0M1)\Theta_{M^{\prime}}=(\{\rho^{\prime}_{m}\}_{m=0}^{M^{\prime}-1},\{E^{\prime}_{m}\}_{m=0}^{M^{\prime}-1}). Since this is true for every Λ𝒩𝒪RN(κ,𝒩)\Lambda^{\mathcal{N}^{\prime}}\in\mathcal{O}_{R}^{N}(\kappa,\mathcal{N}^{\prime}), we conclude the following with Eq. (6):

ps(ΘM,𝒩)1M×2[PDmax(𝒩)+Γκ(𝒩)+κ].\displaystyle p_{s}\left(\Theta_{M},\mathcal{N}^{\prime}\right)\leq\frac{1}{M}\times 2^{\left[P_{D_{\rm max}}(\mathcal{N}^{\prime})+\Gamma_{\kappa}(\mathcal{N}^{\prime})+\kappa\right]}. (34)

Now we use the estimate |ps(ΘM,𝒩)ps(ΘM,𝒩)|12𝒩𝒩\left|p_{s}\left(\Theta_{M},\mathcal{N}^{\prime}\right)-p_{s}\left(\Theta_{M},\mathcal{N}\right)\right|\leq\frac{1}{2}\left\|\mathcal{N}^{\prime}-\mathcal{N}\right\|_{\diamond} Takagi2019-3 , where supA,ρSA(A)(ρSA)1\left\|\mathcal{E}\right\|_{\diamond}\coloneqq\sup_{{\rm A},\rho_{\rm SA}}\left\|(\mathcal{E}\otimes\mathcal{I}_{\rm A})(\rho_{\rm SA})\right\|_{1} is the diamond norm. This can be seen by the following computation

ps(ΘM,𝒩)ps(ΘM,𝒩)\displaystyle p_{s}\left(\Theta_{M},\mathcal{N}^{\prime}\right)-p_{s}\left(\Theta_{M},\mathcal{N}\right) =1Mm=0M1tr[Em(𝒩𝒩)(ρm)]\displaystyle=\frac{1}{M}\sum_{m=0}^{M-1}{\rm tr}\left[E_{m}(\mathcal{N}^{\prime}-\mathcal{N})(\rho_{m})\right]
12Mm=0M1𝒩𝒩\displaystyle\leq\frac{1}{2M}\sum_{m=0}^{M-1}\left\|\mathcal{N}^{\prime}-\mathcal{N}\right\|_{\diamond}
=12𝒩𝒩,\displaystyle=\frac{1}{2}\left\|\mathcal{N}^{\prime}-\mathcal{N}\right\|_{\diamond}, (35)

which follows from the estimate supρsup0E𝕀2tr[E()(ρ)]\sup_{\rho}\sup_{0\leq E\leq\mathbb{I}}2{\rm tr}[E(\mathcal{E}^{\prime}-\mathcal{E})(\rho)]\leq\left\|\mathcal{E}^{\prime}-\mathcal{E}\right\|_{\diamond} Takagi2019-3 for arbitrary channels ,\mathcal{E},\mathcal{E}^{\prime}. Gathering the above ingredients, we conclude that for every channel 𝒩\mathcal{N}^{\prime} satisfying 𝒩𝒩2δ\left\|\mathcal{N}-\mathcal{N}^{\prime}\right\|_{\diamond}\leq 2\delta and MM-code ΘM\Theta_{M} achieving ps(ΘM,𝒩)1ϵp_{s}\left(\Theta_{M},\mathcal{N}\right)\geq 1-\epsilon, we have

1ϵ\displaystyle 1-\epsilon ps(ΘM,𝒩)\displaystyle\leq p_{s}\left(\Theta_{M},\mathcal{N}\right)
ps(ΘM,𝒩)+δ\displaystyle\leq p_{s}\left(\Theta_{M},\mathcal{N}^{\prime}\right)+\delta
1M×2[PDmax(𝒩)+Γκ(𝒩)+κ]+δ.\displaystyle\leq\frac{1}{M}\times 2^{\left[P_{D_{\rm max}}(\mathcal{N}^{\prime})+\Gamma_{\kappa}(\mathcal{N}^{\prime})+\kappa\right]}+\delta. (36)

In other words, for every given ϵ,δ0& 0<κ<1\epsilon,\delta\geq 0\;\&\;0<\kappa<1 satisfying ϵ+δ<1\epsilon+\delta<1 we have

C(1)ϵ(𝒩)\displaystyle C_{(1)}^{\epsilon}(\mathcal{N}) log211ϵδ+κ+inf𝒩𝒩2δ[PDmax(𝒩)+Γκ(𝒩)]\displaystyle\leq\log_{2}\frac{1}{1-\epsilon-\delta}+\kappa+\inf_{\left\|\mathcal{N}-\mathcal{N}^{\prime}\right\|_{\diamond}\leq 2\delta}\left[P_{D_{\rm max}}(\mathcal{N}^{\prime})+\Gamma_{\kappa}(\mathcal{N}^{\prime})\right]
log22κ1ϵδ+inf𝒩𝒩2δPDmax(𝒩)+Γκδ(𝒩),\displaystyle\leq\log_{2}\frac{2^{\kappa}}{1-\epsilon-\delta}+\inf_{\left\|\mathcal{N}-\mathcal{N}^{\prime}\right\|_{\diamond}\leq 2\delta}P_{D_{\rm max}}(\mathcal{N}^{\prime})+\Gamma_{\kappa}^{\delta}(\mathcal{N}), (37)

and the result follows. ∎

C.2 Remark

Note that in some cases Eq. (7) can be simplified. For instance, when the free states are isotropic states Horodecki1999-2 (i.e., RR is asymmetry of the group UUU\otimes U^{*}), then Γκδ(𝒩)log2(2×d2d21)\Gamma_{\kappa}^{\delta}(\mathcal{N})\leq\log_{2}\left(2\times\frac{d^{2}}{d^{2}-1}\right) footnote:IsotropicStates . When d1d\gg 1, Theorem 1 implies C(1)ϵ(𝒩)PDmaxδ(𝒩)+log211ϵδ+1C_{\rm(1)}^{\epsilon}(\mathcal{N})\lesssim P_{D_{\rm max}}^{\delta}(\mathcal{N})+\log_{2}\frac{1}{1-\epsilon-\delta}+1, and the additional degrees of freedom of (UU)(U\otimes U^{*})-asymmetry allow performance better than isotropic states. Another example is when RR is athermality, which implies Corollary 2 that will be proved in the following sub-section.

C.3 Proof of Corollary 2

Proof.

First, from Eq. (80) in Ref. Hsieh2020-1 we learn that PDmax|γ()=supρDmax[(ρ)γ]P_{D_{\rm max}|\gamma}(\mathcal{E})=\sup_{\rho}D_{\rm max}[\mathcal{E}(\rho)\|\gamma]. This means that (ρ)2PDmax|γ()γρ\mathcal{E}(\rho)\leq 2^{P_{D_{\rm max}|\gamma}(\mathcal{E})}\gamma\;\forall\rho and hence

Γκ(𝒩)\displaystyle\Gamma_{\kappa}(\mathcal{N}) log2infΛ𝒩𝒪RN(κ;𝒩)supΘMm=0M1tr[EmΛ𝒩(ρm)]\displaystyle\coloneqq\log_{2}\inf_{\Lambda^{\mathcal{N}}\in\mathcal{O}_{R}^{N}(\kappa;\mathcal{N})}\sup_{\Theta_{M}}\sum_{m=0}^{M-1}{\rm tr}\left[E_{m}\Lambda^{\mathcal{N}}(\rho_{m})\right]
log2infΛ𝒩𝒪RN(κ;𝒩)2PDmax|γ(Λ𝒩)supΘMm=0M1tr(Emγ)\displaystyle\leq\log_{2}\inf_{\Lambda^{\mathcal{N}}\in\mathcal{O}_{R}^{N}(\kappa;\mathcal{N})}2^{P_{D_{\rm max}|\gamma}\left(\Lambda^{\mathcal{N}}\right)}\sup_{\Theta_{M}}\sum_{m=0}^{M-1}{\rm tr}\left(E_{m}\gamma\right)
=infΛ𝒩𝒪RN(κ;𝒩)PDmax|γ(Λ𝒩),\displaystyle=\inf_{\Lambda^{\mathcal{N}}\in\mathcal{O}_{R}^{N}(\kappa;\mathcal{N})}P_{D_{\rm max}|\gamma}\left(\Lambda^{\mathcal{N}}\right), (38)

and the result follows. ∎

Appendix D Proof of Theorem 3

Proof.

We follow the proof of Theorem 2 in Ref. Korzekwa2019 . First, a GG-twirling channel, which is an operation used to symmetrize all input states with respect to a unitary group G{U(g)}g=1|G|G\coloneqq\left\{U^{(g)}\right\}_{g=1}^{|G|}, is defined by

𝒯G()1|G|g=1|G|U(g)()U(g),.\displaystyle\mathcal{T}_{G}(\cdot)\coloneqq\frac{1}{|G|}\sum_{g=1}^{|G|}U^{(g)}(\cdot)U^{(g),\dagger}. (39)

When the group is infinite, one can replace the summation by integration equipped with the Haar measure: 𝒯G()UGU()U𝑑U.\mathcal{T}_{G}(\cdot)\coloneqq\int_{U\in G}U(\cdot)U^{\dagger}dU. We focus on the finite case to illustrate the proof.

With a given state ρ\rho and a given codebook 𝒞\mathcal{C} (that is, a mapping, mgmm\mapsto g_{m}, from the classical information {m}m=0M1\{m\}_{m=0}^{M-1} to the set {1,2,,|G|}\{1,2,...,|G|\}Korzekwa2019 , consider the encoding

{σgm|ρU(gm)ρU(gm),}m=0M1.\displaystyle\left\{\sigma_{g_{m}|\rho}\coloneqq U^{(g_{m})}\rho U^{(g_{m}),\dagger}\right\}_{m=0}^{M-1}. (40)

To construct the decoding, consider the following MM elements of POVM (which is a pretty good measurement scheme):

{Em𝒞|ρSσgm|ρS}m=0M1,\displaystyle\left\{E_{m}^{\mathcal{C}|\rho}\coloneqq S\sigma_{g_{m}|\rho}S\right\}_{m=0}^{M-1}, (41)

where S(m=0M1σgm|ρ)12S\coloneqq\left(\sum_{m=0}^{M-1}\sigma_{g_{m}|\rho}\right)^{-\frac{1}{2}}. Note that for a positive semi-definite operator AA, the notation A1A^{-1} is the inverse of AA restricted to the support of AA Beigi2014 . This means A1A=AA1A^{-1}A=AA^{-1} will be the projection onto the support of AA, and we have A1A=AA1𝕀A^{-1}A=AA^{-1}\leq\mathbb{I} in general. Hence, {Em𝒞|ρ}m=0M1\left\{E_{m}^{\mathcal{C}|\rho}\right\}_{m=0}^{M-1} is not a POVM in general, since m=0M1Em𝒞|ρ\sum_{m=0}^{M-1}E_{m}^{\mathcal{C}|\rho} will be the projection onto the support of m=0M1σgm|ρ\sum_{m=0}^{M-1}\sigma_{g_{m}|\rho}. Recently, Korzekwa et al. (see Eqs. (44), (45) and (51) in Ref. Korzekwa2019 ) show that for 0<κ<10<\kappa<1 we have

𝔼𝒞1Mm=0M1tr[Em𝒞|ρσgm|ρ](1κ)(1MeDsκ[ρ𝒯G(ρ)]),\displaystyle\mathbb{E}_{\mathcal{C}}\frac{1}{M}\sum_{m=0}^{M-1}{\rm tr}\left[E_{m}^{\mathcal{C}|\rho}\sigma_{g_{m}|\rho}\right]\geq(1-\kappa)\left(1-Me^{-D_{s}^{\kappa}[\rho\,\|\,\mathcal{T}_{G}(\rho)]}\right), (42)

where 𝔼𝒞\mathbb{E}_{\mathcal{C}} indicates the average over randomly chosen codebook 𝒞\mathcal{C} (following Ref. Korzekwa2019 , each mm is independently and uniformly at random encoded into the integer gmg_{m}, which means {gm}m=0M1\{g_{m}\}_{m=0}^{M-1} can be interpreted as independent and identically distributed random variables with uniform distribution). For a GG-covariant channel 𝒩\mathcal{N}, consider the MM-code given by ΘM𝒞|ρ({σgm|ρ}m=0M1,{E~m}m=0M1)\Theta_{M}^{\mathcal{C}|\rho}\coloneqq\left(\{\sigma_{g_{m}|\rho}\}_{m=0}^{M-1},\{\widetilde{E}_{m}\}_{m=0}^{M-1}\right), where E~mEm𝒞|𝒩(ρ)\widetilde{E}_{m}\coloneqq E_{m}^{\mathcal{C}|\mathcal{N}(\rho)} for m>0m>0 and E~0E0𝒞|𝒩(ρ)+(𝕀m=0M1Em𝒞|𝒩(ρ))\widetilde{E}_{0}\coloneqq E_{0}^{\mathcal{C}|\mathcal{N}(\rho)}+\left(\mathbb{I}-\sum_{m=0}^{M-1}E_{m}^{\mathcal{C}|\mathcal{N}(\rho)}\right). Note that m=0M1Em𝒞|𝒩(ρ)\sum_{m=0}^{M-1}E_{m}^{\mathcal{C}|\mathcal{N}(\rho)} is the projection onto the support of m=0M1σgm|𝒩(ρ)\sum_{m=0}^{M-1}\sigma_{g_{m}|\mathcal{N}(\rho)}, which means m=0M1Em𝒞|𝒩(ρ)𝕀\sum_{m=0}^{M-1}E_{m}^{\mathcal{C}|\mathcal{N}(\rho)}\leq\mathbb{I}. Then we have

supρ𝔼𝒞ps(ΘM𝒞|ρ,𝒩)\displaystyle\sup_{\rho}\mathbb{E}_{\mathcal{C}}p_{s}\left(\Theta_{M}^{\mathcal{C}|\rho},\mathcal{N}\right) =supρ𝔼𝒞1M(tr[(E~0E0𝒞|𝒩(ρ))𝒩(σgm|ρ)]+m=0M1tr[Em𝒞|𝒩(ρ)𝒩(σgm|ρ)])\displaystyle=\sup_{\rho}\mathbb{E}_{\mathcal{C}}\frac{1}{M}\left({\rm tr}\left[\left(\widetilde{E}_{0}-E_{0}^{\mathcal{C}|\mathcal{N}(\rho)}\right)\mathcal{N}\left(\sigma_{g_{m}|\rho}\right)\right]+\sum_{m=0}^{M-1}{\rm tr}\left[E_{m}^{\mathcal{C}|\mathcal{N}(\rho)}\mathcal{N}\left(\sigma_{g_{m}|\rho}\right)\right]\right)
supρ𝔼𝒞1Mm=0M1tr[Em𝒞|𝒩(ρ)𝒩(σgm|ρ)]\displaystyle\geq\sup_{\rho}\mathbb{E}_{\mathcal{C}}\frac{1}{M}\sum_{m=0}^{M-1}{\rm tr}\left[E_{m}^{\mathcal{C}|\mathcal{N}(\rho)}\mathcal{N}\left(\sigma_{g_{m}|\rho}\right)\right]
=supρ𝔼𝒞1Mm=0M1tr[Em𝒞|𝒩(ρ)σgm|𝒩(ρ)]\displaystyle=\sup_{\rho}\mathbb{E}_{\mathcal{C}}\frac{1}{M}\sum_{m=0}^{M-1}{\rm tr}\left[E_{m}^{\mathcal{C}|\mathcal{N}(\rho)}\sigma_{g_{m}|\mathcal{N}(\rho)}\right]
(1κ)(1MesupρDsκ[𝒩(ρ)𝒯G𝒩(ρ)]).\displaystyle\geq(1-\kappa)\left(1-Me^{-\sup_{\rho}D_{s}^{\kappa}[\mathcal{N}(\rho)\,\|\,\mathcal{T}_{G}\circ\mathcal{N}(\rho)]}\right). (43)

Since E~0E0𝒞|𝒩(ρ)=𝕀m=0M1Em𝒞|𝒩(ρ)\widetilde{E}_{0}-E_{0}^{\mathcal{C}|\mathcal{N}(\rho)}=\mathbb{I}-\sum_{m=0}^{M-1}E_{m}^{\mathcal{C}|\mathcal{N}(\rho)} is non-negative, the second line follows. The third line is because 𝒩\mathcal{N} is GG-covariant and U(gm)GU^{(g_{m})}\in G, so we have 𝒩(σgm|ρ)=𝒩(U(gm)ρU(gm),)=U(gm)𝒩(ρ)U(gm),=σgm|𝒩(ρ)\mathcal{N}(\sigma_{g_{m}|\rho})=\mathcal{N}(U^{(g_{m})}\rho U^{(g_{m}),\dagger})=U^{(g_{m})}\mathcal{N}\left(\rho\right)U^{(g_{m}),\dagger}=\sigma_{g_{m}|\mathcal{N}(\rho)}. The last line is a direct application of Eq. (42) by replacing the role of ρ\rho by 𝒩(ρ)\mathcal{N}(\rho). This means when 1ϵ<(1κ)(1MesupρDsκ[𝒩(ρ)𝒯G𝒩(ρ)])1-\epsilon<(1-\kappa)\left(1-Me^{-\sup_{\rho}D_{s}^{\kappa}[\mathcal{N}(\rho)\,\|\,\mathcal{T}_{G}\circ\mathcal{N}(\rho)]}\right), there must exist an MM-code ΘM𝒞|ρ\Theta_{M}^{\mathcal{C}|\rho} with some ρ\rho and 𝒞\mathcal{C} achieving ps(ΘM𝒞|ρ,𝒩)1ϵp_{s}(\Theta_{M}^{\mathcal{C}|\rho},\mathcal{N})\geq 1-\epsilon. Let log2M=C(1)ϵ(𝒩)\log_{2}M_{*}=C_{\rm(1)}^{\epsilon}(\mathcal{N}). Because no (M+1)(M_{*}+1)-code can achieve success probability ps1ϵp_{s}\geq 1-\epsilon, we must have

1ϵ(1κ)(1(M+1)esupρDsκ[𝒩(ρ)𝒯G𝒩(ρ)]).\displaystyle 1-\epsilon\geq(1-\kappa)\left(1-(M_{*}+1)e^{-\sup_{\rho}D_{s}^{\kappa}[\mathcal{N}(\rho)\,\|\,\mathcal{T}_{G}\circ\mathcal{N}(\rho)]}\right). (44)

Following Ref. Korzekwa2019 , we set κ=ϵδ\kappa=\epsilon-\delta. Since log2nlog2(n+1)1\log_{2}n\geq\log_{2}(n+1)-1 for all positive integer nn, we conclude that

log2M\displaystyle\log_{2}{M_{*}} >1ln2supρDsϵδ[𝒩(ρ)𝒯G𝒩(ρ)]+log2δ1\displaystyle>\frac{1}{\ln 2}{\sup_{\rho}D_{s}^{\epsilon-\delta}[\mathcal{N}(\rho)\,\|\,\mathcal{T}_{G}\circ\mathcal{N}(\rho)]}+\log_{2}\delta-1
1ln2infΛ𝒪RNsupρDsϵδ[𝒩(ρ)Λ(ρ)]+log2δ1,\displaystyle\geq\frac{1}{\ln 2}{\inf_{\Lambda\in\mathcal{O}_{R}^{N}}\sup_{\rho}D_{s}^{\epsilon-\delta}[\mathcal{N}(\rho)\,\|\,\Lambda(\rho)]}+\log_{2}\delta-1, (45)

where the first line is a direct consequence of Eq. (44), and the second line is because 𝒯G𝒩\mathcal{T}_{G}\circ\mathcal{N} is an GG-covariant channel that can only output symmetric states. ∎

Appendix E Collision Model for Thermalization

Following Ref. Sparaciari2019 , consider a finite dimensional system S with Hamiltonian HSH_{\rm S} and a bath B{\rm B}, which is assumed to be n1n-1 copies of the system S{\rm S}, and each copy has Hamiltonian HSH_{\rm S}. Labeling the bath as B1B2Bn1{\rm B}_{1}{\rm B_{2}}...{\rm B}_{n-1}, this means HB=k=1n1𝕀B1𝕀Bk1HS𝕀Bk+1𝕀Bn1H_{\rm B}=\sum_{k=1}^{n-1}\mathbb{I}_{{\rm B}_{1}}\otimes...\otimes\mathbb{I}_{{\rm B}_{k-1}}\otimes H_{\rm S}\otimes\mathbb{I}_{{\rm B}_{k+1}}\otimes...\otimes\mathbb{I}_{{\rm B}_{n-1}}. The collision model introduced in Ref. Sparaciari2019 used to depict thermalization processes is given by

ρSB(t)t=kλk[USB(k)ρSB(t)USB(k),ρSB(t)],\displaystyle\frac{\partial\rho_{\rm SB}(t)}{\partial t}=\sum_{k}\lambda_{k}\left[U_{\rm SB}^{(k)}\rho_{\rm SB}(t)U_{\rm SB}^{(k),\dagger}-\rho_{\rm SB}(t)\right], (46)

where ρSB(t)\rho_{\rm SB}(t) is the global state on SB{\rm SB} at time tt, USB(k)U_{\rm SB}^{(k)} represents an energy-preserving unitary on SB{\rm SB}; i.e., [USB(k),HS+HB]=0[U_{\rm SB}^{(k)},H_{\rm S}+H_{\rm B}]=0, and λk\lambda_{k} is the rate for USB(k)U_{\rm SB}^{(k)} to occur (see also Eqs. (A2) and (A3) in Appendix A of Ref. Sparaciari2019 ). Roughly speaking, each USB(k)U_{\rm SB}^{(k)} models an elastic collision between certain subsystems of SB{\rm SB}. We refer the reader to Ref. Sparaciari2019 for the details of the model and its physical reasoning. In this work, we use the notation 𝒞n\mathcal{C}_{n} to denote the set of channels \mathcal{E} on SB{\rm SB} such that, for every ρ\rho in S{\rm S}, (ργ(n1))\mathcal{E}\left(\rho\otimes\gamma^{\otimes(n-1)}\right) can be realized by Eq. (46) at a time point t>0t>0 [i.e., ρSB(t)=(ργ(n1))\rho_{\rm SB}(t)=\mathcal{E}\left(\rho\otimes\gamma^{\otimes(n-1)}\right); note that we also allow t=t=\infty] with ρSB(0)=ργ(n1)\rho_{\rm SB}(0)=\rho\otimes\gamma^{\otimes(n-1)}. See also Ref. Hsieh2020-1 for its connection with resource preservability.

Appendix F Proof of Theorem 4

Theorem 4 is a consequence of the combination of Theorem 1 and Theorem 4 in Ref. Hsieh2020-1 , which is formally stated as follows (here we implicitly assume the system Hamiltonian is the one realizing the thermal state γ\gamma with some temperature):

Theorem F.1.

Hsieh2020-1 Given a Gibbs-preserving channel 𝒩\mathcal{N}, 0ϵ<10\leq\epsilon<1, and a full-rank thermal state γ\gamma. If 𝒩\mathcal{N} is coherence-annihilating and the system Hamiltonian satisfies the energy subspace condition, then we have

2PDmax|γ(𝒩)γϵ(𝒩)+2ϵpmin(γ)+1,\displaystyle 2^{P_{D_{\rm max}|\gamma}}(\mathcal{N})\leq\mathcal{B}^{\epsilon}_{\gamma}(\mathcal{N})+\frac{2\sqrt{\epsilon}}{p_{\rm min}(\gamma)}+1, (47)

where pmin(γ)p_{\rm min}(\gamma) is the smallest eigenvalue of γ\gamma.

We remark that being coherence-annihilating is required by the proof given in Ref. Sparaciari2019 (specifically, it is crucial for the proof of Lemma 17 in Appendix C of Ref. Sparaciari2019 ), which explains the assumption made in Theorem 4. Combining Corollary 2 (and hence Theorem 1) and Theorem F.1, we are now in the position to prove Theorem 4:

Proof.

Consider coherence as the state resource (R=CohR={\rm Coh}) in Corollary 2, we obtain

C(1)ϵ(𝒩)PDmax|Coh(𝒩)+PDmax|γ(Λ𝒩)+log22κ1ϵ\displaystyle C_{\rm(1)}^{\epsilon}(\mathcal{N})\leq P_{D_{\rm max}|{\rm Coh}}(\mathcal{N})+P_{D_{\rm max}|\gamma}\left(\Lambda^{\mathcal{N}}\right)+\log_{2}\frac{2^{\kappa}}{1-\epsilon} (48)

for every Λ𝒩𝒪CohN(κ,𝒩)\Lambda^{\mathcal{N}}\in\mathcal{O}_{\rm Coh}^{N}(\kappa,\mathcal{N}). Hence, Λ𝒩\Lambda^{\mathcal{N}} is coherence-annihilating. Use Theorem F.1 (consequently, we need to assume all conditions made in the statement of Theorem F.1), we conclude that for 0ϵ,δ<10\leq\epsilon,\delta<1 we have

C(1)ϵ(𝒩)PDmax|Coh(𝒩)+log2(γδ(Λ𝒩)+2δpmin(γ)+1)+log22κ1ϵ,\displaystyle C_{\rm(1)}^{\epsilon}(\mathcal{N})\leq P_{D_{\rm max}|{\rm Coh}}(\mathcal{N})+\log_{2}\left(\mathcal{B}^{\delta}_{\gamma}(\Lambda^{\mathcal{N}})+\frac{2\sqrt{\delta}}{p_{\rm min}(\gamma)}+1\right)+\log_{2}\frac{2^{\kappa}}{1-\epsilon}, (49)

which completes the proof. ∎

Appendix G Proof of Theorem 5

Before the proof, let us recall a crucial tool called fully entangled fraction (FEF) Horodecki1999-2 ; Albeverio2002 . For a bipartite state ρ\rho with equal local dimension dd, its FEF is defined by

max(ρ)max|ΦdΦd|ρΦd,\displaystyle\mathcal{F}_{\rm max}(\rho)\coloneqq\max_{|\Phi_{d}\rangle}\langle\Phi_{d}|\rho|\Phi_{d}\rangle, (50)

which maximizes over all maximally entangled states |Φd|\Phi_{d}\rangle with local dimension dd. FEF is well-known for characterizing different forms of inseparability Horodecki1999-2 ; Albeverio2002 ; Zhao2010 ; Cavalcanti2013 ; Bell-RMP ; Quintino2016 ; Hsieh2016 ; Hsieh2018E ; Ent-RMP ; Liang2019 ; Hsieh2020-2 . For instance, max(ρ)>1d\mathcal{F}_{\rm max}(\rho)>\frac{1}{d} implies ρ\rho is free entangled Ent-RMP ; Horodecki1999-1 , useful for teleporation Horodecki1999-2 , multi-copy nonlocal Palazuelos2012 ; Cavalcanti2013 , and multi-copy steerable Hsieh2016 ; Quintino2016 (see also Appendix B.2).

Following the proof of Theorem 1, we will prove a lemma and have the main theorem as a corollary. Given a state resource RR, recall from Sec. II that 𝔽R\mathbb{F}_{R} is the set of free operations of RR-preservability given by Hsieh2020-1 Λ+(Λ~)Λ\mathcal{E}\mapsto\Lambda_{+}\circ(\mathcal{E}\otimes\widetilde{\Lambda})\circ\Lambda_{-}, where Λ+,Λ𝒪R\Lambda_{+},\Lambda_{-}\in\mathcal{O}_{R} are free operations and Λ~𝒪~RN\widetilde{\Lambda}\in\widetilde{\mathcal{O}}_{R}^{N} is an absolutely resource annihilating channel. Since now, we will always assume that for every 𝔉𝔽R\mathfrak{F}\in\mathbb{F}_{R} and 𝒩𝒪R\mathcal{N}\in\mathcal{O}_{R}, the input and output systems of 𝔉(𝒩)\mathfrak{F}(\mathcal{N}) are both bipartite with finite equal local dimension. Also, we will use the notation R(d×d)\mathcal{F}_{R}^{(d\times d)} to denote the set of free states of RR in bipartite systems with equal local dimension dd.

Lemma G.1.

Given ϵ,δ>0\epsilon,\delta>0 satisfying ϵ+δ<1\epsilon+\delta<1. For 𝒩𝒪R\mathcal{N}\in\mathcal{O}_{R} and 𝔉𝔽R\mathfrak{F}\in\mathbb{F}_{R}, if MM achieves ps|ME[M,𝔉(𝒩)]1ϵp_{s|{\rm ME}}\left[M,\mathfrak{F}(\mathcal{N})\right]\geq 1-\epsilon, then we have

0PDmaxδ(𝒩)+log211ϵδ+log2FR(d),\displaystyle 0\leq P_{D_{\rm max}}^{\delta}(\mathcal{N})+\log_{2}\frac{1}{1-\epsilon-\delta}+\log_{2}F^{R}(d), (51)

where

FR(d)supηR(d×d)max(η),\displaystyle F^{R}(d)\coloneqq\sup_{\eta\in\mathcal{F}_{R}^{(d\times d)}}\mathcal{F}_{\rm max}(\eta), (52)

and dd is the local dimension of the output bipartite system of 𝔉(𝒩)\mathfrak{F}(\mathcal{N}).

Proof.

For every positive integer kk and every 𝒩\mathcal{N}^{\prime} such that 𝒩𝒩2δ\left\|\mathcal{N}^{\prime}-\mathcal{N}\right\|_{\diamond}\leq 2\delta, Eq. (30) implies the existence of a value λk0\lambda_{k}\geq 0 and a RR-annihilating channel Λk𝒪RN\Lambda_{k}\in\mathcal{O}_{R}^{N} such that (i) |PDmax(𝒩)log2λk|1k\left|P_{D_{\rm max}}(\mathcal{N}^{\prime})-\log_{2}\lambda_{k}\right|\leq\frac{1}{k}, and (ii) 𝒫kΛ~A\mathcal{P}_{k}\otimes\widetilde{\Lambda}_{\rm A} is a positive map for every ancillary system A{\rm A} and Λ~A𝒪~RN\widetilde{\Lambda}_{\rm A}\in\widetilde{\mathcal{O}}_{R}^{N}, where 𝒩+𝒫k=λkΛk\mathcal{N}^{\prime}+\mathcal{P}_{k}=\lambda_{k}\Lambda_{k}. Write 𝔉()=Λ+(Λ~)Λ\mathfrak{F}(\mathcal{E})=\Lambda_{+}\circ(\mathcal{E}\otimes\widetilde{\Lambda})\circ\Lambda_{-} with Λ+,Λ𝒪R\Lambda_{+},\Lambda_{-}\in\mathcal{O}_{R} and Λ~𝒪~RN\widetilde{\Lambda}\in\widetilde{\mathcal{O}}_{R}^{N}. This means for every positive integer MM we have (note that |Φm|\Phi^{\prime}_{m}\rangle’s are all staying in the output space of 𝔉(𝒩)\mathfrak{F}(\mathcal{N}), which is a bipartite system with equal local dimension dd):

ps|ME[M,𝔉(𝒩)]\displaystyle p_{s|{\rm ME}}[M,\mathfrak{F}(\mathcal{N}^{\prime})] sup{|Φm}m=0M1,{|Φm}m=0M11Mm=0M1Φm|Λ+(𝒩Λ~)Λ(|ΦmΦm|)|Φm\displaystyle\coloneqq\sup_{\{|\Phi_{m}\rangle\}_{m=0}^{M-1},\{|\Phi^{\prime}_{m}\rangle\}_{m=0}^{M-1}}\frac{1}{M}\sum_{m=0}^{M-1}\langle\Phi^{\prime}_{m}|\Lambda_{+}\circ(\mathcal{N}^{\prime}\otimes\widetilde{\Lambda})\circ\Lambda_{-}(|\Phi_{m}\rangle\langle\Phi_{m}|)|\Phi^{\prime}_{m}\rangle
2[PDmax(𝒩)+1k]M×sup{|Φm}m=0M1,{|Φm}m=0M1m=0M1Φm|Λ+(ΛkΛ~)Λ(|ΦmΦm|)|Φm\displaystyle\leq\frac{2^{\left[P_{D_{\rm max}}(\mathcal{N}^{\prime})+\frac{1}{k}\right]}}{M}\times\sup_{\{|\Phi_{m}\rangle\}_{m=0}^{M-1},\{|\Phi^{\prime}_{m}\rangle\}_{m=0}^{M-1}}\sum_{m=0}^{M-1}\langle\Phi^{\prime}_{m}|\Lambda_{+}\circ(\Lambda_{k}\otimes\widetilde{\Lambda})\circ\Lambda_{-}(|\Phi_{m}\rangle\langle\Phi_{m}|)|\Phi^{\prime}_{m}\rangle
2[PDmax(𝒩)+1k]M×sup{ηm}m=0M1R(d×d),{|Φm}m=0M1m=0M1Φm|ηm|Φm\displaystyle\leq\frac{2^{\left[P_{D_{\rm max}}(\mathcal{N}^{\prime})+\frac{1}{k}\right]}}{M}\times\sup_{\{\eta_{m}\}_{m=0}^{M-1}\subseteq\mathcal{F}_{R}^{(d\times d)},\{|\Phi^{\prime}_{m}\rangle\}_{m=0}^{M-1}}\sum_{m=0}^{M-1}\langle\Phi^{\prime}_{m}|\eta_{m}|\Phi^{\prime}_{m}\rangle
2[PDmax(𝒩)+1k]×FR(d).\displaystyle\leq 2^{\left[P_{D_{\rm max}}(\mathcal{N}^{\prime})+\frac{1}{k}\right]}\times F^{R}(d). (53)

The second line is because Λ+(𝒫kΛ~)Λ\Lambda_{+}\circ(\mathcal{P}_{k}\otimes\widetilde{\Lambda})\circ\Lambda_{-} is a positive map. The third line is because Λ+(ΛkΛ~)Λ𝒪RN\Lambda_{+}\circ(\Lambda_{k}\otimes\widetilde{\Lambda})\circ\Lambda_{-}\in\mathcal{O}_{R}^{N}, and the fact that the output system is a bipartite system with equal local dimension dd. From here we conclude that

ps|ME[M,𝔉(𝒩)]2PDmax(𝒩)×FR(d).\displaystyle p_{s|{\rm ME}}[M,\mathfrak{F}(\mathcal{N}^{\prime})]\leq 2^{P_{D_{\rm max}}(\mathcal{N}^{\prime})}\times F^{R}(d). (54)

Hence, for every MM achieving ps|ME[M,𝔉(𝒩)]1ϵp_{s|{\rm ME}}[M,\mathfrak{F}(\mathcal{N}^{\prime})]\geq 1-\epsilon and for every 𝒩\mathcal{N}^{\prime} achieving 𝒩𝒩2δ\left\|\mathcal{N}^{\prime}-\mathcal{N}\right\|_{\diamond}\leq 2\delta, we have

1ϵ\displaystyle 1-\epsilon ps|ME[M,𝔉(𝒩)]\displaystyle\leq p_{s|{\rm ME}}[M,\mathfrak{F}(\mathcal{N})]
ps|ME[M,𝔉(𝒩)]+δ\displaystyle\leq p_{s|{\rm ME}}[M,\mathfrak{F}(\mathcal{N}^{\prime})]+\delta
2PDmax(𝒩)×FR(d)+δ.\displaystyle\leq 2^{P_{D_{\rm max}}(\mathcal{N}^{\prime})}\times F^{R}(d)+\delta. (55)

The second line is a consequence of Eqs. (C.1) and (B.1). The desired result follows. ∎

As a direct observation on Lemma G.1, once FR(d)F^{R}(d) has an explicit dependency on dd, one could conclude an upper bound on log2M\log_{2}M. This is the case for various resources, and this fact allows us to prove Theorem 5 as follows.

Proof.

In the first case, consider R=R= athermality with the thermal state γ\gamma, which is in a bipartite system with equal local dimension. Then it suffices to notice that in this case Eq. (G) becomes

ps|ME[M,𝔉(𝒩)]\displaystyle p_{s|{\rm ME}}[M,\mathfrak{F}(\mathcal{N}^{\prime})] 2[PDmax(𝒩)+1k]M×sup{|Φm}m=0M1m=0M1Φm|γ|Φm\displaystyle\leq\frac{2^{\left[P_{D_{\rm max}}(\mathcal{N}^{\prime})+\frac{1}{k}\right]}}{M}\times\sup_{\{|\Phi^{\prime}_{m}\rangle\}_{m=0}^{M-1}}\sum_{m=0}^{M-1}\langle\Phi^{\prime}_{m}|\gamma|\Phi^{\prime}_{m}\rangle
2[PDmax(𝒩)+1k]M.\displaystyle\leq\frac{2^{\left[P_{D_{\rm max}}(\mathcal{N}^{\prime})+\frac{1}{k}\right]}}{M}. (56)

This means we have CME,(1)ϵ[𝔉(𝒩)]PDmaxδ(𝒩)+log211ϵδC_{{\rm ME},(1)}^{\epsilon}[\mathfrak{F}(\mathcal{N})]\leq P_{D_{\rm max}}^{\delta}(\mathcal{N})+\log_{2}\frac{1}{1-\epsilon-\delta} for all 𝔉𝔽R\mathfrak{F}\in\mathbb{F}_{R}, and the desired bound follows.

Now we recall that a bipartite state ρ\rho with equal finite local dimension dd is free entangled Ent-RMP ; Horodecki1998 , multi-copy nonlocal Palazuelos2012 ; Cavalcanti2013 , and multi-copy steerable Hsieh2016 ; Quintino2016 if max(ρ)>1d\mathcal{F}_{\rm max}(\rho)>\frac{1}{d}. This means for these resources we have FR(d)1dF^{R}(d)\leq\frac{1}{d}. Given 𝔉𝔽R\mathfrak{F}\in\mathbb{F}_{R} and 𝒩𝒪R\mathcal{N}\in\mathcal{O}_{R}, Lemma G.1 implies that for every MM achieving ps|ME[M,𝔉(𝒩)]1ϵp_{s|{\rm ME}}\left[M,\mathfrak{F}(\mathcal{N})\right]\geq 1-\epsilon, we have [in what follows we again use dd to denote the local dimension of the output bipartite system of 𝔉(𝒩)\mathfrak{F}(\mathcal{N})]

0\displaystyle 0 PDmaxδ(𝒩)+log211ϵδ+log2FR(d)\displaystyle\leq P_{D_{\rm max}}^{\delta}(\mathcal{N})+\log_{2}\frac{1}{1-\epsilon-\delta}+\log_{2}F^{R}(d)
PDmaxδ(𝒩)+log211ϵδlog2M,\displaystyle\leq P_{D_{\rm max}}^{\delta}(\mathcal{N})+\log_{2}\frac{1}{1-\epsilon-\delta}-\log_{2}\sqrt{M}, (57)

where the second line is because for any such MM the output space of 𝔉(𝒩)\mathfrak{F}(\mathcal{N}) contains MM mutually orthonormal maximally entangled states, which means Md2M\leq d^{2} and hence 1d1M\frac{1}{d}\leq\frac{1}{\sqrt{M}}. From here we conclude that

12×log2MPDmaxδ(𝒩)+log211ϵδ,\displaystyle\frac{1}{2}\times\log_{2}M\leq P_{D_{\rm max}}^{\delta}(\mathcal{N})+\log_{2}\frac{1}{1-\epsilon-\delta}, (58)

which is the desired bound by considering all possible MM and 𝔉𝔽R\mathfrak{F}\in\mathbb{F}_{R}. ∎

Appendix H Implications of Theorem 5

Theorem 5 gives further implications to superdense coding and also connects FEF and resource preservability. We briefly summarize these remarks in this appendix.

H.1 Maintaining Orthogonal Maximal Entanglement and Superdense Coding

Theorem 5 allows an interpretation for superdense coding, and we briefly introduce the setup here to illustrate this. Consider two agents, Alice and Bob, sharing a maximally entangled state |Φ01di=0d1|ii|\Phi_{0}\rangle\coloneqq\frac{1}{\sqrt{d}}\sum_{i=0}^{d-1}|ii\rangle with local dimension dd. First, Alice encodes the classical information mm in her local system (a qudit) by applying UmU_{m}, the unitary operator achieving (Um𝕀B)|Φ0=|Φm(U_{m}\otimes\mathbb{I}_{\rm B})|\Phi_{0}\rangle=|\Phi_{m}\rangle with a given set of orthogonal maximally entangled states {|Φm}m=0M1\{|\Phi_{m}\rangle\}_{m=0}^{M-1}. After this, she sends her qudit to Bob, and both Alice’s and Bob’s qudits undergo a dynamics modeled by a bipartite channel 𝒩\mathcal{N}. After receiving Alice’s qudit, Bob decodes the classical information mm from 𝒩(|ΦmΦm|)\mathcal{N}(|\Phi_{m}\rangle\langle\Phi_{m}|) by a bipartite measurement. We call such task a dd dimensional superdense coding through 𝒩\mathcal{N}. It is the conventional superdesne coding when 𝒩=\mathcal{N}=\mathcal{I}, where Bob can apply a d2d^{2} dimensional Bell measurement to perfectly decode d2d^{2} classical data when only one qudit has been sent. In general, different channels have different abilities to admit superdense coding, and sup𝔉𝔽RCME,(1)ϵ[𝔉(𝒩)]\sup_{\mathfrak{F}\in\mathbb{F}_{R}}C^{\epsilon}_{{\rm ME},(1)}\left[\mathfrak{F}(\mathcal{N})\right] is the highest amount of classical information allowed by a dd dimensional superdense coding through a channel 𝒩𝒪R\mathcal{N}\in\mathcal{O}_{R} even with all possible assistance structures constrained by RR, i.e., 𝔉𝔽R\mathfrak{F}\in\mathbb{F}_{R}. In this sense, sup𝔉𝔽RCME,(1)ϵ[𝔉(𝒩)]\sup_{\mathfrak{F}\in\mathbb{F}_{R}}C^{\epsilon}_{{\rm ME},(1)}\left[\mathfrak{F}(\mathcal{N})\right] can be understood as the superdense coding ability of 𝒩\mathcal{N}, and Theorem 5 estimates the optimal performance of superdense coding.

H.2 Fully Entangled Fraction and Resource Preservability

It is worth mentioning that the proof of Theorem 5 largely relies on FEF. Once an FEF threshold with an explicit dependency of local dimension exists, a result similar to Theorem 5 can be obtained. For example, from Ref. Hsieh2016 we learn that ρ\rho is (two-way) steerable if max(ρ)>d1+d+1dd+1\mathcal{F}_{\rm max}(\rho)>\frac{d-1+\sqrt{d+1}}{d\sqrt{d+1}}. This means when R=R= (two-way) steerability and Md2M\leq d^{2}, we have FR(d)d1+d+1dd+11d+1d1M14+1M.F^{R}(d)\leq\frac{d-1+\sqrt{d+1}}{d\sqrt{d+1}}\leq\frac{1}{\sqrt{d}}+\frac{1}{d}\leq\frac{1}{M^{\frac{1}{4}}}+\frac{1}{\sqrt{M}}. Hence, when R=R= (two-way) steerability, we have

14×sup𝔉𝔽RCME,(1)ϵ[𝔉(𝒩)]PDmaxδ(𝒩)+log211ϵδ+1.\displaystyle\frac{1}{4}\times\sup_{\mathfrak{F}\in\mathbb{F}_{R}}C^{\epsilon}_{{\rm ME},(1)}\left[\mathfrak{F}(\mathcal{N})\right]\leq P_{D_{\rm max}}^{\delta}(\mathcal{N})+\log_{2}\frac{1}{1-\epsilon-\delta}+1. (59)

It turns out that resource preservability can be related to FEF as follows

Proposition H.1.

Given a resource RR, then 𝒩𝒪R\mathcal{N}\in\mathcal{O}_{R} cannot maintain any maximally entangled state with an average error less than ϵ\epsilon if

PDmax(𝒩)<log21ϵFR(d),\displaystyle P_{D_{\rm max}}(\mathcal{N})<\log_{2}\frac{1-\epsilon}{F^{R}(d)}, (60)

where dd is the local dimension of the output bipartite space of 𝒩\mathcal{N}.

Proof.

Applying Lemma G.1 with δ=0\delta=0, we learn that there exists no MM that can achieve ps(M,𝒩)1ϵp_{s}(M,\mathcal{N})\geq 1-\epsilon if

PDmax(𝒩)<log2FR(d)+log2(1ϵ).\displaystyle P_{D_{\rm max}}(\mathcal{N})<-\log_{2}F^{R}(d)+\log_{2}(1-\epsilon). (61)

In other words, 𝒩𝒪R\mathcal{N}\in\mathcal{O}_{R} cannot maintain any maximally entangled state with an average error lower than ϵ\epsilon when this inequality is satisfied. ∎

Proposition H.1 gives a new way to understand FEF: Once a channel’s resource preservability is not strong enough compared with a threshold induced by FEF, it is impossible to maintain maximal entanglement to the desired level. For examples, suppose 𝒩\mathcal{N} is a free operation of free entanglement (or, similarly, multi-copy nonlocality or multi-copy steerability; note that we need to assume Assumptions 1 and 2), then Proposition H.1 implies that 𝒩\mathcal{N} cannot maintain any maximally entangled state with an average error lower than ϵ\epsilon if PDmax(𝒩)<log2d(1ϵ).P_{D_{\rm max}}(\mathcal{N})<\log_{2}d(1-\epsilon).

Appendix I Proof of Theorem 6

To demonstrate the proof, we will first provide a generalized version of the entanglement preserving local thermalization introduced in Ref. Hsieh2020-2 . After that, Theorem 6 can be proved by using it and the capacity-like measure CME,(1)C_{\rm ME,(1)} defined in Eq. (13). Before proceeding, we remark that, in this appendix, we will use subscripts to denote the located subsystems for both states and channels; e.g., ρX\rho_{\rm X} and X\mathcal{E}_{\rm X} are a state and a channel in the system X{\rm X}, respectively. Also, following Ref. Hsieh2020-2 , we assume no energy degeneracy for all subsystems, and all thermal states are assumed to be full-rank.

To start with, we recall from Ref. Hsieh2020-2 the following alternative definition of a local thermalization with a given pair of thermal states (γA,γB)(\gamma_{\rm A},\gamma_{\rm B}) on the subsystem A,B{\rm A,B}, respectively [recall from Eq. (22) the definition of local operation and pre-shared randomness (LOSR) channels]:

Definition I.1.

Hsieh2020-2 A bipartite channel AB\mathcal{E}_{\rm AB} on AB{\rm AB} is called a local thermalization to (γA,γB)(\gamma_{\rm A},\gamma_{\rm B}) if

  1. 1.

    AB\mathcal{E}_{\rm AB} is a LOSR channel in AB{\rm AB} bipartition.

  2. 2.

    trBAB(ρAB)=γA{\rm tr}_{\rm B}\circ\mathcal{E}_{\rm AB}(\rho_{\rm AB})=\gamma_{\rm A} and trAAB(ρAB)=γB{\rm tr}_{\rm A}\circ\mathcal{E}_{\rm AB}(\rho_{\rm AB})=\gamma_{\rm B} for all states ρAB\rho_{\rm AB}.

As remarked in Ref. Hsieh2020-2 , a local thermalization is local in the sense that it is a LOSR channel that can locally thermalize all inputs. This definition is equivalent to the one given in the main text, and we refer the reader to Definition 1, Definition 2, and Theorem 2 in Ref. Hsieh2020-2 for the detailed reasoning.

Now, consider a tripartite system ABC{\rm ABC} with finite local dimensions d,d,d2+1d,d,d^{2}+1, respectively. As mentioned in the main text, we will focus on the bipartition A|BC{\rm A|BC}, and the subsystem C{\rm C} can be treated as an ancillary system possessed by the local agent in B{\rm B}. In what follows, γX\gamma_{{\rm X}} is the given thermal state of the subsystem X=A,B,C{\rm X=A,B,C} with temperature TXT_{\rm X} and Hamiltonian HXH_{\rm X}. Suppose {|nX}n=0DX1\{|n\rangle_{{\rm X}}\}_{n=0}^{D_{\rm X}-1} is the orthonormal energy eigenbasis of HXH_{\rm X} with local dimension DXD_{\rm X}. Then we define the following channel, which has a schematic interpretation given in Fig. 2. First, consider a given ordered set of local unitary operators in B{\rm B} denoted by 𝐕{VB(n)}n=0d2,{\bf V}\coloneqq\left\{V^{(n)}_{\rm B}\right\}_{n=0}^{d^{2}}, where VB(d2)𝕀BV^{(d^{2})}_{\rm B}\coloneqq\mathbb{I}_{\rm B} is always assumed to be the maximally mixed state, and other elements are arbitrary. Then, for every κ[0,1]\kappa\in[0,1], we define

A|BC𝐕,κ(𝒟ABκC)(ABC𝐕)(𝒯ABC),\displaystyle\mathcal{E}_{\rm A|BC}^{{\bf V},\kappa}\coloneqq\left(\mathcal{D}_{\rm AB}^{\kappa}\otimes\mathcal{I}_{\rm C}\right)\circ\left(\mathcal{I}_{\rm A}\otimes\mathcal{L}_{\rm BC}^{\bf V}\right)\circ\left(\mathcal{T}_{\rm AB}\otimes\mathcal{I}_{\rm C}\right), (62)

whose components are defined as follows. First, in a bipartite system with finite equal local dimension,

𝒯()(UU)()(UU)𝑑U\displaystyle\mathcal{T}(\cdot)\coloneqq\int(U\otimes U^{*})(\cdot)(U\otimes U^{*})^{\dagger}dU (63)

is the (UU)(U\otimes U^{*})-twirling operation Horodecki1999-1 ; Bennett1996 used to symmetrize the local states. To encode classical information into the correlation shared by A,B{\rm A,B}, we define the following local channel in BC{\rm BC}:

BC𝐕()trC[n=0d2(VB(n)|nn|C)()(VB(n)|nn|C)]γC.\displaystyle\mathcal{L}_{\rm BC}^{\bf V}(\cdot)\coloneqq{\rm tr}_{\rm C}\left[\sum_{n=0}^{d^{2}}\left(V^{(n)}_{\rm B}\otimes|n\rangle\langle n|_{\rm C}\right)(\cdot)\left(V^{(n)}_{\rm B}\otimes|n\rangle\langle n|_{\rm C}\right)^{\dagger}\right]\otimes\gamma_{\rm C}. (64)

Using operator-sum representation QCI-book , one can check that this is indeed a channel. Finally,

𝒟ABκ()(1κ)γ~Aκγ~Bκ+κ×()\displaystyle\mathcal{D}_{\rm AB}^{\kappa}(\cdot)\coloneqq(1-\kappa)\widetilde{\gamma}_{\rm A}^{\kappa}\otimes\widetilde{\gamma}_{\rm B}^{\kappa}+\kappa\times(\cdot) (65)

where

γ~XκγX+κ1κ(γX𝕀Xd)\displaystyle\widetilde{\gamma}_{\rm X}^{\kappa}\coloneqq\gamma_{\rm X}+\frac{\kappa}{1-\kappa}\left(\gamma_{\rm X}-\frac{\mathbb{I}_{\rm X}}{d}\right) (66)

is an alternative thermal state for X=A,B{\rm X=A,B}. Let pmin|ABmin{pmin(γA);pmin(γB)}p_{\rm min|AB}\coloneqq\min\{p_{\rm min}(\gamma_{{\rm A}});p_{\rm min}(\gamma_{{\rm B}})\}, which is the smallest eigenvalue among γA,γB\gamma_{\rm A},\gamma_{\rm B}. Write γBCγBγC\gamma_{\rm BC}\coloneqq\gamma_{\rm B}\otimes\gamma_{\rm C}, then we have the following result:

Theorem I.2.

For every d<d<\infty and 𝐕{\bf V}, we have

  1. 1.

    A|BC𝐕,κ\mathcal{E}_{\rm A|BC}^{{\bf V},\kappa} is a local thermalization to (γA,γBC)(\gamma_{\rm A},\gamma_{\rm BC}) for all 0κκdpmin|AB0\leq\kappa\leq\kappa_{*}\coloneqq dp_{\rm min|AB}.

  2. 2.

    A|BC𝐕,κ=κ\mathcal{E}_{\rm A|BC}^{{\bf V},\kappa=\kappa_{*}} is an entanglement preserving local thermalization to (γA,γBC)(\gamma_{\rm A},\gamma_{\rm BC}) for all full-rank γA\gamma_{{\rm A}} and γB\gamma_{{\rm B}}.

Proof.

Following Lemma 1 in Ref. Hsieh2020-2 , we conclude that the local output of A|BC𝐕,κ\mathcal{E}_{\rm A|BC}^{{\bf V},\kappa} in A{\rm A} and BC{\rm BC} will always be γA\gamma_{{\rm A}} and γBC\gamma_{{\rm BC}}, respectively. Also, 0κκ0\leq\kappa\leq\kappa_{*} guarantees that both γ~Aκ,γ~Bκ\widetilde{\gamma}_{{\rm A}}^{\kappa},\widetilde{\gamma}_{{\rm B}}^{\kappa} are legal quantum states. This implies that A|BC𝐕,κ\mathcal{E}_{\rm A|BC}^{{\bf V},\kappa} is a legal channel, which is by definition a LOSR channel in the A|BC{\rm A|BC} bipartition. Hence, for all 𝐕{\bf V}, A|BC𝐕,κ\mathcal{E}_{\rm A|BC}^{{\bf V},\kappa} is a local thermalization when κ\kappa is in the desired range. Finally, when γA\gamma_{{\rm A}} and γB\gamma_{{\rm B}} are both full-rank (and hence non-pure), Theorem 1 in Ref. Hsieh2020-2 implies that A|BC𝐕,κ=κ(|Ψd+Ψd+|AB|d2d2|C)\mathcal{E}_{\rm A|BC}^{{\bf V},\kappa=\kappa_{*}}(|\Psi_{d}^{+}\rangle\langle\Psi_{d}^{+}|_{\rm AB}\otimes|d^{2}\rangle\langle d^{2}|_{\rm C}) will be entangled in the A|BC{\rm A|BC} bipartition since in the subsystem AB{\rm AB} it is identical to the entanglement preserving local thermalization constructed in Ref. Hsieh2020-2 . This proves the desired claim. ∎

Refer to caption
Figure 2: Schematic interpretation of Eq. (62) and Theorem I.2. We use κ=κ\kappa=\kappa_{*} as an example. As shown in the figure, the channel A|BC𝐕,κ=κ\mathcal{E}_{\rm A|BC}^{{\bf V},\kappa=\kappa_{*}} can be decomposed into three steps. First, a (UU)(U\otimes U^{*})-twirling operation 𝒯AB\mathcal{T}_{\rm AB} is applied, which only requires local unitary operations and shared randomness (the thick dashed line). After that, the channel BC=BC𝐕\mathcal{L}_{\rm BC}=\mathcal{L}_{\rm BC}^{\bf V} is applied locally in the subsystem BC{\rm BC}. Finally, using shared randomness again, the channel 𝒟AB=𝒟ABκ\mathcal{D}_{\rm AB}=\mathcal{D}_{\rm AB}^{\kappa_{*}} can be achieved. The resulting global channel has the property that, when one only observes it locally in A{\rm A} and BC{\rm BC}, its is identical to a full thermalization channel; namely, a state preparation channel ΦγX()γXtr()\Phi_{\gamma_{\rm X}}(\cdot)\coloneqq\gamma_{\rm X}{\rm tr}(\cdot) of the given thermal state in the subsystem X{\rm X}, where X=A{\rm X=A} or BC{\rm BC}. Still, as shown in this section, this channel is able to preserve entanglement in the AB{\rm AB} bipartition, and it is also possible to transmit classical information by encoding the information into the global correlation shared by A{\rm A} and B{\rm B}.

We leave a schematic interpretation in Fig. 2, and now we proceed to prove Theorem 6 by using Theorem I.2:

Proof.

Note that by assuming each subsystem Hamiltonian HXH_{{\rm X}} to be non-degenerate and finite-energy, the corresponding thermal state γX\gamma_{{\rm X}} is non-pure and full-rank if and only if TX>0T_{{\rm X}}>0 Hsieh2020-2 . Hence, for every TX>0T_{{\rm X}}>0, we conclude that A|BC𝐕,κ\mathcal{E}_{\rm A|BC}^{{\bf V},\kappa_{*}} is an entanglement preserving local thermalization to (γA,γBC)(\gamma_{\rm A},\gamma_{\rm BC}) according to Theorem I.2. Now we note that the one-shot classical capacity of A|BC𝐕,κ\mathcal{E}_{\rm A|BC}^{{\bf V},\kappa_{*}} can be estimated by the ability of 𝒟ABκ\mathcal{D}_{\rm AB}^{\kappa_{*}} to maintain maximally entangled bases (see Sec. V). More precisely, for a given Md2M\leq d^{2} and sets of orthonormal maximally entangled states (in the A|B{\rm A|B} bipartition) {|Φm}m=0M1,{|Φm}m=0M1\left\{|\Phi_{m}\rangle\right\}_{m=0}^{M-1},\left\{|\Phi^{\prime}_{m}\rangle\right\}_{m=0}^{M-1}, the corresponding success probability of 𝒟ABκ\mathcal{D}_{\rm AB}^{\kappa_{*}} reads 1Mm=0M1Φm|𝒟ABκ(|ΦmΦm|)|Φm\frac{1}{M}\sum_{m=0}^{M-1}\langle\Phi^{\prime}_{m}|\mathcal{D}_{\rm AB}^{\kappa_{*}}(|\Phi_{m}\rangle\langle\Phi_{m}|)|\Phi^{\prime}_{m}\rangle. Being maximally entangled, there exist MM unitary operators in B{\rm B}, {UB(m)}m=0M1\{U^{(m)}_{\rm B}\}_{m=0}^{M-1}, such that |Φm=(𝕀AUB(m))|Ψd+|\Phi_{m}\rangle=\left(\mathbb{I}_{\rm A}\otimes U^{(m)}_{\rm B}\right)|\Psi_{d}^{+}\rangle, where |Ψd+i=0d11d|ii|\Psi_{d}^{+}\rangle\coloneqq\sum_{i=0}^{d-1}\frac{1}{\sqrt{d}}|ii\rangle. Define the set 𝐔={UB(m)}m=0d2{\bf U}=\{{U}_{\rm B}^{(m)}\}_{m=0}^{d^{2}} with UB(m)=𝕀B{U}^{(m)}_{\rm B}=\mathbb{I}_{\rm B} for all mMm\geq M. Now we choose the encoding {ρm}m=0M1\{\rho_{m}\}_{m=0}^{M-1} as

ρm|Ψd+Ψd+|AB|mm|C\displaystyle\rho_{m}\coloneqq|\Psi_{d}^{+}\rangle\langle\Psi_{d}^{+}|_{\rm AB}\otimes|m\rangle\langle m|_{\rm C} (67)

and decoding POVM {Em}m=0M1\left\{E_{m}\right\}_{m=0}^{M-1} as

Em|ΦmΦm|AB𝕀Cm<M1;\displaystyle E_{m}\coloneqq|\Phi^{\prime}_{m}\rangle\langle\Phi^{\prime}_{m}|_{\rm AB}\otimes\mathbb{I}_{\rm C}\quad\forall\;m<M-1; (68)
EM1𝕀ABCm=0M2Em=(𝕀ABm=0M2|ΦmΦm|AB)𝕀C.\displaystyle E_{M-1}\coloneqq\mathbb{I}_{\rm ABC}-\sum_{m=0}^{M-2}E_{m}=\left(\mathbb{I}_{\rm AB}-\sum_{m=0}^{M-2}|\Phi^{\prime}_{m}\rangle\langle\Phi^{\prime}_{m}|_{\rm AB}\right)\otimes\mathbb{I}_{\rm C}. (69)

Then we have

1Mm=0M1Φm|𝒟ABκ(|ΦmΦm|)|Φm\displaystyle\frac{1}{M}\sum_{m=0}^{M-1}\langle\Phi^{\prime}_{m}|\mathcal{D}_{\rm AB}^{\kappa_{*}}(|\Phi_{m}\rangle\langle\Phi_{m}|)|\Phi^{\prime}_{m}\rangle =1Mm=0M1tr[(|ΦmΦm|AB𝕀C)A|BC𝐔,κ(ρm)]\displaystyle=\frac{1}{M}\sum_{m=0}^{M-1}{\rm tr}\left[\left(|\Phi^{\prime}_{m}\rangle\langle\Phi^{\prime}_{m}|_{\rm AB}\otimes\mathbb{I}_{\rm C}\right)\mathcal{E}_{\rm A|BC}^{{\bf U},\kappa_{*}}(\rho_{m})\right]
1Mm=0M1tr[EmA|BC𝐔,κ(ρm)],\displaystyle\leq\frac{1}{M}\sum_{m=0}^{M-1}{\rm tr}\left[E_{m}\mathcal{E}_{\rm A|BC}^{{\bf U},\kappa_{*}}(\rho_{m})\right], (70)

where the inequality is due to the fact that |ΦM1ΦM1|AB𝕀CEM1|\Phi^{\prime}_{M-1}\rangle\langle\Phi^{\prime}_{M-1}|_{\rm AB}\otimes\mathbb{I}_{\rm C}\leq E_{M-1}. This means [see Eq. (13) for the definition of CME,(1)ϵC_{\rm ME,(1)}^{\epsilon}]

sup𝐕C(1)ϵ(A|BC𝐕,κ)CME,(1)ϵ(𝒟ABκ),\displaystyle\sup_{{\bf V}}C_{\rm(1)}^{\epsilon}\left(\mathcal{E}_{\rm A|BC}^{{\bf V},\kappa_{*}}\right)\geq C_{\rm ME,(1)}^{\epsilon}\left(\mathcal{D}_{\rm AB}^{\kappa_{*}}\right), (71)

where the optimization is taken over every possible ordered set of d2+1d^{2}+1 unitary operators 𝐕{VB(n)}n=0d2{\bf V}\coloneqq\left\{V^{(n)}_{\rm B}\right\}_{n=0}^{d^{2}} with VB(d2)𝕀BV^{(d^{2})}_{\rm B}\coloneqq\mathbb{I}_{\rm B}. On the other hand, we have

1Mm=0M1Φm|𝒟ABκ(|ΦmΦm|)|Φm=1κMm=0M1Φm|(γ~Aκγ~Bκ)|Φm+κ.\displaystyle\frac{1}{M}\sum_{m=0}^{M-1}\langle\Phi_{m}|\mathcal{D}_{\rm AB}^{\kappa_{*}}(|\Phi_{m}\rangle\langle\Phi_{m}|)|\Phi_{m}\rangle=\frac{1-\kappa_{*}}{M}\sum_{m=0}^{M-1}\langle\Phi_{m}|\left(\widetilde{\gamma}_{\rm A}^{\kappa_{*}}\otimes\widetilde{\gamma}_{\rm B}^{\kappa_{*}}\right)|\Phi_{m}\rangle+\kappa_{*}. (72)

Note that 𝒟ABκ\mathcal{D}_{\rm AB}^{\kappa_{*}} has output system dimension d2d^{2}. Hence, when M=d2M=d^{2}, the success probability is given by 1κd2+κ.\frac{1-\kappa_{*}}{d^{2}}+\kappa_{*}. This means that CME,(1)ϵ(𝒟ABκ)=log2d2C_{\rm ME,(1)}^{\epsilon}\left(\mathcal{D}_{\rm AB}^{\kappa_{*}}\right)=\log_{2}d^{2} if 1κd2+κ1ϵ\frac{1-\kappa_{*}}{d^{2}}+\kappa_{*}\geq 1-\epsilon, or, equivalently,

ϵ(11d2)(1κ).\displaystyle\epsilon\geq\left(1-\frac{1}{d^{2}}\right)(1-\kappa_{*}). (73)

Consequently, by using Eq. (71), we learn that when Eq. (73) holds, there must exists an 𝐕{\bf V} such that C(1)ϵ(A|BC𝐕,κ)log2d2C_{\rm(1)}^{\epsilon}\left(\mathcal{E}_{\rm A|BC}^{{\bf V},\kappa_{*}}\right)\geq\log_{2}d^{2}. The result follows. ∎

References

  • (1) E. Chitambar and G. Gour, Quantum resource theories, Rev. Mod. Phys. 91, 025001 (2019).
  • (2) R, Horodecki, P, Horodecki, M, Horodecki, and K, Horodecki, Quantum entanglement, Rev. Mod. Phys. 81, 865 (2009).
  • (3) M. A. Nielsen and I. L. Chuang, Quantum Computation and Quantum Information, (Cambridge University Press, Cambridge, UK, 2000).
  • (4) L. Wang and R. Renner, One-shot classical-quantum capacity and hypothesis testing, Phys. Rev. Lett. 108, 200501 (2012).
  • (5) N. Datta, M. Mosonyi, M.-H. Hsieh, and F. G. S. L. Branda~\tilde{\rm a}o, A smooth entropy approach to quantum hypothesis testing and the classical capacity of quantum channels, IEEE Trans. Inf. Theory 59, 8014 (2013).
  • (6) V. Vedral, M. B. Plenio, M. A. Rippin, and P. L. Knight, Quantifying entanglement, Phys. Rev. Lett. 78, 2275 (1997).
  • (7) G. Vidal and R. F. Werner, Computable measure of entanglement, Phys. Rev. A 65, 032314 (2002).
  • (8) A. Streltsov, G. Adesso, and M. B. Plenio, Colloquium: Quantum coherence as a resource, Rev. Mod. Phys. 89, 041003 (2017).
  • (9) T. Baumgratz, M. Cramer, and M. B. Plenio, Quantifying coherence, Phys. Rev. Lett. 113, 140401 (2014).
  • (10) J. S. Bell, On the Einstein Podolsky Rosen paradox, Physics Physique Fizika 1, 195 (1964).
  • (11) N. Brunner, D. Cavalcanti, S. Pironio, V. Scarani, and S. Wehner, Bell nonlocality, Rev. Mod. Phys. 86, 419 (2014).
  • (12) E. Wolfe, D. Schmid, A. B. Sainz, R. Kunjwal, and R. W. Spekkens, Quantifying Bell: The resource theory of nonclassicality of common-cause boxes, Quantum 4, 280 (2020).
  • (13) H. M. Wiseman, S. J. Jones, and A. C. Doherty, Steering, entanglement, nonlocality, and the Einstein-Podolsky-Rosen paradox, Phys. Rev. Lett. 98, 140402 (2007).
  • (14) S. J. Jones, H. M. Wiseman, and A. C. Doherty, Entanglement, Einstein-Podolsky-Rosen correlations, Bell nonlocality, and steering, Phys. Rev. A 76, 052116 (2007).
  • (15) D. Cavalcanti and P. Skrzypczyk, Quantum steering: A review with focus on semidefinite programming, Rep. Prog. Phys. 80, 024001 (2017).
  • (16) R. Uola, A. C. S. Costa, H. C. Nguyen, and O. Gu¨\ddot{\rm u}hne, Quantum steering, Rev. Mod. Phys. 92, 015001 (2020).
  • (17) P. Skrzypczyk, M. Navascués, and D. Cavalcanti, Quantifying Einstein-Podolsky-Rosen steering, Phys. Rev. Lett. 112, 180404 (2014).
  • (18) M. Piani and J. Watrous, Necessary and sufficient quantum information characterization of Einstein-Podolsky-Rosen steering, Phys. Rev. Lett. 114, 060404 (2015).
  • (19) R. Gallego and L. Aolita, Resource theory of steering, Phys. Rev. X 5, 041008 (2015).
  • (20) G. Gour and R. W. Spekkens, The resource theory of quantum reference frames: Manipulations and monotones, New J. Phys. 10, 033023 (2008).
  • (21) I. Marvian and R. W. Spekkens, How to quantify coherence: Distinguishing speakable and unspeakable notions, Phys. Rev. A 94, 052324 (2016).
  • (22) R. Takagi, Skew informations from an operational view via resource theory of asymmetry, Sci. Rep. 9, 14562 (2019).
  • (23) F. G. S. L. Branda~\tilde{\rm a}o, M. Horodecki, J. Oppenheim, J. M. Renes, and R. W. Spekkens, Resource theory of quantum states out of thermal equilibrium, Phys. Rev. Lett. 111, 250404 (2013).
  • (24) M. Horodecki and J. Oppenheim, Fundamental limitations for quantum and nanoscale thermodynamics, Nat. Commun. 4, 2059 (2013).
  • (25) F. G. S. L. Branda~\tilde{\rm a}o, M. Horodecki, N. Ng, J. Oppenheim, and S. Wehner, The second laws of quantum thermodynamics, Proc. Natl. Acad. Sci. U.S.A. 112, 3275 (2015).
  • (26) M. Lostaglio, An introductory review of the resource theory approach to thermodynamics, Rep. Prog. Phys. 82, 114001 (2019).
  • (27) V. Narasimhachar, S. Assad, F. C. Binder, J. Thompson, B. Yadin, and M. Gu, Thermodynamic resources in continuous-variable quantum systems, arXiv:1909.07364.
  • (28) A. Serafini, M. Lostaglio, S. Longden, U. Shackerley-Bennett, C.-Y. Hsieh, and G. Adesso, Gaussian thermal operations and the limits of algorithmic cooling, Phys. Rev. Lett. 124, 010602 (2020).
  • (29) M. Horodecki and J. Oppenheim, (Quantumness in the context of) resource theories, Int. J. Mod. Phys. B 27, 1345019 (2013).
  • (30) F. G. S. L. Branda~\tilde{\rm a}o and G. Gour, Reversible framework for quantum resource theories, Phys. Rev. Lett. 115, 070503 (2015).
  • (31) L. del Rio, L. Kraemer, and R. Renner, Resource theories of knowledge, arXiv:1511.08818.
  • (32) B. Coecke, T. Fritz, and R. W. Spekkens, A mathematical theory of resources, Inf. Comput. 250, 59 (2016).
  • (33) G. Gour, Quantum resource theories in the single-shot regime, Phys. Rev. A 95, 062314 (2017).
  • (34) Z.-W. Liu, X. Hu, and S. Lloyd, Resource destroying maps, Phys. Rev. Lett. 118, 060502 (2017).
  • (35) A. Anshu, M.-H. Hsieh, and R. Jain, Quantifying resources in general resource theory with catalysts, Phys. Rev. Lett. 121, 190504 (2018).
  • (36) L. Lami, B. Regula, X. Wang, R. Nichols, A. Winter, and G. Adesso, Gaussian quantum resource theories, Phys. Rev. A 98, 022335 (2018).
  • (37) B. Regula, Convex geometry of quantum resource quantification, J. Phys. A: Math. Theor. 51, 045303 (2018).
  • (38) Z.-W. Liu, K. Bu, and R. Takagi, One-shot operational quantum resource theory, Phys. Rev. Lett. 123, 020401 (2019).
  • (39) K. Fang and Z.-W. Liu, No-go theorems for quantum resource purification, Phys. Rev. Lett. 125, 060405 (2020).
  • (40) R. Takagi and B. Regula, General resource theories in quantum mechanics and beyond: Operational characterization via discrimination tasks, Phys. Rev. X 9, 031053 (2019).
  • (41) R. Takagi, B. Regula, K. Bu, Z.-W. Liu, and G. Adesso, Operational advantage of quantum resources in subchannel discrimination, Phys. Rev. Lett. 122, 140402 (2019).
  • (42) L. Li, K. Bu and Z.-W. Liu, Quantifying the resource content of quantum channels: An operational approach, Phys. Rev. A 101, 022335 (2020).
  • (43) K. Korzekwa, Z. Puchała, M. Tomamichel, and K. Z˙\dot{\rm Z}yczkowski, Encoding classical information into quantum resources, arXiv:1911.12373.
  • (44) Channels, or quantum channels, also known as completely-positive trace-preserving maps QCI-book ), are general mathematical forms used to depict physical processes and dynamics.
  • (45) J.-H. Hsieh, S.-H. Chen, and C.-M. Li, Quantifying quantum-mechanical processes, Scientific Reports 7, 13588 (2017).
  • (46) C.-C. Kuo, S.-H. Chen, W.-T. Lee, H.-M. Chen, H. Lu, and C.-M. Li, Quantum process capability, Scientific Reports 9, 20316 (2019).
  • (47) K. B. Dana, M. G. Díaz, M. Mejatty, and A. Winter, Resource theory of coherence: Beyond states, Phys. Rev. A 95, 062327 (2017).
  • (48) S. Pirandola, R. Laurenza, C. Ottaviani, and L. Banchi, Fundamental limits of repeaterless quantum communications, Nat. Commun. 8, 15043 (2017).
  • (49) M. G. Díaz, K. Fang, X. Wang, M. Rosati, M. Skotiniotis, J. Calsamiglia, and A. Winter, Using and reusing coherence to realize quantum processes, Quantum 2, 100 (2018).
  • (50) D. Rosset, F. Buscemi, and Y.-C. Liang, A resource theory of quantum memories and their faithful verification with minimal assumptions, Phys. Rev. X 8, 021033 (2018).
  • (51) M. M. Wilde, Entanglement cost and quantum channel simulation, Phys. Rev. A 98, 042338 (2018).
  • (52) Q. Zhuang, P. W. Shor, and J. H. Shapiro, Resource theory of non-Gaussian operations, Phys. Rev. A 97, 052317 (2018).
  • (53) S. Ba¨\ddot{\rm a}uml, S. Das, X. Wang, and M. M. Wilde, Resource theory of entanglement for bipartite quantum channels, arXiv:1907.04181.
  • (54) J. R. Seddon and E. Campbell, Quantifying magic for multi-qubit operations, Proc. R. Soc. A 475, 20190251 (2019).
  • (55) Z.-W. Liu and A. Winter, Resource theories of quantum channels and the universal role of resource erasure, arXiv:1904.04201.
  • (56) Y. Liu and X. Yuan, Operational resource theory of quantum channels, Phys. Rev. Research 2, 012035(R) (2020).
  • (57) G. Gour, Comparison of quantum channels by superchannels, IEEE Trans. Inf. Theory 65, 5880 (2019).
  • (58) G. Gour and A. Winter, How to quantify a dynamical resource? Phys. Rev. Lett. 123, 150401 (2019).
  • (59) G. Gour and C. M. Scandolo, The entanglement of a bipartite channel, arXiv:1907.02552.
  • (60) R. Takagi, K. Wang, M. Hayashi, Application of the resource theory of channels to communication scenarios, Phys. Rev. Lett. 124, 120502 (2020).
  • (61) T. Theurer, D. Egloff, L. Zhang, and M. B. Plenio, Quantifying operations with an application to coherence, Phys. Rev. Lett. 122, 190405 (2019).
  • (62) X. Wang and M. M. Wilde, Resource theory of asymmetric distinguishability for quantum channels, Phys. Rev. Research 1, 033169 (2019).
  • (63) G. D. Berk, A. J. P. Garner, B. Yadin, K. Modi, and F. A. Pollock, Resource theories of multi-time processes: A window into quantum non-Markovianity, arXiv:1907.07003.
  • (64) G. Saxena, E. Chitambar,and G. Gour, Dynamical resource theory of quantum coherence, Phys. Rev. Research 2, 023298 (2020).
  • (65) Y. Zhang, R. A. Bravo, V. O. Lorenz, and E. Chitambar, Channel activation of CHSH nonlocality, New J Phys. 22, 043003 (2020).
  • (66) C.-Y. Hsieh, Resource preservability, Quantum 4, 244 (2020).
  • (67) H. Kristjánsson, G. Chiribella, S. Salek, D. Ebler, and M. Wilson, Resource theories of communication, New J. Phys. 22, 073014 (2020).
  • (68) C. Sparaciari, M. Goihl, P. Boes, J. Eisert, and N. Ng, Bounding the resources for thermalizing many-body localized systems, arXiv:1912.04920.
  • (69) R. Landauer, Irreversibility and Heat Generation in the Computing Process, IBM J. Res. Develop., 5, 183 (1961).
  • (70) To clarify different notions, note that resource annihilating channels introduced in Ref. Hsieh2020-1 is different from resource destroying maps Liu2017 , where the latter keep free states invariant and map resourceful states to some free states. For instance, a state preparation channel preparing a fixed free state is a legal resource annihilating channel, while it is not a resource destroying map. As another remark, when the resource is entanglement, the notion of resource annihilating channel coincides with entanglement-annihilating channel Moravcikova2010 , which is different from entanglement-breaking channel Horodecki2003 .
  • (71) C. Palazuelos, Superactivation of quantum nonlocality, Phys. Rev. Lett. 109, 190401 (2012).
  • (72) D. Cavalcanti, A. Acin, N. Brunner, and T. Vertesi, All quantum states useful for teleportation are nonlocal resources, Phys. Rev. A 87, 042104 (2013).
  • (73) C.-Y. Hsieh, Y.-C. Liang, and R.-K. Lee, Quantum steerability: Characterization, quantification, superactivation and unbounded amplification, Phys. Rev. A 94, 062120 (2016).
  • (74) M. T. Quintino, M. Huber, and N. Brunner, Superactivation of quantum steering, Phys. Rev. A 94, 062123 (2016).
  • (75) G. Chiribella, G. M. D’Ariano, and P. Perinotti, Transforming quantum operations: quantum supermaps, EPL (Europhysics Letters) 83, 30004 (2008).
  • (76) G. Chiribella, G. M. D’Ariano, and P. Perinotti, Quantum circuit architecture, Phys. Rev. Lett. 101, 060401 (2008).
  • (77) N. Datta, Min- and max-relative entropies and a new entanglement monotone, IEEE Trans. Inf. Theory 55, 2816 (2009).
  • (78) M. Horodecki, P. Horodecki, and R. Horodecki, General teleportation channel, singlet fraction, and quasidistillation, Phys. Rev. A 60, 1888 (1999).
  • (79) In a bipartite system with equal local dimension dd, isotropic states are given by ρiso(p)=p|Ψd+Ψd+|+(1p)𝕀d2\rho_{\rm iso}(p)=p|\Psi_{d}^{+}\rangle\langle\Psi_{d}^{+}|+(1-p)\frac{\mathbb{I}}{d^{2}}, where p[1d21,1]p\in\left[-\frac{1}{d^{2}-1},1\right] due to the positivity of quantum states Horodecki1999-2 . From here we conclude that, for every MM, POVM in dd-dimensional system {Em(d)}m=0M1\{E_{m}^{(d)}\}_{m=0}^{M-1}, and isotropic states {ρiso(pm)}m=0M1\{\rho_{\rm iso}(p_{m})\}_{m=0}^{M-1}, we have m=0M1tr[Em(d)ρiso(pm)]=m=0M1pmtr(Em(d)|Ψd+Ψd+|)+m=0M1(1pm)tr(Em(d)𝕀d2)m=0M1tr(Em(d)|Ψd+Ψd+|)+(1+1d21)m=0M1tr(Em(d)𝕀d2)=2+1d21<2×d2d21\sum_{m=0}^{M-1}{\rm tr}\left[E_{m}^{(d)}\rho_{\rm iso}(p_{m})\right]=\sum_{m=0}^{M-1}p_{m}{\rm tr}\left(E_{m}^{(d)}|\Psi_{d}^{+}\rangle\langle\Psi_{d}^{+}|\right)+\sum_{m=0}^{M-1}(1-p_{m}){\rm tr}\left(E_{m}^{(d)}\frac{\mathbb{I}}{d^{2}}\right)\leq\sum_{m=0}^{M-1}{\rm tr}\left(E_{m}^{(d)}|\Psi_{d}^{+}\rangle\langle\Psi_{d}^{+}|\right)+\left(1+\frac{1}{d^{2}-1}\right)\sum_{m=0}^{M-1}{\rm tr}\left(E_{m}^{(d)}\frac{\mathbb{I}}{d^{2}}\right)=2+\frac{1}{d^{2}-1}<2\times\frac{d^{2}}{d^{2}-1}. Hence, we have Γκδ(𝒩)log2(2×d2d21)\Gamma_{\kappa}^{\delta}(\mathcal{N})\leq\log_{2}\left(2\times\frac{d^{2}}{d^{2}-1}\right) for every 𝒩,κ,δ\mathcal{N},\kappa,\delta.
  • (80) M. Hayashi and H. Nagaoka, General formulas for capacity of classical-quantum channels, IEEE Trans. Inf. Theory, 49, 1753 (2003).
  • (81) M. Tomamichel and M. Hayashi, A hierarchy of information quantities for finite block length analysis of quantum tasks, IEEE Trans. Inf. Theory, 59,7693 (2013).
  • (82) Felix Leditzky, Relative entropies and their use in quantum information theory, PhD thesis, University of Cambridge, 2016.
  • (83) Let {|ϕi}i\{|\phi_{i}\rangle\}_{i} be the basis defining the coherence. One way to construct coherence-annihilating channels is to consider 𝒟\mathcal{D}\circ\mathcal{E}, where \mathcal{E} is an arbitrary channel and 𝒟()i|ϕiϕi|()|ϕiϕi|\mathcal{D}(\cdot)\coloneqq\sum_{i}|\phi_{i}\rangle\langle\phi_{i}|(\cdot)|\phi_{i}\rangle\langle\phi_{i}| is the dephasing channel with respect to the given basis {|ϕi}i\{|\phi_{i}\rangle\}_{i} Saxena2019 . Then the channel 𝒟\mathcal{D}\circ\mathcal{E} can only output states diagonalized in the given basis, thereby being incoherent states. See Ref. Saxena2019 for more details. In this work, we use the energy eigenbasis of the given Hamiltonian to define coherence.
  • (84) L. del Rio, J. Åberg, R. Renner, O. Dahlsten, and V. Vedral, The thermodynamic meaning of negative entropy, Nature 474, 61 (2011).
  • (85) C. H. Bennett, G. Brassard, C. Crépeau, R. Jozsa, A. Peres, and W. K. Wootters, Teleporting an unknown quantum state via dual classical and Einstein-Podolsky-Rosen channels, Phys. Rev. Lett. 70, 1895 (1993).
  • (86) We implicitly assume that both input and output systems of 𝒩\mathcal{N} are bipartite systems with equal local dimensions. The local dimension for the input and output spaces can be different. Also, note that |Φm|\Phi^{\prime}_{m}\rangle’s are all stay in the output system of the information processing channel.
  • (87) M. Horodecki, P. Horodecki, and R. Horodecki, Mixed-state entanglement and distillation: Is there a “bound” entanglement in nature?, Phys. Rev. Lett. 80, 5239 (1998).
  • (88) S. Albeverio, S.-M. Fei, and W.-L. Yang, Optimal teleportation based on bell measurements, Phys. Rev. A 66, 012301 (2002).
  • (89) C.-Y. Hsieh, M. Lostaglio, and A. Acín, Entanglement preserving local thermalization, Phys. Rev. Research 2, 013379 (2020).
  • (90) C.-Y. Hsieh, M. Lostaglio, and A. Acín, Quantum channel marginal problem, arXiv:2102.10926.
  • (91) M. Horodecki, P. W. Shor, and M. B. Ruskai, Entanglement breaking channels, Rev. Math. Phys. 15, 629 (2003).
  • (92) S. Beigi and A. Gohari, Quantum achievability proof via collision relative entropy IEEE Trans. Inf. Theory, 60, 7980 (2014).
  • (93) D. Cavalcanti, P. Skrzypczyk, and I. Sˇ\check{\rm S}upić, All entangled states can demonstrate nonclassical teleportation, Phys. Rev. Lett. 119, 110501 (2017).
  • (94) M.-J. Zhao, Z.-G. Li, S.-M. Fei, and Z.-X. Wang, A note on fully entangled fraction, J. Phys. A: Math. Theor. 43, 275203 (2010).
  • (95) C.-Y. Hsieh and R.-K. Lee, Work extraction and fully entangled fraction, Phys. Rev. A 96, 012107 (2017); Phys. Rev. A 97, 059904(E) (2018).
  • (96) Y.-C. Liang, Y.-H. Yeh, P. E. M. F. Mendonça, R. Y. Teh, M. D. Reid, and P. D. Drummond, Quantum fidelity measures for mixed states, Rep. Prog. Phys. 82, 076001 (2019).
  • (97) M. Horodecki and P. Horodecki, Reduction criterion of separability and limits for a class of distillation protocols, Phys. Rev. A 59, 4206 (1999).
  • (98) C. H. Bennett, G. Brassard, S. Popescu, B. Schumacher, J. A. Smolin, and W. K. Wootters, Purification of noisy entanglement and faithful teleportation via noisy channels, Phys. Rev. Lett. 76, 722 (1996).
  • (99) L. Moravcˇ\check{\rm c}íková and M. Ziman, Entanglement-annihilating and entanglement-breaking channels, J Phys. A: Math. Theor. 43, 275306 (2010).
  • (100) Note that not every resource theory has convexity. Non-Markovianity, for instance, is an example, since a convex combination of two Markovian evolutions can be non-Markovian (see, e.g., Refs. Wolf2008 ; Regula2020 ; DeSantis2021 ; Chruscinski2014 ).
  • (101) I. Marvian, Symmetry, asymmetry and quantum information, Ph.D. thesis (2012).
  • (102) I. Marvian and R. Spekkens, The theory of manipulations of pure state asymmetry: I. Basic tools, equivalence classes and single copy transformations, New. J. Phys. 15, 033001 (2013).
  • (103) S. Bartlett, T. Rudolph, and R. Spekkens, Reference frames, superselection rules, and quantum information, Rev. Mod. Phys. 79, 55 (2007).
  • (104) We remark that in general the faithfulness of PDP_{D} depends on the topological properties of the set 𝒪RN\mathcal{O}_{R}^{N} and the generalized distance measure DD. See also Theorem 2 in Ref. Hsieh2020-1 . Note that this is satisfied by several commonly used options, such as DmaxD_{\rm max} used in this work. Here we aim to keep the generality of the statement for every DD.
  • (105) M. M. Wolf, J. Eisert, T. S. Cubitt, and J. I. Cirac, Assessing non-Markovian quantum dynamics, Phys. Rev. Lett. 101, 150402 (2008).
  • (106) D. Chruściński and F. A. Wudarski, Non-Markovianity degree for random unitary evolution, Phys. Rev. A 91, 012104 (2014).
  • (107) B. Regula, K. Bu, R. Takagi, and Z.-W. Liu, Benchmarking one-shot distillation in general quantum resource theories, Phys. Rev. A 101, 062315 (2020).
  • (108) D. De Santis and V. Giovannetti, Measuring non-Markovianity via incoherent mixing with Markovian dynamics, Phys. Rev. A 103, 012218 (2021).