arxiv: v2 [quant-ph] 18 May 2018

Similar documents
arxiv: v2 [quant-ph] 8 Mar 2018

arxiv: v1 [quant-ph] 3 Jan 2008

Exact entanglement cost of quantum states and channels under PPT-preserving operations

Semidefinite programming strong converse bounds for quantum channel capacities

Semidefinite programming strong converse bounds for quantum channel capacities

Lecture 6: Quantum error correction and quantum capacity

Lecture 19 October 28, 2015

Squashed entanglement

Recoverability in quantum information theory

Compression and entanglement, entanglement transformations

Universal recoverability in quantum information theory

Converse bounds for private communication over quantum channels

Capacity Estimates of TRO Channels

arxiv: v4 [quant-ph] 28 Oct 2009

A Holevo-type bound for a Hilbert Schmidt distance measure

Strong converse theorems using Rényi entropies

Multiplicativity of Maximal p Norms in Werner Holevo Channels for 1 < p 2

Classical and Quantum Channel Simulations

Attempting to reverse the irreversible in quantum physics

On the complementary quantum capacity of the depolarizing channel

On the complementary quantum capacity of the depolarizing channel

Single-shot Quantum State Merging

Teleportation of Quantum States (1993; Bennett, Brassard, Crepeau, Jozsa, Peres, Wootters)

Entanglement Measures and Monotones

Limitations on separable measurements by convex optimization

Bound entangled states with secret key and their classical counterpart

Entanglement: Definition, Purification and measures

Lecture 11 September 30, 2015

Maximal Entanglement A New Measure of Entanglement

Entanglement: concept, measures and open problems

Entropy in Classical and Quantum Information Theory

Lecture 8: Semidefinite programs for fidelity and optimal measurements

Distinguishing multi-partite states by local measurements

arxiv: v1 [quant-ph] 28 May 2011

Fundamental rate-loss tradeoff for optical quantum key distribution

On balance of information in bipartite quantum communication systems: entanglement-energy analogy

Lecture 21: HSP via the Pretty Good Measurement

The first law of general quantum resource theories

Quantum state discrimination with post-measurement information!

Is Entanglement Sufficient to Enable Quantum Speedup?

The Minimax Fidelity for Quantum Channels: Theory and Some Applications

THE fundamental physical resource for performing various

Limitations on Quantum Key Repeaters

Degradable Quantum Channels

Quantum Teleportation Pt. 3

Quantum rate distortion, reverse Shannon theorems, and source-channel separation

arxiv:quant-ph/ v1 9 May 2005

arxiv: v5 [quant-ph] 5 Jun 2018

Sending Quantum Information with Zero Capacity Channels

The Quantum Reverse Shannon Theorem Based on One-Shot Information Theory

Quantum Hadamard channels (II)

Application of Structural Physical Approximation to Partial Transpose in Teleportation. Satyabrata Adhikari Delhi Technological University (DTU)

Fidelity of Quantum Teleportation through Noisy Channels

Technical Report Communicating Secret Information Without Secret Messages

arxiv:quant-ph/ v1 13 Mar 2007

Probabilistic exact cloning and probabilistic no-signalling. Abstract

arxiv: v3 [quant-ph] 5 Jun 2015

Quantum key distribution based on private states

On Composite Quantum Hypothesis Testing

Teleporting an Unknown Quantum State Via Dual Classical and Einstein Podolsky Rosen Channels 1

Asymptotic Analysis of a Three State Quantum Cryptographic Protocol

Unambiguous Discrimination Between Linearly Dependent States With Multiple Copies

Lecture 7: Semidefinite programming

Multivariate trace inequalities. David Sutter, Mario Berta, Marco Tomamichel

Some Bipartite States Do Not Arise from Channels

Entanglement and information

Quantum Information Processing and Diagrams of States

Permutations and quantum entanglement

Strong converse theorems using Rényi entropies

Multiparty Quantum Remote Control

Quantum entanglement and symmetry

1 Quantum states and von Neumann entropy

Distinguishing multi-partite states by local measurements

Quantum Error Correcting Codes and Quantum Cryptography. Peter Shor M.I.T. Cambridge, MA 02139

1. Basic rules of quantum mechanics

THE NUMBER OF ORTHOGONAL CONJUGATIONS

Local cloning of entangled states

Lecture 21: Quantum communication complexity

Trading classical communication, quantum communication, and entanglement in quantum Shannon theory

A Hierarchy of Information Quantities for Finite Block Length Analysis of Quantum Tasks

ACENTRAL goal of information theory is the (quantitative)

Separation between quantum Lovász number and entanglement-assisted zero-error classical capacity

Lecture 4: Purifications and fidelity

CS/Ph120 Homework 4 Solutions

Theory of Quantum Entanglement

Instantaneous Nonlocal Measurements

Quantum Entanglement- Fundamental Aspects

Quantum Information Theory Tutorial

Channel Steering. Marco Piani 1, 2. University of Waterloo, N2L 3G1 Waterloo ON, Canada. compiled: December 23, 2014

Basics on quantum information

Approximate reversal of quantum Gaussian dynamics

arxiv:quant-ph/ v2 17 Jun 1996

Geometry of Entanglement

On PPT States in C K C M C N Composite Quantum Systems

Probabilistic Teleportation of an Arbitrary Two-Qubit State via Positive Operator-Valued Measurement with Multi Parties

Quantum Gates, Circuits & Teleportation

. Here we are using the standard inner-product over C k to define orthogonality. Recall that the inner-product of two vectors φ = i α i.

Lecture: Quantum Information

Homework 3 - Solutions

On positive maps in quantum information.

Transcription:

Amortization does not enhance the max-rains information of a quantum channel Mario Berta Mark M. Wilde arxiv:1709.04907v2 [quant-ph] 18 May 2018 May 21, 2018 Abstract Given an entanglement measure E, the entanglement of a quantum channel is defined as the largest amount of entanglement E that can be generated from the channel, if the sender and receiver are not allowed to share a quantum state before using the channel. The amortized entanglement of a quantum channel is defined as the largest net amount of entanglement E that can be generated from the channel, if the sender and receiver are allowed to share an arbitrary state before using the channel. Our main technical result is that amortization does not enhance the entanglement of an arbitrary quantum channel, when entanglement is quantified by the max-rains relative entropy. We prove this statement by employing semi-definite programming (SDP) duality and SDP formulations for the max-rains relative entropy and a channel s max- Rains information, found recently in [Wang et al., arxiv:1709.00200]. The main application of our result is a single-letter, strong-converse, and efficiently computable upper bound on the capacity of a quantum channel for transmitting qubits when assisted by positive-partialtranspose preserving (PPT-P) channels between every use of the channel. As the class of local operations and classical communication (LOCC) is contained in PPT-P, our result establishes a benchmark for the LOCC-assisted quantum capacity of an arbitrary quantum channel, which is relevant in the context of distributed quantum computation and quantum key distribution. 1 Introduction One of the main goals of quantum information theory is to understand the fundamental limitations on communication when a sender and receiver are connected by a quantum communication channel [Hol12, Hay06, Wil16a]. Since it might be difficult to transmit information reliably by making use of a channel just once, a practically relevant setting is when the sender and receiver use the channel multiple times, with the goal being to maximize the rate of communication subject to a constraint on the error probability. The capacity of a quantum channel is defined to be the maximum rate of reliable communication, such that the error probability tends to zero in the limit when the channel is utilized an arbitrary number of times. Among the various capacities of a quantum channel N, the LOCC-assisted quantum capacity Q (N ) [BDSW96] is particularly relevant for tasks such as distributed quantum computation. In Department of Computing, Imperial College London, London SW7 2AZ, UK, and Institute for Quantum Information and Matter, California Institute of Technology, Pasadena, California 91125, USA Hearne Institute for Theoretical Physics, Department of Physics and Astronomy, Center for Computation and Technology, Louisiana State University, Baton Rouge, Louisiana 70803, USA 1

the setting corresponding to this capacity, the sender and receiver are allowed to perform arbitrary LOCC (local operations and classical communication) between every use of the channel, and the capacity is equal to the maximum rate, measured in qubits per channel use, at which qubits can be transmitted reliably from the sender to the receiver [BDSW96]. Due to the teleportation protocol [BBC + 93], this rate is equal to the maximum rate at which shared entangled bits (Bell pairs) can be generated reliably between the sender and the receiver [BDSW96]. The LOCC-assisted quantum capacity of certain channels such as the quantum erasure channel has been known for some time [BDS97], but in general, it remains an open question to characterize Q (N ). One can address this question by establishing either lower bounds or upper bounds on Q (N ). In this paper, we are interested in placing upper bounds on the LOCC-assisted quantum capacity, and one way of simplifying the mathematics behind this task is to relax the class of free operations that the sender and receiver are allowed to perform between each channel use. With this in mind, we follow the approach of [Rai99, Rai01] and relax the set LOCC to a larger class of operations known as PPT-preserving (PPT-P), standing for channels that are positive partial transpose preserving. The resulting capacity is then known as the PPT-P-assisted quantum capacity Q PPT-P, (N ), and it is equal to the maximum rate at which qubits can be communicated reliably from a sender to a receiver, when they are allowed to use a PPT-preserving channel in between every use of the actual channel N. Figure 1 provides a visualization of such a PPT-P-assisted quantum communication protocol. Due to the containment LOCC PPT-P [Rai99, Rai01], the inequality Q (N ) Q PPT-P, (N ) (1) holds for all channels N. Thus, if we find an upper bound on Q PPT-P, (N ), then by (1), such an upper bound also bounds the physically relevant LOCC-assisted quantum capacity Q (N ). A general approach for bounding these assisted capacities of a quantum channel has been developed recently in [KW18] (see [BHLS03, LHL03, CMH17, BDGDMW17, RKB + 18] for related notions). The starting point is to consider an entanglement measure E(A; B) ρ [HHHH09], which is evaluated for a bipartite state ρ AB. Given such an entanglement measure, one can define the entanglement E(N ) of a channel N in terms of it by taking an optimization over all pure, bipartite states that could be input to the channel: E(N ) = sup ψ RA E(R; B) ω, (2) where ω RB = N A B (ψ RA ). The channel s entanglement E(N ) characterizes the amount of entanglement that a sender and receiver can generate by using the channel if they do not share entanglement prior to its use. Due to the properties of an entanglement measure and the well known Schmidt decomposition theorem, it suffices to take system R isomorphic to the channel input system A and furthermore to optimize over pure states ψ RA. One can alternatively consider the amortized entanglement E A (N ) of a channel N as the following optimization [KW18]: E A (N ) = sup ρ A AB [ E(A ; BB ) τ E(A A; B ] ) ρ, (3) where τ A BB = N A B(ρ A AB ) and ρ A AB is a state. The supremum is with respect to all states ρ A AB and the systems A B are finite-dimensional but could be arbitrarily large (so that the supremum might never be achieved for any particular finite-dimensional A B, but only in the limit 2

A1 A2 A n M A PPT-P A1 N B1 PPT-P A2 N B2 PPT-P PPT-P A n N B n PPT-P B1 B2 B n M B Figure 1: A protocol for PPT-P-assisted quantum communication that uses a quantum channel n times. Every channel use is interleaved by a PPT-preserving channel. The goal of such a protocol is to produce an approximate maximally entangled state in the systems M A and M B, where Alice possesses system M A and Bob system M B. of unbounded dimension). Thus, E A (N ) is not known to be computable in general. The amortized entanglement quantifies the net amount of entanglement that can be generated by using the channel N, if the sender and receiver are allowed to begin with some initial entanglement in the form of the state ρ A AB. That is, E(A A; B ) ρ quantifies the entanglement of the initial state ρ A AB, and E(A ; BB ) τ quantifies the final entanglement of the state after the channel acts. As observed in [KW18], the inequality E(N ) E A (N ) (4) always holds for any entanglement measure E and for any channel N, simply because one could take the B system trivial in the optimization for E A (N ), which is the same as not allowing entanglement between the sender and receiver before the channel acts. It is nontrivial if the opposite inequality E A (N )? E(N ) (5) holds, which is known to occur generally for certain entanglement measures [TGW14, CMH17, KW18] or for certain channels with particular symmetries [KW18]. One of the main observations of [KW18], connected to earlier developments in [BHLS03, LHL03, CMH17, BDGDMW17, RKB + 18], is that the amortized entanglement of a channel serves as an upper bound on the entanglement of the final state ω AB generated by an LOCC- or PPT-P-assisted quantum communication protocol that uses the channel n times: E(A; B) ω ne A (N ). (6) The basic intuition for why this bound holds is that, after a given channel use, the sender and receiver are allowed to perform a free operation such as LOCC or PPT, and thus the state that they share before the next channel use could have some entanglement. So the amount of entanglement generated by each channel use cannot exceed the amortized entanglement E A (N ), and if the channel is used n times in such a protocol, then the entanglement of the final state ω AB cannot exceed the channel s amortized entanglement multiplied by the number n of channel uses. Such a general bound can then be used to derive particular upper bounds on the assisted quantum capacities, such as strong converse bounds. Clearly, if the inequality in (5) holds, then E A (N ) = E(N ) and the upper bound becomes much simpler because the channel entanglement E(N ) is simpler than 3

the amortized entanglement E A (N ). Thus, one of the main contributions of [KW18] was to reduce the physical question of determining meaningful upper bounds on the assisted capacities of N to a purely mathematical question of whether amortization can enhance the entanglement of a channel, i.e., whether the equality E A (N )? = E(N ) (7) holds for a given entanglement measure E and/or channel N. Furthermore, it was shown in [KW18] how to incorporate the previous results of [BDSW96, MH12, Pir17] into the amortization framework of [KW18]. In this paper, we solve the mathematical question posed above for the max-rains information R max (N ) of a quantum channel N, by proving that amortization does not enhance it; i.e., we prove that R max,a (N ) = R max (N ), (8) for all channels N, where R max,a (N ) denotes the amortized max-rains information. Note that R max (N ) and R max,a (N ) are respectively defined by taking the entanglement measure E in (2) and (3) to be the max-rains relative entropy, which we define formally in the next section. We note here that the equality in (8) solves an open question posed in the conclusion of [CMH17], and we set our result in the context of the prior result of [CMH17] and other literature in Section 6. The max-rains information of a quantum channel is a special case of a quantity known as the sandwiched Rényi-Rains information [TWW17] and was recently shown to be equal to an information quantity discussed in [WD16b, WFD17] and based on semi-definite programming. To prove our main technical result (the equality in (8)), we critically make use of the tools and framework developed in the recent works [WD16b, WD16a, WFD17]. In particular, we employ semi-definite programming duality [BV04] and the well known Choi isomorphism to establish our main result, with the proof consisting of just a few lines once the framework from [WD16b, WD16a, WFD17] is set in place. The main application of the equality in (8) is an efficiently computable, single-letter, strong converse bound on Q PPT-P, (N ), the PPT-P-assisted quantum capacity of an arbitrary channel N. Due to (1), this is also an upper bound on the physically relevant LOCC-assisted quantum capacity Q (N ). To arrive at this result, we simply apply the general inequality in (6) along with the equality in (8). For the benefit of the reader, we give technical details of this application in Section 4. The quantity R max (N ) has already been shown in [WFD17] to be efficiently computable via a semi-definite program, and in Section 4, we explain how R max (N ) is both single-letter and strong converse. The usefulness of the upper bound given in our paper is ultimately related with the importance of PPT-preserving channels. This is because the set of PPT-preserving channels contains the set of separable channels, and the set of separable channels strictly contains the set of LOCC channels, as shown in [BDF + 99] and then in [KANI13] for a classical scenario. Moreover, there is an entanglement monotone that can be increased by separable channels [CD09]. Thus, in general, PPT-preserving channels can increase entanglement, although this increase is not detectable by the max-rains information. Thus, in this sense, the max-rains information might be considered a rough measure for bounding LOCC-assisted quantum capacity. Therefore, as stressed earlier, the usefulness of our bound on the PPT-P assisted quantum capacity is directly related to PPTpreserving channels. Our paper is organized as follows. In the next section, we review some background material before starting with the main development. Section 3 gives a short proof of our main technical 4

result, and Section 4 discusses its application as an efficiently computable, single-letter, strong converse bound on Q PPT, (N ). In Section 5, we revisit a result from [CMH17], in which it was shown that amortization does not enhance a channel s max-relative entropy of entanglement. The authors of [CMH17] proved this statement by employing complex interpolation theory [BL76]. We prove the main inequality underlying this statement using a method different from that used in [CMH17], but along the lines of that given for our proof of (8) (i.e., convex programming duality), and we suspect that our alternative approach could be useful in future applications. In Section 6, we discuss how our result fits into the prior literature on assisted quantum capacities and strong converses. We conclude with a brief summary in Section 7. 2 Background and notation In this section, we provide background on the Choi isomorphism, partial transpose, positive partial transpose (PPT) states, separable states, PPT-preserving channels, max-relative entropy, max- Rains relative entropy, and max-rains information. For basic concepts and standard notation used in quantum information theory, we point the reader to [Wil16a]. The Choi isomorphism represents a well known duality between channels and states, often employed in quantum information theory. Let N A B be a quantum channel, and let Υ RA denote the maximally entangled vector Υ RA = i R i A, (9) i where the Hilbert spaces H R and H A are of the same dimension and { i R } i and { i A } i are fixed orthonormal bases. The Choi operator for a channel N A B is defined as J N RB = (id R N A B )( Υ Υ RA ), (10) where id R denotes the identity map on system R. One can recover the action of the channel N A B on an arbitrary input state ρ SA as follows: Υ A R ρ SA J N RB Υ A R = N A B (ρ SA ), (11) where A is a system isomorphic to the channel input A. The above identity can be understood in terms of a postselected variant [HM04, Ben05] of the quantum teleportation protocol [BBC + 93]. Another identity we recall is that Υ RA (X SR I A ) Υ RA = Tr R {X SR }, (12) for an operator X SR acting on H S H R. For a fixed basis { i B } i, the partial transpose is the following map: (id A T B )(X AB ) = i,j (I A i j B ) X AB (I A i j B ), (13) where X AB is an arbitrary operator acting on a tensor-product Hilbert space H A H B. For simplicity we often employ the abbreviation T B (X AB ) = (id A T B )(X AB ). The partial transpose map plays a role in the following well known transpose trick identity: (X SR I A ) Υ RA = (T A (X SA ) I R ) Υ RA. (14) 5

The partial transpose map plays another important role in quantum information theory because a separable (unentangled) state σ AB = x p(x)τ x A ω x B SEP(A:B), (15) for a distribution p(x) and states τa x and ωx B, stays within the set of separable states under this map [HHH96, Per96]: T B (σ AB ) SEP(A:B). (16) This motivates defining the set of PPT states, which are those states σ AB for which T B (σ AB ) 0. This in turn motivates defining the more general set of positive semi-definite operators [ADMVW02]: PPT (A:B) = {σ AB : σ AB 0 T B (σ AB ) 1 1}, (17) where we have employed the trace norm, defined for an operator X as X 1 = Tr{ X } with X = X X. We then have the containments SEP PPT PPT. An LOCC quantum channel N AB A B consists of an arbitrarily large but finite number of compositions of the following: 1. Alice performs a quantum instrument, which has both a quantum and classical output. She forwards the classical output to Bob, who then performs a quantum channel conditioned on the classical data received. This sequence of actions corresponds to a channel of the following form: FA A x Gx B B, (18) x where {F x A A } x is a collection of completely positive maps such that x F x A A is a quantum channel and {G x B B } x is a collection of quantum channels. 2. The situation is reversed, with Bob performing the initial instrument, who forwards the classical data to Alice, who then performs a quantum channel conditioned on the classical data. This sequence of actions corresponds to a channel of the form in (18), with the A and B labels switched. A quantum channel N AB A B is a PPT-preserving channel if the map T B N AB A B T B is a quantum channel [Rai99, Rai01]. Any LOCC channel is a PPT-preserving channel [Rai99, Rai01]. The max-relative entropy of a state ρ relative to a positive semi-definite operator σ is defined as [Dat09] D max (ρ σ) = inf{λ : ρ 2 λ σ}. (19) If supp(ρ) supp(σ), then D max (ρ σ) =. The max-relative entropy is monotone non-increasing under the action of a quantum channel N [Dat09], in the sense that D max (ρ σ) D max (N (ρ) N (σ)). (20) The above inequality is also called the data-processing inequality for max-relative entropy. The max-rains relative entropy of a state ρ AB is defined as R max (A; B) ρ = min D max (ρ AB σ AB ), (21) σ AB PPT (A:B) 6

and it is monotone non-increasing under the action of a PPT-preserving quantum channel N AB A B [TWW17], in the sense that R max (A; B) ρ R max (A ; B ) ω, (22) for ω A B = N AB A B (ρ AB). The max-rains information of a quantum channel N A B is defined by replacing E in (2) with the max-rains relative entropy R max ; i.e., R max (N ) = max φ SA R max (S; B) ω, (23) where ω SB = N A B (φ SA ) and φ SA is a pure state, with S = A. The amortized max-rains information of a channel, denoted as R max,a (N ), is defined by replacing E in (3) with the max- Rains relative entropy R max. Recently, in [WD16a, Eq. (8)] (see also [WFD17, Eq. (36)]), the max-rains relative entropy of a state ρ AB was expressed as R max (A; B) ρ = log 2 W (A; B) ρ, (24) where W (A; B) ρ is the solution to the following semi-definite program: minimize Tr{C AB + D AB } subject to C AB, D AB 0, T B (C AB D AB ) ρ AB. (25) Similarly, in [WFD17, Eq. (21)], the max-rains information of a quantum channel N A B was expressed as R max (N ) = log Γ(N ), (26) where Γ(N ) is the solution to the following semi-definite program: minimize Tr B {V SB + Y SB } subject to Y SB, V SB 0, T B (V SB Y SB ) J N SB. (27) These formulations of R max (A; B) ρ and R max (N ) are the tools that we use to prove our main technical result, Proposition 1. It is worthwhile to mention that the formulations above follow by employing the theory of semi-definite programming and its duality. 3 Main technical result The following proposition constitutes our main technical result, and an immediate corollary of it is that amortization does not enhance the max-rains information of a quantum channel: Proposition 1 Let ρ A AB be a state and let N A B be a quantum channel. Then R max (A ; BB ) ω R max (N ) + R max (A A; B ) ρ, (28) where ω A BB = N A B(ρ A AB ). (29) 7

Proof. By removing logarithms and applying (24) and (26), the desired inequality is equivalent to the following one: W (A ; BB ) ω Γ(N ) W (A A; B ) ρ, (30) and so we aim to prove this one. Exploiting the identity in (25), we find that subject to the constraints while the identity in (27) gives that subject to the constraints W (A A; B ) ρ = min Tr{C A AB + D A AB }, (31) C A AB, D A AB 0, (32) T B (C A AB D A AB ) ρ A AB, (33) Γ(N ) = min Tr B {V SB + Y SB }, (34) Y SB, V SB 0, (35) T B (V SB Y SB ) J N SB. (36) The identity in (25) implies that the left-hand side of (30) is equal to subject to the constraints W (A ; BB ) ω = min Tr{E A BB + F A BB }, (37) E A BB, F A BB 0, (38) N A B (ρ A AB ) T BB (E A BB F A BB ). (39) With these SDP formulations in place, we can now establish the inequality in (30) by making judicious choices for E A BB and F A BB. Let C A AB and D A AB be optimal for W (A A; B ) ρ, and let Y SB and V SB be optimal for Γ(N ). Let Υ SA be the maximally entangled vector, as defined in (9). Pick E A BB = Υ SAC A AB V SB + D A AB Y SB Υ SA, F A BB = Υ SAC A AB Y SB + D A AB V SB Υ SA. We note that these choices are somewhat similar to those made in the proof of [WFD17, Proposition 6], and they can be understood roughly via (11) as a postselected teleportation of the optimal operators of W (A A; B ) ρ through the optimal operators of Γ(N ), with the optimal operators of W (A A; B ) ρ being in correspondence with the input state ρ A AB through (33) and the optimal operators of Γ(N ) being in correspondence with the Choi operator JSB N through (36). We then have that E A BB, F A BB 0 because C A AB, D A AB, Y SB, V SB 0. Consider that T BB (E A BB F A BB ) = T BB [ Υ SA(C A AB D A AB ) (V SB Y SB ) Υ SA ] = Υ SA T B (C A AB D A AB ) T B(V SB Y SB ) Υ SA Υ SA ρ A AB J SB Υ N SA = N A B (ρ A AB ). (40) 8

The inequality follows from (33) and (36), and the last equality follows from (11). Also consider that Tr{E A BB + F A BB } = Tr{ Υ SA(C A AB + D A AB ) (V SB + Y SB ) Υ SA } = Tr{(C A AB + D A AB )T A(V AB + Y AB )} = Tr{(C A AB + D A AB )T A(Tr B {V AB + Y AB })} Tr{C A AB + D A AB } T A(Tr B {V AB + Y AB }) = Tr{C A AB + D A AB } Tr B {V AB + Y AB } = W (A A; B ) ρ Γ(N ). (41) The second equality follows from (14) and (12). The inequality is a consequence of Hölder s inequality. The final equality follows because the spectrum of an operator is invariant under the action of a (full) transpose (note, in this case, that T A is a full transpose because the operator Tr B {V AB + Y AB } acts only on system A). Thus, we can conclude that our choices of E A BB and F A BB are feasible for W (A ; BB ) ω. Since W (A ; BB ) ω involves a minimization over all E A BB and F A BB satisfying (38) and (39), this concludes our proof of (30). An immediate corollary of Proposition 1 is the following: Corollary 2 Amortization does not enhance the max-rains information of a quantum channel N A B ; i.e., the following equality holds R max,a (N ) = R max (N ). (42) Proof. The inequality R max,a (N ) R max (N ) always holds, as reviewed in (4). The other inequality is an immediate consequence of Proposition 1. Letting ρ A AB denote an arbitrary input state, Proposition 1 implies that R max (A ; BB ) ω R max (A A; B ) ρ R max (N ), (43) where ω A BB = N A B(ρ A AB ). Since the inequality holds for any state ρ A AB, it holds for the supremum over all such input states, leading to R max,a (N ) R max (N ). 4 Application to PPT-P-assisted quantum communication We now give our main application of Proposition 1, which is that the max-rains information is a single-letter, strong-converse upper bound on the PPT-P-assisted quantum capacity of any channel. The term single-letter refers to the fact that the max-rains information requires an optimization over a single use of the channel. As we remarked previously, the max-rains information is efficiently computable via semi-definite programming, as observed in [WD16b, WFD17]. Finally, the bound is a strong converse bound because, as we will show, if the rate of a sequence of PPT-P-assisted quantum communication protocols exceeds the max-rains information, then the error probability of these protocols necessarily tends to one exponentially fast in the number of channel uses. 9

4.1 Protocol for PPT-P-assisted quantum communication We begin by reviewing the structure of a PPT-P-assisted quantum communication protocol, along the lines discussed in [KW18]. In such a protocol, a sender Alice and a receiver Bob are spatially separated and connected by a quantum channel N A B. They begin by performing a PPT-P channel, which leads to a PPT state ρ (1) A 1 A where A 1B 1, 1 and B 1 are systems that are finitedimensional but arbitrarily large. The system A 1 is such that it can be fed into the first channel use. P (1) A 1 A 1B 1 Alice sends system A 1 through the first channel use, leading to a state σ (1) A 1 B 1B 1 N A1 B 1 (ρ (1) A 1 A 1B 1). Alice and Bob then perform the PPT-P channel P (2) A 1 B 1B 1 A 2 A 2B 2, which leads to the state ρ (2) A 2 A P (2) 2B 2 A 1 B 1B 1 A 2 A (σ (1) 2B 2 A 1 B (44) 1B 1). Alice sends system A 2 through the second channel use N A2 B 2, leading to the state σ (2) A 2 B 2B 2 N A2 B 2 (ρ (1) A 2 A This process iterates: the protocol uses the channel n times. In general, we 2B 2). have the following states for all i {2,..., n}: ρ (i) A i A ib i P (i) A i 1 B i 1B i 1 A i A (σ (i 1) ib i A i 1 B i 1B i 1), (45) σ (i) A i B ib i N Ai B i (ρ (i) A i A ib i), (46) where P (i) A i 1 B i 1B i 1 A i A is a PPT channel. The final step of the protocol consists of a PPT-P ib i channel P (n+1) A n BnB n M AM B, which generates the systems M A and M B for Alice and Bob, respectively. The protocol s final state is as follows: ω MA M B P (n+1) A n BnB n M AM B (σ (n) A ). (47) n BnB n Figure 1 depicts such a protocol. The goal of the protocol is that the final state ω MA M B is close to a maximally entangled state. Fix n, M N and ε [0, 1]. The original protocol is an (n, M, ε) protocol if the channel is used n times as discussed above, M A = M B = M, and if F (ω MA M B, Φ MA M B ) = Φ MA M B ω MA M B Φ MA M B (48) 1 ε, (49) where the fidelity F (τ, κ) τ κ 2 1 [Uhl76] and the maximally entangled state Φ M A M B = Φ Φ MA M B is defined from Φ MA M B 1 M M m=1 m MA m MB. (50) A rate R is achievable for PPT-P-assisted quantum communication if for all ε (0, 1], δ > 0, and sufficiently large n, there exists an (n, 2 n(r δ), ε) protocol. The PPT-P-assisted quantum capacity of a channel N, denoted as Q PPT-P, (N ), is equal to the supremum of all achievable rates. On the other hand, a rate R is a strong converse rate for PPT-P-assisted quantum communication if for all ε [0, 1), δ > 0, and sufficiently large n, there does not exist an (n, 2 n(r+δ), ε) 10

protocol. The strong converse PPT-P-assisted quantum capacity Q PPT-P, (N ) is equal to the infimum of all strong converse rates. We say that a channel obeys the strong converse property for PPT-P-assisted quantum communication if Q PPT-P, (N ) = Q PPT-P, (N ). We can also consider the whole development above when we only allow the assistance of LOCC channels instead of PPT channels. In this case, we have similar notions as above, and then we arrive at the LOCC-assisted quantum capacity Q (N ) and the strong converse LOCC-assisted quantum capacity Q (N ). It then immediately follows that because every LOCC channel is a PPT channel. Q (N ) Q PPT-P, (N ), (51) Q (N ) Q PPT-P, (N ) (52) 4.2 Max-Rains information as a strong converse rate for PPT-P-assisted quantum communication We now prove the following upper bound on the communication rate 1 n log 2 M (qubits per channel use) of any (n, M, ε) PPT-P-assisted protocol: Theorem 3 Fix n, M N and ε (0, 1). The following bound holds for an (n, M, ε) protocol for PPT-P-assisted quantum communication over a quantum channel N : ( ) 1 log 2 M nr max (N ) + log 2. (53) 1 ε Proof. For convenience of the reader, we give a complete proof, but we note that some of the essential steps are available in prior works [CMH17, RKB + 18, KW18]. From the assumption in (49), it follows that Tr{Φ MA M B ω MA M B } 1 ε, (54) while [Rai99, Lemma 2] implies that Tr{Φ MA M B σ MA M B } 1 M, (55) for all σ MA M B PPT (M A : M B ). So under an entanglement test, i.e., a measurement of the form {Φ MA M B, I MA M B Φ MA M B } and applying the data processing inequality for the max-relative entropy, we find for all σ MA M B PPT (M A :M B ) that D max (ω MA M B σ MA M B ) D max ({p, 1 p} {q, Tr{σ MA M B } q}) (56) = log 2 max{p/q, (1 p)/(tr{σ MA M B } q)} (57) log 2 (p/q) (58) log 2 [(1 ε) M], (59) where p Tr{Φ MA M B ω MA M B } and q = Tr{Φ MA M B σ MA M B }. Since the above chain of inequalities holds for all σ MA M B PPT (M A :M B ), we conclude that R max (M A ; M B ) ω log 2 [(1 ε) M]. (60) 11

From the monotonicity of the Rains relative entropy with respect to PPT-preserving channels [Rai01, TWW17], we find that R max (M A ; M B ) ω R max (A n; B n B n) σ (n) (61) = R max (A n; B n B n) σ (n) R max (A 1A 1 ; B 1) ρ (1) (62) [ n ] = R max (A n; B n B n) σ (n) + i=2 R max (A ia i ; B i) ρ (i) R max (A ia i ; B i) ρ (i) R max (A 1A 1 ; B 1) ρ (1) (63) n [ ] R max (A i; B i B i) σ (i) R max (A ia i ; B i) ρ (i) (64) i=1 nr max (N ). (65) The first equality follows because the state ρ (1) A 1 A is a PPT state with vanishing max-rains relative 1B 1 entropy. The second equality follows by adding and subtracting terms. The second inequality follows because R max (A i A i; B i ) ρ (i) R max(a i 1 ; B i 1B i 1 ) σ (i 1) for all i {2,..., n}, due to monotonicity of the Rains relative entropy with respect to PPT-P channels. The final inequality follows by applying Proposition 1 to each term R max (A n; B n B n) σ (i) R max (A i A i; B i ) ρ (i). Combining (60) and (65), we arrive at the inequality in (53). Remark 4 The bound in (53) can also be rewritten in the following way: 1 ε 2 n[q Rmax(N )], (66) where we set the rate Q = 1 n log 2 M. Thus, if the communication rate Q is strictly larger than the max-rains information R max (N ), then the fidelity of the transmission (1 ε) decays exponentially fast to zero in the number n of channel uses. An immediate corollary of the above is the following strong converse statement: Corollary 5 The strong converse PPT-P-assisted quantum capacity is bounded from above by the max-rains information: Q PPT-P, (N ) R max (N ). (67) 5 Amortization does not increase a channel s max-relative entropy of entanglement One of the main results of [CMH17] is that amortization does not increase a channel s max-relative entropy of entanglement; i.e., E max,a (N ) = E max (N ), (68) where E max (N ) denotes a channel s max-relative entropy of entanglement (we will define this shortly). The authors of [CMH17] proved (68) by employing the methods of complex interpolation [BL76]. The main application of (68) is that E max (N ) is a strong converse upper bound on the secret-key-agreement capacity of a quantum channel [CMH17] (this is defined as the private capacity 12

of the channel, when arbitrary LOCC is allowed between every channel use see [WTB17] or [CMH17] for a definition). In this section, we provide an alternate proof of (68), which is along the lines of the proofs of Proposition 1 and Corollary 2. We think that this approach brings a different perspective to the result of [CMH17] and could potentially be useful in future applications. To begin with, let us recall the definition of the max-relative entropy of entanglement of a bipartite state ρ AB [Dat09]: E max (A; B) ρ = min D max(ρ AB σ AB ). (69) σ AB SEP(A:B) Let SEP(A : B) denote the cone of all separable operators, i.e., X AB SEP(A : B) if there exists a positive integer L and positive semi-definite operators {PA x} x and {Q x B } x such that X AB = L x=1 P A x Qx B. The arrow in SEP(A : B) is meant to remind the reader of cone and is not intended to indicate any directionality between the A and B systems. In what follows, we sometimes employ the shorthands SEP and SEP when the bipartite cuts are clear from the context. Then we have the following alternative expression for the max-relative entropy of entanglement: Lemma 6 Let ρ AB be a bipartite state. Then E max (A; B) ρ = log 2 W sep (A; B) ρ, (70) where W sep (A; B) ρ = min X AB SEP {Tr{X AB } : ρ AB X AB }. (71) Proof. Employing the definition in (69), consider that min D max(ρ AB σ AB ) = log 2 min σ AB SEP(A:B) This concludes the proof. {µ : ρ AB µσ AB, σ AB SEP} (72) µ,σ AB Tr{X AB } : ρ AB X AB, X AB } SEP. (73) { = log 2 min X AB We can then define a channel s max-relative entropy of entanglement E max (N ) as in (2), by replacing E with E max. We can alternatively write E max (N ) as follows, by employing similar reasoning as given in the proof of [CMW16, Lemma 6]: E max (N ) = max ρ S min D max(ρ 1/2 σ SB SEP S J N SBρ 1/2 S σ SB), (74) where ρ S is a density operator and JSB N is the Choi operator for the channel N, as defined in (10). We now prove the following alternative expression for E max (N ): Lemma 7 Let N A B be a quantum channel. Then E max (N ) = log 2 Σ(N ), (75) where Σ(N ) = { min TrB Y SB {Y SB } : J N } SB Y SB. (76) SEP 13

Proof. Employing (74) and Lemma 6, we find that E max (N ) = log max ρ S min Y SB SEP { Tr{Y SB } : ρ 1/2 S J N SBρ 1/2 S Y SB }. (77) So our aim is to prove that the expression inside the logarithm is equal to Σ(N ). Taking the ansatz that ρ S is an invertible density operator, we find that the condition ρ 1/2 S J SB N ρ1/2 S Y SB is equivalent to the condition JSB N ρ 1/2 S Y SB ρ 1/2 S = Y SB SEP(S : B). Noting that Y SB = ρ 1/2 S Y SB ρ1/2 S, this means that { } max min ρ S Y SB Tr{Y SB } : ρ 1/2 S J SBρ N 1/2 { S Y SB = max min Tr{ρS Y SEP ρ S SB} : J Y SB SB N Y SB} SEP { = min Tr{ρS Y SB} : JSB N Y SB } Y SB SEP = min Y SB SEP = min Y SB SEP max ρ S { Tr{ρS Tr B {Y SB}} : JSB N Y SB } max ρ S { TrB {Y SB} } : JSB N Y SB} = Σ(N ). (78) The second equality follows from the Sion minimax theorem: the sets over which we are optimizing are convex, with the set of density operators additionally being compact, and the objective function Tr{ρ S Y SB } is linear in ρ S and Y SB, and so the Sion minimax theorem applies. The third equality follows from partial trace, and the fourth follows because D = max ρ Tr{Dρ}, when the optimization is with respect to density operators. Finally, we note that the ansatz may be lifted by an appropriate limiting argument. We can now see that the expressions for E max (A; B) ρ in Lemma 6 and E max (N ) in Lemma 7 have a very similar form to those in (24) and (26) for R max (A; B) ρ and R max (N ), respectively. However, the optimization problems for E max (A; B) ρ and E max (N ) are not necessarily efficiently computable because they involve an optimization over the cone of separable operators, which is known to be difficult [HM13] in general. Regardless, due to the forms that we now have for E max (A; B) ρ and E max (N ), we can prove an inequality from [CMH17], analogous to (28), with a proof very similar to that given in the proof of Proposition 1: Proposition 8 ([CMH17]) Let ρ A AB be a state and let N A B be a quantum channel. Then E max (A ; BB ) ω E max (N ) + E max (A A; B ) ρ, (79) where ω A BB = N A B(ρ A AB ). (80) Proof. By removing logarithms and applying Lemmas 6 and 7, the desired inequality is equivalent to the following one: W sep (A ; BB ) ω Σ(N ) W sep (A A; B ) ρ, (81) and so we aim to prove this one. Exploiting the identity in Lemma 6, we find that W sep (A A; B ) ρ = min Tr{C A AB }, (82) 14

subject to the constraints C A AB SEP(A A:B ), (83) C A AB ρ A AB, (84) while the identity in Lemma 7 gives that Σ(N ) = min Tr B {Y SB }, (85) subject to the constraints Y SB SEP(S :B), (86) Y SB J N SB. (87) The identity in Lemma 6 implies that the left-hand side of (81) is equal to W sep (A ; BB ) ω = min Tr{E A BB }, (88) subject to the constraints E A BB SEP(A :BB ), (89) E A BB N A B(ρ A AB ). (90) With these optimizations in place, we can now establish the inequality in (81) by making a judicious choice for E A BB. Let C A AB be optimal for W sep(a A; B ) ρ, and let Y SB be optimal for Σ(N ). Let Υ SA be the maximally entangled vector, as defined in (9). Pick E A BB = Υ SAC A AB Y SB Υ SA. This choice is clearly similar to that in the proof of Proposition 1. We need to prove that E A BB is feasible for W sep (A ; BB ) ω. To this end, consider that Υ SA C A AB Y SB Υ SA Υ SA ρ A AB J SB Υ N SA = N A B (ρ A AB ), (91) which follows from (84), (87), and (11). Now, since C A AB SEP(A A : B ), it can be written as x P A x A Qx B for positive semi-definite P x A A and Qx B. Furthermore, consider that since Y SB SEP(S :B), it can be written as y Ly S M y B for positive semi-definite Ly S and M y B. Then we have that Υ SA C A AB Y SB Υ SA = x,y Υ SA P x A A Qx B Ly S M y B Υ SA = Υ SA PA x A T A(L y A ) Qx B I S M y B Υ SA x,y = x,y Tr A {P x A A T A(L y A )} Qx B M y B SEP(A :BB ). (92) 15

The second equality follows from (14) and the third from (12). The last statement follows because { } Tr A P x A A T A(L y A )} = Tr A { T A (L y A )P A x A T A (L ya ) is positive semi-definite for each x and y. Finally, consider that Tr{E A BB } = Tr{ Υ SAC A AB Y SB Υ SA } = Tr{C A AB T A(Y AB )} = Tr{C A AB T A(Tr B {Y AB })} Tr{C A AB } T A(Tr B {Y AB }) = Tr{C A AB } Tr B {Y AB } = W sep (A A; B ) ρ Σ(N ). (93) The reasoning for this chain is identical to that for (41). Thus, we can conclude that our choice of E A BB is feasible for W (A ; BB ) ω. Since W (A ; BB ) ω involves a minimization over all E A BB satisfying (89) and (90), this concludes our proof of (81). By the same reasoning employed in the proof of Corollary 2, the equality in (68) follows as a consequence of the inequality in Proposition 8. We finally note that max-relative entropy of entanglement is subadditive as a function of quantum channels, in the following sense: E max (N M) E max (N ) + E max (M) (94) where N and M are quantum channels. This follows as a consequence of the equality in (68) and [KW18, Proposition 4], the latter of which states that the amortized entanglement is always subadditive as a function of quantum channels. It is an interesting open question to determine whether the max-relative entropy of entanglement is additive as a function of quantum channels. 6 On converses for quantum and private capacities Here we discuss briefly how our strong converse result stands with respect to prior work on strong converses and quantum and private capacities [HW01, TWW17, MHRW16, WD16b, Pir17, WTB17, CMH17, WFD17]. 6.1 Quantum capacities Let Q(N ) and Q (N ) denote the quantum capacity and the strong converse quantum capacity of a quantum channel N. These quantities are defined similarly to Q PPT-P, (N ) and Q PPT-P, (N ), but there is no PPT assistance allowed. The partial transposition bound was defined in [HW01] as follows: Q Θ (N ) = log 2 T N, (95) where T denotes the transpose map and is the diamond norm. In [HW01], Q Θ (N ) was established as a pretty-strong converse rate, in the sense of [MW14], for the quantum capacity of the channel N. This result was subsequently improved in [MHRW16] to the following strong converse bound: Q (N ) Q PPT-P, (N ) Q Θ (N ). (96) 16

The recent work in [WD16b, WFD17] established the following two bounds: R max (N ) Q Θ (N ), (97) Q (N ) R max (N ). (98) Thus, in light of the above history, it is clear that the natural question was whether Q PPT-P, (N ) R max (N ), and this is the question that our paper affirmatively answers. In summary, we now have that Q(N ) Q (N ) Q PPT-P, (N ) R max (N ) Q Θ (N ). (99) We now mention some other related results. The Rains relative entropy R(A; B) ρ of a bipartite state ρ AB is defined as [Rai99, Rai01, ADMVW02] R(A; B) ρ = min D(ρ AB σ AB ), (100) σ AB PPT (A:B) where D denotes the quantum relative entropy [Ume62, Lin73], defined as D(ω τ) = Tr{ω[log 2 ω log 2 τ]} whenever supp(ω) supp(τ) and + otherwise. Then the Rains information R(N ) of a quantum channel N is defined by replacing E in (2) with R(A; B) ρ [TWW17]. One can also define the amortized Rains information R A (N ) via the recipe in (3). Due to the inequality D(ω τ) D max (ω τ) [Dat09], the following inequality holds The following bound is known from [TWW17] and it is open to determine whether R(N ) R max (N ). (101) Q (N ) R(N ), (102) Q PPT-P, (N )? R(N ). (103) This latter inequality is known to hold if the channel N has sufficient symmetry [TWW17]. The squashed entanglement E sq (A; B) ρ of a quantum state ρ AB is defined as [CW04] E sq (A; B) ρ = 1 2 inf ρ ABE {I(A; B E) ρ : Tr E {ρ ABE } = ρ AB }, (104) where I(A; B E) ρ = H(AE) ρ +H(BE) ρ H(E) ρ H(ABE) ρ and H(F ) σ = Tr{σ F log 2 σ F }. (See also discussions in [Tuc99, Tuc02] for squashed entanglement.) One can also consider the squashed entanglement of a channel E sq (N ) [TGW14], as well as the amortized squashed entanglement E sq,a (N ). Another function of a quantum channel is its entanglement cost [BBCW13], which we write as E C (N ) and for which a definition is given in [BBCW13]. The following bounds and relations are known regarding these quantities: Q (N ) E C (N ), [BBCW13] (105) E sq,a (N ) = E sq (N ), [TGW14] (106) Q (N ) E sq (N ) E C (N ), [TGW14] (107) It is open to determine whether the following inequality holds Q (N )? E sq (N ). (108) 17

6.2 Private capacities One can also consider various private capacities and strong converse private capacities of a quantum channel, denoted as P (N ), P (N ), P (N ), and P (N ). Defining the relative entropy of entanglement E R [VP98] as E R (A; B) ρ = min D(ρ AB σ AB ), (109) σ AB SEP(A:B) and the max-relative entropy of entanglement E max as we did in (69), we can also define their channel versions E R (N ) and E max (N ) and their amortized versions E R,A (N ) and E max,a (N ). For these various quantities, we have that It is not known whether E R (N ) E max (N ), (110) P (N ) E sq (N ), [TGW14, Wil16b] (111) P (N ) E C (N ), [CMH17] (112) E max,a (N ) = E max (N ), [CMH17] (113) P (N ) E max (N ), [CMH17] (114) P (N ) E R (N ). [WTB17] (115) P (N )? E sq (N ), (116) P (N )? E R (N ), (117) but the latter inequality is known to hold for channels with sufficient symmetry [WTB17]. An interesting question is whether the max-rains information of a channel N could serve as an upper bound on one of its private capacities P (N ), P (N ), P (N ), or P (N ). The guiding principle behind many strong converse bounds in quantum information theory is to compare the output of the actual protocol, with respect to a relative-entropy-like measure, to a state or positive semi-definite operator that is useless for the task. By useless, we mean that the state or operator should have a probability of passing a test for the task that is no larger than inversely proportional to the dimension of the system being communicated. For example, this kind of result is known from [Rai99, Lemma 2] for operators in the set PPT (M A : M B ) and the entanglement test, and we used this bound effectively in (55) in order to establish the max-rains information as an upper bound on PPT-P-assisted quantum capacity. Furthermore, this kind of result is known from [HHHO05, HHHO09, WTB17] for separable states and the privacy test, and prior work has used this result to establish upper bounds on various private capacities of a channel [WTB17, CMH17]. However, it is not known how to obtain this kind of result for operators in the set PPT (M A : M B ) and the privacy test, and it is for this reason that we have not been able to establish the max-rains information as an upper bound on private capacity. We doubt whether this would be possible, given that there exist channels that produce PPT states with non-zero distillable secret key [HHH + 08a, HHH + 08b]. In the same spirit, one might wonder about differences between the max-rains relative entropy and the max-relative entropy of entanglement. First, it is clear that the max-relative entropy of entanglement can increase under the action of a PPT-P channel, because there exist states that 18

are PPT and entangled [HHHO05]. Furthermore, the aforementioned is related to the fact that there exist states for which there is a strict separation between the max-rains relative entropy and the max-relative entropy of entanglement. Any state that is PPT and entangled has a max-rains relative entropy equal to zero, while its max-relative entropy of entanglement is non-zero. 6.3 Summary: Channel measures that do not increase under amortization In summary, we know that amortization does not increase 1. the squashed entanglement E sq (N ) [TGW14], 2. the max-relative entropy of entanglement E max (N ) [CMH17], 3. or the max-rains information R max (N ) (Corollary 2). This is the main reason that these information quantities are single-letter converse bounds for assisted capacities. Is there any chance that the same could hold generally for E R (N ) or R(N )? If so, then the known capacity bounds could be improved. 7 Conclusion The main contribution of our paper was to show that the max-rains information of a quantum channel does not increase under amortization. That is, when entanglement is quantified by the max-rains relative entropy, the net entanglement that a channel can generate is the same as the amount of entanglement that it can generate if the sender and receiver do not start with any initial entanglement. This result then implies a single-letter, strong-converse, and efficiently computable bound for the capacity of a quantum channel to communicate qubits along with the assistance of PPT-preserving operations between every channel use. As such, the max-rains information can be easily evaluated and is a general benchmark for this capacity. As we emphasized previously, our upper bound is also an upper bound on the physically relevant LOCC-assisted quantum capacity. The main tool that we used to prove our result is the formulation of the max-rains relative entropy and max-rains information as semi-definite programs [WD16b, WD16a, WFD17] (in particular, we employed semi-definite programming duality we note here that this kind of approach has previously been employed successfully for multiplicativity, additivity, or parallel repetition problems in quantum information theory [BT16, BFT17, VW16]). We also compared our result to other results in the growing literature on the topic of bounds for the assisted capacities of arbitrary quantum channels [TGW14, TWW17, MHRW16, WTB17, CMH17]. We also provided an alternative proof for the fact that amortization does not enhance a channel s max-relative entropy of entanglement [CMH17]: i.e., E max,a (N ) = E max (N ). This statement was proved in [CMH17] by employing the methods of complex interpolation [BL76], but here we found a different proof by establishing alternative expressions for the max-relative entropy of entanglement (Lemma 6) and a channel s max-relative entropy of entanglement (Lemma 7). These alternative expressions then allowed us to employ reasoning similar to that in our proof of Proposition 1 in order to establish a different proof for the equality E max,a (N ) = E max (N ). We suspect that our approach could be useful in future applications. Finally, in [WFD17], it was noted that the max-rains information does not give a good upper bound on the quantum capacity of the qubit depolarizing channel. Our result gives a compelling 19

reason for this observation: the max-rains information finds its natural place as an upper bound on the PPT-P-assisted quantum capacity of the qubit depolarizing channel, and these assisting operations allowed between every channel use could result in a significant increase in capacity. Acknowledgements. We are grateful to Omar Fawzi, Xin Wang, David Reeb, Siddhartha Das, and Andreas Winter for discussions related to the topic of this paper. We also thank the anonymous referee for comments that helped to improve our paper, in particular for comments about the usefulness of PPT-P channels. Part of this work was done during the workshop Beyond I.I.D. in Information Theory, hosted by the Institute for Mathematical Sciences, Singapore, 24-28 July 2017. MB acknowledges funding by the SNSF through a fellowship. MMW acknowledges support from the Office of Naval Research and the National Science Foundation under grant no. 1350397. References [ADMVW02] Koenraad Audenaert, Bart De Moor, Karl Gerd H. Vollbrecht, and Reinhard F. Werner. Asymptotic relative entropy of entanglement for orthogonally invariant states. Physical Review A, 66(3):032310, September 2002. arxiv:quant-ph/0204143. [BBC + 93] [BBCW13] [BDF + 99] Charles H. Bennett, Gilles Brassard, Claude Crépeau, Richard Jozsa, Asher Peres, and William K. Wootters. Teleporting an unknown quantum state via dual classical and Einstein-Podolsky-Rosen channels. Physical Review Letters, 70(13):1895 1899, March 1993. Mario Berta, Fernando G. S. L. Brandao, Matthias Christandl, and Stephanie Wehner. Entanglement cost of quantum channels. IEEE Transactions on Information Theory, 59(10):6779 6795, October 2013. arxiv:1108.5357. Charles H. Bennett, David P. DiVincenzo, Christopher A. Fuchs, Tal Mor, Eric Rains, Peter W. Shor, John A. Smolin, and William K. Wootters. Quantum nonlocality without entanglement. Physical Review A, 59(2):1070 1091, February 1999. arxiv:quant-ph/9804053. [BDGDMW17] Khaled Ben Dana, María García Díaz, Mohamed Mejatty, and Andreas Winter. Resource theory of coherence: Beyond states. Physical Review A, 95(6):062327, June 2017. arxiv:1704.03710. [BDS97] Charles H. Bennett, David P. DiVincenzo, and John A. Smolin. Capacities of quantum erasure channels. Physical Review Letters, 78(16):3217 3220, April 1997. arxiv:quant-ph/9701015. [BDSW96] Charles H. Bennett, David P. DiVincenzo, John A. Smolin, and William K. Wootters. Mixed-state entanglement and quantum error correction. Physical Review A, 54(5):3824 3851, November 1996. arxiv:quant-ph/9604024. [Ben05] Charles H. Bennett. Simulated time travel, teleportation without communication, and how to conduct a romance with someone who has fallen into a black hole. https://www.research.ibm.com/people/b/bennetc/quponbshort. pdf, May 2005. 20

[BFT17] [BHLS03] [BL76] Mario Berta, Omar Fawzi, and Marco Tomamichel. On variational expressions for quantum relative entropies. Letters in Mathematical Physics, 107(12):2239 2265, December 2017. arxiv:1512.02615. Charles H. Bennett, Aram W. Harrow, Debbie W. Leung, and John A. Smolin. On the capacities of bipartite Hamiltonians and unitary gates. IEEE Transactions on Information Theory, 49(8):1895 1911, August 2003. arxiv:quant-ph/0205057. J. Bergh and Jorgen Löfström. Interpolation Spaces. Springer-Verlag Berlin Heidelberg, 1976. [BT16] Mario Berta and Marco Tomamichel. The fidelity of recovery is multiplicative. IEEE Transactions on Information Theory, 62(4):1758 1763, April 2016. arxiv:1502.07973. [BV04] [CD09] Stephen Boyd and Lieven Vandenberghe. Convex Optimization. Cambridge University Press, The Edinburgh Building, Cambridge, CB2 8RU, UK, 2004. Eric Chitambar and Runyao Duan. Nonlocal entanglement transformations achievable by separable operations. Physical Review Letters, 103(11):110502, September 2009. arxiv:0811.3739. [CMH17] Matthias Christandl and Alexander Müller-Hermes. Relative entropy bounds on quantum, private and repeater capacities. Communications in Mathematical Physics, 353(2):821 852, July 2017. arxiv:1604.03448. [CMW16] [CW04] [Dat09] Tom Cooney, Milan Mosonyi, and Mark M. Wilde. Strong converse exponents for a quantum channel discrimination problem and quantum-feedback-assisted communication. Communications in Mathematical Physics, 344(3):797 829, June 2016. arxiv:1408.3373. Matthias Christandl and Andreas Winter. Squashed entanglement : An additive entanglement measure. Journal of Mathematical Physics, 45(3):829 840, March 2004. arxiv:quant-ph/0308088. Nilanjana Datta. Min- and max-relative entropies and a new entanglement monotone. IEEE Transactions on Information Theory, 55(6):2816 2826, June 2009. arxiv:0803.2770. [Hay06] Masahito Hayashi. Quantum Information: An Introduction. Springer, 2006. [HHH96] [HHH + 08a] Michal Horodecki, Pawel Horodecki, and Ryszard Horodecki. Separability of mixed states: necessary and sufficient conditions. Physics Letters A, 223(1-2):1 8, November 1996. arxiv:quant-ph/9605038. Karol Horodecki, Micha l Horodecki, Pawe l Horodecki, Debbie Leung, and Jonathan Oppenheim. Quantum key distribution based on private states: Unconditional security over untrusted channels with zero quantum capacity. IEEE Transactions on Information Theory, 54(6):2604 2620, June 2008. arxiv:quant-ph/0608195. 21