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

The Communication Value of a Quantum Channel

Eric Chitambar, Ian George, Brian Doolittle, Marius Junge E. Chitambar and I. George are with the Department of Electrical and Computer Engineering, University of Illinois Urbana-Champaign, Urbana, IL, 61801 USA, e-mail: [email protected]. Doolittle is with the Department of Physics, University of Illinois Urbana-Champaign, Urbana, IL, 61801 USA.M. Junge is with the Department of Mathematics, University of Illinois Urbana-Champaign, Urbana, IL, 61801 USA.
Abstract

There are various ways to quantify the communication capabilities of a quantum channel. In this work we study the communication value (cv) of channel, which describes the optimal success probability of transmitting a randomly selected classical message over the channel. The cv also offers a dual interpretation as the classical communication cost for zero-error channel simulation using non-signaling resources. We first provide an entropic characterization of the cv as a generalized conditional min-entropy over the cone of separable operators. Additionally, the logarithm of a channel’s cv is shown to be equivalent to its max-Holevo information, which can further be related to channel capacity. We evaluate the cv exactly for all qubit channels and the Werner-Holevo family of channels. While all classical channels are multiplicative under tensor product, this is no longer true for quantum channels in general. We provide a family of qutrit channels for which the cv is non-multiplicative. On the other hand, we prove that any pair of qubit channels have multiplicative cv when used in parallel. Even stronger, all entanglement-breaking channels and the partially depolarizing channel are shown to have multiplicative cv when used in parallel with any channel. We then turn to the entanglement-assisted cv and prove that it is equivalent to the conditional min-entropy of the Choi matrix of the channel. Combining with previous work on zero-error channel simulation, this implies that the entanglement-assisted cv is the classical communication cost for perfectly simulating a channel using quantum non-signaling resources. A final component of this work investigates relaxations of the channel cv to other cones such as the set of operators having a positive partial transpose (PPT). The PPT cv is analytically and numerically investigated for well-known channels such as the Werner-Holevo family and the dephrasure family of channels.

I Introduction

A noisy communication channel prohibits perfect transmission of messages from the sender (Alice) to the receiver (Bob). While there are a number of ways to quantify the noise of a channel, perhaps the simplest is in terms of a guessing game. Suppose that with uniform probability Alice randomly chooses a channel input and sends it to Bob over the channel. Based on the channel output, Bob tries to guess Alice’s input with the greatest probability of success. In this game, Bob’s optimal strategy is to perform maximum likelihood estimation based on the channel’s transition probabilities. To be concrete, suppose that 𝐏:[n][n]\mathbf{P}:[n]\to[n^{\prime}] is a channel mapping set [n]:={1,,n}[n]:=\{1,\cdots,n\} to set [n][n^{\prime}] with transition probabilities P(y|x)P(y|x). We define the channel’s communication value (cv) to be

cv(𝐏)=y[n]maxx[n]P(y|x).\text{cv}(\mathbf{P})=\sum_{y\in[n^{\prime}]}\max_{x\in[n]}P(y|x). (1)

It is then straightforward to see that 1ncv(𝐏)\frac{1}{n}\text{cv}(\mathbf{P}) is the largest success probability of correctly identifying the input xx based on the output yy, when xx is drawn uniformly from [n][n]. The quantity cv(𝐏)\text{cv}(\mathbf{P}) is thus a natural measure for how well a channel 𝐏\mathbf{P} transmits data on the single-copy level. The goal of this paper is to better understand the channel cv in different communication settings.

The channel cv also emerges in the problem of zero-error channel simulation [1, 2, 3, 4]. In the general task of channel simulation, Alice and Bob attempt to generate one channel 𝐏\mathbf{P} using another channel 𝐐\mathbf{Q} combined with pre- and post-processing [5, 6, 7, 8, 9]. Interesting variations to this problem arise when different types of resources are used to coordinate the pre- and post-processing of 𝐐\mathbf{Q}. For example, these resources could be shared randomness [6, 10], shared quantum entanglement [11, 12, 13, 14, 15], or non-signaling side-channel [1, 2, 4]. The latter refers to a general bipartite channel that prohibits communication from one party to the other. When 𝐐=idr\mathbf{Q}=\textrm{id}_{r} is the identity map on [r][r], then the goal is to perfectly simulate 𝐏\mathbf{P} using rr noiseless messages from Alice to Bob, along with any auxiliary resource. For a given class of resource, the smallest number rr needed to accomplish this simulation is called the communication cost of 𝐏\mathbf{P} (also referred to as the signaling dimension of 𝐏\mathbf{P} in Refs. [16, 17]). It turns out that cv(𝐏)\lceil\text{cv}(\mathbf{P})\rceil is a lower bound on the communication cost when Alice and Bob have access to shared randomness [17]. In fact, this lower bound is tight when the Alice and Bob are allowed to use non-signaling resources [1]. Combining this discussion with the previous paragraph, we thus have two dual interpretations of the communication value: 1ncv(𝐏)\frac{1}{n}\text{cv}(\mathbf{P}) as an optimal guessing probability and cv(𝐏)\text{cv}(\mathbf{P}) as an optimal simulation cost. This is no coincidence since Eq. (1) is the dual formulation of the linear program characterizing the communication cost for perfectly simulating 𝐏\mathbf{P} using classical non-signaling resources (see Section VIII for more details).

The goal of this paper is to understand the communication value of quantum channels. Formally, a quantum channel is described by a completely positive trace-preserving (CPTP) map 𝒩\mathcal{N} mapping density operators ρA\rho^{A} on Hilbert space A\mathcal{H}^{A} to density operators 𝒩(ρ)\mathcal{N}(\rho) on Hilbert space B\mathcal{H}^{B}. Every quantum channel is able to generate a family of classical channels by encoding classical data into quantum objects. Namely, for each x[n]x\in[n], Alice prepares a quantum state ρx\rho_{x} and sends it through the channel to Bob’s side. Upon receiving 𝒩(ρx)\mathcal{N}(\rho_{x}), Bob performs a quantum measurement, described by a general positive operator-valued measure (POVM) {Πy}y[n]\{\Pi_{y}\}_{y\in[n^{\prime}]}, and regards his measurement outcome as the decoded classical data. The induced classical channel then has the form

P(y|x)=Tr[Πy𝒩(ρx)].P(y|x)=\textrm{Tr}[\Pi_{y}\mathcal{N}(\rho_{x})]. (2)

How noisy this channel will be depends on the state encoding {ρx}x[n]\{\rho_{x}\}_{x\in[n]} and measurement decoding {Πy}y[n]\{\Pi_{y}\}_{y\in[n^{\prime}]}, and ideally one chooses the states and measurement to minimize the error in data transmission. We define the cv of 𝒩\mathcal{N} in terms of the classical channels it can generate.

Definition 1.

The [n][n][n]\to[n^{\prime}] communication value (cv) of a quantum channel 𝒩CPTP(AB)\mathcal{N}\in\text{CPTP}(A\to B) is

cvnn(𝒩)=max{Πy}y=1n{ρx}x=1n{cv(𝐏)|P(y|x)=Tr[Πy𝒩(ρx)]},\text{cv}^{n\to n^{\prime}}(\mathcal{N})=\max_{\begin{subarray}{c}\{\Pi_{y}\}_{y=1}^{n^{\prime}}\\ \{\rho_{x}\}_{x=1}^{n}\end{subarray}}\{\text{cv}(\mathbf{P})\;|\;P(y|x)=\textrm{Tr}[\Pi_{y}\mathcal{N}(\rho_{x})]\}, (3)

and the cv value of 𝒩\mathcal{N} is defined as

cv(𝒩)=supn,ncvnn(𝒩).\text{cv}(\mathcal{N})=\sup_{n,n\in\mathbb{N}}\text{cv}^{n\to n^{\prime}}(\mathcal{N}). (4)

Analogous to the classical case, 1ncvnn(𝒩)\frac{1}{n}\text{cv}^{n\to n^{\prime}}(\mathcal{N}) quantifies the largest success probability attainable in an nn-input guessing game using the channel 𝒩\mathcal{N}. The quantity cv(𝒩)\text{cv}(\mathcal{N}) also has a dual interpretation as the classical communication cost for simulating any classical channel generated by 𝒩\mathcal{N} when Alice and Bob have access to non-signaling resources.

By taking multiple copies of the channel, one can consider the cv capacity, defined as

𝒞𝒱(𝐏)\displaystyle\mathcal{CV}(\mathbf{P}) =limk1klogcv(𝐏k),\displaystyle=\lim_{k\to\infty}\frac{1}{k}\log\text{cv}(\mathbf{P}^{k}),
𝒞𝒱(𝒩)\displaystyle\mathcal{CV}(\mathcal{N}) =limk1klogcv(𝒩k)\displaystyle=\lim_{k\to\infty}\frac{1}{k}\log\text{cv}(\mathcal{N}^{\otimes k}) (5)

in the classical and quantum cases, respectively. It is not difficult to see that logcv(𝐏)\log\text{cv}(\mathbf{P}) is an additive quantity and so 𝒞𝒱(𝐏)=logcv(𝐏)\mathcal{CV}(\mathbf{P})=\log\text{cv}(\mathbf{P}). On the other hand, as we show below, logcv(𝒩)\log\text{cv}(\mathcal{N}) is non-additive in general for quantum channels. A primary objective of this paper is to understand when additivity of logcv(𝒩)\log\text{cv}(\mathcal{N}) (equivalently multiplicativity of cv(𝒩)\text{cv}(\mathcal{N})) holds and when it does not. One of our main results is that multiplicativity always holds for qubit channels, whereas it does not for qutrits.

One can also study the cv of channels that are enhanced by auxiliary resources shared between the sender and receiver. In the quantum setting, it is natural to consider entanglement-assisted channel communication, as depicted in Fig. 2. This is precisely the setup in the well-known quantum superdense coding protocol [18]. Letting cv(𝒩)\text{cv}^{*}(\mathcal{N}) denote the entangled-assisted cv of 𝒩\mathcal{N}, another main result of ours is that cv(𝒩)\text{cv}^{*}(\mathcal{N}) equals the conditional min-entropy [19] of the Choi matrix of 𝒩\mathcal{N}. Combining with the results of Duan and Winter [2], this implies that cv(𝒩)\lceil\text{cv}^{*}(\mathcal{N})\rceil captures the zero error classical communication cost for simulating 𝒩\mathcal{N} when Alice and Bob have quantum non-signaling resources. Note that by additivity of the min-entropy, it follows that logcv(𝒩)\log\text{cv}^{*}(\mathcal{N}) is additive and therefore 𝒞𝒱(𝒩)=logcv(𝒩)\mathcal{CV}^{*}(\mathcal{N})=\log\text{cv}^{*}(\mathcal{N}). A summary between channel cv and zero-error channel simulation is given in Table I.

cv(𝒩)\lceil\text{cv}(\mathcal{N})\rceil Classical communication cost to perfectly simulate every classical channel induced by 𝒩\mathcal{N} using classical non-signaling resource.
cv(𝒩)\lceil\text{cv}^{*}(\mathcal{N})\rceil Classical communication cost to perfectly simulate 𝒩\mathcal{N} using quantum non-signaling resource.

This paper is structured as follows. We begin in Section II by introducing the notation used in this manuscript and reviewing some preliminary concepts. Section III takes a deeper dive into the definition of channel communication value and relates it to the geometric measure of entanglement and other information-theoretic quantities such as the conditional min-entropy. Section IV-A focuses on qubit channels and provides an analytic expression for cv in terms of the correlation matrix of the channel’s Choi matrix. The Werner-Holveo family of channels is introduced in Section IV-B and the cv is computed. The question of cv multiplicativity is taken up in Section V with examples of both multiplicativity and non-multiplicativity being presented. Notably, the cv capacity is shown to take a single-letter form for entanglement-breaking channels, Pauli qubit channels, and the general depolarizing channel. Section VI introduces the notion of entanglement-assisted communication value and relates it to the conditional min-entropy of the Choi matrix. Different relations to the communication value are considered in Section VII, with a particular focus on the PPT communication value and computable examples that it supports. In Section VII-B we describe a procedure for numerically estimating the cv of a given channel, and we provide a link to our developed software package, which performs this estimation. Finally, Section VIII provides a discussion of our results as they relate to channel capacity and zero error channel simulation.

II Notation and Preliminaries

This paper considers exclusively finite-dimensional quantum systems represented by Hilbert spaces A,B,\mathcal{H}^{A},\mathcal{H}^{B},\cdots etc. The collection of positive operators acting on Hilbert space A\mathcal{H}^{A} will be denoted by Pos(A)\mathrm{Pos}(A), which consists of all hermitian operators Herm(A)\text{Herm}(A) acting on AA with a non-negative eigenvalue spectrum. The subset of these operators having unit trace constitute the collection of density operators for system AA, and we denote this set by 𝒟(A)\mathcal{D}(A). We write \|\cdot\|_{\infty} and 1\|\cdot\|_{1} to indicate the spectral and trace norms of elements in Pos(A)\mathrm{Pos}(A), respectively. For bipartite systems, an operator ΩPos(AB)\Omega\in\mathrm{Pos}(AB) is called separable if it can be expressed as a positive combination of product states, Ω=iti|αiαi||βiβi|\Omega=\sum_{i}t_{i}|\alpha_{i}\rangle\langle\alpha_{i}|\otimes|\beta_{i}\rangle\langle\beta_{i}|, with ti0t_{i}\geq 0, |αiαi|𝒟(A)|\alpha_{i}\rangle\langle\alpha_{i}|\in\mathcal{D}(A), and |βiβi|𝒟(B)|\beta_{i}\rangle\langle\beta_{i}|\in\mathcal{D}(B), and we let SEP(A:B)\text{SEP}(A:B) denote the set of all separable operators on systems AA and BB. Classical systems can be incorporated into this framework by demanding that the density matrix of every classical state be diagonal in a fixed basis. In general, we will label a classical system by XX or YY.

Quantum channels provide the basic building blocks of any dynamical system. Mathematically, they are represented by CPTP maps, and we denote the set of CPTP maps from system AA to BB by CPTP(AB)\text{CPTP}(A\to B). The set CPTP(AB)\text{CPTP}(A\to B) is isomorphic to the subset of Pos(AB)\mathrm{Pos}(AB) consisting of operators whose reduced density operator on system AA is the identity. Specifically, for every 𝒩CPTP(AB)\mathcal{N}\in\text{CPTP}(A\to B) its Choi matrix is the associated operator J𝒩Pos(AB)J_{\mathcal{N}}\in\mathrm{Pos}(AB) given by

J𝒩=id𝒩(ϕdA+),J_{\mathcal{N}}=\textrm{id}\otimes\mathcal{N}(\phi^{+}_{d_{A}}),

where id is the identity map and ϕdA+=|ϕdA+ϕdA+|=i,j=1dA|iijj|\phi^{+}_{d_{A}}=|\phi^{+}_{d_{A}}\rangle\langle\phi^{+}_{d_{A}}|=\sum_{i,j=1}^{d_{A}}|ii\rangle\langle jj|. Note that |ϕdA+|\phi^{+}_{d_{A}}\rangle is proportional to the normalized dAd_{A}-dimensional maximally entangled state, and we write the latter as |ΦdA+:=1dA|ϕdA+|\Phi^{+}_{d_{A}}\rangle:=\frac{1}{\sqrt{d_{A}}}|\phi^{+}_{d_{A}}\rangle. The fact that 𝒩\mathcal{N} is completely positive assures that J𝒩0J_{\mathcal{N}}\geq 0, and the trace-preserving condition means that TrBJ𝒩=𝕀A\textrm{Tr}_{B}J_{\mathcal{N}}=\mathbb{I}^{A}, where 𝕀\mathbb{I} is the identity operator. On the other hand, if TrAJ𝒩=𝕀B\textrm{Tr}_{A}J_{\mathcal{N}}=\mathbb{I}^{B} then 𝒩\mathcal{N} is a unital map, meaning that 𝒩(𝕀A)=𝕀B\mathcal{N}(\mathbb{I}^{A})=\mathbb{I}^{B}. More generally, we say a map is sub-unital if 𝒩(𝕀A)𝕀B\mathcal{N}(\mathbb{I}^{A})\leq\mathbb{I}^{B}.

An important subclass of channels are known as entanglement-breaking (EB). These are characterized by the property that 𝒩ABidC(ρAC)SEP(B:C)\mathcal{N}^{A\to B}\otimes\textrm{id}^{C}(\rho^{AC})\in\text{SEP}(B:C) for all ρACPos(AC)\rho^{AC}\in\mathrm{Pos}(AC). It is not difficult to see that 𝒩CPTP(AB)\mathcal{N}\in\text{CPTP}(A\to B) is EB if and only if J𝒩SEP(A:B)J_{\mathcal{N}}\in\text{SEP}(A:B). For any subset SHerm(A)S\subset\text{Herm}(A), we let

S={ωHerm(A)|ω,τ:=Tr[ωτ]0τS}S^{*}=\{\omega\in\text{Herm}(A)\;|\;\langle\omega,\tau\rangle:=\textrm{Tr}[\omega\tau]\geq 0\;\;\forall\tau\in S\}

denote the dual cone of SS. As a final bit of notation, we write exp(x)\exp(x) and log(x)\log(x) to mean 2x2^{x} and log2x\log_{2}x, respectively.

III Characterizing the Communication Value

Let us begin with a few remarks regarding Definition 1. First, every choice of optimal POVM {Πy}y=1n\{\Pi_{y}\}_{y=1}^{n^{\prime}} and set of signal states {ρx}x=1n\{\rho_{x}\}_{x=1}^{n} is characterized by a labeling function f:[n][n]f:[n^{\prime}]\to[n] such that maxx[n]Tr[Πyρx]=Tr[Πyρf(x)]\max_{x\in[n]}\textrm{Tr}[\Pi_{y}\rho_{x}]=\textrm{Tr}[\Pi_{y}\rho_{f(x)}]. If the range of ff is strictly contained in [n][n], then we can replace the set {ρx}x=1n\{\rho_{x}\}_{x=1}^{n} with a smaller set of signal states such that the map ff is surjective. Similarly, if ff is not one-to-one, then we can coarse-grain the nn^{\prime} POVM elements so that each outcome uniquely identifies a signal state. Hence without changing the cv, we can assume f:[m][m]f:[m]\to[m] is a bijection with m=min{n,n}m=\min\{n,n^{\prime}\}, and so

cvnn(𝒩)=cvmm(𝒩)form=min{n,n}.\text{cv}^{n\to n^{\prime}}(\mathcal{N})=\text{cv}^{m\to m}(\mathcal{N})\qquad\text{for}\qquad m=\min\{n,n^{\prime}\}. (6)

Another observation is that

cvmm(𝒩)cvmm(𝒩)formm,\text{cv}^{m\to m}(\mathcal{N})\leq\text{cv}^{m^{\prime}\to m^{\prime}}(\mathcal{N})\qquad\text{for}\qquad m\leq m^{\prime}, (7)

which follows from the fact that we can always trivially split a POVM to increase outcomes Π12Π+12Π\Pi\to\frac{1}{2}\Pi+\frac{1}{2}\Pi, and we can always increase the size of our input set {ρx}x=1m\{\rho_{x}\}_{x=1}^{m} by adding the same state ρx\rho_{x} multiple times. Finally, we note that

cv(𝒩)=cvdB2dB2(𝒩),\text{cv}(\mathcal{N})=\text{cv}^{d_{B}^{2}\to d_{B}^{2}}(\mathcal{N}), (8)

where dBd_{B} is the dimension of the output system. This follows from the fact that any POVM on a dBd_{B}-dimensional systems can always be decomposed into a convex combination of extremal POVMs, each with at most dB2d_{B}^{2} outcomes [20], and the cv can always be attained with one of these extremal measurements.

It is also not difficult to see that

1cv(𝒩)min{dA,dB}\displaystyle 1\leq\text{cv}(\mathcal{N})\leq\min\{d_{A},d_{B}\} (9)

for any channel 𝒩\mathcal{N}. The lower bound holds by considering a constant input ρx=ρ\rho_{x}=\rho (for all xx) so that xTr[Πx𝒩(ρx)]=xTr[Πx𝒩(ρ)]=1\sum_{x}\textrm{Tr}[\Pi_{x}\mathcal{N}(\rho_{x})]=\sum_{x}\textrm{Tr}[\Pi_{x}\mathcal{N}(\rho)]=1, since 𝒩\mathcal{N} is trace preserving and xΠx=𝕀dB\sum_{x}\Pi_{x}=\mathbb{I}_{d_{B}}. Similarly, the upper bound follows from the inequalities

xTr[Πx𝒩(ρx)]\displaystyle\sum_{x}\textrm{Tr}[\Pi_{x}\mathcal{N}(\rho_{x})] xTr[Πx]=Tr[𝕀dB]=dB;\displaystyle\leq\sum_{x}\textrm{Tr}[\Pi_{x}]=\textrm{Tr}[\mathbb{I}_{d_{B}}]=d_{B}; (10)
xTr[Πx𝒩(ρx)]\displaystyle\sum_{x}\textrm{Tr}[\Pi_{x}\mathcal{N}(\rho_{x})] =xTr[𝒩(Πx)ρx]\displaystyle=\sum_{x}\textrm{Tr}[\mathcal{N}^{\dagger}(\Pi_{x})\rho_{x}]
xTr[𝒩(Πx)]=Tr[𝕀dA]=dA,\displaystyle\leq\sum_{x}\textrm{Tr}[\mathcal{N}^{\dagger}(\Pi_{x})]=\textrm{Tr}[\mathbb{I}_{d_{A}}]=d_{A}, (11)

where 𝒩:BA\mathcal{N}^{\dagger}:B\to A is the adjoint map of 𝒩\mathcal{N}, and so {𝒩(Πx)]}x\{\mathcal{N}^{\dagger}(\Pi_{x})]\}_{x} will always be a valid POVM on Alice’s system.

Notice that when cv(𝒩)=1\text{cv}(\mathcal{N})=1, Bob can do no better than randomly guessing Alice’s input. The following proposition characterizes the type of channel for which this is the case.

Proposition 1.

cv(𝒩)=1\text{cv}(\mathcal{N})=1 iff 𝒩\mathcal{N} is a replacer channel; i.e. there exists a fixed state σ\sigma such that 𝒩(ρ)=σ\mathcal{N}(\rho)=\sigma for all states ρ\rho.

Proof.

If 𝒩\mathcal{N} is a replacer channel, then clearly cv(𝒩)=1\text{cv}(\mathcal{N})=1. On the other hand, suppose that 𝒩\mathcal{N} is not a replacer channel. This means there exists two inputs ρ1\rho_{1} and ρ2\rho_{2} such that Δ=𝒩(ρ1)𝒩(ρ2)0\Delta=\mathcal{N}(\rho_{1})-\mathcal{N}(\rho_{2})\not=0. Hence by performing a Helstrom measurement on the channel output (i.e. projecting onto the ±\pm parts of Δ\Delta) [21], one obtains cv(𝒩)1+12Δ1>1\text{cv}(\mathcal{N})\geq 1+\tfrac{1}{2}\|\Delta\|_{1}>1. ∎

III-A Communication Value via Conic Optimization

In general cv(𝒩)\text{cv}(\mathcal{N}) is difficult to compute. This can be seen more explicitly by casting cv(𝒩)\text{cv}(\mathcal{N}) as an optimization over the separable cone, SEP(A:B)\text{SEP}(A:B), whose membership is NP-Hard to decide [22]. Nevertheless, expressing cv(𝒩)\text{cv}(\mathcal{N}) as an optimization over SEP(A:B)\text{SEP}(A:B) leads to computable upper bounds since there are well-known relaxations to the set SEP(A:B)\text{SEP}(A:B) that are easier to handle analytically.

Proposition 2.

For 𝒩CPTP(AB)\mathcal{N}\in\text{CPTP}(A\to B),

cv(𝒩)=max\displaystyle\text{cv}(\mathcal{N})=\max Tr[ΩABJ𝒩]\displaystyle\;\;\textrm{Tr}[\Omega^{AB}J_{\mathcal{N}}]
subject to TrA[ΩAB]=𝕀B;\displaystyle\;\;\textrm{Tr}_{A}[\Omega^{AB}]=\mathbb{I}^{B};
ΩABSEP(A:B).\displaystyle\;\;\Omega^{AB}\in\text{SEP}(A:B). (12)
Proof.

By Eq. (8) we have

cv(𝒩)\displaystyle\text{cv}(\mathcal{N}) =max{Πx},{ρx}x=1dB2Tr[Πx𝒩(ρx)]\displaystyle=\max_{\{\Pi_{x}\},\{\rho_{x}\}}\sum_{x=1}^{d_{B}^{2}}\textrm{Tr}[\Pi_{x}\mathcal{N}(\rho_{x})]
=max{Πx},{ρx}x=1dB2Tr[ρxTΠx(J𝒩)]\displaystyle=\max_{\{\Pi_{x}\},\{\rho_{x}\}}\sum_{x=1}^{d_{B}^{2}}\textrm{Tr}[\rho_{x}^{T}\otimes\Pi_{x}(J_{\mathcal{N}})]
=maxTr[ΩABJ𝒩],\displaystyle=\max\textrm{Tr}[\Omega^{AB}J_{\mathcal{N}}], (13)

where ΩAB=x=1dB2ρxTΠx\Omega^{AB}=\sum_{x=1}^{d_{B}^{2}}\rho_{x}^{T}\otimes\Pi_{x} satisfies the conditions of Eq. (12). Conversely, any ΩABSEP(A:B)\Omega^{AB}\in\text{SEP}(A:B) can be written as ΩAB=x|ψxψx|AωxB\Omega^{AB}=\sum_{x}|\psi_{x}\rangle\langle\psi_{x}|^{A}\otimes\omega_{x}^{B} with |ψx|\psi_{x}\rangle being a pure state. The condition TrA[ΩAB]=𝕀B\textrm{Tr}_{A}[\Omega^{AB}]=\mathbb{I}^{B} implies that {ωx}x\{\omega_{x}\}_{x} constitutes a POVM. ∎

Note that strong duality holds for the conic program here, and so Proposition 2 can be cast in dual form as

cv(𝒩)=min\displaystyle\text{cv}(\mathcal{N})=\min Tr[ZB]\displaystyle\;\;\textrm{Tr}[Z^{B}]
subject to 𝕀AZBJ𝒩ABSEP(A:B).\displaystyle\;\;\mathbb{I}^{A}\otimes Z^{B}-J_{\mathcal{N}}^{AB}\in\text{SEP}^{*}(A:B). (14)

In Section VIII, we will explore different relaxations to this problem by considering outer approximations of SEP(A:B)\text{SEP}(A:B). For example, the cone PPT(A:B)\text{PPT}(A:B), which consists of all bipartite positive operators having a positive partial transpose, contains SEP(A:B)\text{SEP}(A:B) [23]. Replacing SEP(A:B)\text{SEP}(A:B) with PPT(A:B)\text{PPT}(A:B) in Eq. (12) gives us a semi-definite program (SDP). Furthermore, since SEP(A:B)=PPT(A:B)\text{SEP}(A:B)=\text{PPT}(A:B) whenever dAdB6d_{A}d_{B}\leq 6 [24], we thus obtain the following.

Corollary 1.

Suppose 𝒩CPTP(AB)\mathcal{N}\in\text{CPTP}(A\to B) with dAdB6d_{A}d_{B}\leq 6. Then

cv(𝒩)=max\displaystyle\text{cv}(\mathcal{N})=\max Tr[ΩABJ𝒩]\displaystyle\;\;\textrm{Tr}[\Omega^{AB}J_{\mathcal{N}}]
subject to TrA[ΩAB]=𝕀B\displaystyle\;\;\textrm{Tr}_{A}[\Omega^{AB}]=\mathbb{I}^{B}
ΩABPPT(A:B).\displaystyle\;\;\Omega^{AB}\in\text{PPT}(A:B). (15)

III-B An Entropic Characterization of Communication Value

III-B1 The Conditional Separable Min-Entropy

An alternative but related manner of characterizing the communication value is in terms of the min-entropy or variations of it. Equation (14) might strike the reader as closely resembling the conditional min-entropy of J𝒩J_{\mathcal{N}}. Recall that the conditional min-entropy of a positive bipartite operator ωAB\omega^{AB} is given by

Hmin(A|B)ω=minσB𝒟(B)Dmax(ω𝕀AσB),\displaystyle H_{\min}(A|B)_{\omega}=-\min_{\sigma^{B}\in\mathcal{D}(B)}D_{\max}(\omega\|\mathbb{I}^{A}\otimes\sigma^{B}), (16)

where Dmax(μν)=min{λ|μ2λν}D_{\max}(\mu\|\nu)=\min\{\lambda\;|\;\mu\leq 2^{\lambda}\nu\} [19]. Here \leq denotes a generalized inequality over the convex cone of positive-semidefinite operators; i.e. XYX\leq Y iff YXPos(AB)Y-X\in\mathrm{Pos}(AB). Equivalently, we can combine the two minimizations in the definition of HminH_{\min} to write

exp[Hmin(A|B)ω]=min\displaystyle\exp[-H_{\min}(A|B)_{\omega}]=\min Tr[ZB]\displaystyle\;\;\textrm{Tr}[Z^{B}]
subject to 𝕀AZBωABPos(AB).\displaystyle\;\;\mathbb{I}^{A}\otimes Z^{B}-\omega^{AB}\in\mathrm{Pos}(AB). (17)

Comparing with Eq. (14), we see that cv is recovered by changing the cone from Pos(AB)\mathrm{Pos}(AB) to SEP(A:B)\text{SEP}^{*}(A:B). Let us denote the cone inequality over SEP(A:B)\text{SEP}^{*}(A:B) by SEP\leq_{\text{SEP}^{*}} such that XSEPYX\leq_{\text{SEP}^{*}}\!Y iff YXSEP(A:B)Y-X\in\text{SEP}^{*}(A:B). Then we can introduce a restricted conditional min-entropy.

Definition 2.

The conditional separable min-entropy of a positive bipartite operator ωAB\omega^{AB} is defined as

Hminsep(A|B)ω=minσB𝒟(B)Dmaxsep(ω𝕀AσB),H_{\min}^{\text{sep}}(A|B)_{\omega}=-\min_{\sigma^{B}\in\mathcal{D}(B)}D^{\text{sep}}_{\max}(\omega\|\mathbb{I}^{A}\otimes\sigma^{B}), (18)

where Dmaxsep(μν)=min{λ|μSEP2λν}D^{\text{sep}}_{\max}(\mu\|\nu)=\min\{\lambda\;|\;\mu\leq_{\text{SEP}^{*}}\!2^{\lambda}\nu\}.

By Eq. (14), we therefore have

cv(𝒩)=exp[Hminsep(A|B)J𝒩].\text{cv}(\mathcal{N})=\exp[-H^{\text{sep}}_{\min}(A|B)_{J_{\mathcal{N}}}]. (19)

The separable min-entropy enjoys a data-processing inequality under one-way LOCC from Bob to Alice. The latter consists of any bipartite map ΦCPTP(ABAB)\Phi\in\text{CPTP}(AB\to A^{\prime}B^{\prime}) having the form Φ=i𝒩ii\Phi=\sum_{i}\mathcal{N}_{i}\otimes\mathcal{M}_{i}, where 𝒩iCPTP(AA)\mathcal{N}_{i}\in\text{CPTP}(A\to A^{\prime}) and iiCPTP(BB)\sum_{i}\mathcal{M}_{i}\in\text{CPTP}(B\to B^{\prime}) with each individual i\mathcal{M}_{i} being CP. In fact we can prove the data-processing inequality under an even larger class of operations.

Proposition 3.

Let Φ:Pos(AB)Pos(AB)\Phi:\mathrm{Pos}(AB)\to\mathrm{Pos}(A^{\prime}B^{\prime}) be any positive map whose adjoint is non-entangling (i.e. Φ:SEP(A:B)SEP(A:B)\Phi^{\dagger}:\text{SEP}(A^{\prime}:B^{\prime})\to\text{SEP}(A:B)) and that further satisfies Φ(𝕀AσB)𝕀Aϕ(σB)\Phi(\mathbb{I}^{A}\otimes\sigma^{B^{\prime}})\leq\mathbb{I}^{A^{\prime}}\otimes\phi(\sigma^{B^{\prime}}) for some trace-preserving map ϕ:Pos(B)Pos(B)\phi:\mathrm{Pos}(B)\to\mathrm{Pos}(B^{\prime}). Then Hminsep(A|B)PHminsep(A|B)Φ(P)H^{\text{sep}}_{\min}(A|B)_{P}\leq H^{\text{sep}}_{\min}(A|B)_{\Phi(P)} for all PPos(AB)P\in\mathrm{Pos}(AB).

Proof.

Since Φ\Phi^{\dagger} preserves separability, we must have Φ(Q)SEP(A:B)\Phi(Q)\in\text{SEP}^{*}(A^{\prime}:B^{\prime}) for all QSEP(A:B)Q\in\text{SEP}^{*}(A:B). Hence,

2λ𝕀σPSEP0\displaystyle 2^{\lambda}\mathbb{I}\otimes\sigma-P\geq_{\text{SEP}^{*}}0\;  2λΦ(𝕀σ)Φ(P)SEP0\displaystyle\Rightarrow\;2^{\lambda}\Phi(\mathbb{I}\otimes\sigma)-\Phi(P)\geq_{\text{SEP}^{*}}0
 2λ𝕀ϕ(σ)Φ(P)SEP0.\displaystyle\Rightarrow\;2^{\lambda}\mathbb{I}\otimes\phi(\sigma)-\Phi(P)\geq_{\text{SEP}^{*}}0.

In other words, any feasible pair (σ,λ)(\sigma,\lambda) in the minimization of Hminsep(A|B)PH_{\min}^{\text{sep}}(A|B)_{P} also leads to a feasible pair for Hminsep(A|B)Φ(P)H_{\min}^{\text{sep}}(A|B)_{\Phi(P)}. The 1-1 factor in the definition of HminsepH_{\min}^{\text{sep}} then implies the proposition. ∎

The maps of Proposition 3 include those of the form Φ=𝒩\Phi=\mathcal{M}\otimes\mathcal{N}, where \mathcal{M} is sub-unital and 𝒩\mathcal{N} is CPTP. These maps are known to satisfy the data-processing inequality for the standard min-entropy [25]. However we suspect that Proposition 3 includes maps for which the standard min-entropy data-processing inequality does not hold.

We can apply Proposition 3 to the processing of Choi matrices. However, in this case not all maps Φ\Phi satisfying the conditions of Proposition 3 are physically meaningful. Specifically, we require the additional condition that TrBΦ(P)=𝕀A\textrm{Tr}_{B^{\prime}}\Phi(P)=\mathbb{I}^{A^{\prime}} for all operators PP in which TrBP=𝕀A\textrm{Tr}_{B}P=\mathbb{I}^{A}. This assures that Φ\Phi maps Choi matrices to Choi matrices. One particular class of maps having this form are those in which Φ\Phi is a product of a positive unital map and a CPTP map, i.e. Φ=prepost\Phi=\mathcal{E}_{\text{pre}}^{\dagger}\otimes\mathcal{E}_{\text{post}}. In this case, pre\mathcal{E}_{\text{pre}} and post\mathcal{E}_{\text{post}} are pre- and post-processing maps for a given channel, respectively. As a consequence of Proposition 3 we therefore observe the following corollary, which can also be seen directly from the definition of communication value.

Corollary 2.

Communication value is non-increasing under pre- and post-processing of the channel.

Note that for classical systems XX and YY, we have Pos(XB)=SEP(X:B)\mathrm{Pos}(XB)=\text{SEP}(X:B) and Pos(AY)=SEP(A:Y)\mathrm{Pos}(AY)=\text{SEP}(A:Y). These correspond to classical-to-quantum and quantum-to-classical channels, respectively, and in these cases Eq. (19) reduces to

cv(𝒩XB)\displaystyle\text{cv}(\mathcal{N}^{X\to B}) =exp[Hmin(X|B)J𝒩]\displaystyle=\exp[-H_{\min}(X|B)_{J_{\mathcal{N}}}] (20)
cv(𝒩AY)\displaystyle\text{cv}(\mathcal{N}^{A\to Y}) =exp[Hmin(A|Y)J𝒩].\displaystyle=\exp[-H_{\min}(A|Y)_{J_{\mathcal{N}}}]. (21)

III-B2 The max-Holevo Information

The cv can be further related to the max-Holevo information of a channel, χmax(𝒩)\chi_{\max}(\mathcal{N}). This quantity has been introduced in the study of “sandwiched” Rényi divergences [26, 27] and is defined as

χmax(𝒩)=maxρXAminσBDmax(ρXB||ρXσB),\displaystyle\chi_{\max}(\mathcal{N})=\max_{\rho^{XA}}\min_{\sigma^{B}}D_{\max}(\rho^{XB}||\rho^{X}\otimes\sigma^{B}),

where the maximization is taken over all cq states ρXA=xp(x)|xx|ρxA\rho^{XA}=\sum_{x}p(x)|x\rangle\langle x|\otimes\rho_{x}^{A} and

ρXB:=xp(x)|xx|𝒩(ρxA).\rho^{XB}:=\sum_{x}p(x)|x\rangle\langle x|\otimes\mathcal{N}(\rho_{x}^{A}).

In fact, since DmaxD_{\max} is quasi-convex (i.e. Dmax(ip(i)ρiip(i)σi)maxiDmax(ρiσi)D_{\max}(\sum_{i}p(i)\rho_{i}\|\sum_{i}p(i)\sigma_{i})\leq\max_{i}D_{\max}(\rho_{i}\|\sigma_{i}) [28]), if follows that we can restrict attention to pure ρxA=|ψxψx|A\rho_{x}^{A}=|\psi_{x}\rangle\langle\psi_{x}|^{A} in the definition of χmax\chi_{\max}. Letting UU be the unitary such that U|x=|ψxU|x\rangle=|\psi_{x}\rangle, the maximization over ρXA\rho^{XA} can then be replaced by a maximization over UU such that

ρXA=(𝕀U)xp(x)|xxxx|(𝕀U).\rho^{XA}=(\mathbb{I}\otimes U)\sum_{x}p(x)|xx\rangle\langle xx|(\mathbb{I}\otimes U)^{\dagger}. (22)

We use this simplification to prove a relationship between channel χmax\chi_{\max} and conditional HminsepH^{\text{sep}}_{\min}.

Theorem 1.

For any channel 𝒩AB\mathcal{N}^{A\to B},

χmax(𝒩)=logcv(𝒩).\chi_{\max}(\mathcal{N})=\log\text{cv}(\mathcal{N}). (23)
Proof.

Using Eq. (22) we have

ρXB=xp(x)|xx|ρx,\rho^{XB}=\sum_{x}p(x)|x\rangle\langle x|\otimes\rho_{x},

where ρx=𝒩(U|xx|U)\rho_{x}=\mathcal{N}(U|x\rangle\langle x|U^{\dagger}). Since ρX=xp(x)|xx|\rho^{X}=\sum_{x}p(x)|x\rangle\langle x|, the definition of DmaxD_{\max} yields

Dmax(ρXBρXσB)\displaystyle D_{\max}(\rho^{XB}\|\rho^{X}\otimes\sigma^{B}) =min{λ|ρx2λσ,x}\displaystyle=\min\{\lambda\;|\;\rho_{x}\leq 2^{\lambda}\sigma,\;\forall x\}
=Dmax(ρ~XB𝕀σB)\displaystyle=D_{\max}(\widetilde{\rho}^{XB}\|\mathbb{I}\otimes\sigma^{B})
=Dmaxsep(ρ~XB𝕀σB),\displaystyle=D_{\max}^{\text{sep}}(\widetilde{\rho}^{XB}\|\mathbb{I}\otimes\sigma^{B}), (24)

where

ρ~XB\displaystyle\widetilde{\rho}^{XB} =x|xx|𝒩(U|xx|U)\displaystyle=\sum_{x}|x\rangle\langle x|\otimes\mathcal{N}(U|x\rangle\langle x|U^{\dagger})
=ΔUTidB(J𝒩),\displaystyle=\Delta_{U^{T}}\otimes\textrm{id}^{B}(J_{\mathcal{N}}), (25)

and ΔUT(τ)=x|xx|UT(τ)U|xx|\Delta_{U^{T}}(\tau)=\sum_{x}|x\rangle\langle x|U^{T}(\tau)U^{*}|x\rangle\langle x| is a completely dephasing map after applying the rotation UTU^{T}. The last equality in Eq. (24) follows from the fact that Pos(XB)=SEP(X:B)\mathrm{Pos}(XB)=\text{SEP}(X:B), as noted above. Then by data-processing (Proposition 3), we have

Dmaxsep(ΔUTidB(J𝒩)𝕀σB)Dmaxsep(J𝒩𝕀σB)\displaystyle D_{\max}^{\text{sep}}(\Delta_{U^{T}}\otimes\textrm{id}^{B}(J_{\mathcal{N}})\|\mathbb{I}\otimes\sigma^{B})\leq D_{\max}^{\text{sep}}(J_{\mathcal{N}}\|\mathbb{I}\otimes\sigma^{B})

for any σB\sigma^{B} and unitary UU on system AA. Hence from the definitions it follows that

χmax(𝒩)Hminsep(A|B)J𝒩=logcv(𝒩).\chi_{\max}(\mathcal{N})\leq-H_{\min}^{\text{sep}}(A|B)_{J_{\mathcal{N}}}=\log\text{cv}(\mathcal{N}). (26)

To prove the reverse inequality, for arbitrary σB\sigma^{B} let λ0=Dmaxsep(J𝒩𝕀σB)\lambda_{0}=D_{\max}^{\text{sep}}(J_{\mathcal{N}}\|\mathbb{I}\otimes\sigma^{B}). Hence

2λ0𝕀σBJ𝒩SEP(A:B),2^{\lambda_{0}}\mathbb{I}\otimes\sigma^{B}-J_{\mathcal{N}}\in\text{SEP}^{*}(A:B), (27)

and since λ0\lambda_{0} is a minimizer, there must exist some product state |α|β|\alpha\rangle|\beta\rangle such that

2λ0β|σB|β=β|𝒩(|αα|)|β.2^{\lambda_{0}}\langle\beta|\sigma^{B}|\beta\rangle=\langle\beta|\mathcal{N}(|\alpha^{*}\rangle\langle\alpha^{*}|)|\beta\rangle. (28)

Let UU be any unitary that rotates {|x}x=1dA\{|x\rangle\}_{x=1}^{d_{A}} such that U|1=|αU|1\rangle=|\alpha^{*}\rangle. Therefore

2λ0α,β|𝕀σB|α,β=α,β|ΔUTidB(J𝒩)|α,β,\displaystyle 2^{\lambda_{0}}\langle\alpha,\beta|\mathbb{I}\otimes\sigma^{B}|\alpha,\beta\rangle=\langle\alpha,\beta|\Delta_{U^{T}}\otimes\textrm{id}^{B}(J_{\mathcal{N}})|\alpha,\beta\rangle, (29)

which means that Dmaxsep(ΔUTidB(J𝒩)𝕀σB)D_{\max}^{\text{sep}}(\Delta_{U^{T}}\otimes\textrm{id}^{B}(J_{\mathcal{N}})\|\mathbb{I}\otimes\sigma^{B}) can be no less than λ0=Dmaxsep(J𝒩𝕀σB)\lambda_{0}=D_{\max}^{\text{sep}}(J_{\mathcal{N}}\|\mathbb{I}\otimes\sigma^{B}). Since this holds for all σB\sigma^{B} and we are maximizing over UU, we have

χmax(𝒩)Hminsep(A|B)J𝒩=logcv(𝒩).\chi_{\max}(\mathcal{N})\geq-H_{\min}^{\text{sep}}(A|B)_{J_{\mathcal{N}}}=\log\text{cv}(\mathcal{N}). (30)

We close this section by providing an alternative proof of Theorem 1. Instead of going through the Choi matrix, the following argument relies on a characterization of cv in terms of maximizing the min-entropy over encodings. In some sense this is intuitive as the communication value is optimizing minimal error discrimination, which min-entropy characterizes [19]. For this reason, the conceptual underpinning of this alternative derivation may be of interest in other applications.

Let {ρxA}\{\rho_{x}^{A}\} denote a subset of states for some alphabet 𝒳\mathcal{X}, ρXA\rho^{XA} be a cq state defined using {ρxA}\{\rho_{x}^{A}\}, ρU\rho_{U} be the maximally mixed state on the relevant space, and ρXB:=(idX𝒩)(ρXA)\rho^{XB}:=(\mathrm{id}^{X}\otimes\mathcal{N})(\rho^{XA}). Starting from (14),

cv(𝒩)=\displaystyle\text{cv}(\mathcal{N})= min{Tr[ZB]:𝕀AZBSEPJ𝒩}\displaystyle\min\{\textrm{Tr}[Z^{B}]:\mathbb{I}^{A}\otimes Z^{B}\geq_{\text{SEP}^{*}}J_{\mathcal{N}}\}
=\displaystyle= sup{ρxA}min{Tr[ZB]:ZB𝒩(ρxA)x𝒳}\displaystyle\sup_{\{\rho_{x}^{A}\}}\min\{\textrm{Tr}[Z^{B}]:Z^{B}\geq\mathcal{N}(\rho_{x}^{A})\,\forall x\in\mathcal{X}\}
=\displaystyle= supρXA:ρX=ρU|𝒳|min{Tr[Z~B]:𝕀AZ~BρXB}\displaystyle\sup_{\begin{subarray}{c}\rho^{XA}:\\ \rho^{X}=\rho_{U}\end{subarray}}|\mathcal{X}|\min\{\textrm{Tr}[\widetilde{Z}^{B}]:\mathbb{I}^{A}\otimes\widetilde{Z}^{B}\geq\rho^{XB}\}
=\displaystyle= supρXA:ρX=ρU|𝒳|exp(Hmin(X|B)ρXB)\displaystyle\sup_{\begin{subarray}{c}\rho^{XA}:\\ \rho^{X}=\rho_{U}\end{subarray}}|\mathcal{X}|\exp(-H_{\min}(X|B)_{\rho^{XB}})
=\displaystyle= sup{ρxA}minσBλmax(x|xx|σ1/2ρxBσ1/2)\displaystyle\sup_{\{\rho_{x}^{A}\}}\min_{\sigma^{B}}\lambda_{\max}(\sum_{x}|x\rangle\langle x|\otimes\sigma^{-1/2}\rho^{B}_{x}\sigma^{-1/2})
=\displaystyle= supρXAminσBexp(Dmax(ρXB||ρXσB))\displaystyle\sup_{\rho^{XA}}\min_{\sigma^{B}}\exp(D_{\max}(\rho^{XB}||\rho^{X}\otimes\sigma^{B}))
=\displaystyle= exp(χmax),\displaystyle\exp(\chi_{\max})\ , (31)

where the second equality is using XABSEPα|β|X|α|βX^{AB}\in\text{SEP}^{*}\Leftrightarrow\langle\alpha|\langle\beta|X|\alpha\rangle|\beta\rangle for all unit vectors α,β\alpha,\beta and the action of the channel in terms of the Choi, the third is by using uniform probability on 𝒳\mathcal{X}, the fourth is by definition of min-entropy, the fifth is using Dmax(ρ||σ)=λmax(σ1/2ρσ1/2)D_{\max}(\rho||\sigma)=\lambda_{\max}(\sigma^{-1/2}\rho\sigma^{-1/2}), the sixth is using that p(x)1/2p(x)p(x)1/2=1p(x)^{-1/2}p(x)p(x)^{-1/2}=1, and the final equality is by definition.

III-C Communication Value in Terms of Singlet Fraction

Another advantage of viewing cv(𝒩)\text{cv}(\mathcal{N}) in terms of a restricted min-entropy is that it provides an alternative operational interpretation of the communication value in terms of the singlet fraction, which it inherits from the min-entropy conic program. Recall that for a bipartite density matrix ωAB\omega^{AB}, its dAd_{A}-dimensional singlet fraction is defined as

FdA+(ω)=maxUΦdA+|(𝕀AUB)ωAB(𝕀AUB)ΦdA+,F^{+}_{d_{A}}(\omega)=\max_{U}\langle\Phi^{+}_{d_{A}}|(\mathbb{I}^{A}\otimes U^{B})\omega^{AB}(\mathbb{I}^{A}\otimes U^{B})^{\dagger}|\Phi^{+}_{d_{A}}\rangle,

where the maximization is taken over all unitaries applied to system BB [29].

Proposition 4.

cv(𝒩)\text{cv}(\mathcal{N}) is the maximum singlet fraction achievable using an entanglement-breaking channel after the action of 𝒩\mathcal{N} on |ΦdA+|\Phi^{+}_{d_{A}}\rangle.

Proof.

This follows the proof of the operational interpretation of the min-entropy [19], and we walk through the argument again here to exemplify that the only change is in restricting to the separable cone. Proposition 2 shows that cv(𝒩)\text{cv}(\mathcal{N}) is the maximum value Tr[ΩABJ𝒩]\textrm{Tr}[\Omega^{AB}J_{\mathcal{N}}], where ΩAB\Omega^{AB} is the Choi matrix of a unital (entanglement-breaking) map, i.e. ΩAB=J\Omega^{AB}=J_{\mathcal{M}} for some entanglement-breaking unital map \mathcal{M}. Thus,

cv(𝒩)\displaystyle\text{cv}(\mathcal{N}) =J,J𝒩\displaystyle=\langle J_{\mathcal{M}},J_{\mathcal{N}}\rangle
=\displaystyle= dA2(id)(Φ^+),(id𝒩)(Φ^+)\displaystyle d_{A}^{2}\langle(\textrm{id}\otimes\mathcal{M})(\widehat{\Phi}^{+}),(\textrm{id}\otimes\mathcal{N})(\widehat{\Phi}^{+})\rangle
=\displaystyle= dA2Φ^+,(id𝒩)(Φ+),\displaystyle d_{A}^{2}\langle\widehat{\Phi}^{+},(\textrm{id}\otimes\mathcal{M}^{\ast}\circ\mathcal{N})(\Phi^{+})\rangle\ ,

where we have used the definitions of the Choi matrix and adjoint map, and Φ^+\widehat{\Phi}^{+} is the normalized maximally entangled state. Noting the adjoint of an entanglement-breaking map is entanglement-breaking, and the adjoint of a unital map is trace-preserving, we have

cv(𝒩)\displaystyle\text{cv}(\mathcal{N}) =dA2maxEB(BA)FdA+(id𝒩(ΦdA+))\displaystyle=d_{A}^{2}\max_{\mathcal{E}\in\text{EB}(B\to A)}F^{+}_{d_{A}}\left(\textrm{id}\otimes\mathcal{E}\circ\mathcal{N}(\Phi_{d_{A}}^{+})\right)
=dA2maxEB(BA)ΦdA+|id𝒩(ΦdA+)ΦdA+,\displaystyle=d_{A}^{2}\max_{\mathcal{E}\in\text{EB}(B\to A)}\langle\Phi^{+}_{d_{A}}|\textrm{id}\otimes\mathcal{E}\circ\mathcal{N}(\Phi_{d_{A}}^{+})|\Phi^{+}_{d_{A}}\rangle, (32)

where the last line follows from the fact that the maximization over local unitaries in the definition of FdA+F^{+}_{d_{A}} can be included in the maximization over entanglement-breaking channels. ∎

|Φ+|\Phi^{+}\rangle𝒩\mathcal{N}Ψ\PsiidVersus|Φ+|\Phi^{+}\rangleidid
Figure 1: The channel value is a measure of the maximum singlet fraction achievable using an entanglement-breaking (EB) channel to recover the singlet after the action of 𝒩\mathcal{N}. In this setting, non-multiplicativity means that the optimal EB channel Ψ\Psi changes when using 𝒩\mathcal{N} in parallel such that the achievable singlet fraction increases. This suggests that cv(𝒩)\text{cv}(\mathcal{N}) is a measurement of entanglement preservation.

Equation (32) yields the interpretation of cv(𝒩)\text{cv}(\mathcal{N}) as how well a maximally entangled state can be recovered after Alice sends one half of |ΦdA+|\Phi^{+}_{d_{A}}\rangle to Bob over the channel 𝒩\mathcal{N}, and he is limited to performing an entanglement-breaking channel as post-processing error correction (see Fig. 1). In Section VI, it will be shown that the entanglement-assisted communication value is characterized by exp(Hmin(A|B)J𝒩)\exp(-H_{\min}(A|B)_{J_{\mathcal{N}}}), and it thus has a similar operational interpretation except Bob is now able to perform an arbitrary quantum channel to try and recover the maximally entangled state. Moreover, in Section VII, when we consider relaxations of cv to other cones, we will find that the PPT min-entropy HminPPTH_{\min}^{\text{PPT}} also retains this operational interpretation, but with recovery being relaxed to the use of co-positive maps.

III-D The Geometric Measure of Entanglement and Maximum Output Purity

The channel cv is closely related to the geometric measure of entanglement (GME) [30, 31]. For a bipartite positive operator ωAB\omega^{AB}, its GME is defined as G(ω)=logΛ2(ω)G(\omega)=-\log\Lambda^{2}(\omega), where

Λ2(ω)=max|α|βα,β|ωα,β.\displaystyle\Lambda^{2}(\omega)=\max_{|\alpha\rangle|\beta\rangle}\langle\alpha,\beta|\omega|\alpha,\beta\rangle. (33)

We can phrase this as the SDP

Λ2(ω)=max\displaystyle\Lambda^{2}(\omega)=\max Tr[σABω]\displaystyle\;\;\textrm{Tr}[\sigma^{AB}\omega]
subject to Tr[σAB]=1\displaystyle\;\;\textrm{Tr}[\sigma^{AB}]=1
σABSEP(A:B).\displaystyle\;\;\sigma^{AB}\in\text{SEP}(A:B). (34)

For a channel 𝒩\mathcal{N} with Choi matrix J𝒩J_{\mathcal{N}}, this SDP can be expressed in dual form as

Λ2(J𝒩)=min\displaystyle\Lambda^{2}(J_{\mathcal{N}})=\min λ\displaystyle\;\;\lambda
subject to λ𝕀ABJ𝒩ABSEP(A:B).\displaystyle\;\;\lambda\mathbb{I}^{AB}-J_{\mathcal{N}}^{AB}\in\text{SEP}^{*}(A:B). (35)

For any dual feasible λ\lambda in Eq. (35), we can take Z=λ𝕀BZ=\lambda\mathbb{I}^{B} in Eq. (14) to obtain

cv(𝒩)dBΛ2(J𝒩).\displaystyle\text{cv}(\mathcal{N})\leq d_{B}\Lambda^{2}(J_{\mathcal{N}}). (36)

On the other hand, suppose that ZZ is dual feasible in Eq. (14). Then since 𝕀A(λ𝕀BZ)SEP(A:B)\mathbb{I}^{A}\otimes(\lambda\mathbb{I}^{B}-Z)\in\text{SEP}^{*}(A:B) for λ=Z\lambda=\|Z\|_{\infty}, we have that

λ𝕀A𝕀BJ𝒩AB\displaystyle\lambda\mathbb{I}^{A}\otimes\mathbb{I}^{B}-J_{\mathcal{N}}^{AB} =𝕀A(λ𝕀Z)B+𝕀AZBJ𝒩AB\displaystyle=\mathbb{I}^{A}\otimes(\lambda\mathbb{I}-Z)^{B}+\mathbb{I}^{A}\otimes Z^{B}-J_{\mathcal{N}}^{AB}
SEP(A:B).\displaystyle\in\text{SEP}^{*}(A:B). (37)

Hence λ𝕀AB\lambda\mathbb{I}^{AB} is dual feasible in Eq. (35). Since λTr[Z]\lambda\leq\textrm{Tr}[Z] for every ZZ, we conclude

Λ2(J𝒩)cv(𝒩)dBΛ2(J𝒩).\displaystyle\Lambda^{2}(J_{\mathcal{N}})\leq\text{cv}(\mathcal{N})\leq d_{B}\Lambda^{2}(J_{\mathcal{N}}). (38)

While these relationships are somewhat obvious from the formulation of the problem, it is nice to see them explicitly falling out of the two conic programs.

It is known that Λ2(J𝒩)\Lambda^{2}(J_{\mathcal{N}}) is equal to the maximum output purity of the channel 𝒩\mathcal{N} [32, 33] which is defined as

ν(𝒩):=supρ𝒟(A)𝒩(ρ).\nu_{\infty}(\mathcal{N}):=\underset{\rho\in\mathcal{D}(A)}{\sup}\|\mathcal{N}(\rho)\|_{\infty}.

To see the equivalence, first note that this supremum is attained for a pure-state input due to convexity of the operator norm. Then

ν(𝒩)\displaystyle\nu_{\infty}(\mathcal{N}) =sup|α𝒩(|αα|)\displaystyle=\underset{|\alpha\rangle}{\sup}\|\mathcal{N}(|\alpha\rangle\langle\alpha|)\|_{\infty}
=sup|αsup|ββ|𝒩(|αα|)|β\displaystyle=\underset{|\alpha\rangle}{\sup}\sup_{|\beta\rangle}\langle\beta|\mathcal{N}(|\alpha\rangle\langle\alpha|)|\beta\rangle
=sup|α,|βα,β|J𝒩|α,β\displaystyle=\sup_{|\alpha\rangle,|\beta\rangle}\langle\alpha^{*},\beta|J_{\mathcal{N}}|\alpha^{*},\beta\rangle
=Λ2(J𝒩).\displaystyle=\Lambda^{2}(J_{\mathcal{N}}). (39)

An alternative way to prove this equality is using [32]. In that work they establish that ν(𝒩)=Λ2(|𝒩)\nu_{\infty}(\mathcal{N})=\Lambda^{2}(|\mathcal{N}\rangle) where |𝒩|\mathcal{N}\rangle is an un-normalized vector induced by the Kraus representation of 𝒩\mathcal{N}. One can in fact show that |𝒩|\mathcal{N}\rangle is the purification of the Choi matrix, though this has not been stated previously to the best of our knowledge. As the GME of a pure state is the same as the GME of the pure state with a single register traced off [34], one can conclude ν(𝒩)=Λ2(J𝒩)\nu_{\infty}(\mathcal{N})=\Lambda^{2}(J_{\mathcal{N}}). Despite Λ2(J𝒩)=ν(𝒩)\Lambda^{2}(J_{\mathcal{N}})=\nu_{\infty}(\mathcal{N}) being a lower bound of cv(𝒩)\text{cv}(\mathcal{N}) in Eq. (38), in general this bound will not be tight. Hence communication value is capturing property of a quantum channel that is distinct from maximum output purity. In fact, we have the following.

Proposition 5.

ν(𝒩)=cv(𝒩)\nu_{\infty}(\mathcal{N})=\text{cv}(\mathcal{N}) iff 𝒩\mathcal{N} is a replacer channel.

Proof.

If 𝒩\mathcal{N} is a replacer channel, say, 𝒩(ρ)=|ββ|\mathcal{N}(\rho)=|\beta\rangle\langle\beta| ρ\forall\rho, then clearly ν(𝒩)=cv(𝒩)=1\nu_{\infty}(\mathcal{N})=\text{cv}(\mathcal{N})=1. On the other hand, suppose that ν(𝒩)=cv(𝒩)\nu_{\infty}(\mathcal{N})=\text{cv}(\mathcal{N}) and let |α|β:=argmax(Λ2(J𝒩))|\alpha\rangle|\beta\rangle:=\mathrm{argmax}(\Lambda^{2}(J_{\mathcal{N}})). Then for an arbitrary state ρA𝒟(A)\rho^{A}\in\mathcal{D}(A) consider the operator

ΩAB=|αα||ββ|+ρ(𝕀|ββ|).\displaystyle\Omega^{AB}=|\alpha\rangle\langle\alpha|\otimes|\beta\rangle\langle\beta|+\rho\otimes(\mathbb{I}-|\beta\rangle\langle\beta|).

Note that since TrAΩAB=𝕀B\textrm{Tr}_{A}\Omega^{AB}=\mathbb{I}^{B} and ΩABSEP(A:B)\Omega^{AB}\in\text{SEP}(A:B), it is a feasible solution for the optimization of cv(𝒩)\text{cv}(\mathcal{N}) in Proposition 2. Hence for all ρ\rho we have

cv(𝒩)\displaystyle\text{cv}(\mathcal{N}) Tr[ΩABJ𝒩]\displaystyle\geq\textrm{Tr}[\Omega^{AB}J_{\mathcal{N}}]
=α,β|J𝒩|α,β+Tr[𝒩(ρ)(𝕀|ββ|)]\displaystyle=\langle\alpha,\beta|J_{\mathcal{N}}|\alpha,\beta\rangle+\textrm{Tr}[\mathcal{N}(\rho^{*})(\mathbb{I}-|\beta\rangle\langle\beta|)]
=ν(𝒩)+Tr[𝒩(ρ)(𝕀|ββ|)].\displaystyle=\nu_{\infty}(\mathcal{N})+\textrm{Tr}[\mathcal{N}(\rho^{*})(\mathbb{I}-|\beta\rangle\langle\beta|)]. (40)

But by the assumption ν(𝒩)=cv(𝒩)\nu_{\infty}(\mathcal{N})=\text{cv}(\mathcal{N}), the second term must vanish for all ρ\rho. This means that 𝒩\mathcal{N} is a replacer channel, outputting |ββ||\beta\rangle\langle\beta| for all its trace-one inputs. ∎

IV Examples

Having characterized the channel cv in a variety of different ways, we now focus on the problem of computing it. In general, this is a challenging task. Here we provide closed-form solutions for arbitrary qubit channels and the family of Holevo-Werner channels. The latter will also provide a useful case study when we study relaxations to the communication value in Section VII.

IV-A Qubit Channels

Every qubit channel 𝒩\mathcal{N} induces an affine transformation on the Bloch vector of the input state. In more detail, every positive operator ρ\rho can be written as ρ=γ(𝕀+𝐫σ^)\rho=\gamma(\mathbb{I}+\mathbf{r}\cdot\hat{\sigma}), where 𝐫3\mathbf{r}\in\mathbb{R}^{3} has norm no greater than one. Then when 𝒩\mathcal{N} acts on ρ\rho, it induces an affine transformation 𝐫A𝐫+𝐜\mathbf{r}\mapsto A\mathbf{r}+\mathbf{c}, with AA being some 3×33\times 3 matrix and 𝐜3\mathbf{c}\in\mathbb{R}^{3}. Now, let σ=kαkTβk\sigma=\sum_{k}\alpha^{T}_{k}\otimes\beta_{k} be an arbitrary two-qubit separable operator with Tr[αk]=1\textrm{Tr}[\alpha_{k}]=1 and kβk=𝕀\sum_{k}\beta_{k}=\mathbb{I}. We given them Bloch sphere representations αk=12(𝕀+𝐚kσ)\alpha_{k}=\frac{1}{2}(\mathbb{I}+\mathbf{a}_{k}\cdot\vec{\sigma}) and βk=γk(𝕀+𝐛kσ)\beta_{k}=\gamma_{k}(\mathbb{I}+\mathbf{b}_{k}\cdot\vec{\sigma}) so that

Tr[σJ𝒩]\displaystyle\textrm{Tr}[\sigma J_{\mathcal{N}}] =kTr[βk𝒩(αk)]\displaystyle=\sum_{k}\textrm{Tr}[\beta_{k}\mathcal{N}(\alpha_{k})]
=kγk2Tr[(𝕀+𝐛kσ)(𝕀+(A𝐚k+𝐜)σ)]\displaystyle=\sum_{k}\frac{\gamma_{k}}{2}\textrm{Tr}[(\mathbb{I}+\mathbf{b}_{k}\cdot\vec{\sigma})(\mathbb{I}+(A\mathbf{a}_{k}+\mathbf{c})\cdot\vec{\sigma})]
=1+kγk𝐛k(A𝐚k+𝐜)\displaystyle=1+\sum_{k}\gamma_{k}\mathbf{b}_{k}\cdot(A\mathbf{a}_{k}+\mathbf{c})
=1+kγk𝐛kTA𝐚k,\displaystyle=1+\sum_{k}\gamma_{k}\mathbf{b}_{k}^{T}A\mathbf{a}_{k}, (41)

where the last equality follows from the fact that kγk𝐛k=0\sum_{k}\gamma_{k}\mathbf{b}_{k}=0 since kβk=𝕀\sum_{k}\beta_{k}=\mathbb{I}. Our task then is to maximize kγk𝐛kTA𝐚k\sum_{k}\gamma_{k}\mathbf{b}_{k}^{T}A\mathbf{a}_{k} under the constraints that (i) kγk𝐛k=0\sum_{k}\gamma_{k}\mathbf{b}_{k}=0, (ii) kγk=1\sum_{k}\gamma_{k}=1, and (iii) 𝐛k,𝐚k1\|\mathbf{b}_{k}\|,\|\mathbf{a}_{k}\|\leq 1. It is easy to see that this maximization is attained by taking 𝐛1\mathbf{b}_{1} and 𝐛2\mathbf{b}_{2} as anti-parallel unit vectors aligned with the left singular vectors of AA corresponding to its largest singular value, and likewise for 𝐚1\mathbf{a}_{1} and 𝐚2\mathbf{a}_{2} with respect to the right singular vector. Additionally, taking γk=12\gamma_{k}=\frac{1}{2} for k=1,2k=1,2 satisfies all the conditions. Hence we have the following.

Theorem 2.

For a qubit channel 𝒩\mathcal{N}, let AA be the 3×33\times 3 correlation matrix of J𝒩J_{\mathcal{N}}; i.e. Aij=12Tr[(σiσj)J𝒩]A_{ij}=\frac{1}{2}\textrm{Tr}[(\sigma_{i}\otimes\sigma_{j})J_{\mathcal{N}}]. Then

cv(𝒩)=1+σmax(A)\text{cv}(\mathcal{N})=1+\sigma_{\max}(A) (42)

where σmax(A)\sigma_{\max}(A) is the largest singular value of AA.

Remark.

For a unital channel 𝒩\mathcal{N}, the Bloch vector 𝐜\mathbf{c} is zero. Since we are able to obtain the largest value of γk𝐛kTA𝐚k\gamma_{k}\mathbf{b}_{k}^{T}A\mathbf{a}_{k} for each value kk, it follows that

cv(𝒩)=2Λ2(J𝒩);\text{cv}(\mathcal{N})=2\Lambda^{2}(J_{\mathcal{N}}); (43)

i.e. the upper bound in Eq. (38) is tight.

Example: Pauli Channels. As a nice example of Theorem 2, consider the family of Pauli channels, which consists of any qubit channel having the form

𝒩(ρ)=p0ρ+p1XρX+p2ZρZ+p3YρY,\mathcal{N}(\rho)=p_{0}\rho+p_{1}X\rho X+p_{2}Z\rho Z+p_{3}Y\rho Y, (44)

where {X,Y,Z}\{X,Y,Z\} are the standard Pauli matrices and i=03pi=1\sum_{i=0}^{3}p_{i}=1. We can write the Choi matrix as

J𝒩=2i=03pi|Φi+Φi+|,J_{\mathcal{N}}=2\sum_{i=0}^{3}p_{i}|\Phi^{+}_{i}\rangle\langle\Phi^{+}_{i}|, (45)

where |Φi+|\Phi^{+}_{i}\rangle denotes the four Bell states. It is easy to see that the correlation matrix of J𝒩J_{\mathcal{N}} is diagonal with entries {p0+p1p2p3,p0+p1p2+p3,p0p1p2+p3}\{p_{0}+p_{1}-p_{2}-p_{3},-p_{0}+p_{1}-p_{2}+p_{3},p_{0}-p_{1}-p_{2}+p_{3}\}. Therefore, by Theorem 2 we can conclude that

cv(𝒩)=2(p3+p2),\displaystyle\text{cv}(\mathcal{N})=2(p^{\downarrow}_{3}+p^{\downarrow}_{2}), (46)

where p3p^{\downarrow}_{3} and p2p^{\downarrow}_{2} are the two largest probabilities of Pauli gates.

Notice that cv(𝒩)\text{cv}(\mathcal{N}) will equal its largest value of two if and only if there are no more than two Pauli gates applied with nonzero probability in 𝒩\mathcal{N}. In particular, when p3=p2=12p^{\downarrow}_{3}=p^{\downarrow}_{2}=\frac{1}{2} the channel is entanglement-breaking; in fact it is a classical channel. Hence, this example shows that a channel’s communication value captures a property distinct from its ability to transmit entanglement.

IV-B Werner-Holevo Channels

The Werner-Holevo family of channels [32, 35] is defined by

𝒲d,λ:=λΦ0(X)+(1λ)Φ1(X),\displaystyle\mathcal{W}_{d,\lambda}:=\lambda\Phi_{0}(X)+(1-\lambda)\Phi_{1}(X)\ , (47)

where

Φ0(X)\displaystyle\Phi_{0}(X) =1n+1((Tr(X))𝕀+XT)\displaystyle=\frac{1}{n+1}\left((\textrm{Tr}(X))\mathbb{I}+X^{\textrm{T}}\right)
Φ1(X)\displaystyle\Phi_{1}(X) =1n1((Tr(X))𝕀XT).\displaystyle=\frac{1}{n-1}\left((\textrm{Tr}(X))\mathbb{I}-X^{\textrm{T}}\right)\ .

This implies the Choi matrix is given by

J𝒲d,λ=λ2d+1Π++(1λ)2d1Π,\displaystyle J_{\mathcal{W}_{d,\lambda}}=\lambda\frac{2}{d+1}\Pi_{+}+(1-\lambda)\frac{2}{d-1}\Pi_{-}, (48)

where Π+=12(𝕀+𝔽)\Pi_{+}=\frac{1}{2}\left(\mathbb{I}+\mathbb{F}\right) and Π=𝕀Π+\Pi_{-}=\mathbb{I}-\Pi_{+}.

Proposition 6.

The communication value of the Werner-Holevo channel is given by

cv(𝒲d,λ)={d(d+12λ)d21λ1+d2d2dλ1+dλ>1+d2d\displaystyle\text{cv}(\mathcal{W}_{d,\lambda})=\begin{cases}\frac{d(d+1-2\lambda)}{d^{2}-1}&\lambda\leq\frac{1+d}{2d}\\ \frac{2d\lambda}{1+d}&\lambda>\frac{1+d}{2d}\end{cases}
Proof.

Let 𝒰(d)\mathcal{U}(\mathbb{C}^{d}) denote the group of d×dd\times d unitary operators. Since J𝒲d,λJ_{\mathcal{W}_{d,\lambda}} enjoys UUU\otimes U invariance under conjugation for every U𝒰(d)U\in\mathcal{U}(\mathbb{C}^{d}) [36], we can apply the “twirling map”

𝒯UU(X)=𝒰(UU)X(UU)𝑑U\mathcal{T}_{UU}(X)=\int_{\mathcal{U}}(U\otimes U)X(U\otimes U)^{\dagger}dU (49)

to the ABA^{\prime}B^{\prime} systems of σ\sigma while leaving the cv invariant:

Tr[𝒯UU(J𝒲)ΩAB]=Tr[J𝒲𝒯UU(ΩAB)].\textrm{Tr}[\mathcal{T}_{UU}(J_{\mathcal{W}})\Omega^{AB}]=\textrm{Tr}[J_{\mathcal{W}}\mathcal{T}_{UU}(\Omega^{AB})].

Furthermore, since 𝒯UU\mathcal{T}_{UU} preserves the constraints on ΩAB\Omega^{AB}, we can conclude without loss of generality the optimizer is given by X:=𝒯UU(ΩAB)=x𝕀AB+y𝔽X:=\mathcal{T}_{UU}(\Omega^{AB})=x\mathbb{I}^{AB}+y\mathbb{F} for some choice of x,yx,y, i.e. it is an element of the UUU\otimes U-invariant space of operators.

As the space of PPT and SEP UUUU-invariant operators are the same [37], we can relax the optimization program to only requires XPPTX\in\text{PPT}. As is shown in (102), this means we require XX satisfies X0,ΓB(X)0,TrA[Ω]=𝕀BX\geq 0,\Gamma^{B}(X)\geq 0,\textrm{Tr}_{A}[\Omega]=\mathbb{I}^{B}. We will therefore convert these linear constraints into linear constraints on x,yx,y.

Note that {Π+,Π}\{\Pi_{+},\Pi_{-}\} define an orthogonal basis for the space spanned by {𝕀,𝔽}\{\mathbb{I},\mathbb{F}\}. Therefore we can write X=(x+y)Π++(xy)ΠX=(x+y)\Pi_{+}+(x-y)\Pi_{-}, and the positivity constraints on XX are given by

x±y0.\displaystyle x\pm y\geq 0\ . (50)

Similarly, ΓB(X)=x𝕀AB+yΦ+\Gamma^{B}(X)=x\mathbb{I}^{AB}+y\Phi^{+}, where Φ+\Phi^{+} is the unnormalized maximally entangled state. An orthogonal basis for the space spanned by {𝕀AB,Φ+}\{\mathbb{I}^{AB},\Phi^{+}\} is given by {Φ:=d𝕀Φ+,Φ+}\{\Phi^{\perp}:=d\mathbb{I}-\Phi^{+},\Phi^{+}\}. Therefore, XΓB=xd(Φ+Φ+)+yΦ+X^{\Gamma^{B}}=\frac{x}{d}(\Phi^{\perp}+\Phi^{+})+y\Phi^{+}. It follows the partial transpose positivity constraints simplify to

x0x+yd0\displaystyle x\geq 0\hskip 28.45274ptx+yd\geq 0 (51)

The objective function is given by

Tr[J𝒩X]\displaystyle\textrm{Tr}[J_{\mathcal{N}}X]
=\displaystyle= Tr[(2λd+1Π++2(1λ)d1Π)X]\displaystyle\textrm{Tr}\left[\left(\frac{2\lambda}{d+1}\Pi_{+}+\frac{2(1-\lambda)}{d-1}\Pi_{-}\right)X\right]
=\displaystyle= λd(x+y)+(1λ)d(xy)\displaystyle\lambda d(x+y)+(1-\lambda)d(x-y)
=\displaystyle= d(x+(2λ1)y).\displaystyle d(x+(2\lambda-1)y)\ . (52)

Lastly, the trace condition is given by

xTrA[𝕀AB]+yTrA[𝔽]=𝕀Bxd+y=1.\displaystyle x\textrm{Tr}_{A}[\mathbb{I}^{AB}]+y\textrm{Tr}_{A}[\mathbb{F}]=\mathbb{I}^{B}\Rightarrow xd+y=1\ . (53)

Combining these, we have the linear program

maximize d(x+(2λ1)y)\displaystyle d(x+(2\lambda-1)y)
subject to xd+y=1\displaystyle xd+y=1 (54)
x+yd0\displaystyle x+yd\geq 0 (55)
x+y0\displaystyle x+y\geq 0 (56)
xy0\displaystyle x-y\geq 0 (57)
x0.\displaystyle x\geq 0\ . (58)

(54) implies y=1xdy=1-xd. (55),(57) imply 1d+1xdd21\frac{1}{d+1}\leq x\leq\frac{d}{d^{2}-1}. (56)\eqref{eqn:WH-LP-2} implies x1d1x\leq\frac{1}{d-1}, which is always satisfied if xdd21x\leq\frac{d}{d^{2}-1}. Finally, if xx satisfies these constraints,

x+yd=x+(1xd)ddd+1+dd2d+10,x+yd=x+(1-xd)d\geq\frac{d}{d+1}+d-\frac{d^{2}}{d+1}\geq 0\ ,

so (58) is also satisfied. Thus we have reduced the LP to

maximize d(x+(2λ1)(1xd))\displaystyle d(x+(2\lambda-1)(1-xd)) (59)
subject to 1d+1x1d1.\displaystyle\frac{1}{d+1}\leq x\leq\frac{1}{d-1}\ .

Taking the derivative of the objective function one finds that for λ1+d2d\lambda\leq\frac{1+d}{2d}, the derivative is positive. Therefore,

x={dd21λ1+d2d1d+1 otherwise.\displaystyle x^{*}=\begin{cases}\frac{d}{d^{2}-1}&\lambda\leq\frac{1+d}{2d}\\ \frac{1}{d+1}&\text{ otherwise}\ .\end{cases}

Plugging xx^{*} into (59) completes the proof. ∎

In Section VII, we generalize this derivation to determine the PPT relaxation of cv for the nn-fold Werner-Holevo channels.

V Multiplicativity of cv

We next consider how the communication value behaves when we combine two or more channels. The cv is multiplicative for two channels 𝒩\mathcal{N} and \mathcal{M} if

cv(𝒩)=cv(𝒩)cv().\text{cv}(\mathcal{N}\otimes\mathcal{M})=\text{cv}(\mathcal{N})\text{cv}(\mathcal{M}). (60)

When multiplicativity holds, it means that an optimal strategy for guessing channel inputs involves using uncorrelated inputs and measurements across the two channels. A concrete example of non-multiplicativity is given by the Holevo-Werner family of channels, as proven in Section V-D. In general, it is a hard problem to decide whether two channels have a multiplicative communication value. More progress can be made when relaxing this problem to the PPT cone, and we conduct such an analysis in Section VII. Here we resolve on the question of multiplicativity for a few special cases.

V-A Entanglement-Breaking Channels

Our first result shows that non-multiplicativity arises only if the channel is capable of transmitting entanglement.

Theorem 3.

If 𝒩\mathcal{N} is an entanglement-breaking channel, then cv(𝒩)=cv(𝒩)cv()\text{cv}(\mathcal{N}\otimes\mathcal{M})=\text{cv}(\mathcal{N})\text{cv}(\mathcal{M}) for an arbitrary channel \mathcal{M}.

Proof.

Since 𝒩\mathcal{N} is EB, its Choi matrix has the form J𝒩=xΠxρxJ_{\mathcal{N}}=\sum_{x}\Pi_{x}\otimes\rho_{x} for some POVM {Πx}x\{\Pi_{x}\}_{x}. The dual optimization of cv (i.e. Eq. (14)) can then be expressed as

cv(𝒩)=minTr[ZBB]\displaystyle\text{cv}(\mathcal{N}\otimes\mathcal{M})=\min\;\;\textrm{Tr}[Z^{BB^{\prime}}]
subject toZBBxTr[ΠxρA]|xx|(σx)\displaystyle\text{subject to}\;\;Z^{BB^{\prime}}\geq\sum_{x}\textrm{Tr}[\Pi_{x}\rho^{A}]|x\rangle\langle x|\otimes\mathcal{M}(\sigma_{x})
σx:=TrA[(ΠxA𝕀A)ρAA]/Tr[ΠxρA],ρAA.\displaystyle\qquad\sigma_{x}:=\textrm{Tr}_{A}[(\Pi^{A}_{x}\otimes\mathbb{I}^{A^{\prime}})\rho^{AA^{\prime}}]/\textrm{Tr}[\Pi_{x}\rho^{A}],\;\;\forall\rho^{AA^{\prime}}. (61)

Suppose that ZB𝒩(ρ)Z^{B}\geq\mathcal{N}(\rho) for all ρ\rho and ZB(σ)Z^{B^{\prime}}\geq\mathcal{M}(\sigma). Then

xTr[ΠxρA]|xx|(σx)\displaystyle\sum_{x}\textrm{Tr}[\Pi_{x}\rho^{A}]|x\rangle\langle x|\otimes\mathcal{M}(\sigma_{x}) xTr[ΠxρA]|xx|ZB\displaystyle\leq\sum_{x}\textrm{Tr}[\Pi_{x}\rho^{A}]|x\rangle\langle x|\otimes Z^{B^{\prime}}
ZBZB.\displaystyle\leq Z^{B}\otimes Z^{B^{\prime}}.

Thus ZBZBZ^{B}\otimes Z^{B^{\prime}} is feasible in Eq. (V-A). By choosing ZBZ^{B} and ZBZ^{B^{\prime}} to be the dual optimizers for 𝒩\mathcal{N} and \mathcal{M}, respectively, we have cv(𝒩)cv(𝒩)cv()\text{cv}(\mathcal{N}\otimes\mathcal{M})\leq\text{cv}(\mathcal{N})\text{cv}(\mathcal{M}). Since the opposite inequality trivially holds, we have the equality cv(𝒩)=cv(𝒩)cv()\text{cv}(\mathcal{N}\otimes\mathcal{M})=\text{cv}(\mathcal{N})\text{cv}(\mathcal{M}). ∎

By applying Theorem 3 iteratively across nn copies of an entanglement-breaking channel, we obtain a single-letter formulation of the cv capacity.

Corollary 3.

If 𝒩\mathcal{N} is an entanglement-breaking channel, then

𝒞𝒱(𝒩)=cv(𝒩).\mathcal{CV}(\mathcal{N})=\text{cv}(\mathcal{N}). (62)

V-B Covariant Channels

We next turn to channels that have a high degree of symmetry. To study the question of multiplicativity, it will be helpful to use the relationship between cv and GME. The following is a powerful result proven in Ref. [33] regarding multiplicativity of the GME. We say an operator ρAB\rho^{AB} is component-wise non-negative if there exists an orthonormal product basis {|i,j}i,j\{|i,j\rangle\}_{i,j} such that i,j|ρ|i,j0\langle i,j|\rho|i^{\prime},j^{\prime}\rangle\geq 0 for all i,j,i,ji,j,i^{\prime},j^{\prime}.

Lemma 1 ([33]).

If ρAB\rho^{AB} is component-wise non-negative and σAB\sigma^{A^{\prime}B^{\prime}} is any other density operator, then Λ2(ρσ)=Λ2(ρ)Λ2(σ)\Lambda^{2}(\rho\otimes\sigma)=\Lambda^{2}(\rho)\Lambda^{2}(\sigma).

For example, the Choi matrix of the identity channel, ϕd+=Jidd\phi^{+}_{d^{\prime}}=J_{\textrm{id}_{d^{\prime}}}, is component-wise non-negative in the computational basis. Therefore by the previous lemma we have

Λ(J𝒩idd)=Λ(J𝒩)Λ(idd)=Λ(J𝒩)\Lambda(J_{\mathcal{N}}\otimes\textrm{id}_{d^{\prime}})=\Lambda(J_{\mathcal{N}})\Lambda(\textrm{id}_{d^{\prime}})=\Lambda(J_{\mathcal{N}})

for any channel 𝒩\mathcal{N}. As we will now show, this sort of multipicativity can be readily extended to the communication value for channels with symmetry.

Let 𝒢\mathcal{G} be any group with an irreducible unitary representation on d\mathbb{C}^{d}. Then, as we did in Eq. (49), let 𝒯UU\mathcal{T}_{UU} denote the bipartite group twirling map with respect to 𝒢\mathcal{G},

𝒯UU(ρAB)=𝒢𝑑U(UU)ρ(UU).\displaystyle\mathcal{T}_{UU}(\rho^{AB})=\int_{\mathcal{G}}dU(U\otimes U)\rho(U\otimes U)^{\dagger}. (63)

A channel 𝒩\mathcal{N} is called 𝒢\mathcal{G}-covariant if 𝒩(UgρUg)=Ug𝒩(ρ)Ug\mathcal{N}(U_{g}\rho U_{g}^{\dagger})=U_{g}\mathcal{N}(\rho)U_{g}^{\dagger} for all g𝒢g\in\mathcal{G} and all ρ\rho. On the level of Choi matrices, this is equivalent to 𝒯U¯U(J𝒩)=J𝒩\mathcal{T}_{\overline{U}U}(J_{\mathcal{N}})=J_{\mathcal{N}}, where U¯\overline{U} denote complex conjugation. Note that ΓA𝒯ΓA\Gamma_{A}\circ\mathcal{T}\circ\Gamma_{A} is the CPTP twirling map 𝒯U¯U\mathcal{T}_{\overline{U}U}, where ΓA\Gamma_{A} is the partial transpose on system AA. Likewise, the map Γ𝒩Γ\Gamma\circ\mathcal{N}\circ\Gamma is 𝒢\mathcal{G}-covariant if 𝒯UU(J𝒩)=J𝒩\mathcal{T}_{UU}(J_{\mathcal{N}})=J_{\mathcal{N}}, where Γ\Gamma denotes the transpose map.

Theorem 4.

Let 𝒢\mathcal{G} and 𝒢\mathcal{G}^{\prime} have irreducible unitary representations on d\mathbb{C}^{d} and d\mathbb{C}^{d^{\prime}} respectively. Suppose either 𝒩\mathcal{N} or Γ𝒩Γ\Gamma\circ\mathcal{N}\circ\Gamma is 𝒢\mathcal{G}-covariant and likewise either 𝒩\mathcal{N}^{\prime} or Γ𝒩Γ\Gamma\circ\mathcal{N}^{\prime}\circ\Gamma is 𝒢\mathcal{G}’-covariant. Further suppose that J𝒩J_{\mathcal{N}} is component-wise non-negative. Then

cv(𝒩𝒩)=cv(𝒩)cv(𝒩).\displaystyle\text{cv}(\mathcal{N}\otimes\mathcal{N}^{\prime})=\text{cv}(\mathcal{N})\text{cv}(\mathcal{N}^{\prime}). (64)
Proof.

Let |αα|A|ββ|B|\alpha\rangle\langle\alpha|^{A}\otimes|\beta\rangle\langle\beta|^{B} be a product operator with trace equaling dd and satisfying dΛ2(J𝒩)=α,β|J𝒩|α,βd\Lambda^{2}(J_{\mathcal{N}})=\langle\alpha,\beta|J_{\mathcal{N}}|\alpha,\beta\rangle. Suppose now that either 𝒩\mathcal{N} or Γ𝒩Γ\Gamma\circ\mathcal{N}\circ\Gamma is 𝒢\mathcal{G}-covariant. In either case we have

α,β|J𝒩|α,β\displaystyle\langle\alpha,\beta|J_{\mathcal{N}}|\alpha,\beta\rangle =α,β|𝒯(J𝒩)|α,β\displaystyle=\langle\alpha,\beta|\mathcal{T}(J_{\mathcal{N}})|\alpha,\beta\rangle
=Tr[J𝒩𝒯(|α,βα,β|)]\displaystyle=\textrm{Tr}\left[J_{\mathcal{N}}\mathcal{T}^{\dagger}\left(|\alpha,\beta\rangle\langle\alpha,\beta|\right)\right]
=Tr[J𝒩ΩAB],\displaystyle=\textrm{Tr}[J_{\mathcal{N}}\Omega^{AB}], (65)

where ΩAB=𝒯(|α,βα,β|)\Omega^{AB}=\mathcal{T}^{\dagger}(|\alpha,\beta\rangle\langle\alpha,\beta|). Note that ΩAB\Omega^{AB} has trace equaling dd, and since UgU_{g} is an irrep, we have TrAΩAB=𝕀\textrm{Tr}_{A}\Omega^{AB}=\mathbb{I}. Hence cv(𝒩)dΛ2(J𝒩)\text{cv}(\mathcal{N})\geq d\Lambda^{2}(J_{\mathcal{N}}), and an analogous argument for 𝒩\mathcal{N}^{\prime} establishes that cv(𝒩)dΛ2(J𝒩)\text{cv}(\mathcal{N}^{\prime})\geq d^{\prime}\Lambda^{2}(J_{\mathcal{N}^{\prime}}). Therefore,

cv(𝒩𝒩)\displaystyle\text{cv}(\mathcal{N}\otimes\mathcal{N}^{\prime}) cv(𝒩)cv(𝒩)\displaystyle\geq\text{cv}(\mathcal{N})\text{cv}(\mathcal{N}^{\prime})
ddΛ2(J𝒩)Λ2(J𝒩)\displaystyle\geq dd^{\prime}\Lambda^{2}(J_{\mathcal{N}})\Lambda^{2}(J_{\mathcal{N}^{\prime}})
=ddΛ2(J𝒩J𝒩),\displaystyle=dd^{\prime}\Lambda^{2}(J_{\mathcal{N}}\otimes J_{\mathcal{N}^{\prime}}), (66)

where the last equality follows from Lemma 1. However, by the upper bound in Eq. (38) this inequality must be tight, which implies the desired multiplicativity.

Using this theorem, we can compute the cv capacity for certain channels.

Corollary 4.

Let 𝒢\mathcal{G} have irreducible unitary representation on d\mathbb{C}^{d}. Suppose that either 𝒩\mathcal{N} or Γ𝒩Γ\Gamma\circ\mathcal{N}\circ\Gamma is 𝒢\mathcal{G}-covariant and J𝒩J_{\mathcal{N}} is component-wise non-negative. Then

𝒞𝒱(𝒩)=cv(𝒩).\displaystyle\mathcal{CV}(\mathcal{N})=\text{cv}(\mathcal{N}). (67)
Proof.

It suffices to prove that cv(𝒩n)=ncv(𝒩)\text{cv}(\mathcal{N}^{\otimes n})=n\text{cv}(\mathcal{N}). This follows directly from Theorem 4 by letting 𝒩=𝒩n1\mathcal{N}^{\prime}=\mathcal{N}^{\otimes n-1} and 𝒢=𝒢n\mathcal{G}^{\prime}=\mathcal{G}^{\otimes n}. ∎

For example, in a qubit system all Pauli channels satisfy the conditions of Corollary 4. These are channels of the form

𝒩Pauli(ρ)=p0ρ+p1σxρσx+p2σzρσz+p3σyρσy,\displaystyle\mathcal{N}_{\text{Pauli}}(\rho)=p_{0}\rho+p_{1}\sigma_{x}\rho\sigma_{x}+p_{2}\sigma_{z}\rho\sigma_{z}+p_{3}\sigma_{y}\rho\sigma_{y}, (68)

and they are covariant with respect to the Pauli group. Moreover, J𝒩PauliJ_{\mathcal{N}_{\text{Pauli}}} can always be converted into a matrix with non-negative entries by local unitaries. Hence using Eq. (46) we have

𝒞𝒱(𝒩Pauli)=2(p3+p2).\mathcal{CV}(\mathcal{N}_{\text{Pauli}})=2(p_{3}^{\downarrow}+p_{2}^{\downarrow}). (69)

As another example, consider the dd-dimensional partially depolarizing channel 𝒟d,λ\mathcal{D}_{d,\lambda} given by

𝒟d,λ(ρ):=λρ+(1λ)𝕀d,0λ1.\mathcal{D}_{d,\lambda}(\rho):=\lambda\rho+(1-\lambda)\frac{\mathbb{I}}{d},\qquad 0\leq\lambda\leq 1. (70)

The channel Γ𝒟d,λΓ\Gamma\circ\mathcal{D}_{d,\lambda}\circ\Gamma is 𝒢\mathcal{G}-covariant with respect to the full unitary group on d\mathbb{C}^{d} [38]. The Choi matrix is given by J𝒟d,λ=λϕd++(1λ)𝕀𝕀/dJ_{\mathcal{D}_{d,\lambda}}=\lambda\phi^{+}_{d}+(1-\lambda)\mathbb{I}\otimes\mathbb{I}/d, which is clearly component-wise non-negative. Thus by Corollary 4 we have

𝒞𝒱(𝒟d,λ)=cv(𝒟d,λ)=λd+(1λ).\displaystyle\mathcal{CV}(\mathcal{D}_{d,\lambda})=\text{cv}(\mathcal{D}_{d,\lambda})=\lambda d+(1-\lambda). (71)

We remark that the Werner-Holevo family of channels introduced in Section IV-B fail to satisfy Corollary 4 since they are not component-wise non-negative. In fact, their cv is non-multiplicative, as we will see below. Nevertheless, Theorem 4 can be applied to a Werner-Holevo channel by using in parallel with another channel that is component-wise non-negative. For example, when trivially embedding 𝒲d,λ\mathcal{W}_{d,\lambda} into a larger system we have multiplicativity:

cv(𝒲d,λidd)=dcv(𝒲d,λ).\text{cv}(\mathcal{W}_{d,\lambda}\otimes\textrm{id}_{d^{\prime}})=d^{\prime}\text{cv}(\mathcal{W}_{d,\lambda}). (72)

This result is perhaps surprising since mutliplicativity appears to not hold when we relax the cv optimization to the cone of PPT operators, as is shown in Section VII-C1 (See Fig. 5).

V-C Qubit Channels

In Section IV-A we derived an explicit formula for the communication value of qubit channels. Namely, cv(𝒩)=1+σmax(N)\text{cv}(\mathcal{N})=1+\sigma_{\max}(N), where NN is the correlation matrix of J𝒩J_{\mathcal{N}}. Here we show that the cv is multiplicative when using two qubit channels in parallel.

Theorem 5.

Suppose ,𝒩CPTP(AB)\mathcal{M},\mathcal{N}\in\text{CPTP}(A\to B) with dA=dB=2d_{A}=d_{B}=2. Then

cv(𝒩)=cv()cv(𝒩).\text{cv}(\mathcal{M}\otimes\mathcal{N})=\text{cv}(\mathcal{M})\text{cv}(\mathcal{N}). (73)
Proof.

For \mathcal{M} and 𝒩\mathcal{N} consider their Choi matrices

J\displaystyle J_{\mathcal{M}} =12(𝕀(𝕀+𝐜σ)+imiσiσi),\displaystyle=\frac{1}{2}\bigg{(}\mathbb{I}\otimes(\mathbb{I}+\mathbf{c}\cdot\vec{\sigma})+\sum_{i}m_{i}\sigma_{i}\otimes\sigma_{i}\bigg{)},
J𝒩\displaystyle J_{\mathcal{N}} =12(𝕀(𝕀+𝐝σ)+iniσiσi).\displaystyle=\frac{1}{2}\bigg{(}\mathbb{I}\otimes(\mathbb{I}+\mathbf{d}\cdot\vec{\sigma})+\sum_{i}n_{i}\sigma_{i}\otimes\sigma_{i}\bigg{)}. (74)

Notice that their correlation matrices M=diag[m1,m2,m3]M=\text{diag}[m_{1},m_{2},m_{3}] and N=diag[n1,n2,n3]N=\text{diag}[n_{1},n_{2},n_{3}] are diagonal. An arbitrary channel can always be converted into this form by performing appropriate pre- and post SU(2)SU(2) rotations on the channel, which do not change the communication value. Define the operator

ZBB=14(\displaystyle Z^{BB^{\prime}}=\frac{1}{4}\bigg{(} (𝕀+𝐜σ)(𝕀+𝐝σ)+(σmax(M)\displaystyle(\mathbb{I}+\mathbf{c}\cdot\vec{\sigma})\otimes(\mathbb{I}+\mathbf{d}\cdot\vec{\sigma})+(\sigma_{\max}(M)
+σmax(N)+σmax(M)σmax(N))𝕀𝕀).\displaystyle+\sigma_{\max}(N)+\sigma_{\max}(M)\sigma_{\max}(N))\mathbb{I}\otimes\mathbb{I}\bigg{)}. (75)

Since Tr[ZBB]=cv()cv(𝒩)\textrm{Tr}[Z^{BB^{\prime}}]=\text{cv}(\mathcal{M})\text{cv}(\mathcal{N}), by the dual characterization of cv given in Eq. (14), we will prove that cv(𝒩)cv()cv(𝒩)\text{cv}(\mathcal{M}\otimes\mathcal{N})\leq\text{cv}(\mathcal{M})\text{cv}(\mathcal{N}) if we can show that

ZBB𝒩(ρT).Z^{BB^{\prime}}\geq\mathcal{M}\otimes\mathcal{N}(\rho^{T}). (76)

for an arbitrary two-qubit state

ρAA=14(𝕀𝕀+𝐫σ𝕀+𝕀𝐬σ+i,jcijσiσj).\displaystyle\rho^{AA^{\prime}}=\frac{1}{4}\bigg{(}\mathbb{I}\otimes\mathbb{I}+\mathbf{r}\cdot\vec{\sigma}\otimes\mathbb{I}+\mathbb{I}\otimes\mathbf{s}\cdot\vec{\sigma}+\sum_{i,j}c_{ij}\sigma_{i}\otimes\sigma_{j}\bigg{)}.

Note that we have the action (𝕀)=𝕀+𝐜σ\mathcal{M}(\mathbb{I})=\mathbb{I}+\mathbf{c}\cdot\vec{\sigma}, (σx)=mxσx\mathcal{M}(\sigma_{x})=m_{x}\sigma_{x}, (σy)=myσy\mathcal{M}(\sigma_{y})=-m_{y}\sigma_{y}, (σz)=mzσz\mathcal{M}(\sigma_{z})=m_{z}\sigma_{z}, and likewise for the action of 𝒩\mathcal{N}. Hence

𝒩(ρT)=14\displaystyle\mathcal{M}\otimes\mathcal{N}(\rho^{T})=\frac{1}{4} ((𝕀+𝐜σ)(𝕀+𝐝σ)+M𝐫σ𝕀\displaystyle\bigg{(}(\mathbb{I}+\mathbf{c}\cdot\vec{\sigma})\otimes(\mathbb{I}+\mathbf{d}\cdot\vec{\sigma})+M\mathbf{r}\cdot\vec{\sigma}\otimes\mathbb{I}
+𝕀N𝐬σ+i,jminjcijσiσj).\displaystyle+\mathbb{I}\otimes N\mathbf{s}\cdot\vec{\sigma}+\sum_{i,j}m_{i}n_{j}c_{ij}\sigma_{i}\otimes\sigma_{j}\bigg{)}. (77)

When comparing with Eq. (V-C), we see that Eq. (76) reduces to

\displaystyle\| M𝐫σ𝕀+𝕀N𝐬σ+i,jminjcijσiσj\displaystyle M\mathbf{r}\cdot\vec{\sigma}\otimes\mathbb{I}+\mathbb{I}\otimes N\mathbf{s}\cdot\vec{\sigma}+\sum_{i,j}m_{i}n_{j}c_{ij}\sigma_{i}\otimes\sigma_{j}\|_{\infty}
σmax(M)+σmax(N)+σmax(M)σmax(N).\displaystyle\leq\sigma_{\max}(M)+\sigma_{\max}(N)+\sigma_{\max}(M)\sigma_{\max}(N). (78)

To prove this inequality, we note that effectively the non-unital components of \mathcal{M} and 𝒩\mathcal{N} do not appear here. That is, let ~\widetilde{\mathcal{M}} and 𝒩~\widetilde{\mathcal{N}} be the unital CPTP maps defined by the Choi matrices

J~\displaystyle J_{\widetilde{\mathcal{M}}} =12(𝕀𝕀+imiσiσi),\displaystyle=\frac{1}{2}\bigg{(}\mathbb{I}\otimes\mathbb{I}+\sum_{i}m_{i}\sigma_{i}\otimes\sigma_{i}\bigg{)},
J𝒩~\displaystyle J_{\widetilde{\mathcal{N}}} =12(𝕀𝕀+iniσiσi).\displaystyle=\frac{1}{2}\bigg{(}\mathbb{I}\otimes\mathbb{I}+\sum_{i}n_{i}\sigma_{i}\otimes\sigma_{i}\bigg{)}. (79)

Letting |ψ|\psi\rangle denote an eigenvector of largest eigenvalue for the operator on the LHS of Eq. (V-C), we have

𝕀𝕀+M𝐫σ\displaystyle\|\mathbb{I}\otimes\mathbb{I}+M\mathbf{r}\cdot\vec{\sigma}\otimes 𝕀+𝕀N𝐬σ+i,jminjcijσiσj\displaystyle\mathbb{I}+\mathbb{I}\otimes N\mathbf{s}\cdot\vec{\sigma}+\sum_{i,j}m_{i}n_{j}c_{ij}\sigma_{i}\otimes\sigma_{j}\|_{\infty}
=4φ|~𝒩~(ρT)|φ\displaystyle=4\langle\varphi|\widetilde{\mathcal{M}}\otimes\widetilde{\mathcal{N}}(\rho^{T})|\varphi\rangle
=4Tr[ρAA|φφ|BBJ~J𝒩~]\displaystyle=4\textrm{Tr}\left[\rho^{AA^{\prime}}\otimes|\varphi\rangle\langle\varphi|^{BB^{\prime}}J_{\widetilde{\mathcal{M}}}\otimes J_{\widetilde{\mathcal{N}}}\right]
4Λ2(J~J𝒩~)\displaystyle\leq 4\Lambda^{2}(J_{\widetilde{\mathcal{M}}}\otimes J_{\widetilde{\mathcal{N}}})
=2Λ2(J~)2Λ2(J𝒩~)\displaystyle=2\Lambda^{2}(J_{\widetilde{\mathcal{M}}})\cdot 2\Lambda^{2}(J_{\widetilde{\mathcal{N}}})
=(1+σmax(M))(1+σmax(N)),\displaystyle=(1+\sigma_{\max}(M))(1+\sigma_{\max}(N)), (80)

where we have used the fact that the GME is multiplicative for unital qubit channels (Lemma 1), along with Eq. (43). This proves Eq. (V-C).

A natural question is whether Theorem 5 can generalized to the case in which only one of the channels is a qubit channel. Unfortunately the proof of Theorem 5 relies heavily on the Pauli representation of qubit channels, and we therefore only conjecture that qubit channels possess an even stronger form of multiplicativity.

Conjecture. If CPTP(AB)\mathcal{M}\in\text{CPTP}(A\to B) with dA=dB=2d_{A}=d_{B}=2, then

cv(𝒩)=cv()cv(𝒩)\text{cv}(\mathcal{M}\otimes\mathcal{N})=\text{cv}(\mathcal{M})\text{cv}(\mathcal{N})

for any other channel 𝒩\mathcal{N}.

V-D Non-Multiplicativity in Qutrits

In the previous sections we identified examples of channels for which the communication value is multiplicative. We now provide an example of channels that demonstrate non-multiplicativity. Our construction is the Werner-Holevo channels, which were previously known to exemplify non-additivity of a channel’s minimum output purity [32], [33]. Specifically, the channel 𝒲d,0\mathcal{W}_{d,0} has a Choi matrix proportional to the anti-symmetric subspace projector,

J𝒲d,0=1d1(𝕀𝕀𝔽).J_{\mathcal{W}_{d,0}}=\frac{1}{d-1}(\mathbb{I}\otimes\mathbb{I}-\mathbb{F}). (81)

The entanglement properties of this operator have been well-studied [39, 40, 33]. In particular, Zhu et al. have computed its one and two-copy geometric measures of entanglement to be

max|αA|βBα,β|J𝒲d,0α,β\displaystyle\max_{|\alpha\rangle^{A}|\beta\rangle^{B}}\langle\alpha,\beta|J_{\mathcal{W}_{d,0}}|\alpha,\beta\rangle =1d1\displaystyle=\frac{1}{d-1} (82)
max|αAA|βBBα,β|J𝒲d,0J𝒲d,0α,β\displaystyle\max_{|\alpha\rangle^{AA^{\prime}}|\beta\rangle^{BB^{\prime}}}\langle\alpha,\beta|J_{\mathcal{W}_{d,0}}\otimes J_{\mathcal{W}_{d,0}}|\alpha,\beta\rangle =2d(d1).\displaystyle=\frac{2}{d(d-1)}. (83)

Equation (83) is strictly larger than the square of Eq. (82) whenever d3d\geq 3. Furthermore, the maximization in Eq. (83) is attained whenever |αAA|\alpha\rangle^{AA^{\prime}} and |βBB|\beta\rangle^{BB^{\prime}} are maximally entangled states. Thus, we consider the separable operator

σAA:BB=k=1d2|φk+φk+|AA|φk+φk+|BB,\sigma^{AA^{\prime}:BB^{\prime}}=\sum_{k=1}^{d^{2}}|\varphi_{k}^{+}\rangle\langle\varphi_{k}^{+}|^{AA^{\prime}}\otimes|\varphi_{k}^{+}\rangle\langle\varphi_{k}^{+}|^{BB^{\prime}}, (84)

where {|φk+}k=1d2\{|\varphi_{k}^{+}\rangle\}_{k=1}^{d_{2}} is an orthonormal basis consisting of maximally entangled states for dd\mathbb{C}^{d}\otimes\mathbb{C}^{d}. This satisfies the conditions of Proposition 2. Hence, we conclude

cv(𝒲d,0)\displaystyle\text{cv}(\mathcal{W}_{d,0}) =dd1\displaystyle=\frac{d}{d-1} (85)
cv(𝒲d,0𝒲d,0)\displaystyle\text{cv}(\mathcal{W}_{d,0}\otimes\mathcal{W}_{d,0}) =2dd1,\displaystyle=\frac{2d}{d-1}, (86)

which yields cv(𝒲d,0)2<cv(𝒲d,0𝒲d,0)\text{cv}(\mathcal{W}_{d,0})^{2}<\text{cv}(\mathcal{W}_{d,0}\otimes\mathcal{W}_{d,0}) when d3d\geq 3. Most notably, for d=3d=3 we have cv(𝒲d,0𝒲d,0)=3\text{cv}(\mathcal{W}_{d,0}\otimes\mathcal{W}_{d,0})=3 while cv(𝒲d,0)2=2.25\text{cv}(\mathcal{W}_{d,0})^{2}=2.25.

Remark.

In Section VII, we use that the spaces of PPT and SEP UUVVUUVV-invariant operators are equivalent [37], to determine the range of λ\lambda that satisfy non-multiplicativity for cv(𝒲d,λ𝒲d,λ)\text{cv}(\mathcal{W}_{d,\lambda}\otimes\mathcal{W}_{d,\lambda}) numerically.

VI Entanglement-Assisted CV

Refer to caption
Figure 2: Entanglement-assisted communication value scenario.

We next generalize the communication scenario and allow the sender and receiver to share entanglement. Remarkably, this added resource simplifies the problem immensely. In what follows, we will allow Alice and Bob to share an entangled state φAB\varphi^{A^{\prime}B^{\prime}} that can be used to increase the channel cv. The most general entanglement-assisted protocol is as follows (see Fig. 2). For input x[n]x\in[n], Alice performs a CPTP map xCPTP(AA)\mathcal{E}_{x}\in\text{CPTP}(A^{\prime}\to A) on her half of the entangled state φAB\varphi^{A^{\prime}B^{\prime}}. System AA is then fed into the channel, and Bob finally performs a POVM {ΠyBB}y[n]\{\Pi_{y}^{BB^{\prime}}\}_{y\in[n^{\prime}]} on systems BBBB^{\prime}. The induced channel has transition probabilities given by

P(y|x)=Tr[ΠyBB(𝒩ABxAAidB[φAB])].P(y|x)=\textrm{Tr}\left[\Pi_{y}^{BB^{\prime}}\left(\mathcal{N}^{A\to B}\circ\mathcal{E}_{x}^{A^{\prime}\to A}\otimes\textrm{id}^{B^{\prime}}[\varphi^{A^{\prime}B^{\prime}}]\right)\right]. (87)

Note that this scenario corresponds to the one used in superdense coding [18]. The entanglement-assisted channel cv can now be defined.

Definition 3.

The entanglement-assisted communication value (ea cv) of a quantum channel 𝒩CPTP(AB)\mathcal{N}\in\text{CPTP}(A\to B), denoted cv(𝒩)\text{cv}^{*}(\mathcal{N}), is

supφABmax{cv(𝐏)|P(y|x) given by Eq. (87)},\displaystyle\sup_{\varphi^{A^{\prime}B^{\prime}}}\max\{\text{cv}(\mathbf{P})\;|\;\text{$P(y|x)$ given by Eq. \eqref{Eq:Ea-cv}}\}\,, (88)

where the supremum is taken over all entangled states φAB\varphi^{A^{\prime}B^{\prime}} (and all dimensions ABA^{\prime}B^{\prime}), while the maximization considers all n,nn,n^{\prime}\in\mathbb{N} along with arbitrary states {ρx}x[n]\{\rho_{x}\}_{x\in[n]} and POVMs {Πy}y[n]\{\Pi_{y}\}_{y\in[n^{\prime}]}.

Theorem 6.

For an arbitrary channel 𝒩CPTP(AB)\mathcal{N}\in\text{CPTP}(A\to B),

cv(𝒩)=max\displaystyle\text{cv}^{*}(\mathcal{N})=\max Tr[σABJ𝒩]\displaystyle\;\;\textrm{Tr}[\sigma^{AB}J_{\mathcal{N}}]
TrA[σAB]=𝕀B\displaystyle\;\;\textrm{Tr}_{A}[\sigma^{AB}]=\mathbb{I}^{B}
σABPos(A:B),\displaystyle\;\;\sigma^{AB}\in\text{Pos}(A:B), (89)

where Pos(A:B)\text{Pos}(A:B) denotes the positive cone on ABAB. Moreover, cv(𝒩)\text{cv}^{*}(\mathcal{N}) is attained using a (dA)(d_{A})-dimensional maximally entangled state.

In other words, the restriction of σAB\sigma^{AB} to the separable cone (cf Eq. (12)) is removed when considering the entanglement-assisted problem.

Proof.

It is clear that we need to only consider pure states in the supremum since cv(𝐏)\text{cv}(\mathbf{P}) is convex-linear w.r.t. φAB\varphi^{A^{\prime}B^{\prime}}. Let |φAB|\varphi\rangle^{A^{\prime}B^{\prime}} be arbitrary. We first show that without loss of generality we can take |φ|\varphi\rangle to be maximally entangled. Recall Nielsen’s Theorem [41] (see also [42]), which ensures the existence of an LOCC transformation |ΦdA+AA~|φAB|\Phi^{+}_{d_{A^{\prime}}}\rangle^{A^{\prime}\tilde{A^{\prime}}}\to|\varphi\rangle^{A^{\prime}B^{\prime}}, where |ΦdA+AA~|\Phi^{+}_{d_{A^{\prime}}}\rangle^{A^{\prime}\tilde{A^{\prime}}} is a maximally entangled state on AA~A^{\prime}\tilde{A^{\prime}}, with A~A\tilde{A^{\prime}}\cong A^{\prime}. Explicitly, there exists a measurement on Bob’s side with Kraus operators {Mk}k\{M_{k}\}_{k} and correcting unitaries {Uk}k\{U_{k}\}_{k} on Alice’s side such that UkAMkA~B|ΦdA+=p(k)|φkU^{A^{\prime}}_{k}\otimes M^{\tilde{A^{\prime}}\to B^{\prime}}_{k}|\Phi^{+}_{d_{A^{\prime}}}\rangle=\sqrt{p(k)}|\varphi_{k}\rangle. Using that cv is achieved by minimal error discrimination, i.e. cv()=xP(x|x)\text{cv}^{*}(\mathbb{P})=\sum_{x}P(x|x),

cv(𝐏)\displaystyle\text{cv}^{*}(\mathbf{P})
=\displaystyle= x,kTr[ΠxBB(𝒩ABxAAidB[p(k)|φkφk|])].\displaystyle\sum_{x,k}\textrm{Tr}\left[\Pi_{x}^{BB^{\prime}}\left(\mathcal{N}^{A\to B}\circ\mathcal{E}_{x}^{A^{\prime}\to A}\otimes\textrm{id}^{B^{\prime}}\left[p(k)|\varphi_{k}\rangle\langle\varphi_{k}|\right]\right)\right]. (90)

where by construction

p(k)|φkφk|=[(UkMk)Φ+AA~(UkMk)].p(k)|\varphi_{k}\rangle\langle\varphi_{k}|=[(U_{k}\otimes M_{k})\Phi^{+A^{\prime}\tilde{A^{\prime}}}(U_{k}\otimes M_{k})^{\dagger}]\ .

Notice that {(𝕀Mk)Πx(𝕀Mk)}k,x\{(\mathbb{I}\otimes M_{k})^{\dagger}\Pi_{x}(\mathbb{I}\otimes M_{k})\}_{k,x} constitutes a set of POVM elements on BA~B\tilde{A^{\prime}}. This follows from the fact that {Mk}k\{M_{k}\}_{k} are Kraus operators for a CPTP map, and so the dual of this map, XkMk(X)MkX\to\sum_{k}M_{k}^{\dagger}(X)M_{k} , is unital. Likewise, letting 𝒰k():=Uk()Uk\mathcal{U}_{k}(\cdot):=U_{k}(\cdot)U_{k}^{\dagger} denote a unitary channel, the collection {x𝒰k}x,k\{\mathcal{E}_{x}\circ\mathcal{U}_{k}\}_{x,k} forms a family of encoding maps. Therefore, we can express Eq. (VI) as

cv(𝐏)\displaystyle\text{cv}^{*}(\mathbf{P}) =zTr[Π^zBA~(𝒩AB^zAAidA~[Φ+AA~])],\displaystyle=\sum_{z}\textrm{Tr}\!\left[\hat{\Pi}_{z}^{B\tilde{A^{\prime}}}\!\left(\mathcal{N}^{A\to B}\circ\hat{\mathcal{E}}_{z}^{A^{\prime}\to A}\otimes\textrm{id}^{\tilde{A^{\prime}}}[\Phi^{+A^{\prime}\tilde{A^{\prime}}}]\right)\right], (91)

where the ^z\hat{\mathcal{E}}_{z} and Π^z\hat{\Pi}_{z} are the concatenated encoders and decoder. This shows that we can restrict attention just to shared maximally entangled states. Furthermore, without loss of generality, we can assume that dAdAd_{A^{\prime}}\geq d_{A}. The reason is that the transformation |ΦdA′′+A′′A′′~|φAB|\Phi^{+}_{d_{A^{\prime\prime}}}\rangle^{A^{\prime\prime}\tilde{A^{\prime\prime}}}\to|\varphi\rangle^{A^{\prime}B^{\prime}} is always possible for any dA′′dAd_{A^{\prime\prime}}\geq d_{A^{\prime}}; so we could have just as well used the same argument with system A′′A^{\prime\prime} and arrived at Φ+A′′A′′~\Phi^{+A^{\prime\prime}\tilde{A^{\prime\prime}}} in Eq. (91).

We next take Kraus-operator decompositions z()=kNz,k()Nz,k\mathcal{E}_{z}(\cdot)=\sum_{k}N_{z,k}(\cdot)N_{z,k}^{\dagger} with each Nz,k:dAdAN_{z,k}:\mathbb{C}^{d_{A^{\prime}}}\to\mathbb{C}^{d_{A}}. Since dAdAd_{A^{\prime}}\geq d_{A}, we can use the “ricochet” property Nz,k𝕀|ϕdA+AA~=𝕀Nz,kT|ϕdA+AA~N_{z,k}\otimes\mathbb{I}|\phi_{d_{A^{\prime}}}^{+}\rangle^{A^{\prime}\tilde{A^{\prime}}}=\mathbb{I}\otimes N_{z,k}^{T}|\phi_{d_{A}}^{+}\rangle^{A\tilde{A}} to obtain

cv(𝐏)=\displaystyle\text{cv}^{*}(\mathbf{P})= 1dAzkTr[P~z,k(𝒩ABidA~[ϕdA+AA~])]\displaystyle\frac{1}{d_{A^{\prime}}}\sum_{z}\sum_{k}\textrm{Tr}\left[\widetilde{P}_{z,k}\left(\mathcal{N}^{A\to B}\otimes\textrm{id}^{\tilde{A}}[\phi_{d_{A}}^{+A\tilde{A}}]\right)\right]
=\displaystyle= Tr[ΩABJ𝒩AB],\displaystyle\textrm{Tr}[\Omega^{AB}J_{\mathcal{N}}^{AB}], (92)

where P~z,k:=(𝕀BNz,k)Π^zBA~(𝕀BNz,kT)\widetilde{P}_{z,k}:=(\mathbb{I}^{B}\otimes N_{z,k}^{*})\hat{\Pi}_{z}^{B\tilde{A^{\prime}}}(\mathbb{I}^{B}\otimes N_{z,k}^{T}), we have swapped the ordering of the systems to match earlier notation, and

ΩAB\displaystyle\Omega^{AB} =1dAzk(Nz,k𝕀B)Π^zAB(Nz,kT𝕀B)\displaystyle=\frac{1}{d_{A^{\prime}}}\sum_{z}\sum_{k}(N_{z,k}^{*}\otimes\mathbb{I}^{B})\hat{\Pi}_{z}^{A^{\prime}B}(N_{z,k}^{T}\otimes\mathbb{I}^{B})
=1dAzzAAidB(Π^zAB),\displaystyle=\frac{1}{d_{A^{\prime}}}\sum_{z}\mathcal{E}^{*A^{\prime}\to A}_{z}\otimes\textrm{id}^{B}\left(\hat{\Pi}_{z}^{A^{\prime}B}\right), (93)

in which z():=kNz,k()Nz,kT\mathcal{E}^{*}_{z}(\cdot):=\sum_{k}N^{*}_{z,k}(\cdot)N_{z,k}^{T}. Since each z\mathcal{E}_{z}^{*} is trace-preserving, we have

TrAΩAB\displaystyle\textrm{Tr}_{A}\Omega^{AB} =1dAzTrA[zAAidB(Π^zAB)]\displaystyle=\frac{1}{d_{A^{\prime}}}\sum_{z}\textrm{Tr}_{A}\left[\mathcal{E}^{*A^{\prime}\to A}_{z}\otimes\textrm{id}^{B}\left(\hat{\Pi}_{z}^{A^{\prime}B}\right)\right]
=1dAzTrA(Π^zAB)\displaystyle=\frac{1}{d_{A^{\prime}}}\sum_{z}\textrm{Tr}_{A^{\prime}}\left(\hat{\Pi}_{z}^{A^{\prime}B}\right)
=1dATrA(𝕀A𝕀B)=𝕀B.\displaystyle=\frac{1}{d_{A^{\prime}}}\textrm{Tr}_{A^{\prime}}\left(\mathbb{I}^{A^{\prime}}\otimes\mathbb{I}^{B}\right)=\mathbb{I}^{B}. (94)

Hence TrAΩAB=𝕀B\textrm{Tr}_{A}\Omega^{AB}=\mathbb{I}^{B} is a necessary condition on the operator ΩAB\Omega^{AB} such that cv(𝐏)=Tr[ΩABJ𝒩AB]\text{cv}^{*}(\mathbf{P})=\textrm{Tr}[\Omega^{AB}J_{\mathcal{N}}^{AB}]. Let us now sure that it is also sufficient.

Consider any positive operator ΩAB\Omega^{AB} such that TrAΩAB=𝕀B\textrm{Tr}_{A}\Omega^{AB}=\mathbb{I}^{B}. Introduce the generalized Pauli operators on system AA, explicitly given by Um,n=k=0dA1eimk2π/dA|mnm|U_{m,n}=\sum_{k=0}^{d_{A}-1}e^{imk2\pi/d_{A}}|m\oplus n\rangle\langle m|, where m,n=0,,dA1m,n=0,\dots,d_{A}-1 and addition is taken modulo dAd_{A}. It is easy to see that Δ():=1dAm,nUm,n()Um,n\Delta(\cdot):=\frac{1}{d_{A}}\sum_{m,n}U_{m,n}(\cdot)U_{m,n}^{\dagger} is a completely depolarizing map; i.e. Ω(X)=Tr[X]𝕀\Omega(X)=\textrm{Tr}[X]\mathbb{I}. Hence,

ΔAidB[ΩAB]=𝕀ATrAΩAB=𝕀A𝕀B.\Delta^{A}\otimes\textrm{id}^{B}[\Omega^{AB}]=\mathbb{I}^{A}\otimes\textrm{Tr}_{A}\Omega^{AB}=\mathbb{I}^{A}\otimes\mathbb{I}^{B}.

This implies that the elements {𝒰m,nAidB(ΩAB)}m,n\{\mathcal{U}^{A}_{m,n}\otimes\textrm{id}^{B}(\Omega^{AB})\}_{m,n} form a valid POVM on ABAB. Therefore, we can construct an entanglement-assisted protocol as follows. Let Alice and Bob share a maximally entangled state |ΦdA+A~A|\Phi^{+}_{d_{A}}\rangle^{\tilde{A}A}. Alice applies the unitary encoding map on system AA given by 𝒰m,nT():=Um,nT()Um,n\mathcal{U}_{m,n}^{T}(\cdot):=U_{m,n}^{T}(\cdot)U_{m,n}^{*}, and sends her system through the channel 𝒩\mathcal{N}. When Bob performs the POVM just described on systems A~B\tilde{A}B, the obtained score is

m,nP(m,n|m,n)\displaystyle\sum_{m,n}P(m,n|m,n)
=\displaystyle= 1dAm,nTr[(𝒰m,nA~idB[ΩA~B])(idA~𝒩)\displaystyle\frac{1}{d_{A}}\sum_{m,n}\textrm{Tr}\bigg{[}\left(\mathcal{U}^{\tilde{A}}_{m,n}\otimes\textrm{id}^{B}\left[\Omega^{\tilde{A}B}\right]\right)(\textrm{id}^{\tilde{A}}\otimes\mathcal{N})
(idA~𝒰m,nTA[ΦdA+A~A])]\displaystyle\hskip 106.69783pt\left(\textrm{id}^{\tilde{A}}\otimes\mathcal{U}^{TA}_{m,n}\left[\Phi^{+\tilde{A}A}_{d_{A}}\right]\right)\bigg{]}
=\displaystyle= 1(dA)2m,nTr[ΩA~B(idA~𝒩[ϕdA+A~A])]\displaystyle\frac{1}{(d_{A})^{2}}\sum_{m,n}\textrm{Tr}\left[\Omega^{\tilde{A}B}\left(\textrm{id}^{\tilde{A}}\otimes\mathcal{N}\left[\phi_{d_{A}}^{+\tilde{A}A}\right]\right)\right]
=\displaystyle= Tr[ΩJ𝒩].\displaystyle\textrm{Tr}[\Omega J_{\mathcal{N}}]. (95)

The key idea in this equation is that the unitary encoding Um,nU_{m,n} performed on Alice’s side is canceled by exactly one POVM element on Bob’s side. This completes the proof of Theorem 6. ∎

Remark.

The achievability protocol in the previous proof is essentially the original superdense coding protocol applied on a dAd_{A}-dimensional input channel.

Theorem 6 shows that the ea cv can be computed using semi-definite programming. Here we provide a family of channels in which it can be computed even easier.

Corollary 5.

Let 𝒩CPTP(AB)\mathcal{N}\in\text{CPTP}(A\to B) be any channel such that J𝒩J_{\mathcal{N}} has an eigenvector |φAB|\varphi\rangle^{AB} with largest eigenvalue λmax(J𝒩)\lambda_{\max}(J_{\mathcal{N}}) such that φB=𝕀/dB\varphi^{B}=\mathbb{I}/d_{B}. Then

cv(𝒩)=dBλmax(J𝒩).\text{cv}^{*}(\mathcal{N})=d_{B}\lambda_{\max}(J_{\mathcal{N}}). (96)
Proof.

Choose σAB=dB|φφ|AB\sigma^{AB}=d_{B}|\varphi\rangle\langle\varphi|^{AB} in Theorem 6. Clearly this choice is optimal. ∎

In addition, a solution can easily be deduced for all qubit channels.

Theorem 7.

For a qubit channel 𝒩\mathcal{N}, let AA be the 3×33\times 3 correlation matrix of J𝒩J_{\mathcal{N}}; i.e. Aij=12Tr[(σiσj)J𝒩]A_{ij}=\frac{1}{2}\textrm{Tr}[(\sigma_{i}\otimes\sigma_{j})J_{\mathcal{N}}]. Then

cv(𝒩)=1+A1\text{cv}^{*}(\mathcal{N})=1+\|A\|_{1} (97)

where A1=TrAA\|A\|_{1}=\textrm{Tr}\sqrt{A^{\dagger}A}.

Proof.

Using Theorem 6, we can write Ω=12((𝕀+𝐫σ)𝕀+i,jtijσiσj)\Omega=\frac{1}{2}\left((\mathbb{I}+\mathbf{r}\cdot\vec{\sigma})\otimes\mathbb{I}+\sum_{i,j}t_{ij}\sigma_{i}\otimes\sigma_{j}\right). On the other hand, up to local unitaries, the Choi matrix of a channel 𝒩\mathcal{N} can be expressed as J𝒩=12(𝕀(𝕀+𝐬σ)+iaiσiσi)J_{\mathcal{N}}=\frac{1}{2}(\mathbb{I}\otimes(\mathbb{I}+\mathbf{s}\cdot\vec{\sigma})+\allowbreak\sum_{i}a_{i}\sigma_{i}\otimes\sigma_{i}). Hence

Tr[ΩJ𝒩]=1+i=13aitii1+i=13|ai|,\displaystyle\textrm{Tr}[\Omega J_{\mathcal{N}}]=1+\sum_{i=1}^{3}a_{i}t_{ii}\leq 1+\sum_{i=1}^{3}|a_{i}|, (98)

where the last inequality follows form the fact that |tii|1|t_{ii}|\leq 1 since Ω0\Omega\geq 0. The theorem is proven by recalling that the |ai||a_{i}| are the singular values of the correlation matrix AA. ∎

It is worthwhile to compare Theorems 2 and 7. Since A13σmax(A)\|A\|_{1}\leq 3\sigma_{\max}(A) and σmax(A)1\sigma_{\max}(A)\leq 1, we have

cv(𝒩)=1+A12+2σmax(A)=2cv(𝒩).\displaystyle\text{cv}^{*}(\mathcal{N})=1+\|A\|_{1}\leq 2+2\sigma_{\max}(A)=2\text{cv}(\mathcal{N}). (99)

Hence, the shared entanglement between sender and receiver cannot offer a multiplicative enhancement in the cv larger than the dimension. In general, we conjecture the following.

Conjecture: For any channel 𝒩CPTP(AB)\mathcal{N}\in\text{CPTP}(A\to B),

cv(𝒩)dBcv(𝒩).\text{cv}^{*}(\mathcal{N})\leq d_{B}\cdot\text{cv}(\mathcal{N}). (100)

VII Relaxations on the Communication Value

In previous sections we have made use of the fact the communication value can be expressed as a conic optimization problem (Proposition 2). It was noted in generality this problem would be hard to solve, but if the dimension the Choi matrix was sufficiently small, we could relax the cone, SEP(A:B)\text{SEP}(A:B), to the PPT cone, PPT(A:B)\text{PPT}(A:B), and still determine cv(𝒩)\text{cv}(\mathcal{N}) (Corollary 1). Moreover, in Section III-B, we used the optimization program of HminH_{\min} to justify characterizing cv(𝒩)\text{cv}(\mathcal{N}) by a restricted min-entropy, and in Section VI we saw the relationship between HminH_{\min} and cv\text{cv}^{*}. In all of these cases, we have considered the same optimization program and simply varied the cone to which the variable was restricted. That is, we have considered the general conic program

maximize: Tr[XΩAB]\displaystyle\textrm{Tr}[X\Omega^{AB}] (101)
subject to: TrA(X)=𝕀B\displaystyle\textrm{Tr}_{A}(X)=\mathbb{I}^{B}
ΩAB𝒦\displaystyle\Omega^{AB}\in\mathcal{K}

where cv(𝒩)\text{cv}(\mathcal{N}) corresponds to 𝒦=SEP(A:B)\mathcal{K}=\text{SEP}(A:B) and cv\text{cv}^{*} corresponds to 𝒦=Pos(AB)\mathcal{K}=\mathrm{Pos}(A\otimes B). It follows whenever we pick a cone 𝒦\mathcal{K} such that SEP(A:B)𝒦\text{SEP}(A:B)\subset\mathcal{K}, we obtain an upper bound on cv(𝒩)\text{cv}(\mathcal{N}). Throughout the rest of this section, when considering relaxation SEP(A:B)𝒦\text{SEP}(A:B)\subset\mathcal{K}, we denote the value of the optimization program by cv𝒦(𝒩)\text{cv}^{\mathcal{K}}(\mathcal{N}). In this section we primarily consider the PPT relaxation, 𝒦=PPT(A:B)\mathcal{K}=\text{PPT}(A:B). We also discuss the relaxation to the kk-symmetric cone, which is known to converge to the separable cone as kk goes to infinity [43], making it particularly relevant.

VII-A Multiplicativity of Tensored PPT Operators over the PPT cone

We begin with the relaxation to the PPT cone. The primary advantage of this relaxation is that the problem becomes a semidefinite program and so pre-existing software may be used to find the optimal value. One may derive the primal and dual problems to be:

Primal problem

maximize: Tr[XΩAB]\displaystyle\textrm{Tr}[X\Omega^{AB}] (102)
subject to: TrA(ΩAB)=𝕀B\displaystyle\textrm{Tr}_{A}(\Omega^{AB})=\mathbb{I}^{B}
Γ(ΩAB)0\displaystyle\Gamma(\Omega^{AB})\geq 0
ΩAB0.\displaystyle\Omega^{AB}\geq 0\ .

Dual problem

minimize: Tr(Y1)\displaystyle\textrm{Tr}(Y_{1}) (103)
𝕀AY1ΓB(Y2)σ\displaystyle\mathbb{I}^{A}\otimes Y_{1}-\Gamma^{B}(Y_{2})\geq\sigma
Y20\displaystyle Y_{2}\geq 0
Y1Herm(B),\displaystyle Y_{1}\in\mathrm{Herm}(B)\ ,

where ΓB\Gamma^{B} is the partial transpose map on the BB space. This SDP satisfies strong duality as can be verified using Slater’s condition.

With this established, we will now present a special multiplicativity property of the PPT relaxation, cvPPT\text{cv}^{\text{PPT}}.

Theorem 8.

Let RPPT(A1:B1),QPPT(A2:B2)R\in\text{PPT}(A_{1}:B_{1}),Q\in\text{PPT}(A_{2}:B_{2}). Then

cvPPT(RQ)=cvPPT(R)cvPPT(Q).\text{cv}^{\text{PPT}}(R\otimes Q)=\text{cv}^{\text{PPT}}(R)\,\text{cv}^{\text{PPT}}(Q)\ .
Proof.

Let RPPT(A1:B1),QPPT(A2:B2)R\in\text{PPT}(A_{1}:B_{1}),Q\in\text{PPT}(A_{2}:B_{2}). Let (Y1,Y2),(Y¯1,Y¯2)(Y_{1},Y_{2}),(\overline{Y}_{1},\overline{Y}_{2}) be the dual optimizers for R,QR,Q respectively. From (103), we have

𝕀A1Y1R+ΓB1(Y2)\displaystyle\mathbb{I}^{A_{1}}\otimes Y_{1}\geq R+\Gamma^{B_{1}}(Y_{2}) (104)
𝕀A2Y¯1Q+ΓB2(Y¯2).\displaystyle\mathbb{I}^{A_{2}}\otimes\overline{Y}_{1}\geq Q+\Gamma^{B_{2}}(\overline{Y}_{2})\ .

Define R:=ΓB1(R),Q:=ΓB2(Q)R^{\prime}:=\Gamma^{B_{1}}(R),\,Q^{\prime}:=\Gamma^{B_{2}}(Q), which are both positive operators by assumption. Then we have

(𝕀A1Y1)(𝕀A2Y¯1)\displaystyle(\mathbb{I}^{A_{1}}\otimes Y_{1})\otimes(\mathbb{I}^{A_{2}}\otimes\overline{Y}_{1})
\displaystyle\geq (R+ΓB1(Y2))(Q+ΓB2(Y¯2))\displaystyle(R+\Gamma^{B_{1}}(Y_{2}))\otimes(Q+\Gamma^{B_{2}}(\overline{Y}_{2}))
=\displaystyle= RQ+RΓB2(Y¯2)\displaystyle R\otimes Q+R\otimes\Gamma^{B_{2}}(\overline{Y}_{2})
+ΓB1(Y2)Q+ΓB1(Y2)ΓB2(Y¯2)\displaystyle\quad+\Gamma^{B_{1}}(Y_{2})\otimes Q+\Gamma^{B_{1}}(Y_{2})\otimes\Gamma^{B_{2}}(\overline{Y}_{2})
=\displaystyle= RQ+ΓB1B2(RY¯2)\displaystyle R\otimes Q+\Gamma^{B_{1}B_{2}}(R^{\prime}\otimes\overline{Y}_{2})
+ΓB1B2(Y2Q)+ΓB1B2(Y2Y¯2)\displaystyle\quad+\Gamma^{B_{1}B_{2}}(Y_{2}\otimes Q^{\prime})+\Gamma^{B_{1}B_{2}}(Y_{2}\otimes\overline{Y}_{2})
=\displaystyle= RQ+ΓB1B2(RY¯2+Y2Q+Y2Y¯2),\displaystyle R\otimes Q+\Gamma^{B_{1}B_{2}}(R^{\prime}\otimes\overline{Y}_{2}+Y_{2}\otimes Q^{\prime}+Y_{2}\otimes\overline{Y}_{2})\ ,

where the first line follows from (104), the third is because of how the partial transpose over multiple systems may be decomposed, and the fourth is by linearity. Note that R,QR^{\prime},Q^{\prime} are positive as R,QR,Q are PPT. Moreover Y2,Y¯2Y_{2},\overline{Y}_{2} are positive by (103). Thus the whole argument of ΓB1B2\Gamma^{B_{1}B_{2}} is a positive semidefinite operator. Therefore (Y1,new=Y1Y¯1,Y2,new=RY¯2+Y2Q+Y2Y¯2)(Y_{1,new}=Y_{1}\otimes\overline{Y}_{1},Y_{2,new}=R^{\prime}\otimes\overline{Y}_{2}+Y_{2}\otimes Q^{\prime}+Y_{2}\otimes\overline{Y}_{2}) is a feasible point of the dual problem for RQR\otimes Q, and it achieves an optimal value of Tr(Y1)Tr(Y2)\textrm{Tr}(Y_{1})\textrm{Tr}(Y_{2}). If we let X1,X2X_{1},X_{2} be the optimizers for the primal problem for R,QR,Q respectively, then X1X2X_{1}\otimes X_{2} is clearly a feasible point for the primal problem for RQR\otimes Q that achieves optimal value Tr(RX1)Tr(QX2)\textrm{Tr}(RX_{1})\textrm{Tr}(QX_{2}). Using the strong duality of the program, we have Tr(RX1)Tr(QX2)=Tr(Y1)Tr(Y2)\textrm{Tr}(RX_{1})\textrm{Tr}(QX_{2})=\textrm{Tr}(Y_{1})\textrm{Tr}(Y_{2}), so by strong duality we know our proposed optimizers are optimal and this completes the proof. ∎

Corollary 6.

Given any two co-positive maps, 𝒩,\mathcal{N},\mathcal{M}, cvPPT(𝒩)=cvPPT(𝒩)cvPPT()\text{cv}^{\text{PPT}}(\mathcal{N}\otimes\mathcal{M})=\text{cv}^{\text{PPT}}(\mathcal{N})\text{cv}^{\text{PPT}}(\mathcal{M}).

It is interesting to note that we do not know that if only one of the channels is co-positive, then cvPPT\text{cv}^{\text{PPT}} is multiplicative, which would be a stronger claim. This is relevant because we conjecture that cv(𝒩)\text{cv}(\mathcal{N}\otimes\mathcal{M}) is multiplicative if either of the channels is entanglement breaking, which is known to hold for maximal pp-norms for p1p\geq 1 [44], but even the weaker case of multiplicativity where both channels are entanglement-breaking remains open and would mirror Corollary 6, but for separable Choi matrixs and the separable cone optimization.

Relation to kk-Symmetric Extendable Cone

Given the multiplicativity of tensors of PPT operators for cvPPT\text{cv}^{\text{PPT}}, one might hope this property might hope this property extends to cvSymk\text{cv}^{\text{Sym}_{k}} with kk-symmetrically extendable operators, where an operator RPos(AB)R\in\mathrm{Pos}(A\otimes B) is kk-symmetrically extendable if there exists R~AB1kPos(ABk)\tilde{R}^{AB^{k}_{1}}\in\mathrm{Pos}(A\otimes B^{\otimes k}) such that

  1. 1.

    R~=(𝕀AWπ)R~(𝕀AWπ)\tilde{R}=(\mathbb{I}_{A}\otimes W_{\pi})\tilde{R}(\mathbb{I}_{A}\otimes W_{\pi})^{\ast} for all π𝒮k\pi\in\mathcal{S}_{k}

  2. 2.

    TrB2k(R~)=R\textrm{Tr}_{B^{k}_{2}}(\tilde{R})=R

  3. 3.

    (𝕀ATB𝕀B¯2k)R~0(\mathbb{I}_{A}\otimes T^{B}\otimes\mathbb{I}_{\overline{B}^{k}_{2}})\tilde{R}\geq 0  .

Note that the kk-symmetric extendable operators form a cone defined by semidefinite constraints. Moreover, it is known limkSymk=SEP(A:B)\underset{k\to\infty}{\lim}\text{Sym}_{k}=\text{SEP}(A:B) [43]. One can then attempt to extend Theorem 8 in this setting. One can do this by deriving the dual program for cvSymk\text{cv}^{\text{Sym}_{k}}:

min: Tr(W)\displaystyle\textrm{Tr}(W) (105)
𝕀AW𝕀B¯2k+j=1k!(YjΦπj11(Yj))\displaystyle\mathbb{I}_{A}\otimes W\otimes\mathbb{I}_{\overline{B}^{k}_{2}}+\sum_{j=1}^{k!}\left(Y_{j}-\Phi_{\pi^{-1}_{j-1}}(Y_{j})\right)
σ𝕀B2k+ΓB1(Z)\displaystyle\hskip 85.35826pt\succeq\sigma\otimes\mathbb{I}_{B_{2}^{k}}+\Gamma^{B_{1}}(Z)
YjHerm(AB1k)j[k!]\displaystyle Y_{j}\in\mathrm{Herm}(AB_{1}^{k})\quad\forall j\in[k!]
Z0\displaystyle Z\geq 0
WHerm(B),\displaystyle W\in\text{Herm}(B)\ ,

where the indexing of πj\pi_{j} is given by a chosen bijection between the index set [k!][k!] and the permutations in 𝒮k\mathcal{S}_{k}. However, the proof method for Theorem 8 does not seem to naturally extend due to the permutations of the spaces.

VII-B Numerical Evaluation of the Communication Value

To numerically support this work, we developed the CVChannel.jl software package which is publicly available on Github [45]. This Julia [46] software package provides tools for bounding the communication value of quantum channels and certifying their non-multiplicativity. Our software is built upon the disciplined convex programming package, Convex.jl [47], and our numerical results are produced using the splitting conic solver (SCS) [48]. For more details, the curious reader should review the software documentation and source code found on our Github repository [45].

The communication value is difficult to compute in general, but it can be bounded with relative efficiently. CVChannel.jl provides the following methods for bounding cv(𝒩)\text{cv}(\mathcal{N}). An upper bound on cv(𝒩)\text{cv}(\mathcal{N}) is computed via the dual formulation of the PPT relaxation of the communication value Eq. (103),

cv(𝒩)cvPPT(𝒩).\text{cv}(\mathcal{N})\leq\text{cv}^{\text{PPT}}(\mathcal{N}). (106)

While cvPPT(𝒩)\text{cv}^{\text{PPT}}(\mathcal{N}) is a natural upper bound of cv(𝒩)\text{cv}(\mathcal{N}), we consider the dual specifically so that we take a conservative approach to numerical error. That is, numerical error in minimizing the dual will result in a looser upper bound. In general, when considering upper bounds we work with the dual problem and when considering lower bounds we work with the primal problem. While the SDP satisfies strong duality, this guarantees minimizing false positives

For a lower bound on cv(𝒩)\text{cv}(\mathcal{N}), we take a biconvex optimization approach to the problem

cv(𝒩)=max{Πx},{ρx}x=1dB2Tr[Πx𝒩(ρx)].\text{cv}(\mathcal{N})=\max_{\{\Pi_{x}\},\{\rho_{x}\}}\sum_{x=1}^{d_{B}^{2}}\textrm{Tr}[\Pi_{x}\mathcal{N}(\rho_{x})]. (107)

This “see-saw” technique is applied to similar problems in [49, 50], although our implementation remains distinct. To begin, an ensemble of pure quantum states {ρx}x=1dB2\{\rho_{x}\}_{x=1}^{d_{B}^{2}} are initialized at random according to the Haar measure. Then, the following procedure is iterated:

  1. 1.

    With the states fixed, the POVM measurement is numerically optimized as a semidefinite program

    max{Πy}y=1dB2x=1dB2Tr[Πx𝒩(ρx)].\max_{\{\Pi_{y}\}_{y=1}^{d_{B}^{2}}}\sum_{x=1}^{d_{B}^{2}}\textrm{Tr}[\Pi_{x}\mathcal{N}(\rho_{x})]. (108)
  2. 2.

    With optimal measurement as {Πy}\{\Pi_{y}^{\star}\}, we compute the optimal ensemble of quantum states {ρx}x=1db2\{\rho_{x}^{\star}\}_{x=1}^{d_{b}^{2}} as

    ρx=𝒩(Πx),\rho_{x}^{\star}=||\mathcal{N}^{\dagger}(\Pi_{x})||_{\infty}, (109)

    where 𝒩\mathcal{N}^{\dagger} is the adjoint channel and ||||||\cdot||_{\infty} denotes the largest eigenvalue.

Repeating this procedure results in a set of optimized states {ρx}y=1dB2\{\rho_{x}^{\star}\}_{y=1}^{d_{B}^{2}} and measurement {Πy}y=1dB2\{\Pi_{y}^{\star}\}_{y=1}^{d_{B}^{2}} such that

cvSeeSaw(𝒩)=x=1dB2Tr[Πx𝒩(ρx)]cv(𝒩).\text{cv}^{SeeSaw}(\mathcal{N})=\sum_{x=1}^{d_{B}^{2}}\textrm{Tr}[\Pi_{x}^{\star}\mathcal{N}(\rho_{x}^{\star})]\leq\text{cv}(\mathcal{N}). (110)

To improve the see-saw optimization, the procedure is simply performed many times with randomly initialized states. Combining these techniques, we numerically bound the communication value,

cvSeeSaw(𝒩)cv(𝒩)cvPPT(𝒩,dual).\text{cv}^{SeeSaw}(\mathcal{N})\leq\text{cv}(\mathcal{N})\leq\text{cv}^{\text{PPT}}(\mathcal{N},\;\text{dual}). (111)

To numerically certify that quantum channels 𝒩\mathcal{N} and \mathcal{M} are non-multiplicative, we need to compute a lower bound on cv(𝒩)\text{cv}(\mathcal{N}\otimes\mathcal{M}) and upper bound on cv(𝒩)\text{cv}(\mathcal{N}) and cv(𝒩)\text{cv}(\mathcal{N}). CVChannel.jl computes the lower bound as cvSeeSaw(𝒩)cv(𝒩)\text{cv}^{SeeSaw}(\mathcal{N}\otimes\mathcal{M})\leq\text{cv}(\mathcal{N}\otimes\mathcal{M}) and the upper bound as cv(𝒩)cvPPT(𝒩,dual)\text{cv}(\mathcal{N})\leq\text{cv}^{\text{PPT}}(\mathcal{N},\;\text{dual}). Non-multiplicativity is numerically confirmed when

cvSeeSaw(𝒩)cvPPT(𝒩)cvPPT()>ε\text{cv}^{SeeSaw}(\mathcal{N}\otimes\mathcal{M})-\text{cv}^{\text{PPT}}(\mathcal{N})\text{cv}^{\text{PPT}}(\mathcal{M})>\varepsilon (112)

where ε>0\varepsilon>0 is a conservative bound to the numerical error. One drawback of this procedure is its susceptibility to false negatives due to the fact that the PPT Relaxation is a loose upper bound of the communication value.

VII-C Examples

Having established properties of the PPT relaxation of the communication value, we investigate channels which are known in other settings to admit non-multiplicative behaviour. In particular, we look at the family of Werner-Holevo channels [32], the dephrasure channel [51], and the Siddhu channel [52]. We see that the Werner-Holevo channel is not multiplicative over a range of parameters, but the dephrasure and Siddhu channel which are known for their superactivation of coherent information are always multiplicative for the communication value. In some sense this should not be surprising as communication value captures a notion of using the quantum channel to transmit classical information whereas the coherent information measures the ability to transfer quantum information. However, it exemplifies how different the coherent information and communication values are as measures.

Werner-Holevo Channels

In Section IV-B, we showed how to determine the cv for the Werner-Holevo channels. In this section we extend the method to obtain this result to the construction of a linear program (LP) for determining the cvPPT\text{cv}^{\text{PPT}} for nn Werner-Holevo channels ran in parallel. We then use this to show non-multiplicativity for cv(𝒲d,λ𝒲d,λ)\text{cv}(\mathcal{W}_{d,\lambda}\otimes\mathcal{W}_{d,\lambda}) as a function of λ\lambda, as well as the non-multiplicativity of cvPPT\text{cv}^{\text{PPT}} for more copies of the channel. We note our derivation assumes the dimension is the same for all channels, but a generalization is straightforward.

Proposition 7.

Considering nn Werner-Holevo channels, there is a linear program

max{a,c:Ac0,Bc0,g,c=1},\max\{\langle a,c\rangle\,:\,Ac\geq 0,\,Bc\geq 0,\,\langle g,c\rangle=1\}\ ,

which obtains the value of cvPPT(i=1nJ(𝒲d,λi))\text{cv}^{\text{PPT}}(\otimes_{i=1}^{n}J(\mathcal{W}_{d,\lambda_{i}})). Moreover, there exists an algorithm to generate the constraints a,A,B,ga,A,B,g which takes at most 𝒪(n22n)\mathcal{O}(n2^{2n}) steps.

Derivation of Constraints.

Let Π0:=Π+\Pi_{0}:=\Pi_{+}, Π1:=Π\Pi_{1}:=\Pi_{-}. This labelling will simplify notation. We are interested in cvPPT(i=1nJ(𝒲d,λi))\text{cv}^{\text{PPT}}(\bigotimes_{i=1}^{n}J(\mathcal{W}_{d,\lambda_{i}})). Recalling the objective function of (102) is

Tr[i=1nJ(𝒲d,λi)ΩAnBn],\textrm{Tr}[\bigotimes_{i=1}^{n}J(\mathcal{W}_{d,\lambda_{i}})\Omega^{A^{n}B^{n}}]\ ,

we can twirl Ω\Omega by moving the symmetry of the Holevo channels onto Ω\Omega. This results in Ω=s{0,1}ncsRs\Omega=\sum_{s\in\{0,1\}^{n}}c_{s}R_{s} where

Rs=i=1n𝔽s(i)=j{0,1}n(i[n](1)s(i)j(i)Πj(i)),R_{s}=\bigotimes_{i=1}^{n}\mathbb{F}^{s(i)}=\sum_{j\in\{0,1\}^{n}}\left(\bigotimes_{i\in[n]}(-1)^{s(i)\wedge j(i)}\Pi_{j(i)}\right)\ ,

where the constraint on the sign is because 𝔽0=𝕀=Π0+Π1\mathbb{F}^{0}=\mathbb{I}=\Pi_{0}+\Pi_{1} and 𝔽1=Π0Π1\mathbb{F}^{1}=\Pi_{0}-\Pi_{1}, so a term is negative iff s(i)=j(i)=1s(i)=j(i)=1. Combining these, we can express Ω\Omega as a linear combination of orthogonal subspaces with coefficients stored in a vector cc:

Ω=s{0,1}ncsj{0,1}n(i[n](1)s(i)j(i)Πj(i)).\displaystyle\Omega=\sum_{s\in\{0,1\}^{n}}c_{s}\sum_{j\in\{0,1\}^{n}}\left(\bigotimes_{i\in[n]}(-1)^{s(i)\wedge j(i)}\Pi_{j(i)}\right)\ . (113)

With the state simplified into mutually orthogonal subspaces, we just need to convert the constraints of (102) to constraints on c2nc\in\mathbb{R}^{2^{n}}.

Guaranteeing positivity of Ω\Omega is equivalent to guaranteeing the weight of each orthogonal subspace in (113) is non-negative. As multiple elements of cc can have weight on multiple subspaces, the constraint is that the relevant linear combination of cc is non-negative for each subspace. Thus the positivity constraints may be written as Ac0Ac\geq 0 where A2n×2nA\in\mathbb{R}^{2^{n}\times 2^{n}} matrix storing the sign information (1)s(i)j(i)(-1)^{s(i)\wedge j(i)} for all ss,jj.

The PPT constraints correspond to ΩΓ0\Omega^{\Gamma}\geq 0. Noting that 𝔽Γ=Φ+\mathbb{F}^{\Gamma}=\Phi^{+}, the unnormalized maximally entangled state. We have ΩΓ=s{0,1}ncsi[n]Xs(i),\Omega^{\Gamma}=\sum_{s\in\{0,1\}^{n}}c_{s}\bigotimes_{i\in[n]}X^{s(i)}\ , where

Xs(i):={d1(Φ+Φ+)s(i)=0Φ+s(i)=1,X^{s(i)}:=\begin{cases}d^{-1}(\Phi^{\perp}+\Phi^{+})&s(i)=0\\ \Phi^{+}&s(i)=1\end{cases}\ ,

where Φ=d𝕀Φ+\Phi^{\perp}=d\mathbb{I}-\Phi^{+}. In other words, we have decomposed ΩΓ\Omega^{\Gamma} into linear combinations of a set of orthogonal subspaces.111We note this implies d1(Φ+Φ+)=𝕀d^{-1}(\Phi^{\perp}+\Phi^{+})=\mathbb{I}. The choice of presentation is to make it clear we are considering two orthogonal subspaces. Again, we only need to store the constraints on cc which in this case is the order of dd and if the coefficient is zero. By the definition of Xs(i)X^{s(i)}, there is not weight of a subspace for csc_{s} iff the ithi^{th} element in the tensor is Φ\Phi^{\perp} and s(i)=1s(i)=1, and otherwise the weight is given by d(2nw(s))d^{-(2^{n}-w(s))} where w()w(\cdot) is the Hamming weight of the string ss. Thus the PPT constraints may be written as Bc0Bc\geq 0 where B2n×2nB\in\mathbb{R}^{2^{n}\times 2^{n}}.

Recalling Ω=s{0,1}ncsRs\Omega=\sum_{s\in\{0,1\}^{n}c_{s}R_{s}} and TrA(F)=𝕀B\textrm{Tr}_{A}(F)=\mathbb{I}^{B}, TrA(𝕀)=d𝕀B\textrm{Tr}_{A}(\mathbb{I})=d\mathbb{I}^{B}, the partial trace condition is reduced to g,c=1\langle g,c\rangle=1 where g2ng\in\mathbb{R}^{2^{n}} and g(s)=dnw(s)g(s)=d^{n-w(s)}.

Finally, we have the objective function. We write i=1nJ(𝒲d,λi)=s{0,1}n(iζi(s(i))Πs(i))\bigotimes_{i=1}^{n}J(\mathcal{W}_{d,\lambda_{i}})=\sum_{s\in\{0,1\}^{n}}\left(\bigotimes_{i}\zeta_{i}(s(i))\Pi_{s({i})}\right), where ζi(s(i))={λif0i=0(1λi)f1i=1\zeta_{i}(s(i))=\begin{cases}\lambda_{i}f_{0}&i=0\\ (1-\lambda_{i})f_{1}&i=1\end{cases} where fif_{i} is the normalization constant in front of the projector. Calculating Tr[i=1nJ(𝒲d,λi)Ω]\textrm{Tr}[\bigotimes_{i=1}^{n}J(\mathcal{W}_{d,\lambda_{i}})\Omega] using the above expression along with (113), one can simplify the objective function to

s{0,1}ncs(j{0,1}n[i[n](1)s(i)j(i)φi(j(i))]),\sum_{s\in\{0,1\}^{n}}c_{s}\left(\sum_{j\in\{0,1\}^{n}}\left[\prod_{i\in[n]}(-1)^{s(i)\wedge j(i)}\varphi_{i}(j(i))\right]\right)\ ,

where φi(j(i))\varphi_{i}(j(i)) is the same as ζi(j(i))\zeta_{i}(j(i)), except without the normalization constant. Thus we may define aa as the argument of the large parentheses. This completes the derivation of the LP. Finally we note to construct the constraints one needs to loop through nested loops of sizes 2n,2n,n2^{n},2^{n},n which results in the 𝒪(n22n)\mathcal{O}(n2^{2^{n}}) steps in the algorithm. ∎

Using these numerics, we can look at the behaviour of the PPT-relaxation of the communication value of the nn-fold Werner Holevo Channel (Fig. 3). We can see that the non-multiplicativity over the PPT cone grows exponentially (Fig. 3) and that all non-multiplicativity dies out at λ=0.3\lambda=0.3 in all cases. We note it is known that for the tensor product of two Werner states, the space of PPT operators is the same as the space of separable operators. In this case, we see the non-additivity of the true communication value for the Werner-Holevo channels.

Refer to caption
Figure 3: The cvPPT\text{cv}^{\text{PPT}} value of the nn-fold Werner Holevo channel for all values of channel parameter λ[0,1]\lambda\in[0,1].
Refer to caption
Figure 4: Here we see the non-multiplicativity of tensoring the Werner-Holevo channel with itself. We note that the blue line characterizes the multiplicativity of cv rather than just cvPPT\text{cv}^{\text{PPT}}.

VII-C1 PPT Relaxation of Werner-Holevo with the Identity

An immediate corollary of Theorem 4 is that the Werner-Holevo channel when ran in parallel with the identity channel of any dimension is multiplicative. That is, cv(𝒲d,λidd)=dcv(𝒲d,λ)\text{cv}(\mathcal{W}_{d,\lambda}\otimes id_{d^{\prime}})=d^{\prime}\cdot\text{cv}(\mathcal{W}_{d,\lambda}). However, here we find that this is not the case for cvPPT\text{cv}^{PPT} which is non-multiplicative, exhibiting a clear separation between the cv and its relaxation. This separation is given for the 𝒲d,0idd\mathcal{W}_{d,0}\otimes id_{d^{\prime}} in Fig. 5. It is determined using the following proposition.

Proposition 8.

The PPT communication value of the Werner-Holevo channel ran in parallel with an identity channel, cvPPT(𝒲d,λidd)\text{cv}^{\text{PPT}}(\mathcal{W}_{d,\lambda}\otimes\textrm{id}_{d^{\prime}}), is given by the linear program

max\displaystyle\max dd[w+yd+(2λ1)(x+zd)]\displaystyle\;\;dd^{\prime}[w+yd^{\prime}+(2\lambda-1)(x+zd^{\prime})] (114)
  0wx+dydz\displaystyle\;\;0\leq w-x+d^{\prime}y-d^{\prime}z
  0wx\displaystyle\;\;0\leq w-x
  0w+x+dy+dz\displaystyle\;\;0\leq w+x+d^{\prime}y+d^{\prime}z
  0w+x\displaystyle\;\;0\leq w+x
  0w+dxydz\displaystyle\;\;0\leq w+dx-y-dz
  0wy\displaystyle\;\;0\leq w-y
  0w+dx+y+dz\displaystyle\;\;0\leq w+dx+y+dz
  0w+y\displaystyle\;\;0\leq w+y
  1=ddw+dx+dy+z.\displaystyle\;\;1=dd^{\prime}w+d^{\prime}x+dy+z.
Derivation.

The derivation is similar to that of the previous cvPPT\text{cv}^{PPT} LP derivations, we just also consider V¯V\overline{V}V covariance for the identity channel. Let us consider the channel 𝒲d,λidd\mathcal{W}_{d,\lambda}\otimes\textrm{id}_{d^{\prime}}, where 𝒲d,λ\mathcal{W}_{d,\lambda} is defined in (47). Then 𝒥𝒲ϕd+\mathcal{J}_{\mathcal{W}}\otimes\phi^{+}_{d^{\prime}} is UUV¯VUU\overline{V}V-covariant, and so for any feasible operator σAB\sigma^{AB} in the cvPPT\text{cv}^{PPT} SDP, we have

Tr[σAA:BBJ𝒲Jidd]\displaystyle\textrm{Tr}[\sigma^{AA:BB^{\prime}}J_{\mathcal{W}}\otimes J_{\textrm{id}_{d^{\prime}}}]
=\displaystyle= Tr[σAA:BB𝒯UU(𝒥𝒲)𝒯V¯V(ϕd+)]\displaystyle\textrm{Tr}[\sigma^{AA^{\prime}:BB^{\prime}}\mathcal{T}_{UU}(\mathcal{J}_{\mathcal{W}})\otimes\mathcal{T}_{\overline{V}V}(\phi^{+}_{d^{\prime}})]
=\displaystyle= Tr[𝒯UU𝒯V¯V(σAA:BB)J𝒲ϕd+].\displaystyle\textrm{Tr}[\mathcal{T}_{UU}\otimes\mathcal{T}_{\overline{V}V}(\sigma^{AA^{\prime}:BB^{\prime}})J_{\mathcal{W}}\otimes\phi^{+}_{d^{\prime}}].

Note that 𝒯UU𝒯V¯V(σ)\mathcal{T}_{UU}\otimes\mathcal{T}_{\overline{V}V}(\sigma) is still a feasible operator, and so without loss of generality we can assume that σAA:BB\sigma^{AA^{\prime}:BB^{\prime}} is itself UUV¯VUU\overline{V}V-covariant. Thus, we can parametrize σ\sigma as

w𝕀AB𝕀AB+x𝔽d𝕀+y𝕀ϕd++z𝔽dϕd+.\displaystyle w\mathbb{I}^{AB}\otimes\mathbb{I}^{A^{\prime}B^{\prime}}+x\mathbb{F}_{d}\otimes\mathbb{I}+y\mathbb{I}\otimes\phi^{+}_{d^{\prime}}+z\mathbb{F}_{d}\otimes\phi^{+}_{d^{\prime}}.

The space of UUV¯VUU\overline{V}V operators is spanned by the set of four orthogonal operators

{Πdϕd+Πd(d𝕀ϕd+)Πd+ϕd+Πd+(d𝕀ϕd+)}\left\{\begin{array}[]{cc}\Pi_{d}^{-}\otimes\phi_{d^{\prime}}^{+}&\Pi_{d}^{-}\otimes(d^{\prime}\mathbb{I}-\phi_{d^{\prime}}^{+})\\[5.69054pt] \Pi_{d}^{+}\otimes\phi_{d^{\prime}}^{+}&\Pi_{d}^{+}\otimes(d^{\prime}\mathbb{I}-\phi_{d^{\prime}}^{+})\end{array}\right\}

Positivity then amounts to the conditions

wx+dydz\displaystyle w-x+d^{\prime}y-d^{\prime}z 0\displaystyle\geq 0 (115)
wx\displaystyle w-x 0\displaystyle\geq 0
w+x+dy+dz\displaystyle w+x+d^{\prime}y+d^{\prime}z 0\displaystyle\geq 0
w+x\displaystyle w+x 0.\displaystyle\geq 0.

The partial transpose of σ\sigma, σΓBB\sigma^{\Gamma_{BB^{\prime}}} is given by

w𝕀AB𝕀AB+xϕd+𝕀+y𝕀𝔽d+zϕd+𝔽d.\displaystyle w\mathbb{I}^{AB}\otimes\mathbb{I}^{A^{\prime}B^{\prime}}+x\phi^{+}_{d}\otimes\mathbb{I}+y\mathbb{I}\otimes\mathbb{F}_{d^{\prime}}+z\phi^{+}_{d}\otimes\mathbb{F}_{d^{\prime}}.

To check positivity, we now use just need to swap the orthogonal basis operators:

{ϕd+Πd(d𝕀ϕd+)Πdϕd+Πd+(d𝕀ϕd+)Πd+}\left\{\begin{array}[]{cc}\phi_{d}^{+}\otimes\Pi_{d^{\prime}}^{-}&(d\mathbb{I}-\phi_{d}^{+})\otimes\Pi_{d^{\prime}}^{-}\\[5.69054pt] \phi_{d}^{+}\otimes\Pi_{d^{\prime}}^{+}&(d\mathbb{I}-\phi_{d}^{+})\otimes\Pi_{d^{\prime}}^{+}\end{array}\right\}

This yields the conditions

w+dxydz\displaystyle w+dx-y-dz 0\displaystyle\geq 0 (116)
wy\displaystyle w-y 0\displaystyle\geq 0
w+dx+y+dz\displaystyle w+dx+y+dz 0\displaystyle\geq 0
w+y\displaystyle w+y 0.\displaystyle\geq 0.

Finally, we compute the objective function

Tr[σAA:BBJ𝒲ϕd+]\displaystyle\textrm{Tr}[\sigma^{AA^{\prime}:BB^{\prime}}J_{\mathcal{W}}\otimes\phi_{d^{\prime}}^{+}]
=\displaystyle= d(wTr[J𝒲]+xTr[𝔽J𝒲])+d2(yTr[J𝒲]+zTr[𝔽J𝒲])\displaystyle d^{\prime}(w\textrm{Tr}[J_{\mathcal{W}}]+x\textrm{Tr}[\mathbb{F}J_{\mathcal{W}}])+{d^{\prime}}^{2}(y\textrm{Tr}[J_{\mathcal{W}}]+z\textrm{Tr}[\mathbb{F}J_{\mathcal{W}}])
=\displaystyle= d(wd+xd(2λ1))+d2(yd+zd(2λ1))\displaystyle d^{\prime}(wd+xd(2\lambda-1))+{d^{\prime}}^{2}(yd+zd(2\lambda-1))
=\displaystyle= dd[w+yd+(2λ1)(x+zd)],\displaystyle dd^{\prime}[w+yd^{\prime}+(2\lambda-1)(x+zd^{\prime})], (117)

and the partial trace condition

TrAA[σAA:BB]\displaystyle\textrm{Tr}_{AA^{\prime}}[\sigma^{AA^{\prime}:BB^{\prime}}] =(ddw+dx+dy+z)𝕀BB.\displaystyle=(dd^{\prime}w+d^{\prime}x+dy+z)\mathbb{I}^{BB^{\prime}}. (118)

Combining (115) – (118) completes the derivation. ∎

Refer to caption
Figure 5: This shows the gap between cv(𝒩WH,d1,λidd2)=d2cv(𝒩WH,d,λ)\text{cv}(\mathcal{N}_{WH,d_{1},\lambda}\otimes\textrm{id}_{d_{2}})=d_{2}\text{cv}(\mathcal{N}_{WH,d,\lambda}) and cvPPT(𝒩WH,d1,λidd2)\text{cv}^{PPT}(\mathcal{N}_{WH,d_{1},\lambda}\otimes\textrm{id}_{d_{2}}) for λ=0\lambda=0.

Dephrasure Channel

We next consider the dephrasure channel,

𝒩p,q(X):=(1q)((1p)ρ+pZρZ)+qTr(X)|ee|,\mathcal{N}_{p,q}(X):=(1-q)\left((1-p)\rho+pZ\rho Z\right)+q\textrm{Tr}(X)|e\rangle\langle e|\ ,

where p,q[0,1]p,q\in[0,1]. The interesting aspect of the dephrasure channel is that in some parameter regime it admits superadditivity of coherent information [51]. We will first present it’s communication value.

Lemma 2.

cv(𝒩p,q)=2q\text{cv}(\mathcal{N}_{p,q})=2-q.

Proof.

We are going to prove this by constructing feasible operators in the primal and dual which achieve this value. First we note the Choi matrix:

J(𝒩p,q)=\displaystyle J(\mathcal{N}_{p,q})= (1q)(|0000|+|1111|)\displaystyle(1-q)\left(|00\rangle\langle 00|+|11\rangle\langle 11|\right)
+γ(|0011|+|1100|)\displaystyle+\gamma\left(|00\rangle\langle 11|+|11\rangle\langle 00|\right)
+q(|0e0e|+|1e1e|),\displaystyle\hskip 28.45274pt+q\left(|0e\rangle\langle 0e|+|1e\rangle\langle 1e|\right)\ ,

where γ:=(1q)(12p)\gamma:=(1-q)(1-2p). Then for the primal problem, we may choose

X=|0000|+|1111|+1/2(|0e0e|+|1e1e|).X=|00\rangle\langle 00|+|11\rangle\langle 11|+1/2\left(|0e\rangle\langle 0e|+|1e\rangle\langle 1e|\right)\ .

This clearly satisfies TrA(X)=𝕀B\textrm{Tr}_{A}(X)=\mathbb{I}^{B}, it is PPT as it is diagonal, and X,J(𝒩p,q)=2q\langle X,J(\mathcal{N}_{p,q})\rangle=2-q. For the dual problem, let

Y1\displaystyle Y_{1} =(1q)(|00|+|11|)+q|ee|\displaystyle=(1-q)(|0\rangle\langle 0|+|1\rangle\langle 1|)+q|e\rangle\langle e|
Y2\displaystyle Y_{2} =κ(|0101|+|1010|)γ(|0110|+|1001|),\displaystyle=\kappa(|01\rangle\langle 01|+|10\rangle\langle 10|)-\gamma(|01\rangle\langle 10|+|10\rangle\langle 01|)\ ,

where (1q)κ|γ|=(1q)|(12p)|(1-q)\geq\kappa\geq|\gamma|=(1-q)|(1-2p)|. Note this interval is never empty as |(12p)|[0,1]|(1-2p)|\in[0,1] for all p[0,1]p\in[0,1].

Then Y1Y_{1} is clearly Hermitian, and Y20Y_{2}\succeq 0 as it’s eigenvalues are κ±γ|γ|±γ0\kappa\pm\gamma\geq|\gamma|\pm\gamma\geq 0 and 0 with multiplicity 44. Then, one may calculate from these expressions that

𝕀AY1Γ(Y2)J(𝒩p,q)\displaystyle\mathbb{I}_{A}\otimes Y_{1}-\Gamma(Y_{2})-J(\mathcal{N}_{p,q})
=\displaystyle= ((1qκ)[|0011|+|1100|],\displaystyle\left((1-q-\kappa\right)\left[|00\rangle\langle 11|+|11\rangle\langle 00|\right]\ ,

Therefore we have constructed a feasible choice. Finally, Tr(Y1)=2q\textrm{Tr}(Y_{1})=2-q completes the proof. ∎

Note what the above implies is the ‘dephasing’ property of the dephrasure is irrelevant. This is in some sense intuitive as the dephasing cannot hurt the classical information if the optimal strategy is sending data in the classical basis. Indeed, it is easy to see the above value may be achieved by using the signal states {|00|,|11|}\{|0\rangle\langle 0|,|1\rangle\langle 1|\} and the projective measurement decoder {|00|+1/2|ee|,|11|+1/2|ee|}\{|0\rangle\langle 0|+1/2|e\rangle\langle e|,|1\rangle\langle 1|+1/2|e\rangle\langle e|\} as then for both signal states you will guess correctly (1q)+q/2(1-q)+q/2 conditioned on the state sent. As one might expect, in such a situation the communication value of the channel would be multiplicative with itself. As we require an upper bound, we verify this by an exhaustive numerical search using the dual problem of cvPPT\text{cv}^{\text{PPT}}.

Theorem 9.

cv(𝒩p,q2)=cv(𝒩p,q)2\text{cv}(\mathcal{N}_{p,q}^{\otimes 2})=\text{cv}(\mathcal{N}_{p,q})^{2}, i.e. the dephrasure channel’s communication value is multiplicative.

Proof.

A search over the dual problem cvPPT(𝒩p,q2)\text{cv}^{\text{PPT}}(\mathcal{N}_{p,q}^{\otimes 2}) for p,q[0,0.01,,1]p,q\in[0,0.01,...,1] is always within numerical error of cv(𝒩p,q)2\text{cv}(\mathcal{N}_{p,q})^{2}. As the dual problem always obtains an upper bound on cvPPT\text{cv}^{\text{PPT}}, and cvPPT\text{cv}^{\text{PPT}} is an upper bound on cv, we may conclude that the dephrasure channel is multiplicative. ∎

Siddhu Channel

Finally we consider the following family of channels:

𝒩s(X):=i=01KiXKi,\displaystyle\mathcal{N}_{s}(X):=\sum_{i=0}^{1}K_{i}XK_{i}^{\dagger}\ ,

where

K0=s|00|+|21|K1=1s|10|+|22|,\displaystyle K_{0}=\sqrt{s}|0\rangle\langle 0|+|2\rangle\langle 1|\quad K_{1}=\sqrt{1-s}|1\rangle\langle 0|+|2\rangle\langle 2|\ ,

where s[0,1/2]s\in[0,1/2]. This channel is known to have non-additive coherent information over its entire parameter range when tensored with itself. However, we will now show the communication value of the channel is multiplicative with itself over the whole range.

Lemma 3.

cv(𝒩s)=2\text{cv}(\mathcal{N}_{s})=2 for all s[0,1/2]s\in[0,1/2].

Proof.

Like the dephrasure channel, we prove this by constructing upper and lower bounds that are the same.

For a lower bound on cv(𝒩s)\text{cv}(\mathcal{N}_{s}), consider the encoding {|00|,|11|,|22|}\{|0\rangle\langle 0|,|1\rangle\langle 1|,|2\rangle\langle 2|\} and the decoding {|00|+|11|,|22|}\{|0\rangle\langle 0|+|1\rangle\langle 1|,|2\rangle\langle 2|\}. Note that for all s[0,1/2]s\in[0,1/2], 𝒩s(|00|)=s|00|+(1s)|11|\mathcal{N}_{s}(|0\rangle\langle 0|)=s|0\rangle\langle 0|+(1-s)|1\rangle\langle 1| and 𝒩s(|11|)=𝒩s(|22|)=|22|\mathcal{N}_{s}(|1\rangle\langle 1|)=\mathcal{N}_{s}(|2\rangle\langle 2|)=|2\rangle\langle 2|. Thus, with this encoding and decoding, we induce the conditional probability distribution 1=𝐏(0|1)=𝐏(1|2)=𝐏(1|3)1=\mathbf{P}(0|1)=\mathbf{P}(1|2)=\mathbf{P}(1|3) and zero otherwise. Thus we have 2cv32(𝒩s)cv(𝒩s)2\leq\text{cv}^{3\to 2}(\mathcal{N}_{s})\leq\text{cv}(\mathcal{N}_{s}).

For an upper bound, we consider the dual problem of cvPPT\text{cv}^{\text{PPT}} (103). First note that we can write the Choi matrix as:

J(𝒩s)=[s0000s00001s0000001s000000000000000000000000000s0000100000000000000000000001s0000001].\displaystyle J(\mathcal{N}_{s})=\begin{bmatrix}s&0&0&0&0&\sqrt{s}&0&0&0\\ 0&1-s&0&0&0&0&0&0&\sqrt{1-s}\\ 0&0&0&0&0&0&0&0&0\\ 0&0&0&0&0&0&0&0&0\\ 0&0&0&0&0&0&0&0&0\\ \sqrt{s}&0&0&0&0&1&0&0&0\\ 0&0&0&0&0&0&0&0&0\\ 0&0&0&0&0&0&0&0&0\\ 0&\sqrt{1-s}&0&0&0&0&0&0&1\\ \end{bmatrix}\ .

Then we let Y1=s|00|+(1s)|11|+|22|Y_{1}=s|0\rangle\langle 0|+(1-s)|1\rangle\langle 1|+|2\rangle\langle 2| and

Y2=[000000000000000000001α000β000αs000γ000000000000000100000000000000βγ0001s0000000001],\displaystyle Y_{2}=\begin{bmatrix}0&0&0&0&0&0&0&0&0\\ 0&0&0&0&0&0&0&0&0\\ 0&0&1&-\alpha&0&0&0&-\beta&0\\ 0&0&-\alpha&s&0&0&0&\gamma&0\\ 0&0&0&0&0&0&0&0&0\\ 0&0&0&0&0&1&0&0&0\\ 0&0&0&0&0&0&0&0&0\\ 0&0&-\beta&\gamma&0&0&0&1-s&0\\ 0&0&0&0&0&0&0&0&1\\ \end{bmatrix}\ ,

where α=s,β=1s,γ=s(1s)\alpha=\sqrt{s},\beta=\sqrt{1-s},\gamma=\sqrt{s(1-s)}, which is positive semidefinite as it has eigenvalues 22 and 0 with multiplicity eight. It is then easy to determine

𝕀AY1Γ(Y2)J(𝒩s)\displaystyle\mathbb{I}_{A}\otimes Y_{1}-\Gamma(Y_{2})-J(\mathcal{N}_{s})
=\displaystyle= [00000δ00000000000ϵ00000000000000000000001s0γ00δ000000000000γ0s000000000000ϵ0000000],\displaystyle\begin{bmatrix}0&0&0&0&0&\delta&0&0&0\\ 0&0&0&0&0&0&0&0&\epsilon\\ 0&0&0&0&0&0&0&0&0\\ 0&0&0&0&0&0&0&0&0\\ 0&0&0&0&1-s&0&-\gamma&0&0\\ \delta&0&0&0&0&0&0&0&0\\ 0&0&0&0&-\gamma&0&s&0&0\\ 0&0&0&0&0&0&0&0&0\\ 0&\epsilon&0&0&0&0&0&0&0\\ \end{bmatrix}\ ,

where δ=αs,ϵ=β1s\delta=\alpha-\sqrt{s},\epsilon=\beta-\sqrt{1-s}. One may verify that this has eigenvalues of 11 and 0 with multiplicity eight. Thus it is feasible and Tr(Y1)=2\textrm{Tr}(Y_{1})=2. Noting that cv(𝒩s)cvPPT(𝒩s)\text{cv}(\mathcal{N}_{s})\leq\text{cv}^{\text{PPT}}(\mathcal{N}_{s}), we have 2cv(𝒩s)22\leq\text{cv}(\mathcal{N}_{s})\leq 2, which completes the proof. ∎

Theorem 10.

For all s[0,1/2]s\in[0,1/2], cv(𝒩s2)=cv(𝒩s)2\text{cv}(\mathcal{N}_{s}^{\otimes 2})=\text{cv}(\mathcal{N}_{s})^{2}, i.e. the communication value is always multiplicative.

Proof.

A numerical search of cvPPT(𝒩s2)\text{cv}^{\text{PPT}}(\mathcal{N}_{s}^{\otimes 2}) for s[0,0.01,,0.5]s\in[0,0.01,...,0.5] finds the value equals four within an error of 3×106\leq 3\times 10^{-6}. As cvPPT\text{cv}^{\text{PPT}} upper bounds cv, the channel is multiplicative. ∎

VIII Relationship to Capacities and No-Signalling

As noted in the introduction, cv(𝒩)\lceil\text{cv}(\mathcal{N})\rceil captures the classical communication cost to perfectly simulate every classical channel induced by 𝒩\mathcal{N} using non-signalling (NS) resources. This is because for a classical channel 𝐏\mathbf{P}, the one-shot classical communication cost for zero-error simulation with classical NS, κ0NS\kappa_{0}^{\text{NS}}, is given by ymaxxp(y|x)\lceil\sum_{y}\max_{x}p(y|x)\rceil [1, Theorem 16]. Noting that cv(𝐏)=ymaxxp(y|x)\text{cv}(\mathbf{P})=\sum_{y}\max_{x}p(y|x), it follows κ0NS=cv(𝐏)\kappa_{0}^{\text{NS}}=\lceil\text{cv}(\mathbf{P})\rceil. Furthermore, due to the multiplicativity of cv for classical channels, the no-signalling assisted zero-error simulation capacity is also given by κ0NS\kappa_{0}^{\text{NS}}, as was remarked in the original paper. Moreover, it is easy to show the classical capacity of a classical channel is bounded by cv(𝐏)\text{cv}(\mathbf{P}) [1, Remark 17]:

C(𝐏)logcv(𝐏)=χmax(𝐏),C(\mathbf{P})\leq\log\text{cv}(\mathbf{P})=\chi_{\max}(\mathbf{P})\ ,

where we have used Theorem 1 in the last equality. Losing the single-letter property, it is easy to generalize this to arbitrary quantum channels by using the Holevo-Schumacher-Westmoreland theorem [53, 54],

C(𝒩)=\displaystyle C(\mathcal{N})= limk1kχ(𝒩k)\displaystyle\underset{k\to\infty}{\lim}\frac{1}{k}\chi(\mathcal{N}^{\otimes k})
\displaystyle\leq limk1kχmax(𝒩k)\displaystyle\underset{k\to\infty}{\lim}\frac{1}{k}\chi_{\max}(\mathcal{N}^{\otimes k})
=\displaystyle= limk1klog(cv(𝒩k)),\displaystyle\underset{k\to\infty}{\lim}\frac{1}{k}\log(\text{cv}(\mathcal{N}^{\otimes k}))\ ,

and whenever 𝒩\mathcal{N} satisfies weak multiplicativity for cv, such as for entanglement-breaking channels, this reduces to a single-letter upper bound.

In the entanglement-assisted regime, the relationships persist. First we recall that the SDP for min-entropy is multiplicative, and so 𝒞𝒱(𝒩)=cv(𝒩)\mathcal{CV}^{*}(\mathcal{N})=\text{cv}^{*}(\mathcal{N}) for arbitrary quantum channel 𝒩\mathcal{N}. This aligns with the fact the entanglement-assisted capacity of a quantum channel, CE(𝒩)C_{E}(\mathcal{N}), is single-letter but the unassisted capacity is not. Continuing the parallels, cv(𝒩)\lceil\text{cv}^{*}(\mathcal{N})\rceil gives the classical communication cost to perfectly simulate 𝒩\mathcal{N} with a quantum no-signalling resource [2]. Given the above, a natural question is then if one can find bounds on the entanglement-assisted capacity, CE(𝒩)C_{E}(\mathcal{N}), in terms of cv(𝒩)\text{cv}^{*}(\mathcal{N}). Indeed, this can be done by using the definition of cv\text{cv}^{*} and the fact that cv\text{cv}^{*} is characterized by minimal error discrimination (as in Eq. (31)),

cv(𝒩)=supρXAA|𝒳|exp(Hmin(X|BC)(idX𝒩idA)(ρ)),\text{cv}^{*}(\mathcal{N})=\underset{\rho_{XAA^{\prime}}}{\sup}|\mathcal{X}|\exp(-H_{\min}(X|BC)_{(id_{X}\otimes\mathcal{N}\otimes id_{A^{\prime}})(\rho)})\ ,

where the supremum is over ρXAA\rho_{XAA^{\prime}} such that ρX\rho_{X} is uniform and the state is homogenous on register AA^{\prime} [55, 56]. It follows by the same manipulations used in Eq. (31) that

logcv(𝒩)=χE,max(𝒩),\log\text{cv}^{*}(\mathcal{N})=\chi_{E,\max}(\mathcal{N}), (119)

where χE,max\chi_{E,\max} is the entanglement-assisted max\max-Holevo information, which is straightforward to define using [55, 56, 27]. Since the entanglement-assisted capacity equals the regularized entanglement-assisted Holevo information, we can conclude the

CE(𝒩)logcv(𝒩),C_{E}(\mathcal{N})\leq\log\text{cv}^{*}(\mathcal{N})\ ,

where the regularization disappears because cv(𝒩)\text{cv}^{*}(\mathcal{N}) is always multiplicative.

References

  • [1] T. S. Cubitt, D. Leung, W. Matthews, and A. Winter, “Zero-error channel capacity and simulation assisted by non-local correlations,” IEEE Transactions on Information Theory, vol. 57, no. 8, pp. 5509–5523, Aug. 2011.
  • [2] R. Duan and A. Winter, “No-signalling-assisted zero-error capacity of quantum channels and an information theoretic interpretation of the lovász number,” IEEE Transactions on Information Theory, vol. 62, no. 2, pp. 891–914, Feb. 2016.
  • [3] X. Wang and R. Duan, “On the quantum no-signalling assisted zero-error classical simulation cost of non-commutative bipartite graphs,” in 2016 IEEE International Symposium on Information Theory (ISIT).   IEEE, Jul. 2016. [Online]. Available: https://doi.org/10.1109/isit.2016.7541698
  • [4] K. Fang, X. Wang, M. Tomamichel, and M. Berta, “Quantum channel simulation and the channel’s smooth max-information,” IEEE Transactions on Information Theory, vol. 66, no. 4, pp. 2129–2140, Apr. 2020. [Online]. Available: https://doi.org/10.1109/tit.2019.2943858
  • [5] C. Bennett, P. Shor, J. Smolin, and A. Thapliyal, “Entanglement-assisted capacity of a quantum channel and the reverse shannon theorem,” IEEE Transactions on Information Theory, vol. 48, no. 10, pp. 2637–2655, Oct. 2002.
  • [6] C. H. Bennett, I. Devetak, A. W. Harrow, P. W. Shor, and A. Winter, “The quantum reverse shannon theorem and resource tradeoffs for simulating quantum channels,” IEEE Transactions on Information Theory, vol. 60, no. 5, pp. 2926–2959, May 2014. [Online]. Available: https://doi.org/10.1109/tit.2014.2309968
  • [7] M. Berta, M. Christandl, and R. Renner, “The quantum reverse shannon theorem based on one-shot information theory,” Communications in Mathematical Physics, vol. 306, no. 3, pp. 579–615, Aug. 2011. [Online]. Available: https://doi.org/10.1007/s00220-011-1309-7
  • [8] T. Heinosaari and O. Kerppo, “Communication of partial ignorance with qubits,” Journal of Physics A: Mathematical and Theoretical, vol. 52, no. 39, p. 395301, 2019.
  • [9] T. Heinosaari, O. Kerppo, and L. Leppäjärvi, “Communication tasks in operational theories,” Journal of Physics A: Mathematical and Theoretical, vol. 53, no. 43, p. 435302, Oct. 2020.
  • [10] P. E. Frenkel and M. Weiner, “Classical information storage in an n-level quantum system,” Communications in Mathematical Physics, vol. 340, no. 2, pp. 563–574, Sep. 2015.
  • [11] M. Berta, F. G. S. L. Brandao, M. Christandl, and S. Wehner, “Entanglement cost of quantum channels,” IEEE Transactions on Information Theory, vol. 59, no. 10, pp. 6779–6795, Oct. 2013.
  • [12] X. Wang and M. M. Wilde, “Exact entanglement cost of quantum states and channels under ppt-preserving operations,” 2018.
  • [13] M. M. Wilde, “Entanglement cost and quantum channel simulation,” Phys. Rev. A, vol. 98, p. 042338, Oct 2018. [Online]. Available: https://link.aps.org/doi/10.1103/PhysRevA.98.042338
  • [14] G. Gour and C. M. Scandolo, “Entanglement of a bipartite channel,” Phys. Rev. A, vol. 103, p. 062422, Jun 2021. [Online]. Available: https://link.aps.org/doi/10.1103/PhysRevA.103.062422
  • [15] P. E. Frenkel and M. Weiner, “On entanglement assistance to a noiseless classical channel,” 2021.
  • [16] M. Dall’Arno, S. Brandsen, A. Tosini, F. Buscemi, and V. Vedral, “No-hypersignaling principle,” Phys. Rev. Lett., vol. 119, p. 020401, Jul 2017. [Online]. Available: https://link.aps.org/doi/10.1103/PhysRevLett.119.020401
  • [17] B. Doolittle and E. Chitambar, “Certifying the classical simulation cost of a quantum channel,” 2021.
  • [18] C. H. Bennett and S. J. Wiesner, “Communication via one- and two-particle operators on einstein-podolsky-rosen states,” Phys. Rev. Lett., vol. 69, pp. 2881–2884, Nov 1992. [Online]. Available: https://link.aps.org/doi/10.1103/PhysRevLett.69.2881
  • [19] R. Konig, R. Renner, and C. Schaffner, “The operational meaning of min- and max-entropy,” IEEE Transactions on Information Theory, vol. 55, no. 9, pp. 4337–4347, Sep. 2009.
  • [20] E. Davies, “Information and quantum measurement,” IEEE Transactions on Information Theory, vol. 24, no. 5, pp. 596–599, Sep. 1978. [Online]. Available: https://doi.org/10.1109/tit.1978.1055941
  • [21] C. W. Helstrom, Quantum detection and estimation theory.   Academic Press, New York, 1976.
  • [22] L. Gurvits, “Classical deterministic complexity of edmonds’ problem and quantum entanglement,” in Proceedings of the thirty-fifth ACM symposium on Theory of computing - STOC ’03.   ACM Press, 2003. [Online]. Available: https://doi.org/10.1145/780542.780545
  • [23] A. Peres, “Separability criterion for density matrices,” Phys. Rev. Lett., vol. 77, pp. 1413–1415, Aug 1996. [Online]. Available: https://link.aps.org/doi/10.1103/PhysRevLett.77.1413
  • [24] M. Horodecki, P. Horodecki, and R. Horodecki, “Separability of mixed states: necessary and sufficient conditions,” Physics Letters A, vol. 223, no. 1-2, pp. 1–8, Nov. 1996. [Online]. Available: https://doi.org/10.1016/s0375-9601(96)00706-2
  • [25] M. Tomamichel, Quantum information processing with finite resources: mathematical foundations.   Springer, 2015, vol. 5.
  • [26] M. M. Wilde, A. Winter, and D. Yang, “Strong converse for the classical capacity of entanglement-breaking and hadamard channels via a sandwiched rényi relative entropy,” Communications in Mathematical Physics, vol. 331, no. 2, pp. 593–622, Jul. 2014. [Online]. Available: https://doi.org/10.1007/s00220-014-2122-x
  • [27] S. Beigi, “Sandwiched rényi divergence satisfies data processing inequality,” Journal of Mathematical Physics, vol. 54, no. 12, p. 122202, Dec. 2013.
  • [28] N. Datta, “Min- and max-relative entropies and a new entanglement monotone,” IEEE Transactions on Information Theory, vol. 55, no. 6, pp. 2816–2826, Jun. 2009.
  • [29] C. H. Bennett, D. P. DiVincenzo, J. A. Smolin, and W. K. Wootters, “Mixed-state entanglement and quantum error correction,” Phys. Rev. A, vol. 54, pp. 3824–3851, Nov 1996. [Online]. Available: https://link.aps.org/doi/10.1103/PhysRevA.54.3824
  • [30] A. Shimony, “Degree of entanglement,” Annals of the New York Academy of Sciences, vol. 755, no. 1, pp. 675–679, Apr. 1995.
  • [31] T.-C. Wei and P. M. Goldbart, “Geometric measure of entanglement and applications to bipartite and multipartite quantum states,” Phys. Rev. A, vol. 68, p. 042307, Oct 2003.
  • [32] R. F. Werner and A. S. Holevo, “Counterexample to an additivity conjecture for output purity of quantum channels,” Journal of Mathematical Physics, vol. 43, no. 9, pp. 4353–4357, Sep. 2002.
  • [33] H. Zhu, L. Chen, and M. Hayashi, “Additivity and non-additivity of multipartite entanglement measures,” New Journal of Physics, vol. 13, no. 1, p. 019501, Jan. 2011.
  • [34] E. Jung, M.-R. Hwang, H. Kim, M.-S. Kim, D. Park, J.-W. Son, and S. Tamaryan, “Reduced state uniquely defines the groverian measure of the original pure state,” Physical Review A, vol. 77, no. 6, p. 062317, 2008.
  • [35] D. Leung and W. Matthews, “On the power of ppt-preserving and non-signalling codes,” IEEE Transactions on Information Theory, vol. 61, no. 8, pp. 4486–4499, 2015.
  • [36] R. F. Werner, “Quantum states with einstein-podolsky-rosen correlations admitting a hidden-variable model,” Physical Review A, vol. 40, no. 8, pp. 4277–4281, Oct. 1989. [Online]. Available: https://doi.org/10.1103/physreva.40.4277
  • [37] K. G. H. Vollbrecht and R. F. Werner, “Entanglement measures under symmetry,” Physical Review A, vol. 64, no. 6, p. 062307, 2001.
  • [38] M. Horodecki, P. Horodecki, and R. Horodecki, “General teleportation channel, singlet fraction, and quasidistillation,” Phys. Rev. A, vol. 60, pp. 1888–1898, Sep 1999.
  • [39] M. Christandl, N. Schuch, and A. Winter, “Entanglement of the antisymmetric state,” Communications in Mathematical Physics, vol. 311, no. 2, pp. 397–422, Mar. 2012.
  • [40] R. Hübener, M. Kleinmann, T.-C. Wei, C. González-Guillén, and O. Gühne, “Geometric measure of entanglement for symmetric states,” Phys. Rev. A, vol. 80, p. 032324, Sep 2009.
  • [41] M. A. Nielsen, “Conditions for a class of entanglement transformations,” Phys. Rev. Lett., vol. 83, pp. 436–439, Jul 1999.
  • [42] H.-K. Lo and S. Popescu, “Concentrating entanglement by local actions: Beyond mean values,” Phys. Rev. A, vol. 63, p. 022301, Jan 2001.
  • [43] A. C. Doherty, P. A. Parrilo, and F. M. Spedalieri, “Complete family of separability criteria,” Physical Review A, vol. 69, no. 2, p. 022308, 2004.
  • [44] C. King, “Maximal p-norms of entanglement breaking channels,” Quantum Information and Computation, vol. 3, no. 2, pp. 186–190, 2003.
  • [45] B. Doolittle and I. George, “Cvchannel.jl,” Sep. 2021. [Online]. Available: https://github.com/ChitambarLab/CVChannel.jl
  • [46] J. Bezanson, A. Edelman, S. Karpinski, and V. B. Shah, “Julia: A fresh approach to numerical computing,” SIAM review, vol. 59, no. 1, pp. 65–98, 2017. [Online]. Available: https://doi.org/10.1137/141000671
  • [47] M. Udell, K. Mohan, D. Zeng, J. Hong, S. Diamond, and S. Boyd, “Convex optimization in Julia,” SC14 Workshop on High Performance Technical Computing in Dynamic Languages, 2014.
  • [48] B. O’Donoghue, E. Chu, N. Parikh, and S. Boyd, “SCS: Splitting conic solver, version 2.1.4,” https://github.com/cvxgrp/scs, Nov. 2019.
  • [49] M. Reimpell and R. F. Werner, “Iterative optimization of quantum error correcting codes,” Physical Review Letters, vol. 94, no. 8, Mar. 2005. [Online]. Available: https://doi.org/10.1103/physrevlett.94.080501
  • [50] R. L. Kosut and D. A. Lidar, “Quantum error correction via convex optimization,” Quantum Information Processing, vol. 8, no. 5, pp. 443–459, Jul. 2009. [Online]. Available: https://doi.org/10.1007/s11128-009-0120-2
  • [51] F. Leditzky, D. Leung, and G. Smith, “Dephrasure channel and superadditivity of coherent information,” Physical review letters, vol. 121, no. 16, p. 160501, 2018.
  • [52] V. Siddhu, “Log-singularities for studying capacities of quantum channels,” arXiv preprint arXiv:2003.10367, 2020.
  • [53] A. Holevo, “The capacity of the quantum channel with general signal states,” IEEE Transactions on Information Theory, vol. 44, no. 1, pp. 269–273, 1998. [Online]. Available: https://doi.org/10.1109/18.651037
  • [54] B. Schumacher and M. D. Westmoreland, “Sending classical information via noisy quantum channels,” Phys. Rev. A, vol. 56, pp. 131–138, Jul 1997.
  • [55] A. S. Holevo, “On entanglement-assisted classical capacity,” Journal of Mathematical Physics, vol. 43, no. 9, pp. 4326–4333, 2002.
  • [56] J. Watrous, The Theory of Quantum Information.   Cambridge university press, 2018.