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

General state transitions with exact resource morphisms:
a unified resource-theoretic approach

Wenbin Zhou [email protected] Graduate School of Informatics, Nagoya University, Chikusa-ku, 464-8601 Nagoya, Japan    Francesco Buscemi [email protected] Graduate School of Informatics, Nagoya University, Chikusa-ku, 464-8601 Nagoya, Japan
Abstract

Given a non-empty closed convex subset 𝖥\mathsf{F} of density matrices, we formulate conditions that guarantee the existence of an 𝖥\mathsf{F}-morphism (namely, a completely positive trace-preserving linear map that maps 𝖥\mathsf{F} into itself) between two arbitrarily chosen density matrices. While we allow errors in the transition, the corresponding map is required to be an exact 𝖥\mathsf{F}-morphism. Our findings, though purely geometrical, are formulated in a resource-theoretic language and provide a common framework that comprises various resource theories, including the resource theories of bipartite and multipartite entanglement, coherence, athermality, and asymmetric distinguishability.

We show how, when specialized to some situations of physical interest, our general results are able to unify and extend previous analyses. We also study conditions for the existence of maximally resourceful states, defined here as density matrices from which any other one can be obtained by means of a suitable 𝖥\mathsf{F}-morphism. Moreover, we quantitatively characterize the paradigmatic tasks of optimal resource dilution and distillation, as special transitions in which one of the two endpoints is maximally resourceful.

I Introduction

Many problems in information theory and statistics (and quantum generalizations thereof) can be reformulated as to whether a suitable transition between two objects is possible under suitable constraints. For example, noisy channel coding in information theory can be understood as the problem of transforming a number of uses of a noisy channel into less uses of a less noisy (ideally noiseless) channel, under the constraint that no side communication is available between sender and receiver. Another example is provided in mathematical statistics by the task of extracting statistics from a given sample. The exact same intuition becomes paradigmatic in statistical mechanics, most notably in thermodynamics, where a typical problem, for example, is to whether a thermodynamic process satisfying certain constraints (e.g., an isothermal process, an adiabatic process, etc.) exists between two given states.

Whenever the set of allowed transformations is constrained for some reasons, it is natural to consider the possibility of performing a forbidden transformation as a resource DHW (08); HHHH (09); CFS (16); CG (19). The state space of the system (that is, the set of objects that are being transformed) thus ends up being divided into free states (that is, states that can be reached from any other state by means of allowed transformations) and non-free states. The latter are implicitly defined to be the resourceful states in the theory. Notice that the word “state” here is not to be meant in its strict sense of “state of a physical system”, but denotes more generally the objects that are being transformed. These could comprise states proper, like in conventional thermodynamics, but are not limited to these. One could consider as objects, for example, statistical models (as in mathematical statistics), or noisy channels (as in information theory), or entire generalized operational theories for that matter.

In fact, it turns out that it is often easier to study resource theories, in which the first thing being defined is not the set of allowed transformations, but rather the set of free states, while allowed transformations are implicitly defined only later, as all those transformations that map the free set into itself. Even though sometimes less clear operationally, such an approach (that we may call “geometric”, in order to contrast it with the previous “operational” one that puts constraints on allowed operations) has, from a purely mathematical viewpoint, various advantages: for example, resource theories typically become asymptotically reversible only if the set of allowed transformations is taken to be the maximal one (plus epsilon) compatible with the theory BP (08); BG (15).

In this paper we study general resource theories by following the geometric approach, that is, by building the theory on top of a given set of free states 𝖥\mathsf{F}, that we in particular assume to be non-empty, closed, and convex. Accordingly, we allow any transformation that maps 𝖥\mathsf{F} into itself. For the sake of generality, in this paper we do not even assume any particular rule of composition for free states, so that we essentially work in the single-shot regime, but we still allow the input and the output systems, and their respective free sets, to change as a consequence of the transformation. In this way, even though we cannot say anything about asymptotic rates, we can still discuss in full generality single-shot rates for resource distillation and dilution, for example.

The main aspect that differentiates our work from previous ones is that here we do not focus only on the tasks of resource dilution and distillation. Instead, we consider the more general problem of formulating sufficient (and, in some cases, necessary) conditions for the existence of an 𝖥\mathsf{F}-preserving transformation between an arbitrarily given input–target pair of states. This means that the analysis presented here does not rely on the existence of any privileged maximally resourceful state (like the maximally entangled state in bipartite entanglement theory) and thus applies to quite general resource theories.

Concerning the kind of transformations that we consider in this work, two remarks are in order. First, while we allow for noisy transitions, that is, situations in which the input state is mapped not exactly but “close” to the wanted target, we only allow transitions implemented by operations that are exactly 𝖥\mathsf{F}-preserving. This stands in contrast to other single-shot analyses of general resource theories in which “almost-free” operations are allowed in the finite block-length regime (see, e.g., Ref. BG (15)). Second, we search for conditions that are expressed in terms of resource monotones (a notion to be rigorously introduced in what follows), computed separately for the initial state and for the target state. On the one hand, in this way we are able to “decouple” the initial state from the target state, and to speak of their respective resource’s worth, independently of each other. On the other hand, this means that functions that comprise both states at once, like those that in some situations can be obtained by means of semidefinite linear programs BG (17); GJB+ (18), will not be considered in this work. Another aspect of the problem that is not considered in this work is the computational complexity of computing resource monotones.

The paper is structured as follows. After introducing the notation and reviewing some basic notions in Section II, we present our main results and prove them in detail in Section III. Section IV presents some applications of our general analysis to specific cases of physical interest: applications to the resource theories of athermality and asymmetric distinguishability (Section IV.1); bipartite entanglement (Section IV.2); conditions for the existence of maximally resourceful states, together with necessary and sufficient conditions for dilution and distillation which can give, whenever a dimension scaling is provided, optimal dilution and distillation rates (Section IV.3). Section V concludes the paper.

II Mathematical Preliminaries

We denote by 𝖣(m)\mathsf{D}(\mathbb{C}^{m}) the set of all mm-by-mm complex density matrices ρ\rho, i.e., ρ0\rho\geqslant 0 and Tr{ρ}=1\operatorname{Tr}\left\{\rho\right\}=1, which are used here to represent quantum states of mm-dimensional quantum systems. Within 𝖣(m)\mathsf{D}(\mathbb{C}^{m}), we identify a non-empty closed convex subset 𝖥\mathsf{F} as the set of “free states”. The closure and the convexity of 𝖥\mathsf{F} is crucial in various steps of our proofs, for example, when invoking the closure under convex mixtures, or when applying a variant of the minimax theorem that requires convex domain. Here and throughout this work, resource morphisms (or more precisely 𝖥\mathsf{F}-morphisms111Here we prefer the term “resource morphisms” to the more common “free operations” because it reminds the fact that the foundational concept, in the geometric approach in which we are working, is the free set 𝖥\mathsf{F}, not the set of allowed transformations, which are just defined as all those that map 𝖥\mathsf{F} into itself.) are defined as completely positive, trace-preserving (CPTP) linear maps :𝖣(m)𝖣(m)\mathcal{E}:\mathsf{D}(\mathbb{C}^{m})\to\mathsf{D}(\mathbb{C}^{m}) such that (𝖥)𝖥\mathcal{E}(\mathsf{F})\subseteq\mathsf{F}. More generally, one may consider CPTP maps that change the dimension of the system, for example, :𝖣(m)𝖣(n)\mathcal{E}:\mathsf{D}(\mathbb{C}^{m})\to\mathsf{D}(\mathbb{C}^{n}). Also in this case, whenever the output free set 𝖥\mathsf{F}^{\prime} is also specified, it is possible to define a notion of resource morphisms by the condition that (𝖥)𝖥\mathcal{E}(\mathsf{F})\subseteq\mathsf{F}^{\prime}. However in what follows, for the sake of readability, we will restrict ourselves to the case of equal input and output dimensions and 𝖥=𝖥\mathsf{F}=\mathsf{F}^{\prime}, keeping in mind however that all the results we derive can be straightforwardly extended to the general case. We will go back to the more general setting, with different input and output systems, in Section IV when discussing various applications like the tasks of resource dilution and distillation.

The general structure that we study in this work is the following:

Definition 1 (Resourcefulness Preorder).

Given two density matrices ρ,σ𝖣(m)\rho,\sigma\in\mathsf{D}(\mathbb{C}^{m}), we write ρϵσ\rho\succ_{\epsilon}\sigma whenever there exists a resource morphism \mathcal{E} such that 12σ(ρ)1ϵ\frac{1}{2}\left|\!\left|{\sigma-\mathcal{E}(\rho)}\right|\!\right|_{1}\leqslant\epsilon, where X1:=Tr{XX}\left|\!\left|{X}\right|\!\right|_{1}:=\operatorname{Tr}\left\{\sqrt{X^{\dagger}X}\right\} denotes the trace-norm. In particular, we write ρσ\rho\succ\sigma whenever ρϵ=0σ\rho\succ_{\epsilon=0}\sigma.

In the above definition, the preorder ϵ\succ_{\epsilon} has been introduced with respect to the distance induced by the trace-norm, although it is possible to use any other well-behaved distance measure between density matrices (like the fidelity, for example) without substantially changing the results NC (00); Wil (13). In any case, an important thing to notice in Definition 1 is that, even though errors are allowed in the state transformation, we always require the constraint (𝖥)𝖥\mathcal{E}(\mathsf{F})\subseteq\mathsf{F} to be strictly satisfied.

The resourcefulness preorder naturally lead us to define a maximally resourceful element as follows:

Definition 2 (Maximally resourceful element).

An element α𝖣(m)\alpha\in\mathsf{D}(\mathbb{C}^{m}) is said to be maximally resourceful if ασ\alpha\succ\sigma for any σ𝖣(m)\sigma\in\mathsf{D}(\mathbb{C}^{m}).

Given a general resource theory, an important question is to whether the theory possesses maximally resourceful elements or not. In Section IV.3 we will consider sufficient conditions for their existence. However, we recall that our main results do not rely in any way on the existence of maximally resourceful elements.

II.1 Information-theoretic divergences

In what follows, for any operator ρ𝖣(m)\rho\in\mathsf{D}(\mathbb{C}^{m}) we denote by Πρ\Pi_{\rho} the orthogonal projector onto its support (i.e., the orthogonal complement of its kernel). Moreover, for any ϵ[0,1]\epsilon\in[0,1], we denote by 𝖡ϵ(ρ)\mathsf{B}^{\epsilon}(\rho) the set of operators {ρ𝖣(m):ρρ12ϵ}\{\rho^{\prime}\in\mathsf{D}(\mathbb{C}^{m}):\left|\!\left|{\rho-\rho^{\prime}}\right|\!\right|_{1}\leqslant 2\epsilon\} and by 𝖯ϵ(ρ)\mathsf{P}^{\epsilon}(\rho) the set of operators {P:0P𝟙 and Tr{ρP}1ϵ}\{P:0\leqslant P\leqslant\mathds{1}\text{ and }\operatorname{Tr}\left\{\rho P\right\}\geqslant 1-\epsilon\}. All logarithms are taken in base 2.

Definition 3 (Relative entropies).

Given two density matrices ρ,σ𝖣(m)\rho,\sigma\in\mathsf{D}(\mathbb{C}^{m}), we define

  1. (i)

    the Umegaki relative entropy Ume (62):

    D(ρσ):={Tr{ρ(logρlogσ)}, if ΠσΠρ,+, otherwise;\displaystyle D(\rho\|\sigma):=\begin{cases}\operatorname{Tr}\left\{\rho\ (\log\rho-\log\sigma)\right\}\;,&\text{ if }\Pi_{\sigma}\geqslant\Pi_{\rho}\;,\\ +\infty\;,&\text{ otherwise}\;;\end{cases} (1)
  2. (ii)

    the hypothesis testing relative entropy BD (10): for any ϵ[0,1]\epsilon\in[0,1]

    Dhϵ(ρσ):=logminP𝖯ϵ(ρ)Tr{σP},\displaystyle D_{h}^{\epsilon}(\rho\|\sigma):=-\log\quad\min_{P\in\mathsf{P}^{\epsilon}(\rho)}\quad\operatorname{Tr}\left\{\sigma\ P\right\}\;, (2)

    with the convention log0:=+-\log 0:=+\infty; for ϵ=0\epsilon=0, one recovers the min-divergence, defined as Dat09b

    Dmin(ρσ):=logTr{σΠρ};D_{\min}(\rho\|\sigma):=-\log\operatorname{Tr}\left\{\sigma\ \Pi_{\rho}\right\}\;; (3)
  3. (iii)

    the max-divergence Dat09b :

    Dmax(ρσ):={logmin{λ:λσρ0}, if ΠσΠρ,+, otherwise;D_{\max}(\rho\|\sigma):=\begin{cases}\log\min\{\lambda\in\mathbb{R}:\lambda\sigma-\rho\geqslant 0\}\;,&\text{ if }\Pi_{\sigma}\geqslant\Pi_{\rho}\;,\\ +\infty\;,&\text{ otherwise}\;;\end{cases} (4)

    in this case we also define a “smoothed” version as follows: for any ϵ[0,1]\epsilon\in[0,1],

    Dmaxϵ(ρσ):=infρ𝖡ϵ(ρ)Dmax(ρσ).D_{\max}^{\epsilon}(\rho\|\sigma):=\inf_{\rho^{\prime}\in\mathsf{B}^{\epsilon}(\rho)}D_{\max}(\rho^{\prime}\|\sigma)\;. (5)

A crucial property satisfied by all these divergences is the monotonicity under CPTP linear maps, that is, for example, D((ρ)(σ)D(ρσ)D(\mathcal{E}(\rho)\|\mathcal{E}(\sigma)\leqslant D(\rho\|\sigma) and analogously for the others. In a resource theory characterized by the set of free states 𝖥\mathsf{F}, we also introduce the following:

Definition 4 (Max-divergence of resourcefulness).

Given two density matrices ρ,σ𝖣(m)\rho,\sigma\in\mathsf{D}(\mathbb{C}^{m}) and a non-empty closed convex subset 𝖥𝖣(m)\mathsf{F}\subseteq\mathsf{D}(\mathbb{C}^{m}), the max-divergence relative to 𝖥\mathsf{F} is defined as

Dmax,𝖥(ρσ):=loginf{λ:λσρλ1𝖥},D_{\max,\mathsf{F}}(\rho\|\sigma):=\log\inf\left\{\lambda\in\mathbb{R}:\frac{\lambda\sigma-\rho}{\lambda-1}\in\mathsf{F}\right\}\;, (6)

with the convention that inf=+\inf\varnothing=+\infty. Its “smoothed” version is defined in analogy with (5), that is

Dmax,𝖥ϵ(ρσ):=infρ𝖡ϵ(ρ)Dmax,𝖥(ρσ).D_{\max,\mathsf{F}}^{\epsilon}(\rho\|\sigma):=\inf_{\rho^{\prime}\in\mathsf{B}^{\epsilon}(\rho)}D_{\max,\mathsf{F}}(\rho^{\prime}\|\sigma)\;. (7)

We notice that if 𝖥=𝖣(m)\mathsf{F}=\mathsf{D}(\mathbb{C}^{m}), then Dmax,𝖥(ρσ)=Dmax(ρσ)D_{\max,\mathsf{F}}(\rho\|\sigma)=D_{\max}(\rho\|\sigma), but in general Dmax,𝖥(ρσ)Dmax(ρσ)D_{\max,\mathsf{F}}(\rho\|\sigma)\geqslant D_{\max}(\rho\|\sigma). Moreover, while Dmax,𝖥D_{\max,\mathsf{F}} may fail to be monotonic under general CPTP maps, it is monotonic under the action of resource morphisms, that is, CPTP maps that map 𝖥\mathsf{F} into itself. This can be easily seen by noticing that, if for some λ\lambda, (λ1)1(λσρ)(\lambda-1)^{-1}(\lambda\sigma-\rho) is in 𝖥\mathsf{F}, then, for any resource morphism \mathcal{E}, also (λ1)1(λσρ)=(λ1)1(λ(σ)(ρ))(\lambda-1)^{-1}\mathcal{E}(\lambda\sigma-\rho)=(\lambda-1)^{-1}(\lambda\mathcal{E}(\sigma)-\mathcal{E}(\rho)) is automatically in 𝖥\mathsf{F}, so that Dmax,𝖥((ρ)(σ))D_{\max,\mathsf{F}}(\mathcal{E}(\rho)\|\mathcal{E}(\sigma)) cannot be larger than Dmax,𝖥(ρσ)D_{\max,\mathsf{F}}(\rho\|\sigma).

II.2 Resource monotones

We say that a function f:𝖣(m)[0,+]f:\mathsf{D}(\mathbb{C}^{m})\to[0,+\infty] constitutes a resource monotone if it achieves its global minimum on all elements of 𝖥\mathsf{F}, and it does not increase under the action of resource morphisms, i.e., f(ρ)f((ρ))f(\rho)\geqslant f(\mathcal{E}(\rho)) for any resource morphism \mathcal{E}. More properties can be demanded (and are indeed desirable) in order to fruitfully work with concrete examples of resource monotones. The information-theoretic divergences introduced above can be used to introduce resource monotones that inherit many useful properties from the parent divergence. In our construction, the following quantities play a central role LBT (19).

Definition 5 (Entropic Resource Monotones).

Given a non-empty closed convex set 𝖥𝖣(m)\mathsf{F}\subseteq\mathsf{D}(\mathbb{C}^{m}), for any density matrix ρ𝖣(m)\rho\in\mathsf{D}(\mathbb{C}^{m}) and any ϵ[0,1]\epsilon\in[0,1], we define the following quantities:

  1. (i)

    𝔇(ρ):=infω𝖥D(ρω)\mathfrak{D}(\rho):=\inf_{\omega\in\mathsf{F}}D(\rho\|\omega);

  2. (ii)

    𝔇hϵ(ρ):=logmaxω𝖥minP𝖯ϵ(ρ)Tr{Pω}\mathfrak{D}_{h}^{\epsilon}(\rho):=-\log\max_{\omega\in\mathsf{F}}\min_{P\in\mathsf{P}^{\epsilon}(\rho)}\operatorname{Tr}\left\{P\ \omega\right\}, with the convention log0:=+-\log 0:=+\infty;

  3. (iii)

    𝔇maxϵ(ρ):=infω𝖥Dmaxϵ(ρω)\mathfrak{D}_{\max}^{\epsilon}(\rho):=\inf_{\omega\in\mathsf{F}}D_{\max}^{\epsilon}(\rho\|\omega);

  4. (iv)

    𝔇max,𝖥ϵ(ρ):=infω𝖥Dmax,𝖥ϵ(ρω)\mathfrak{D}_{\max,\mathsf{F}}^{\epsilon}(\rho):=\inf_{\omega\in\mathsf{F}}D^{\epsilon}_{\max,\mathsf{F}}(\rho\|\omega).

In the case ϵ=0\epsilon=0, we simply remove the superscript; the only exception is 𝔇hϵ=0(ρ)\mathfrak{D}_{h}^{\epsilon=0}(\rho), for which we will use the special notation 𝔇min(ρ)\mathfrak{D}_{\min}(\rho).

The above quantities are all well-behaved resource monotones. This fact is a direct consequence of the monotonicity of the parent divergences under the action of resource morphisms.

Definition 6 (Free fraction and generalized free fraction).

Given a non-empty closed convex free set 𝖥𝖣(m)\mathsf{F}\subseteq\mathsf{D}(\mathbb{C}^{m}), the free fraction of a density matrix ρ𝖣(m)\rho\in\mathsf{D}(\mathbb{C}^{m}) is defined by the formula

𝔉(ρ):=max{p[0,1]:ω𝖥 s.t. pρ+(1p)ω𝖥}.\mathfrak{F}(\rho):=\max\{p\in[0,1]:\exists\ \omega\in\mathsf{F}\text{ s.t. }p\rho+(1-p)\omega\in\mathsf{F}\}\;. (8)

When mixing with general ω𝖣(m)\omega\in\mathsf{D}(\mathbb{C}^{m}) instead of ω𝖥\omega\in\mathsf{F} is allowed, one obtains the generalized free fraction, defined as

𝔉g(ρ):=max{p[0,1]:ω𝖣(m) s.t. pρ+(1p)ω𝖥}.\mathfrak{F}_{g}(\rho):=\max\{p\in[0,1]:\exists\ \omega\in\mathsf{D}(\mathbb{C}^{m})\text{ s.t. }p\rho+(1-p)\omega\in\mathsf{F}\}\;. (9)

The free fraction and the generalized free fraction are related to the robustness (ρ)\mathfrak{R}(\rho) VT (99) and the generalized robustness g(ρ)\mathfrak{R}_{g}(\rho) Ste (03), respectively, through the relations 𝔉(ρ)1=1+(ρ)\mathfrak{F}(\rho)^{-1}=1+\mathfrak{R}(\rho) and 𝔉g(ρ)1=1+g(ρ)\mathfrak{F}_{g}(\rho)^{-1}=1+\mathfrak{R}_{g}(\rho), and they are both directly related with the entropic resource monotones in Definition 5 as follows:

log𝔉(ρ)=log(1+(ρ))=𝔇max,𝖥(ρ),\displaystyle-\log{\mathfrak{F}(\rho)}=\log(1+\mathfrak{R}(\rho))=\mathfrak{D}_{\max,\mathsf{F}}(\rho)\;, (10)
log𝔉g(ρ)=log(1+g(ρ))=𝔇max(ρ),\displaystyle-\log{\mathfrak{F}_{g}(\rho)}=\log(1+\mathfrak{R}_{g}(\rho))=\mathfrak{D}_{\max}(\rho)\;, (11)

with the convention log0:=+-\log 0:=+\infty. In particular, we have that 𝔇max(ρ)\mathfrak{D}_{\max}(\rho) coincides with the generalized logarithmic robustness of Dat09b ; Dat09a , while 𝔇max,𝖥(ρ)\mathfrak{D}_{\max,\mathsf{F}}(\rho) coincides with the logarithmic robustness of BD (11). For this reason, in what follows, when speaking of 𝔇max(ρ)\mathfrak{D}_{\max}(\rho) (respectively, 𝔇max,𝖥(ρ)\mathfrak{D}_{\max,\mathsf{F}}(\rho)) we will follow the mainstream convention and call it “generalized log-robustness” (respectively, “log-robustness”) even though, depending on the context, it would be more appropriate to use the term we introduced above, that is, “generalized log-free fraction” (respectively, “log-free fraction”).

Remark.

All resource monotones introduced above would still be well-defined monotones even if the class of resource morphisms were enlarged to comprise also positive, but not completely-positive, linear maps. This seems no coincidence, since at the single-shot level, where no rule for composing system is given yet, there really is no compelling mathematical reason to limit the discussion to CPTP linear maps only. This is a common feature of various problems in quantum statistics, in particular quantum decision theory, where the theory becomes simpler if one works with quantum statistical morphism (which may violate complete positivity) and introduce CPTP maps as special cases, rather than starting from the beginning with fully blown CPTP maps Bus (12). Here we do not investigate further into this point, and simply justify the assumption of CPTP-ness on practical grounds.

II.3 Optimal convex decompositions

Our main results rely on the following construction, whose intuitive picture is given in Fig. 1 below.

Refer to caption
Figure 1: Geometric intuition for the generalized free fraction introduced in Definition 6. Here σ0\sigma_{0} denotes the optimized density matrix, which is able to achieve, by means of convex mixing, the generalized free fraction of σ\sigma.

Given σ𝖣(m)\sigma\in\mathsf{D}(\mathbb{C}^{m}), assuming σ𝖥\sigma\notin\mathsf{F}, let us fix a convex decomposition achieving its generalized free fraction and write it as

σ+=𝔉g(σ)σ+[1𝔉g(σ)]σ0.\displaystyle\sigma_{+}=\mathfrak{F}_{g}(\sigma)\sigma+[1-\mathfrak{F}_{g}(\sigma)]\sigma_{0}\;. (12)

In the above equation, due to the optimality of 𝔉g\mathfrak{F}_{g}, σ+𝖥\sigma_{+}\in\mathsf{F} lies on the border of 𝖥\mathsf{F}, while σ0\sigma_{0} lies on the border of 𝖣(m)\mathsf{D}(\mathbb{C}^{m}), as depicted in Fig. 1. The above decomposition includes the situation in which 𝔉g(σ)=0\mathfrak{F}_{g}(\sigma)=0, that is, σ+=σ0\sigma_{+}=\sigma_{0}. For any decomposition as in (12), another free state σ\sigma_{-} can be uniquely defined using the max-divergence of resourcefulness (Definition 4) as follows:

σ:\displaystyle\sigma_{-}: =2Dmax,𝖥(σσ+)σ+σ2Dmax,𝖥(σσ+)1\displaystyle=\frac{2^{D_{\max,\mathsf{F}}(\sigma\|\sigma_{+})}\sigma_{+}-\sigma}{2^{D_{\max,\mathsf{F}}(\sigma\|\sigma_{+})}-1} (13)
=[𝔉g(σ) 2Dmax,𝖥(σσ+)12Dmax,𝖥(σσ+)1]σ+[1𝔉g(σ) 2Dmax,𝖥(σσ+)12Dmax,𝖥(σσ+)1]σ0,\displaystyle=\left[\frac{\mathfrak{F}_{g}(\sigma)\,2^{D_{\max,\mathsf{F}}(\sigma\|\sigma_{+})}-1}{2^{D_{\max,\mathsf{F}}(\sigma\|\sigma_{+})}-1}\right]\sigma+\left[1-\frac{\mathfrak{F}_{g}(\sigma)\,2^{D_{\max,\mathsf{F}}(\sigma\|\sigma_{+})}-1}{2^{D_{\max,\mathsf{F}}(\sigma\|\sigma_{+})}-1}\right]\sigma_{0}\;, (14)

whenever Dmax,𝖥(σσ+)<+D_{\max,\mathsf{F}}(\sigma\|\sigma_{+})<+\infty, or σ:=σ+\sigma_{-}:=\sigma_{+} otherwise. In order to derive (14) we just plugged (12) into (13) and rearranged terms. Notice that since σ𝖥\sigma\notin\mathsf{F}, we have σσ+\sigma\neq\sigma_{+} and Dmax,𝖥(σσ+)>0D_{\max,\mathsf{F}}(\sigma\|\sigma_{+})>0. It is easy to check that, by construction, σ\sigma_{-}, as σ+\sigma_{+}, lies on the intersection between the border of 𝖥\mathsf{F} and the segment joining σ\sigma with σ0\sigma_{0}. Our main results will originate from a careful evaluation of the relative distances between these four points in state space.

III Main Results

In this section, we state and prove the main results of this paper. Firstly we derive, for any finite-dimensional resource theory in which the set of free states is non-empty closed and convex, sufficient conditions for the existence of a resource morphism between any two states, given in terms of resource monotones. Such conditions are formulated so to allow, in general, non-zero errors in the state transition, while the operation implementing the transition is an exact resource morphism.

Theorem 1.

Let us arbitrarily fix two states, ρ,σ𝖣(m)\rho,\sigma\in\mathsf{D}(\mathbb{C}^{m}), and two values ϵ1,ϵ2[0,1]\epsilon_{1},\epsilon_{2}\in[0,1]. Let us moreover choose σ~𝖡ϵ2(σ)\tilde{\sigma}\in\mathsf{B}^{\epsilon_{2}}(\sigma) and σ~+𝖥\tilde{\sigma}_{+}\in\mathsf{F} so that Dmax(σ~σ~+)=𝔇maxϵ2(σ)D_{\max}(\tilde{\sigma}\|\tilde{\sigma}_{+})=\mathfrak{D}^{\epsilon_{2}}_{\max}(\sigma).

  1. (i)

    If 𝔇hϵ1(ρ)=+\mathfrak{D}^{\epsilon_{1}}_{h}(\rho)=+\infty, then ρϵ1σ\rho\succ_{\epsilon_{1}}\sigma.

  2. (ii)

    If 𝔇maxϵ2(σ)=0\mathfrak{D}^{\epsilon_{2}}_{\max}(\sigma)=0, then ρϵ2σ\rho\succ_{\epsilon_{2}}\sigma.

  3. (iii)

    If 𝔇hϵ1(ρ)<+\mathfrak{D}^{\epsilon_{1}}_{h}(\rho)<+\infty and 𝔇maxϵ2(σ)>0\mathfrak{D}^{\epsilon_{2}}_{\max}(\sigma)>0, then

    1. (a)

      either Dmax,𝖥(σ~σ~+)<+D_{\max,\mathsf{F}}(\tilde{\sigma}\|\tilde{\sigma}_{+})<+\infty; in such a case, ρϵ1+ϵ2σ\rho\succ_{\epsilon_{1}+\epsilon_{2}}\sigma if the following two conditions simultaneously hold:

      𝔇hϵ1(ρ)𝔇maxϵ2(σ)\displaystyle\mathfrak{D}^{\epsilon_{1}}_{h}(\rho)\geqslant\mathfrak{D}^{\epsilon_{2}}_{\max}(\sigma) (15)

      and

      2maxω𝖥Dhϵ1(ρω)2Dmax,𝖥(σ~σ~+)𝔇hϵ1(ρ)12Dmax,𝖥(σ~σ~+)1;\displaystyle 2^{-\max_{\omega\in\mathsf{F}}D^{\epsilon_{1}}_{h}(\rho\|\omega)}\geqslant\frac{2^{D_{\max,\mathsf{F}}(\tilde{\sigma}\|\tilde{\sigma}_{+})-\mathfrak{D}^{\epsilon_{1}}_{h}(\rho)}-1}{2^{D_{\max,\mathsf{F}}(\tilde{\sigma}\|\tilde{\sigma}_{+})}-1}\;; (16)
    2. (b)

      or Dmax,𝖥(σ~σ~+)=+D_{\max,\mathsf{F}}(\tilde{\sigma}\|\tilde{\sigma}_{+})=+\infty; in such a case, ρϵ1+ϵ2σ\rho\succ_{\epsilon_{1}+\epsilon_{2}}\sigma if condition (15) above holds together with

      maxω𝖥Dhϵ1(ρω)=minω𝖥Dhϵ1(ρω).\displaystyle\max_{\omega\in\mathsf{F}}D^{\epsilon_{1}}_{h}(\rho\|\omega)=\min_{\omega\in\mathsf{F}}D^{\epsilon_{1}}_{h}(\rho\|\omega)\;. (17)
Remark.

As discussed in Section II.3, the assumption 𝔇maxϵ2(σ)>0\mathfrak{D}^{\epsilon_{2}}_{\max}(\sigma)>0 in case (iii.a) guarantees that also Dmax,𝖥(σ~σ~+)>0D_{\max,\mathsf{F}}(\tilde{\sigma}\|\tilde{\sigma}_{+})>0, so that the denominator appearing in the right-hand side of (16) is strictly greater than zero. Also, since 𝔇hϵ1(ρ)log(1ϵ1)\mathfrak{D}^{\epsilon_{1}}_{h}(\rho)\geqslant-\log(1-\epsilon_{1}) independently of ρ\rho, the parameter ϵ1\epsilon_{1} can be modulated so to compensate, to some extent, eventual lack of resource in the initial state.

Condition (17) is stronger than condition (16), in the sense that if the former is satisfied, the latter is also satisfied. This is because, by multiplying both sides by 2Dmax,𝖥(σ~σ~+)1>02^{D_{\max,\mathsf{F}}(\tilde{\sigma}\|\tilde{\sigma}_{+})}-1>0 (see preceding remark), condition (16) becomes

2Dmax,𝖥(σ~σ~+)maxω𝖥Dhϵ1(ρω)2maxω𝖥Dhϵ1(ρω)2Dmax,𝖥(σ~σ~+)minω𝖥Dhϵ1(ρω)1,2^{D_{\max,\mathsf{F}}(\tilde{\sigma}\|\tilde{\sigma}_{+})-\max_{\omega\in\mathsf{F}}{D}_{h}^{\epsilon_{1}}(\rho\|\omega)}-2^{-\max_{\omega\in\mathsf{F}}{D}_{h}^{\epsilon_{1}}(\rho\|\omega)}\geqslant 2^{D_{\max,\mathsf{F}}(\tilde{\sigma}\|\tilde{\sigma}_{+})-\min_{\omega\in\mathsf{F}}{D}_{h}^{\epsilon_{1}}(\rho\|\omega)}-1\;,

and this, if maxω𝖥Dhϵ1(ρω)=minω𝖥Dhϵ1(ρω)\max_{\omega\in\mathsf{F}}{D}_{h}^{\epsilon_{1}}(\rho\|\omega)=\min_{\omega\in\mathsf{F}}{D}_{h}^{\epsilon_{1}}(\rho\|\omega), becomes equivalent to

2maxω𝖥Dhϵ1(ρω)1,2^{-\max_{\omega\in\mathsf{F}}{D}_{h}^{\epsilon_{1}}(\rho\|\omega)}\leqslant 1\;,

which is always trivially satisfied, due to the non-negativity of the hypothesis testing relative entropy. In other words, we have shown the following:

Corollary 1.

Given a state ρ𝖣(m)\rho\in\mathsf{D}(\mathbb{C}^{m}), suppose that maxω𝖥Dhϵ1(ρω)=minω𝖥Dhϵ1(ρω)\max_{\omega\in\mathsf{F}}D^{\epsilon_{1}}_{h}(\rho\|\omega)=\min_{\omega\in\mathsf{F}}D^{\epsilon_{1}}_{h}(\rho\|\omega). Then, for any σ\sigma,

𝔇hϵ1(ρ)𝔇maxϵ2(σ)ρ(ϵ1+ϵ2)σ.\mathfrak{D}^{\epsilon_{1}}_{h}(\rho)\geqslant\mathfrak{D}^{\epsilon_{2}}_{\max}(\sigma)\qquad\implies\qquad\rho\succ_{(\epsilon_{1}+\epsilon_{2})}\sigma\;.

Corollary 1, for rank-one ρ\rho and ϵ1=0\epsilon_{1}=0, recovers Theorem 2 in Ref. LBT (19).

Proof of Theorem 1.

Case (i) is easily proved as follows. The condition 𝔇hϵ1(ρ)=+\mathfrak{D}^{\epsilon_{1}}_{h}(\rho)=+\infty guarantees the existence of an operator P𝖯ϵ1(ρ)P\in\mathsf{P}^{\epsilon_{1}}(\rho) such that Tr{Pω}=0\operatorname{Tr}\left\{P\ \omega\right\}=0 for all ω𝖥\omega\in\mathsf{F}. Hence, by constructing a CPTP map as follows:

():=Tr{P}σ+Tr{(𝟙P)}φ,\mathcal{E}(\cdot):=\operatorname{Tr}\left\{P\ \cdot\right\}\sigma+\operatorname{Tr}\left\{(\mathds{1}-P)\ \cdot\right\}\varphi\;,

where φ\varphi is an arbitrarily fixed element of 𝖥\mathsf{F}, we see that \mathcal{E} maps all free states to φ\varphi, so that (𝖥)𝖥\mathcal{E}(\mathsf{F})\subseteq\mathsf{F}, while (ρ)σ12(1Tr{Pρ})2ϵ1\left|\!\left|{\mathcal{E}(\rho)-\sigma}\right|\!\right|_{1}\leqslant 2(1-\operatorname{Tr}\left\{P\rho\right\})\leqslant 2\epsilon_{1}.

Case (ii) follows trivially from the fact that condition 𝔇maxϵ2(σ)=0\mathfrak{D}^{\epsilon_{2}}_{\max}(\sigma)=0 guarantees the existence of at least one free state which is ϵ2\epsilon_{2}-close to σ\sigma. Hence, the sought resource morphism is trivially given by the CPTP map that prepares any one such states.

Now we move on to case (iii). We begin by looking at condition (15), which is the same in both (iii.a) and (iii.b), and rewrite it as follows

logTr{Pω}Dmax(σ~σ~+),\displaystyle-\log\operatorname{Tr}\left\{P^{*}\ \omega^{*}\right\}\geqslant D_{\max}(\tilde{\sigma}\|\tilde{\sigma}_{+})\;, (18)

where

  • the operators P𝖯ϵ1(ρ)P^{*}\in\mathsf{P}^{\epsilon_{1}}(\rho) and ω𝖥\omega^{*}\in\mathsf{F} are chosen to satisfy:

    Tr{Pω}\displaystyle\operatorname{Tr}\left\{P^{*}\ \omega^{*}\right\} =2𝔇hϵ1(ρ)\displaystyle=2^{-\mathfrak{D}_{h}^{\epsilon_{1}}(\rho)} (19)
    :=maxω𝖥minP𝖯ϵ1(ρ)Tr{Pω}\displaystyle:=\max_{\omega\in\mathsf{F}}\min_{P\in\mathsf{P}^{\epsilon_{1}}(\rho)}\operatorname{Tr}\left\{P\ \omega\right\} (20)
    =minP𝖯ϵ1(ρ)maxω𝖥Tr{Pω};\displaystyle=\min_{P\in\mathsf{P}^{\epsilon_{1}}(\rho)}\max_{\omega\in\mathsf{F}}\operatorname{Tr}\left\{P\ \omega\right\}\;; (21)

    the equality in the third line follows from the minimax theorem, for example, in Kakutani’s formulation Kak (41); FKK (04), whose hypotheses are satisfied since both optimizations range over convex sets and the functional to be optimized is linear, and hence both convex and concave, in its arguments;

  • the operators σ~𝖣(m)\tilde{\sigma}\in\mathsf{D}(\mathbb{C}^{m}) and σ~+𝖥\tilde{\sigma}_{+}\in\mathsf{F} are chosen so to satisfy:

    Dmax(σ~σ~+)\displaystyle D_{\max}(\tilde{\sigma}\|\tilde{\sigma}_{+}) =𝔇maxϵ2(σ)\displaystyle=\mathfrak{D}_{\max}^{\epsilon_{2}}(\sigma) (22)
    :=minω𝖥minσ𝖡ϵ(σ)Dmax(σω)\displaystyle:=\min_{\omega\in\mathsf{F}}\min_{\sigma^{\prime}\in\mathsf{B}^{\epsilon}(\sigma)}D_{\max}(\sigma^{\prime}\|\omega) (23)
    =minσ𝖡ϵ(σ)minω𝖥Dmax(σω)\displaystyle=\min_{\sigma^{\prime}\in\mathsf{B}^{\epsilon}(\sigma)}\min_{\omega\in\mathsf{F}}D_{\max}(\sigma^{\prime}\|\omega) (24)
    =𝔇max(σ~),\displaystyle=\mathfrak{D}_{\max}(\tilde{\sigma})\;, (25)

    that is, σ~+\tilde{\sigma}_{+} achieves the generalized free fraction for σ~\tilde{\sigma} as in Eq. (12), namely:

    σ~+=𝔉g(σ~)σ~+(1𝔉g(σ~))σ~0.\displaystyle\tilde{\sigma}_{+}=\mathfrak{F}_{g}(\tilde{\sigma})\tilde{\sigma}+(1-\mathfrak{F}_{g}(\tilde{\sigma}))\tilde{\sigma}_{0}\;. (26)

In Ref. BST (19), condition (15) alone is shown to be sufficient for the existence of a test-and-prepare CPTP linear map \mathcal{E} such that (ρ)σ12(ϵ1+ϵ2)\left|\!\left|{\mathcal{E}(\rho)-\sigma}\right|\!\right|_{1}\leqslant 2(\epsilon_{1}+\epsilon_{2}) and (ω)=σ~+\mathcal{E}(\omega^{*})=\tilde{\sigma}_{+}. Such a map is explicitly given as follows:

()=Tr{P}σ~+Tr{(𝟙P)}Mσ~+σ~M1,\displaystyle\mathcal{E}(\cdot)=\operatorname{Tr}\left\{P^{*}\ \cdot\right\}\tilde{\sigma}+\operatorname{Tr}\left\{(\mathds{1}-P^{*})\ \cdot\right\}\frac{M\tilde{\sigma}_{+}-\tilde{\sigma}}{M-1}\;, (27)

where, for convenience of notation, we have put M:=1/Tr{Pω}=2𝔇hϵ1(ρ)M:=1/\operatorname{Tr}\left\{P^{*}\ {\omega^{*}}\right\}=2^{\mathfrak{D}^{\epsilon_{1}}_{h}(\rho)}. Without loss of generality, we can assume that 1<M<+1<M<+\infty for the following reasons. First of all, notice that the assumption 𝔇hϵ1(ρ)<+\mathfrak{D}^{\epsilon_{1}}_{h}(\rho)<+\infty implies M<+M<+\infty. Moreover, we can also assume that Tr{Pω}<1\operatorname{Tr}\left\{P^{*}\ {\omega^{*}}\right\}<1, that is M>1M>1, otherwise 𝔇hϵ1(ρ)=0\mathfrak{D}^{\epsilon_{1}}_{h}(\rho)=0 and, by (15), 𝔇maxϵ2(σ)=0\mathfrak{D}^{\epsilon_{2}}_{\max}(\sigma)=0, thus making the situation trivial.

As shown in BST (19), the above map is CPTP; in order to show that it is a resource morphism, we only need to show that (𝖥)𝖥\mathcal{E}(\mathsf{F})\subseteq\mathsf{F}. To this end, let us assume that the input to \mathcal{E} is an arbitrary φ𝖥\varphi\in\mathsf{F}. We need to show that (φ)𝖥\mathcal{E}(\varphi)\in\mathsf{F}. By arranging terms, we obtain,

(φ)=(11Tr{Pφ}1Tr{Pω})σ~+(1Tr{Pφ}1Tr{Pω})σ~+.\displaystyle\mathcal{E}(\varphi)=\left(1-\frac{1-\operatorname{Tr}\left\{P^{*}\ \varphi\right\}}{1-\operatorname{Tr}\left\{P^{*}\ {\omega^{*}}\right\}}\right)\tilde{\sigma}+\left(\frac{1-\operatorname{Tr}\left\{P^{*}\ \varphi\right\}}{1-\operatorname{Tr}\left\{P^{*}\ {\omega^{*}}\right\}}\right)\tilde{\sigma}_{+}\;. (28)

Again for convenience of notation, let us put t:=1Tr{Pφ}1Tr{Pω}t:=\frac{1-\operatorname{Tr}\left\{P^{*}\ \varphi\right\}}{1-\operatorname{Tr}\left\{P^{*}\ {\omega^{*}}\right\}} and R:=𝔉g(σ~)R:=\mathfrak{F}_{g}(\tilde{\sigma}). We now recall the optimal decomposition (26): by inserting it into (28) and rearranging terms once more, we arrive at

(φ)=(1t+tR)σ~+(ttR)σ~0.\displaystyle\mathcal{E}(\varphi)=(1-t+tR)\tilde{\sigma}+(t-tR)\tilde{\sigma}_{0}\;. (29)

The above relation tells us that (φ)\mathcal{E}(\varphi) lies somewhere on the affine line passing through both σ~\tilde{\sigma} and σ~0\tilde{\sigma}_{0}. Therefore, in order to have (φ)𝖥\mathcal{E}(\varphi)\in\mathsf{F}, the coefficient (1t+tR)(1-t+tR) weighing σ~\tilde{\sigma} must be carefully bounded both from above and from below, so that (φ)\mathcal{E}(\varphi) is neither too close to σ~\tilde{\sigma} nor too close to σ~0\tilde{\sigma}_{0}, in which case it could end up lying outside 𝖥\mathsf{F} (see Fig. 1 for a schematic picture).

The upper bound is computed as follows. Since the free fraction is exactly defined as the maximum weight of σ~\tilde{\sigma} so that a convex mixture with σ~0\tilde{\sigma}_{0} lies in 𝖥\mathsf{F}, we want to show that the weight of σ~\tilde{\sigma} in (29) does not exceed RR, that is,

1t+tRR,1-t+tR\leqslant R\;,

or, equivalently,

 1t(1t)R.\;1-t\leqslant(1-t)R\;. (30)

Since, starting from Eq. (21),

Tr{Pω}\displaystyle\operatorname{Tr}\left\{P^{*}\ {\omega^{*}}\right\} =minP𝖯ϵ1(ρ)maxω𝖥Tr{Pω}\displaystyle=\min_{P\in\mathsf{P}^{\epsilon_{1}}(\rho)}\max_{\omega\in\mathsf{F}}\operatorname{Tr}\left\{P\ \omega\right\}
=maxω𝖥Tr{Pω}\displaystyle=\max_{\omega\in\mathsf{F}}\operatorname{Tr}\left\{P^{*}\ \omega\right\}
Tr{Pφ}0,\displaystyle\geqslant\operatorname{Tr}\left\{P^{*}\ \varphi\right\}\geqslant 0\;,

we see that t1t\geqslant 1, that is, 1t01-t\leqslant 0, and inequality (30) automatically holds for any R[0,1]R\in[0,1], without the need to invoke any extra condition.

Hence, condition (16) or condition (17) are only required to obtain the correct lower bound, that is, to prevent that (φ)\mathcal{E}(\varphi) crosses the border of 𝖥\mathsf{F} when approaching σ~0\tilde{\sigma}_{0}. In order to derive the lower bound, we resort to the construction introduced in Eq. (13) and depicted in Fig. 1. Once a decomposition achieving the generalized free fraction of σ~\tilde{\sigma} is found, σ~\tilde{\sigma}_{-} is the state on the boundary of 𝖥\mathsf{F}, which is “antipodal” with respect to σ~+\tilde{\sigma}_{+}. If we get past it, we end up outside 𝖥\mathsf{F}: we need to make sure this does not happen.

We begin by assuming that Dmax,𝖥(σ~σ~+)<+D_{\max,\mathsf{F}}(\tilde{\sigma}\|\tilde{\sigma}_{+})<+\infty, that is, σ~+σ~\tilde{\sigma}_{+}\neq\tilde{\sigma}_{-}. (We recall that Dmax,𝖥(σ~σ~+)>0D_{\max,\mathsf{F}}(\tilde{\sigma}\|\tilde{\sigma}_{+})>0 is a consequence of the assumption 𝔇maxϵ2(σ)>0\mathfrak{D}^{\epsilon_{2}}_{\max}(\sigma)>0.) In this case, we need to impose that

1t+tRR 2Dmax,𝖥(σ~σ~+)12Dmax,𝖥(σ~σ~+)1.1-t+tR\geqslant\frac{R\,2^{D_{\max,\mathsf{F}}(\tilde{\sigma}\|\tilde{\sigma}_{+})}-1}{2^{D_{\max,\mathsf{F}}(\tilde{\sigma}\|\tilde{\sigma}_{+})}-1}\;. (31)

Before proceeding, we notice that the above inequality, if satisfied, implies in particular 1t+tR01-t+tR\geqslant 0, because R 2Dmax,𝖥(σ~σ~+)=2Dmax,𝖥(σ~σ~+)Dmax(σ~σ~+)1R\,2^{D_{\max,\mathsf{F}}(\tilde{\sigma}\|\tilde{\sigma}_{+})}=2^{D_{\max,\mathsf{F}}(\tilde{\sigma}\|\tilde{\sigma}_{+})-D_{\max}(\tilde{\sigma}\|\tilde{\sigma}_{+})}\geqslant 1.

Condition (31), after writing tt explicitly again, reads as follows:

11Tr{Pφ}1Tr{Pω}+R(1Tr{Pφ}1Tr{Pω})R 2Dmax,𝖥(σ~σ~+)12Dmax,𝖥(σ~σ~+)1.1-\frac{1-\operatorname{Tr}\left\{P^{*}\ \varphi\right\}}{1-\operatorname{Tr}\left\{P^{*}\ {\omega^{*}}\right\}}+R\left(\frac{1-\operatorname{Tr}\left\{P^{*}\ \varphi\right\}}{1-\operatorname{Tr}\left\{P^{*}\ {\omega^{*}}\right\}}\right)\geqslant\frac{R\,2^{D_{\max,\mathsf{F}}(\tilde{\sigma}\|\tilde{\sigma}_{+})}-1}{2^{D_{\max,\mathsf{F}}(\tilde{\sigma}\|\tilde{\sigma}_{+})}-1}\;.

Since we are assuming that Tr{Pω}<1\operatorname{Tr}\left\{P^{*}\ {\omega^{*}}\right\}<1, multiplying both sides by 1Tr{Pω}1-\operatorname{Tr}\left\{P^{*}\ {\omega^{*}}\right\} does not change the inequality, so we obtain the equivalent condition:

(1R)Tr{Pφ}R 2Dmax,𝖥(σ~σ~+)12Dmax,𝖥(σ~σ~+)1(1Tr{Pω})+Tr{Pω}R.(1-R)\operatorname{Tr}\left\{P^{*}\ \varphi\right\}\geqslant\frac{R\,2^{D_{\max,\mathsf{F}}(\tilde{\sigma}\|\tilde{\sigma}_{+})}-1}{2^{D_{\max,\mathsf{F}}(\tilde{\sigma}\|\tilde{\sigma}_{+})}-1}(1-\operatorname{Tr}\left\{P^{*}\ {\omega^{*}}\right\})+\operatorname{Tr}\left\{P^{*}\ {\omega^{*}}\right\}-R\;.

After rearranging the right-hand side, we arrive at

(1R)Tr{Pφ}(1R)Tr{Pω}2Dmax,F(σ~σ~+)12Dmax,F(σ~σ~+)1.(1-R)\operatorname{Tr}\left\{P^{*}\ \varphi\right\}\geqslant(1-R)\frac{\operatorname{Tr}\left\{P^{*}\ {\omega^{*}}\right\}2^{D_{\max,F}(\tilde{\sigma}\|\tilde{\sigma}_{+})}-1}{2^{D_{\max,F}(\tilde{\sigma}\|\tilde{\sigma}_{+})}-1}\;.

Since R<1R<1 (because we assumed that σ~𝖥\tilde{\sigma}\notin\mathsf{F}, that is, 𝔇max(σ~)>0\mathfrak{D}_{\max}(\tilde{\sigma})>0), we can divide both sides by (1R)(1-R) and obtain

Tr{Pφ}1M2Dmax,F(σ~σ~+)12Dmax,F(σ~σ~+)1.\displaystyle\operatorname{Tr}\left\{P^{*}\ \varphi\right\}\geqslant\frac{\frac{1}{M}2^{D_{\max,F}(\tilde{\sigma}\|\tilde{\sigma}_{+})}-1}{2^{D_{\max,F}(\tilde{\sigma}\|\tilde{\sigma}_{+})}-1}\;. (32)

The above condition must be satisfied for any φ𝖥\varphi\in\mathsf{F}. Hence, what we really want is a lower bound on minφ𝖥Tr{Pφ}\min_{\varphi\in\mathsf{F}}\operatorname{Tr}\left\{P^{*}\ \varphi\right\}. Noticing that

minφ𝖥Tr{Pφ}\displaystyle\min_{\varphi\in\mathsf{F}}\operatorname{Tr}\left\{P^{*}\ \varphi\right\} minP𝖯ϵ1(ρ)minφ𝖥Tr{Pφ}\displaystyle\geqslant\min_{P\in\mathsf{P}^{\epsilon_{1}}(\rho)}\min_{\varphi\in\mathsf{F}}\operatorname{Tr}\left\{P\ \varphi\right\} (33)
=minω𝖥minP𝖯ϵ1(ρ)Tr{Pω}\displaystyle=\min_{\omega\in\mathsf{F}}\min_{P\in\mathsf{P}^{\epsilon_{1}}(\rho)}\operatorname{Tr}\left\{P\ \omega\right\} (34)
=2maxω𝖥Dhϵ1(ρω),\displaystyle=2^{-\max_{\omega\in\mathsf{F}}D^{\epsilon_{1}}_{h}(\rho\|\omega)}\;, (35)

condition (32) holds whenever the following, stricter condition holds, that is,

2maxω𝖥Dhϵ1(ρω)\displaystyle 2^{-\max_{\omega\in\mathsf{F}}D^{\epsilon_{1}}_{h}(\rho\|\omega)} 1M2Dmax,F(σ~σ~+)12Dmax,F(σ~σ~+)1\displaystyle\geqslant\frac{\frac{1}{M}2^{D_{\max,F}(\tilde{\sigma}\|\tilde{\sigma}_{+})}-1}{2^{D_{\max,F}(\tilde{\sigma}\|\tilde{\sigma}_{+})}-1}
=2Dmax,𝖥(σ~σ~+)minω𝖥Dhϵ1(ρω)12Dmax,𝖥(σ~σ~+)1.\displaystyle=\frac{2^{D_{\max,\mathsf{F}}(\tilde{\sigma}\|\tilde{\sigma}_{+})-\min_{\omega\in\mathsf{F}}D^{\epsilon_{1}}_{h}(\rho\|\omega)}-1}{2^{D_{\max,\mathsf{F}}(\tilde{\sigma}\|\tilde{\sigma}_{+})}-1}\;.

Hence, condition (16) guarantees that (φ)𝖥\mathcal{E}(\varphi)\in\mathsf{F} for any φ𝖥\varphi\in\mathsf{F}, that is, that the operation \mathcal{E} defined in (27) is a valid resource morphism.

Let us finally consider the case in which Dmax,𝖥(σ~σ~+)=+D_{\max,\mathsf{F}}(\tilde{\sigma}\|\tilde{\sigma}_{+})=+\infty, that is, σ~+=σ~\tilde{\sigma}_{+}=\tilde{\sigma}_{-}. In this case, lower and upper bounds have to coincide, so that the map defined in (27) is a resource morphism if and only if 1t+tR=R1-t+tR=R. This can only happen if R=1R=1 (but this is excluded because σ~𝖥\tilde{\sigma}\notin\mathsf{F}) or if 1t=01-t=0, that is, if t=1t=1 independently of the input φ𝖥\varphi\in\mathsf{F}. This is guaranteed if the operator PP^{*} in (27) has the same trace on all free states, which is exactly the content of (17). ∎


A less general, but simpler, statement stemming from Theorem 1 is the following:

Corollary 2.

With the same notations of Theorem 1, suppose that the following condition holds,

𝔇hϵ1(ρ)Dmax,𝖥(σ~σ~+).\displaystyle\mathfrak{D}^{\epsilon_{1}}_{h}(\rho)\geqslant D_{\max,\mathsf{F}}(\tilde{\sigma}\|\tilde{\sigma}_{+})\;. (36)

Then, ρϵ1+ϵ2σ\rho\succ_{\epsilon_{1}+\epsilon_{2}}\sigma.

Proof.

Assuming (36), if Dmax,𝖥(σ~σ~+)=+D_{\max,\mathsf{F}}(\tilde{\sigma}\|\tilde{\sigma}_{+})=+\infty, then also 𝔇hϵ1(ρ)=+\mathfrak{D}^{\epsilon_{1}}_{h}(\rho)=+\infty. In such a case, we know from Theorem 1, case (i), that ρϵ1σ\rho\succ_{\epsilon_{1}}\sigma, which of course implies also ρ(ϵ1+ϵ2)σ\rho\succ_{(\epsilon_{1}+\epsilon_{2})}\sigma.

On the other hand, if Dmax,𝖥(σ~σ~+)=0D_{\max,\mathsf{F}}(\tilde{\sigma}\|\tilde{\sigma}_{+})=0, we know that σ~𝖥\tilde{\sigma}\in\mathsf{F}, so that, in fact, 𝔇maxϵ2(σ)=0\mathfrak{D}^{\epsilon_{2}}_{\max}(\sigma)=0. In other words, we are in case (ii) of Theorem 1, and again ρ(ϵ1+ϵ2)σ\rho\succ_{(\epsilon_{1}+\epsilon_{2})}\sigma holds.

We are hence left to consider the case

+>𝔇hϵ1(ρ)Dmax,𝖥(σ~σ~+)>0.\displaystyle+\infty>\mathfrak{D}^{\epsilon_{1}}_{h}(\rho)\geqslant D_{\max,\mathsf{F}}(\tilde{\sigma}\|\tilde{\sigma}_{+})>0. (37)

We show that condition (37) alone implies both conditions (15) and (16) of case (iii.a) in Theorem 1.

Since by definition Dmax,𝖥(σ~σ~+)Dmax(σ~σ~+)=𝔇maxϵ2(σ)D_{\max,\mathsf{F}}(\tilde{\sigma}\|\tilde{\sigma}_{+})\geqslant D_{\max}(\tilde{\sigma}\|\tilde{\sigma}_{+})=\mathfrak{D}^{\epsilon_{2}}_{\max}(\sigma), we immediately see that condition (37) implies condition (15). Hence, we only need to show that also condition (16) is implied. In fact, we can show that (37) implies a condition that is even stronger than (16). Such a condition is the following:

02Dmax,𝖥(σ~σ~+)𝔇hϵ1(ρ)12Dmax,𝖥(σ~σ~+)1.0\geqslant\frac{2^{D_{\max,\mathsf{F}}(\tilde{\sigma}\|\tilde{\sigma}_{+})-\mathfrak{D}^{\epsilon_{1}}_{h}(\rho)}-1}{2^{D_{\max,\mathsf{F}}(\tilde{\sigma}\|\tilde{\sigma}_{+})}-1}\;.

If the above is satisfied, also (16) is satisfied, and we can conclude that ρ(ϵ1+ϵ2)σ\rho\succ_{(\epsilon_{1}+\epsilon_{2})}\sigma. The above inequality is satisfied because, as a consequence of Dmax,𝖥(σ~σ~+)>0D_{\max,\mathsf{F}}(\tilde{\sigma}\|\tilde{\sigma}_{+})>0, the denominator in the right-hand side is strictly positive, so that the above inequality is equivalent to

12Dmax,𝖥(σ~σ~+)𝔇hϵ1(ρ),1\geqslant 2^{D_{\max,\mathsf{F}}(\tilde{\sigma}\|\tilde{\sigma}_{+})-\mathfrak{D}^{\epsilon_{1}}_{h}(\rho)}\;,

which is satisfied if and only if condition (37) is satisfied. ∎

A merit of Corollary 2 is to provide a simple compact sufficient condition, free of supplementary caveat like condition (16), which is difficult to interpret operationally. However, the right-hand side of (36) is not yet a valid resource monotone. The following result fills the gap.

Theorem 2.

Given ρ,σ𝖣(m)\rho,\sigma\in\mathsf{D}(\mathbb{C}^{m}) and ϵ1,ϵ2[0,1]\epsilon_{1},\epsilon_{2}\in[0,1], if

𝔇hϵ1(ρ)𝔇max,𝖥ϵ2(σ),\displaystyle\mathfrak{D}^{\epsilon_{1}}_{h}(\rho)\geqslant\mathfrak{D}_{\max,\mathsf{F}}^{\epsilon_{2}}(\sigma)\;, (38)

then ρ(ϵ1+ϵ2)σ\rho\succ_{(\epsilon_{1}+\epsilon_{2})}\sigma.

Theorem 2, when ϵ2=0\epsilon_{2}=0 and σ\sigma is rank-one, recovers Theorem 5 in Ref. LBT (19) (see also Corollary 17 of RBTL (19)).

Theorem 1 and Theorem 2 are independent of each other. This is because, on the one hand, it is possible that 𝔇hϵ1(ρ)𝔇maxϵ2(σ)\mathfrak{D}^{\epsilon_{1}}_{h}(\rho)\geqslant\mathfrak{D}_{\max}^{\epsilon_{2}}(\sigma) even though 𝔇hϵ1(ρ)\centernot𝔇max,𝖥ϵ2(σ)\mathfrak{D}^{\epsilon_{1}}_{h}(\rho)\centernot{\geqslant}\mathfrak{D}_{\max,\mathsf{F}}^{\epsilon_{2}}(\sigma), so that Theorem 2 would be inconclusive. On the other hand, it is possible that 𝔇hϵ1(ρ)𝔇max,𝖥ϵ2(σ)\mathfrak{D}^{\epsilon_{1}}_{h}(\rho)\geqslant\mathfrak{D}_{\max,\mathsf{F}}^{\epsilon_{2}}(\sigma) even though neither condition (16) nor (17) hold, so that Theorem 1 would be inconclusive. In other words, Theorem 1 and Theorem 2 in general apply to two different regimes and are logically independent of each other. Nonetheless, since σ~𝖡ϵ2(σ)\tilde{\sigma}\in\mathsf{B}^{\epsilon_{2}}(\sigma) and σ~+𝖥\tilde{\sigma}_{+}\in\mathsf{F}, we see that Dmax,𝖥(σ~σ~+)𝔇max,𝖥ϵ2(σ)D_{\max,\mathsf{F}}(\tilde{\sigma}\|\tilde{\sigma}_{+})\geqslant\mathfrak{D}_{\max,\mathsf{F}}^{\epsilon_{2}}(\sigma). This implies that Corollary 2 above can be as well derived as a consequence of Theorem 2.

Proof.

We begin by noticing that, if 𝔇hϵ1(ρ)=+\mathfrak{D}^{\epsilon_{1}}_{h}(\rho)=+\infty, we are back to case (i) of Theorem 1. Also, if 𝔇max,𝖥ϵ2(σ)=0\mathfrak{D}_{\max,\mathsf{F}}^{\epsilon_{2}}(\sigma)=0, then also 𝔇maxϵ2(σ)=0\mathfrak{D}_{\max}^{\epsilon_{2}}(\sigma)=0, and we are back to case (ii) of Theorem 1. In what follows we will hence assume that +>𝔇hϵ1(ρ)𝔇max,𝖥ϵ2(σ)>0+\infty>\mathfrak{D}^{\epsilon_{1}}_{h}(\rho)\geqslant\mathfrak{D}_{\max,\mathsf{F}}^{\epsilon_{2}}(\sigma)>0.

Let us define P,ω,σ~,σ~+P^{*},{\omega^{*}},\tilde{\sigma},\tilde{\sigma}_{+} as the optimizers achieving the quantities that appear in condition (38), that is,

𝔇hϵ1(ρ):=minω𝖥Dhϵ1(ρω)=Dhϵ1(ρω)=logTr{Pω}\displaystyle\mathfrak{D}^{\epsilon_{1}}_{h}(\rho):=\min_{\omega\in\mathsf{F}}D_{h}^{\epsilon_{1}}(\rho\|\omega)=D_{h}^{\epsilon_{1}}(\rho\|{\omega^{*}})=-\log\operatorname{Tr}\left\{P^{*}\ {\omega^{*}}\right\} (39)
𝔇max,𝖥ϵ2(σ):=minω𝖥Dmax,𝖥ϵ2(σω)=Dmax,𝖥(σ~σ~+).\displaystyle\mathfrak{D}_{\max,\mathsf{F}}^{\epsilon_{2}}(\sigma):=\min_{\omega\in\mathsf{F}}D_{\max,\mathsf{F}}^{\epsilon_{2}}(\sigma\|\omega)=D_{\max,\mathsf{F}}(\tilde{\sigma}\|\tilde{\sigma}_{+})\;. (40)

Notice that while σ~,σ~+\tilde{\sigma},\tilde{\sigma}_{+} were used in Theorem 1 to denote the optimizers achieving 𝔇maxϵ2(σ)\mathfrak{D}_{\max}^{\epsilon_{2}}(\sigma), for the sake of this proof the same symbols are used to denote the optimizers achieving 𝔇max,𝖥ϵ2(σ)\mathfrak{D}_{\max,\mathsf{F}}^{\epsilon_{2}}(\sigma).

Writing M:=1/Tr{Pω}M:=1/\operatorname{Tr}\left\{P^{*}\ {\omega^{*}}\right\}, that is,

1M=Tr{Pω}=maxω𝖥minP𝖯ϵ1(ρ)Tr{Pω},\frac{1}{M}=\operatorname{Tr}\left\{P^{*}\ {\omega^{*}}\right\}=\max_{\omega\in\mathsf{F}}\min_{P\in\mathsf{P}^{\epsilon_{1}}(\rho)}\operatorname{Tr}\left\{P\ \omega\right\}\;,

we define the map

()=Tr{P}σ~+(1Tr{P})Mσ~+σ~M1.\displaystyle\mathcal{E}(\cdot)=\operatorname{Tr}\left\{P^{*}\ \cdot\right\}\tilde{\sigma}+(1-\operatorname{Tr}\left\{P^{*}\ \cdot\right\})\frac{M\tilde{\sigma}_{+}-\tilde{\sigma}}{M-1}\;. (41)

Notice that, with respect to the map constructed in (27), the above map uses the same operator PP^{*}, but prepares different states depending on the outcome. As before, moreover, it is possible to assume without loss of generality that 1<M<+1<M<+\infty.

Since Dmax,𝖥(σ~σ~+)Dmax(σ~σ~+)D_{\max,\mathsf{F}}(\tilde{\sigma}\|\tilde{\sigma}_{+})\geqslant D_{\max}(\tilde{\sigma}\|\tilde{\sigma}_{+}), condition (38) implies that,

Dhϵ1(ρω)Dmax(σ~σ~+),\displaystyle D_{h}^{\epsilon_{1}}(\rho\|{\omega^{*}})\geqslant D_{\max}(\tilde{\sigma}\|\tilde{\sigma}_{+})\;, (42)

A direct consequence of BST (19) is that condition (42), together with the fact that σ~𝖡ϵ2(σ)\tilde{\sigma}\in\mathsf{B}^{\epsilon_{2}}(\sigma), imply that the map \mathcal{E} defined in (41) is a valid CPTP map such that 12(ρ)σ1ϵ1+ϵ2\frac{1}{2}\left|\!\left|{\mathcal{E}(\rho)-\sigma}\right|\!\right|_{1}\leqslant\epsilon_{1}+\epsilon_{2}. In what follows we show that \mathcal{E} is, in particular, a resource morphism.

Because σ~\tilde{\sigma} and σ~+\tilde{\sigma}_{+} have been chosen as the states that optimize 𝔇max,𝖥ϵ2(σ)\mathfrak{D}_{\max,\mathsf{F}}^{\epsilon_{2}}(\sigma), we have Dmax,𝖥(σ~σ~+)=𝔇max,𝖥(σ~)=log𝔉(σ~)D_{\max,\mathsf{F}}(\tilde{\sigma}\|\tilde{\sigma}_{+})=\mathfrak{D}_{\max,\mathsf{F}}(\tilde{\sigma})=-\log{\mathfrak{F}(\tilde{\sigma})}. Therefore, we obtain the following decomposition of σ+\sigma_{+},

σ~+=𝔉(σ~)σ~+(1𝔉(σ~))σ~0,\displaystyle\tilde{\sigma}_{+}=\mathfrak{F}(\tilde{\sigma})\tilde{\sigma}+(1-\mathfrak{F}(\tilde{\sigma}))\tilde{\sigma}_{0}\;, (43)

with σ~0𝖥\tilde{\sigma}_{0}\in\mathsf{F}. By plugging (43) in (41), and considering as input to the map an arbitrary free state φ𝖥\varphi\in\mathsf{F}, we reach the following

(φ)=(1t+tR)σ~+(ttR)σ~0,\displaystyle\mathcal{E}(\varphi)=(1-t+tR)\tilde{\sigma}+(t-tR)\tilde{\sigma}_{0}\;, (44)

where, for the sake of notation, we put t:=1Tr{Pφ}1Tr{Pω}t:=\frac{1-\operatorname{Tr}\left\{P^{*}\ \varphi\right\}}{1-\operatorname{Tr}\left\{P^{*}\ {\omega^{*}}\right\}} and R:=𝔉(σ)R:=\mathfrak{F}(\sigma). Notice that while the proof of Theorem 1 is obtained by working with the generalized free fraction, in this proof we are mostly working with the free fraction.

We need to show that (φ)𝖥\mathcal{E}(\varphi)\in\mathsf{F}, for all φ𝖥\varphi\in\mathsf{F}. To that end, we only need to show that the weight in front of σ~\tilde{\sigma} in (44) is non-negative and upper bounded by RR.

In order to show that it does not exceed RR, we proceed as follows. In the proof of Theorem 1, we have shown that t1t\geqslant 1, so that 1t+tRR1-t+tR\leqslant R, that is, R(t1)t1R(t-1)\leqslant t-1, holds automatically for any R[0,1]R\in[0,1].

In order to show the weight of σ~\tilde{\sigma} is non-negative, it suffices to show that

R11t.R\geqslant 1-\frac{1}{t}\;.

Since t11Tr{Pω}=MM1t\leqslant\frac{1}{1-\operatorname{Tr}\left\{P^{*}\ {\omega^{*}}\right\}}=\frac{M}{M-1}, we have that 1t11M1M1-t^{-1}\leqslant 1-\frac{M-1}{M}, so that the above is satisfied whenever

R1M=Tr{Pω},R\geqslant\frac{1}{M}=\operatorname{Tr}\left\{P^{*}\ {\omega^{*}}\right\}\;,

that is to say

𝔉(σ~)=2𝔇max,𝖥ϵ2(σ)Tr{Pω}=2𝔇hϵ1(ρ).\mathfrak{F}(\tilde{\sigma})=2^{-\mathfrak{D}_{\max,\mathsf{F}}^{\epsilon_{2}}(\sigma)}\geqslant\operatorname{Tr}\left\{P^{*}\ {\omega^{*}}\right\}=2^{-\mathfrak{D}^{\epsilon_{1}}_{h}(\rho)}\;.

IV Applications and examples

In this section we apply Theorems 1 and 2 to some situations of physical interest, and show how we can not only rederive, but sometimes also strengthen, previous results.

IV.1 Singleton Resource Theories

We begin this section by considering the special case of singleton resource theories, in which the set of free states 𝖥\mathsf{F} comprises only one element. This scenario includes the resource theory of athermality, namely, the case in which free operations are those that preserve the thermal state of the system HO (13); BHO+ (13); BHN+ (15); Bus (15), which in turns provide the backbone of the resource theory of quantum thermodynamics GJB+ (18). More generally, when the output singleton is allowed to differ from the input one, this is referred to as the resource theory of asymmetric distinguishability, whose optimal rates have been studied in Mat (10); BST (19); WW (19).

In the singleton case, the log-robustness typically is infinite, and the applicability of Theorem 2 is quite limited. On the contrary, Theorem 1 can still be useful, even in the case of a singleton 𝖥\mathsf{F}. Indeed, Theorem 1 reduces in the singleton case to Lemma 3.3 of BST (19), which is good enough to serve as the starting point to study optimal asymptotic interconversion rates.

Proposition 1.

Consider an input system, with initial state ρ𝖣(m)\rho\in\mathsf{D}(\mathbb{C}^{m}) and free singleton 𝖥={γ}\mathsf{F}=\{\gamma\}, and an output system, with target state σ𝖣(n)\sigma\in\mathsf{D}(\mathbb{C}^{n}) and free singleton 𝖥={γ}\mathsf{F}^{\prime}=\{\gamma^{\prime}\}. If the following condition holds,

Dhϵ1(ργ)Dmaxϵ2(σγ),D^{\epsilon_{1}}_{h}(\rho\|\gamma)\geqslant D^{\epsilon_{2}}_{\max}(\sigma\|\gamma^{\prime}), (45)

then ρ(ϵ1+ϵ2)σ\rho\succ_{(\epsilon_{1}+\epsilon_{2})}\sigma.

Proof.

We can restrict ourselves to consider only case (iii.b) of Theorem 1, because for a singleton 𝖥={γ}\mathsf{F}^{\prime}=\{\gamma^{\prime}\}, whenever σγ\sigma\neq\gamma^{\prime}, one has Dmax,𝖥(σγ)=+D_{\max,\mathsf{F^{\prime}}}(\sigma\|\gamma^{\prime})=+\infty. But since also the input free set 𝖥\mathsf{F} is a singleton, we have

minω𝖥Dhϵ1(ρω)=maxω𝖥Dhϵ1(ρω),\min_{\omega\in\mathsf{F}}D_{h}^{\epsilon_{1}}(\rho\|\omega)=\max_{\omega\in\mathsf{F}}D_{h}^{\epsilon_{1}}(\rho\|\omega)\;,

and condition (17) is automatically satisfied. ∎

IV.2 Resource Theory of Bipartite Entanglement

Next, we specialize our results to the resource theory of entanglement. We begin by considering bipartite entanglement, namely, the case in which 𝖥\mathsf{F} is the set of all separable states of a given bipartite system. Resource morphisms are given by separability-preserving (or non-entangling) CPTP maps, usually denoted as SEPP. One-shot entanglement distillation and dilution under SEPP have been studied in BD (11). In what follows we show how our Corollary 2 is able to guarantee the existence of a SEPP transition directly mapping ρ\rho to σ\sigma, even in situations in which the results of Ref. BD (11) cannot guarantee the existence of a “distill-and-dilute” transition.

In order to illustrate the point, it is enough to consider the exact case, that is, ϵ1=ϵ2=0\epsilon_{1}=\epsilon_{2}=0. The same conclusions hold also in the approximate case, however, some care must be taken in that while here we use the trace-distance to measure approximations, Ref. BD (11) uses the fidelity. Trace-distance and fidelity are well-known to be equivalent Wil (13); NC (00) , but approximation parameters must be changed: we leave it to the interested reader to work out the exact factors.

By rewriting the main results of BD (11) using our notation, the zero-error one-shot SEPP-distillable entanglement ED,SEPP(1)(ρ)E^{(1)}_{D,\text{\text{SEPP}}}(\rho) and the zero-error one-shot SEPP-entanglement cost EC,SEPP(1)(σ)E^{(1)}_{C,\text{SEPP}}(\sigma) satisfy

ED,SEPP(1)(ρ)𝔇min(ρ)\displaystyle E^{(1)}_{D,\text{SEPP}}(\rho)\geqslant\lfloor\mathfrak{D}_{\min}(\rho)\rfloor (46)

and

EC,SEPP(1)(σ)𝔇max,𝖥(σ)+1,\displaystyle E^{(1)}_{C,\text{SEPP}}(\sigma)\leqslant\mathfrak{D}_{\max,\mathsf{F}}(\sigma)+1\;, (47)

respectively. These two relations together guarantee that it is possible to exactly go from ρ\rho to σ\sigma via SEPP (passing through the maximally entangled state) if

𝔇min(ρ)𝔇max,𝖥(σ)+1,\lfloor\mathfrak{D}_{\min}(\rho)\rfloor\geqslant\mathfrak{D}_{\max,\mathsf{F}}(\sigma)+1\;,

which is more restrictive than what Theorem 2 says, that is,

𝔇min(ρ)𝔇max,𝖥(σ).\mathfrak{D}_{\min}(\rho)\geqslant\mathfrak{D}_{\max,\mathsf{F}}(\sigma)\;.

This is possible because we do not require the transformation to pass through the maximally entangled state, but we allow it to go directly from ρ\rho to σ\sigma.

Remark.

When working within the resource theory of entanglement, especially in the one-shot regime, it is customary to allow the output system to differ from the input one. Consequently, also the set of free states changes from 𝖥\mathsf{F} to 𝖥\mathsf{F}^{\prime}. As already noticed, our bounds can be straightforwardly extended to cover this situation as well: in such a case, all quantities related to the input state ρ\rho will be computed with respect to the input free set 𝖥\mathsf{F}, while all quantities related to the output state σ\sigma will be computed with respect to the output free set 𝖥\mathsf{F}^{\prime}.

IV.3 Existence of a Maximally Resourceful State and Weak-Converse Bounds for Distillation and Dilution

In this section we show how Corollary 1 and Theorem 2 can be used to formulate sufficient conditions that guarantee that an element α\alpha is maximally resourceful, in the sense of Definition 2. We also address the related problem of deciding when Corollary 1 and Theorem 2 are optimal, i.e., when the sufficient conditions they formulate become also necessary. For the sake of the presentation, we focus here on the case of exact transitions, that is, ϵ1=ϵ2=0\epsilon_{1}=\epsilon_{2}=0, keeping in mind, however, that the results of Section III allow us to go beyond the exact case and to speak of, e.g., almost-maximally resourceful elements.

We begin with the following fact (see also Corollary 4 in LBT (19)):

Proposition 2.

The following statements hold.

  1. (i)

    Let α𝖣(d)\alpha\in\mathsf{D}(\mathbb{C}^{d}) be such that 𝔇min(α)=maxρ𝖣(d)𝔇max(ρ)\mathfrak{D}_{\min}(\alpha)=\max_{\rho\in\mathsf{D}(\mathbb{C}^{d})}\mathfrak{D}_{\max}(\rho), and Tr{ωΠα}=constant\operatorname{Tr}\left\{\omega\ \Pi_{\alpha}\right\}=\operatorname{constant}, for any ω𝖥\omega\in\mathsf{F}. Then α\alpha is maximally resourceful in 𝖣(d)\mathsf{D}(\mathbb{C}^{d}).

  2. (ii)

    Let α𝖣(d)\alpha\in\mathsf{D}(\mathbb{C}^{d}) be such that 𝔇min(α)=maxρ𝖣(d)𝔇max,𝖥(ρ)\mathfrak{D}_{\min}(\alpha)=\max_{\rho\in\mathsf{D}(\mathbb{C}^{d})}\mathfrak{D}_{\max,\mathsf{F}}(\rho). Then α\alpha is maximally resourceful in 𝖣(d)\mathsf{D}(\mathbb{C}^{d}).

Proof.

Case (i): being Tr{ωΠα}\operatorname{Tr}\left\{\omega\ \Pi_{\alpha}\right\} constant for any ω𝖥\omega\in\mathsf{F}, the assumptions in Corollary 1 are satisfied with ϵ1=ϵ2=0\epsilon_{1}=\epsilon_{2}=0. The proof then follows trivially, from the assumption that 𝔇min(α)=maxρ𝖣(d)𝔇max(ρ)𝔇max(σ)\mathfrak{D}_{\min}(\alpha)=\max_{\rho\in\mathsf{D}(\mathbb{C}^{d})}\mathfrak{D}_{\max}(\rho)\geqslant\mathfrak{D}_{\max}(\sigma) for any σ𝖣(d)\sigma\in\mathsf{D}(\mathbb{C}^{d}).

Case (ii): in this case we apply Theorem 2, and again the proof follows trivially, from the assumption that 𝔇min(α)=maxρ𝖣(d)𝔇max,𝖥(ρ)𝔇max,𝖥(σ)\mathfrak{D}_{\min}(\alpha)=\max_{\rho\in\mathsf{D}(\mathbb{C}^{d})}\mathfrak{D}_{\max,\mathsf{F}}(\rho)\geqslant\mathfrak{D}_{\max,\mathsf{F}}(\sigma).

Remark.

Since, for any ρ,σ\rho,\sigma, Dmin(ρσ)Dmax(ρσ)Dmax,𝖥(ρσ)D_{\min}(\rho\|\sigma)\leqslant D_{\max}(\rho\|\sigma)\leqslant D_{\max,\mathsf{F}}(\rho\|\sigma), condition (i) in Proposition 2 above implies that 𝔇min(α)=𝔇max(α)=maxρ𝖣(d)𝔇min(ρ)=maxρ𝖣(d)𝔇max(ρ)\mathfrak{D}_{\min}(\alpha)=\mathfrak{D}_{\max}(\alpha)=\max_{\rho\in\mathsf{D}(\mathbb{C}^{d})}\mathfrak{D}_{\min}(\rho)=\max_{\rho\in\mathsf{D}(\mathbb{C}^{d})}\mathfrak{D}_{\max}(\rho); analogously, condition (ii) implies 𝔇min(α)=𝔇max,𝖥(α)=maxρ𝖣(d)𝔇min(ρ)=maxρ𝖣(d)𝔇max,𝖥(ρ)\mathfrak{D}_{\min}(\alpha)=\mathfrak{D}_{\max,\mathsf{F}}(\alpha)=\max_{\rho\in\mathsf{D}(\mathbb{C}^{d})}\mathfrak{D}_{\min}(\rho)=\max_{\rho\in\mathsf{D}(\mathbb{C}^{d})}\mathfrak{D}_{\max,\mathsf{F}}(\rho).

The two sufficient conditions considered in Proposition 2 are independent. For example, both in the resource theory of bipartite entanglement and in the resource theory of coherence a golden state exists, namely, the maximally entangled state and the maximally coherent state, respectively. It is also known that these are both in fact maximally resourceful in their respective theories. However, while the maximally coherent state satisfies condition (i) but not condition (ii), the maximally entangled state satisfies condition (ii) but not (i): see RBTL (19) for the explicit calculation.

Another example is provided by the resource theory of genuine multipartite entanglement, in which the free set is taken to be the set of all biseparable states and resource morphisms correspondingly are defined as biseparability-preserving maps. In this case, it is possible to show by explicit calculation CTPdV (19); RBTL (19) that the generalized GHZ state, that is,

|Ψ𝖦𝖧𝖹(N,d):=1di=1d|iN,|\Psi_{\mathsf{GHZ}}^{(N,d)}\rangle:=\frac{1}{\sqrt{d}}\sum_{i=1}^{d}|i\rangle^{\otimes{N}}\;,

satisfies condition (ii) of Proposition 2. We conclude, therefore, that |Ψ𝖦𝖧𝖹(N,d)|\Psi_{\mathsf{GHZ}}^{(N,d)}\rangle is maximally resourceful.

The following propositions provide sufficient conditions so that the bounds in Corollary 1 and Theorem 2 are optimal. In the following proposition, we make it explicit that the input system (with state space 𝖣(m)\mathsf{D}(\mathbb{C}^{m}) and free set 𝖥\mathsf{F}) in general may differ from the output system (with state space 𝖣(n)\mathsf{D}(\mathbb{C}^{n}) and free set 𝖥\mathsf{F}^{\prime}). A related result is Theorem 2 of Ref. LBT (19).

Proposition 3 (Weak-converse bounds for dilution).

When dealing with transitions from an input system (m,𝖥)(\mathbb{C}^{m},\mathsf{F}) to an output system (n,𝖥)(\mathbb{C}^{n},\mathsf{F}^{\prime}), the following statements hold.

  1. (i)

    Suppose that α𝖣(m)\alpha\in\mathsf{D}(\mathbb{C}^{m}) satisfies 𝔇min(α)=𝔇max(α)\mathfrak{D}_{\min}(\alpha)=\mathfrak{D}_{\max}(\alpha); then, for any σ𝖣(n)\sigma\in\mathsf{D}(\mathbb{C}^{n})

    ασ𝔇min(α)𝔇max(σ).\displaystyle\alpha\succ\sigma\qquad\implies\qquad\mathfrak{D}_{\min}(\alpha)\geqslant\mathfrak{D}_{\max}(\sigma)\;. (48)
  2. (ii)

    Suppose that α𝖣(m)\alpha\in\mathsf{D}(\mathbb{C}^{m}) satisfies 𝔇min(α)=𝔇max,𝖥(α)\mathfrak{D}_{\min}(\alpha)=\mathfrak{D}_{\max,\mathsf{F}}(\alpha); then, for any σ𝖣(n)\sigma\in\mathsf{D}(\mathbb{C}^{n})

    ασ𝔇min(α)𝔇max,𝖥(σ).\displaystyle\alpha\succ\sigma\qquad\implies\qquad\mathfrak{D}_{\min}(\alpha)\geqslant\mathfrak{D}_{\max,\mathsf{F}^{\prime}}(\sigma)\;. (49)
Proof.

Case (i): suppose that ασ\alpha\succ\sigma, so that there exists a resource morphism :𝖣(m)𝖣(n)\mathcal{E}:\mathsf{D}(\mathbb{C}^{m})\to\mathsf{D}(\mathbb{C}^{n}) such that (α)=σ\mathcal{E}(\alpha)=\sigma; then,

𝔇min(α)\displaystyle\mathfrak{D}_{\min}(\alpha) =𝔇max(α)\displaystyle=\mathfrak{D}_{\max}(\alpha)
𝔇max((α))\displaystyle\geqslant\mathfrak{D}_{\max}(\mathcal{E}(\alpha))
=𝔇max(σ),\displaystyle=\mathfrak{D}_{\max}(\sigma)\;,

where the inequality in the second line comes from the fact that 𝔇max\mathfrak{D}_{\max} is a resource monotone.

Case (ii): suppose that ασ\alpha\succ\sigma, then

𝔇min(α)\displaystyle\mathfrak{D}_{\min}(\alpha) =𝔇max,𝖥(α)\displaystyle=\mathfrak{D}_{\max,\mathsf{F}}(\alpha)
𝔇max,𝖥((α))\displaystyle\geqslant\mathfrak{D}_{\max,\mathsf{F}^{\prime}}(\mathcal{E}(\alpha))
=𝔇max,𝖥(σ),\displaystyle=\mathfrak{D}_{\max,\mathsf{F}^{\prime}}(\sigma)\;,

where the inequality in the second line comes from the fact that 𝔇max,𝖥\mathfrak{D}_{\max,\mathsf{F}} is a resource monotone. ∎

An analogous weak converse for distillation is the following (see also Theorem 5 of LBT (19) for a related result).

Proposition 4 (Weak-converse bound for distillation).

Consider an input system (m,𝖥)(\mathbb{C}^{m},\mathsf{F}) and an output system (n,𝖥)(\mathbb{C}^{n},\mathsf{F}^{\prime}), and let α𝖣(n)\alpha\in\mathsf{D}(\mathbb{C}^{n}) be a target state such that 𝔇max,𝖥(α)=𝔇min(α)\mathfrak{D}_{\max,\mathsf{F}^{\prime}}(\alpha)=\mathfrak{D}_{\min}(\alpha). Then, for any ρ𝖣(m)\rho\in\mathsf{D}(\mathbb{C}^{m}),

ρα𝔇min(ρ)𝔇max,𝖥(α).\displaystyle\rho\succ\alpha\qquad\implies\qquad\mathfrak{D}_{\min}(\rho)\geqslant\mathfrak{D}_{\max,\mathsf{F}^{\prime}}(\alpha)\;. (50)
Proof.

If ρα\rho\succ\alpha then 𝔇min(ρ)𝔇min((ρ))=𝔇min(α)=𝔇max,𝖥(α)\mathfrak{D}_{\min}(\rho)\geqslant\mathfrak{D}_{\min}(\mathcal{E}(\rho))=\mathfrak{D}_{\min}(\alpha)=\mathfrak{D}_{\max,\mathsf{F}^{\prime}}(\alpha). ∎

Remark.

By looking at the proofs of Theorem 1 and Theorem 2, we see that the resource morphisms used there have been constructed as test-and-prepare quantum channels. As a consequence, Propositions 3 and 4 above can be interpreted as giving sufficient conditions for which test-and-prepare channels are provably optimal in resource manipulation, despite constituting a very special class among all CPTP maps.

A natural question to ask, at this point, is whether density matrices always exist, for which Propositions 3 and 4 hold, namely, for which test-and-prepare channels provide the optimal resource morphisms. As it turns out, perhaps surprisingly, in any resource theory with non-empty closed and convex 𝖥\mathsf{F}, even if a maximally resourceful element may not exist, a golden state, namely, a rank-one density matrix Ψ+𝖣(d)\Psi_{+}\in\mathsf{D}(\mathbb{C}^{d}) such that maxρ𝖣(d)𝔇min(ρ)=𝔇min(Ψ+)=𝔇max(Ψ+)=maxρ𝖣(d)𝔇max(ρ)\max_{\rho\in\mathsf{D}(\mathbb{C}^{d})}\mathfrak{D}_{\min}(\rho)=\mathfrak{D}_{\min}(\Psi_{+})=\mathfrak{D}_{\max}(\Psi_{+})=\max_{\rho\in\mathsf{D}(\mathbb{C}^{d})}\mathfrak{D}_{\max}(\rho), can always be found LBT (19); RBTL (19). However, before concluding that test-and-prepare morphisms are optimal for golden states, one still needs to verify that, either Ψ+\Psi_{+} satisfies Tr{ωΨ+}=constant\operatorname{Tr}\left\{\omega\ \Psi_{+}\right\}=\text{constant} for all free ω\omega, or 𝔇max(Ψ+)=𝔇max,𝖥(Ψ+)\mathfrak{D}_{\max}(\Psi_{+})=\mathfrak{D}_{\max,\mathsf{F}}(\Psi_{+}) also holds, and both such extra conditions depend on the actual resource theory at hand. The resource theories of coherence and bipartite entanglement again provide two paradigmatic examples in this sense.

V Conclusions: resource comparison without a maximally resourceful state

In this work we have derived sufficient (and, in some cases, necessary) conditions for the existence of a CPTP linear map transforming, up to arbitrary accuracy, an input state ρ\rho into a target state σ\sigma, under the additional condition that a convex non-empty subset of states is mapped into itself. Such a framework is particularly suitable to be applied to generalized resource theories, in which the convex subset represents the set of free states in the theory, so that any transformation that maps free states to free states state is itself free, in the sense that it cannot create resources for free. The conditions that we formulated are expressed in terms of entropic monotones, which are computed independently for the input state and the target state. In this way, we can still speak of the resource’s worth of any state, taken individually, even if a privileged maximally resourceful state does not exist, so that the tasks of resource distillation and dilution (which typically are used to quantify the resource content) cannot be defined. Aspects that we did not cover in this work are the scaling of the interconversion rates in the case in which a composition rule is given, and the complexity of numerically computing the entropic monotones used to compare resources.

Acknowledgements.
The authors are very grateful to Kaifeng Bu, Zi-Wen Liu, Bartosz Regula, and Ryuji Takagi, for insightful discussions about general resource theories, clarifications about the relations between this work and theirs, and a careful reading of a preliminary version of this work. W.Z. thanks Masahito Hayashi for insightful discussions and hospitality during his visit to the Peng Cheng Laboratory, Shenzhen, China. F.B. acknowledges support from the Japan Society for the Promotion of Science (JSPS) KAKENHI, Grants No.19H04066 and No.20K03746.

References