Quick viewing(Text Mode)

Everything You Always Wanted to Know About LOCC

Everything You Always Wanted to Know About LOCC

arXiv:1210.4583v2 [quant-ph] 13 May 2014 rcChitambar, Eric prahdabtaiycoeuigLC,btnevertheless two-qub but a LOCC, using demonstrate close we arbitrarily Finally, approached depolari completely maps. the around implementable ball LOCC oper open quantum an of of subset topolog existence compact not the a is constitutes LOCC LOCC of round set nite the t While those rounds. as communication well o as of terms protocols round in finite classes both captures operational malism related and characte LOCC important to of difficult many scription and in complex operations is of structure mathematical class comm natural classical the and as operations emerges local as known systems sional nti ae esuytesbe fgnrlzdqatmmeas quantum generalized of subset the study we paper this In vrtigYuAwy atdt nwAotLOCC About Know to Wanted Always You Everything nvria u`nm eBreoa S013Blaer ( Bellaterra ES-08193 Barcelona, Aut`onoma de Universitat 6 eateto ahmtc,Uiest fBitl Bristol Bristol, of University Mathematics, of Department 7 etefrQatmTcnlge,Ntoa nvriyo Si of University National Technologies, Quantum for Centre n nttt o unu optn,Uiest fWaterlo of University , Quantum for Institute and 1,2 5 11KthwnRa,Yrtw egt,N 09,USA 10598, NY Heights, Yorktown Road, Kitchawan 1101 CE nttc´ aaaad eec sui Avan¸cats Estudis i Recerca de Instituci´o Catalana – ICREA ebeLeung, Debbie 1 eateto hsc,Suhr liosUniversity, Illinois Southern , of Department 2 g li opns2,000Breoa Spain Barcelona, 08010 23, Companys Lluis Pg. ´ sc eoia nomc´ eoesQu`antics,F´ısica Fenomens Te`orica: Informaci´o i cec rv ,Snaoe174,Singapore 117542, Singapore 3, Drive Science 2 3 h eiee nttt o hoeia Physics, Theoretical for Institute Perimeter The eateto obntrc Optimization & Combinatorics of Department aelo nai 2 G,Canada 3G1, N2L Ontario Waterloo, aelo nai 2 Y,Canada 2Y5, N2L Ontario Waterloo, BtWr fadt Ask) to Afraid Were (But 4 abnae lios691 USA 62901, Illinois Carbondale, B JWto eerhCenter, Research Watson TJ IBM 3 ar Manˇcinska,Laura Abstract 1 3 ie eew rvd rcs de- precise a provide we Here rize. antb mlmne perfectly. implemented be cannot ai Ozols, Maris a tlz nubuddnumber unbounded an utilize hat igmpta osssetrl of entirely consists that map zing unu ntuet.Orfor- Our instruments. quantum f ncto LC) hl LOCC While (LOCC). unication unu nomto ak,its tasks, information quantum tos diinlyw show we Additionally ations. clycoe,w hwta fi- that show we closed, ically tmpwoeato a be can action whose map it rmnso nt dimen- finite on urements acln) Spain Barcelona), S T,U.K. 1TW, BS8 3,4 ngapore, nra Winter Andreas o, , 5,6,7 1 Introduction

The “distant lab” paradigm plays a crucial role in both theoretical and experimental aspects of . Here, a multipartite quantum system is distributed to various parties, and they are restricted to act locally on their respective subsystems by performing measurements and more general quantum operations. However in order to enhance their measurement strategies, the parties are free to communicate any classical data, which includes the sharing of randomness and previous measurement results. Quantum operations implemented in such a manner are known as LOCC (local operations with classical communication), and we can think of LOCC as a spe- cial subset of all physically realizable operations on the global system. This restricted paradigm, motivated by current technological difficulties in communicating quantum data, serves as a tool to study not only quantum correlations and other nonlocal quantum effects, but also resource transformations such as channel capacities. Using LOCC operations to study resource transformation is best illustrated in quantum tele- portation [BBC+93]. Two parties, called Alice and Bob, are separated in distant labs. Equipped with some pre-shared quantum states that characterize their entanglement resource, they are able to transmit quantum states from one location to another using LOCC; specifically, the exchange rate is one quantum () transmitted for one entangled bit (ebit) plus two classical (cbits) consumed. Via teleportation then, LOCC operations become universal in the sense that Alice and Bob can implement any physical evolution of their joint system given a sufficient sup- ply of pre-shared entanglement. Thus entanglement represents a fundamental resource in quan- tum with LOCC being the class of operations that manipulates and consumes this resource [BBPS96, Nie99, BPR+00]. Indeed, the class of non-entangled or separable quantum states are precisely those that can be generated exclusively by the action of LOCC on pure product states [Wer89], and any sensible measure of entanglement must satisfy the crucial property that its expected value is non-increasing under LOCC [BDSW96, VPRK97, HHH00, PV07]. The intricate structure of LOCC was perhaps first realized over 20 years ago by Peres and Wootters who observed that when some classical random variable is encoded into an ensemble of bipartite product states, the accessible information may be appreciably reduced if the decoders - played by Alice and Bob - are restricted to LOCC operations [PW91]. (In fact, results in [PW91] led to the discovery of teleportation.) Several years later, Massar and Popescu analytically con- firmed the spirit of this conjecture by considering pairs of particles that are polarized in the same randomly chosen direction [MP95]. It was shown that when Alice and Bob are limited to a finite number of classical communication exchanges, their LOCC ability to identify the an- gle is strictly less than if they were allowed to make joint measurements on their shared states. This line of research culminated into the phenomenon of quantum data hiding [TDL01, DLT02, EW02]. Here, some classical data is encoded into a bipartite state such that Alice and Bob have arbitrarily small accessible information when restricted to LOCC, however, the data can be perfectly retrieved when the duo measures in the same lab. The examples concerning accessible information mentioned above demonstrate that a gap be- tween the LOCC and globally accessible information exists even in the absence of entanglement. This finding suggeststhat nonlocality and entanglement are two distinct concepts, with the former being more general than the latter. Bennett et al. were able to sharpen this intuition by constructing a set of orthogonal bipartite pure product states that demonstrate “nonlocality without entangle- ment” in the sense that elements of the set could be perfectly distinguished by so-called separable operations (SEP) but not by LOCC [BDF+99]. The significance that this result has on the structure of LOCC becomes most evident when considering the Choi-Jamiołkowski isomorphism between quantum operations and positive operators [Cho75, Jam72]. Separable operations are precisely

2 the class of maps whose Choi matrices are separable, and consequently SEP inherits the relatively well-understood mathematical structure possessed by separable states [HHH96, BCH+02]. The fact that SEP and LOCC are distinct classes means that LOCC lacks this nice mathematical char- acterization, and its structure is therefore much more subtle than SEP. Thus despite having a fairly intuitive physical description, the class of LOCC is notoriously difficult to characterize mathematically [BDF+99, DHR02]. Like all quantum operations, an LOCC measurement can be represented by a trace-preserving completely positive map acting on the space of density operators (or by a “quantum instrument” as we describe below), and the difficulty is in describing the precise structure of these maps. Part of the challenge stems from the way in which LOCC operations combine the globally shared classical information at one instance in time with the particular choice of local measurements at a later time. The potentially unrestricted number of rounds of communication further complicates the analysis. A thorough definition of finite-round LOCC has been presented in Ref. [DHR02] thus formalizing the description given in Ref. [BDF+99]. Recently, there has been a renewed wave of interest in LOCC alongside new discoveries con- cerning asymptotic resources in LOCC processing [Rin04, Chi11, KKB11, CLMO12]. It has now been shown that when an unbounded number of communication rounds are allowed, or when a particular task needs only to be accomplished with an arbitrarily small failure rate (but not per- fectly), more can be accomplished than in the setting of finite rounds and perfect success rates. Consequently, we can ask whether a task can be performed by LOCC, and failing that, whether or not it can be approximated by LOCC, and if so, whether a simple recursive procedure suffices. To make these notions precise, a definition of LOCC and its topological closure is needed. Here, we aim to extend the formalisms developed in [BDF+99, DHR02, KKB11] so to facilitate an analysis of asymptotic resources and a characterization of the most general LOCC protocols. Indeed, we hope that this work will provide a type of “LOCC glossary” for the research community. The approach of this article is to describe LOCC in terms of quantum instruments.1 This will enable us to cleanly introduce the class of infinite-round LOCC protocols as well as the more gen- eral class of LOCC-closure. The explicit definitions for these operational classes are provided in Sect. 2 as well as a discussion on the relationships among them. In Sect. 3, we discuss some topo- logical features possessed by the different LOCC classes. The main results here involve (i) showing that the set of fixed outcome LOCC protocols possess a non-empty interior and (ii) providing an upper bound on the number of measurements needed per round to implement any finite round, finite outcome LOCC protocol. In Sect. 4 we construct a two-qubit separable instrument that can be approached arbitrarily close using LOCC but nevertheless cannot be implemented perfectly via LOCC. This finding represents the first of its kind in the bipartite setting. Finally in Sect. 5 we close with a brief summary of results and discuss some additional open problems. Technical proofs are reserved for the two appendices.

1An alternative characterization of LOCC in terms of physical tasks was described to us privately by the authors of Ref. [KKB11]. Instead of considering how well an LOCC map approximates a target map (by a distance measure on maps), one can define a success measure for a particular task and study the achievable values via LOCC. This is particularly useful when the task does not uniquely define a target map. Here, we define LOCC in terms of quantum instruments, which admits a more precise mathematical description, and a further optimization over possible target maps can be added if one just wishes to focus on success rates.

3 2 How to define LOCC?

2.1 Quantum Instruments Throughout this paper, we consider a finite number of finite-dimensional quantum systems. We denote the associated with and refer to it as the underlying state space of the H system. Let ( ) be the set of bounded linear operators acting on and ( ( )) the set of B H H L B H all bounded linear maps on ( ). A (discrete) quantum instrument J is a family of completely B H positive (CP) maps ( : j Θ) with ( ( )) and Θ a finite or countably infinite index set, Ej ∈ Ej ∈ L B H such that ∑ is trace-preserving [DL70]. When (w.l.o.g.) Θ = 1,2,... , we write the instrument j Ej { } also as an ordered list J = ( , ,...). When it is applied to the state ρ, (ρ) represents the (un- E1 E2 Ej normalized) postmeasurement state associated with the outcome j, which occurs with probability tr( (ρ)). We denote the set of instruments with a given index set Θ as CP[Θ] ( ( ))Θ. For Ej ⊂ L B H Θ = 1,2,..., m we abbreviate CP[Θ] =: CP[m]. If the index set is unimportant or implicitly { } clear from the context we often omit it. Note that the set of quantum instruments is convex, where addition of two instruments and scalar multiplication is defined componentwise. This obviously requires that the instruments we combine are defined over the same index set Θ. However, an instrument J CP[Θ] may naturally ∈ be viewed as an element in Θ Θ by padding J with zeros, i.e. = 0 for j Θ Θ. Also, ′ ⊃ Ej ∈ ′ \ observe that for finite Θ (finite m), the set of instruments over Θ is a closed subset of a finite- dimensional . Given index set Θ, we define a quantum-classical (QC) map over Θ as a trace-preserving Θ completely positive map (TCP) which sends ( ) ( ) (C| |) and is of the form B H → B H ⊗ B C Θ ρ ∑j Θ j(ρ) j j , where the j constitute an orthonormal for | |. In this way, we see 7→ ∈ E ⊗ | ih | | i that for some underlying space , the set CP[Θ] is in a one-to-one correspondence with the set H of QC maps over Θ. For instrument J = ( : j Θ), we denote its corresponding QC map by Ej ∈ [J]( )= ∑ ( ) j j . E · j Ej · ⊗ | ih | Example. Consider a POVM measurement with Kraus operators M1,..., Mn that upon mea- M † † suring ρ and obtaining outcome j gives postmeasurement state MjρMj / Tr(Mj Mjρ). The instru- ment corresponding to this POVM is ( ,..., ) where each (ρ) = M ρM† has only one Kraus E1 En Ej j j operator.

† Example. Consider a TCP map with Kraus operators M1,..., Mn, i.e., (ρ) = ∑i [n] MiρMi N N ∈ where [n] := 1,2,..., n . The instrument corresponding to this TCP map is ( ) with only one { } N CP map (which is necessarily trace-preserving).

Let J = ( : j Θ) and J = ( : k Θ ) be quantum instruments. We say that J is a Ej ∈ ′ Fk ∈ ′ ′ coarse-graining of J if there exists a partition of the index set Θ, given by Θ = k Θ Sk, such ∈ ′ that k = ∑j S j for each k Θ′. Equivalently, and perhaps more intuitively, one can describe F ∈ k E ∈ F this by the action of a coarse-graining map f : Θ Θ , which is simply the function f (k) = j → ′ for k S (the sets S may be infinite). In this picture we are using the coarse-graining map f to ∈ j j post-process the classical information from the instrument J′. Physically, this action corresponds to the discarding of classical information if the coarse-graining is non-trivial. The fully coarse- grained instrument of J corresponds to the TCP map ∑ , obtainable by tracing out the classical j Ej register of [J]. We say that an instrument ( : j Θ) is fine-grained if each of the has action E Ej ∈ Ej of the form ρ M ρM† for some operator M . In this way, the most general instrument can be 7→ j j j implemented by performing a fine-grained instrument followed by coarse-graining.

4 The set of instruments over an index set Θ carries a metric as follows. For instruments J = ( : j Θ) and J˜ = ( ˜ : j Θ), we use the distance measure induced by the diamond norm on Ej ∈ Ej ∈ the associated QC maps:

D (J, J˜) := [J] [J˜ ] = max ∑ ( j ˜j)[ρ] 1 (1) ⋄ E −E 0 ρ I k I⊗E −I⊗ E k ⋄ ≤ ≤ j Θ ∈ where is the diamond norm on superoperators [KSV02, Wat05], A p is the so-called Shatten k·k⋄ k k p-norm of A, is the identity in ( ( )), and ρ ( ). That this series converges follows I L B H ∈ B H⊗H from the bound [J] [J˜ ] 2. To see this observe that [J] [J˜ ] is equal to E −E ⋄ ≤ E −E ⋄

max ( [J])[ρ] ( [J˜ ])[ρ] max ( [J])[ ρ] + max ( [J])[ρ′] . 0 ρ I I⊗E − I⊗E 1 ≤ 0 ρ I I⊗E 1 0 ρ I I⊗E 1 ≤ ≤ ≤ ≤ ≤ ′≤

As both [J] and [J ] are trace-preserving, we have the desired uppper bound. A sequence of E E ′ instruments J CP[Θ], ν = 1,2,..., is said to converge to the instrument J = ( : j Θ) if ν ∈ Ej ∈ limν ∞ D (Jν, J)= 0. → ⋄ If the index set Θ is finite, instrument convergence reduces to the pointwise condition: for all j Θ, limν ∞ ν,j ˜j = 0. In fact, since CP[Θ] is a subset of a finite-dimensional real vector ∈ → kE − E k⋄ space when Θ is finite, the topology is unique, independent of the metric, and CP[Θ] is complete in the sense that every Cauchy sequence converges to an element of CP[Θ] (w.r.t. any norm, such as the one described above). All these statements are no longer true for infinite Θ.

2.2 LOCC Instruments

For an N-partite quantum system, the underlying state space is := A1 A2 AN with AK H(K) H ⊗H ⊗···⊗H being the reduced state space of party K. An instrument J = ( 1, 2,... ) is called one- H F F (J) (K) way local with respect to party K if each of its CP maps has the form j = J=K j j , F 6 T ⊗E where (K) is a CP map on ( AK ), and for each J = K, (J) is some TCP map. Operationally, E B H 6 Tj N  this one-way local operation consists of party K applying an instrument ( , ,... ), broadcasting E1 E2 the classical outcome j to all other parties, and party J applying TCP map (J) after receiving this Tj information. We say that an instrument J is LOCC linked to J =( , ,... ) if there exists a collection of ′ A1 A2 J (Kj) J one-way local instruments j =( 1 j, 2 j,... ) : j = 1,2,... such that ′ is a coarse-graining { B | B | } of the instrument with CP maps j j j. Operationally, after the completion of J, conditioned on B ′| ◦A J (Kj) the measurement outcome j, instrument j is applied followed by coarse-graining (see Fig. 1). Note that we allow the acting party in the conditional instrument Kj to vary according to the previous outcome j. We now define the different classes of LOCC instruments. For a quantum instrument J ∈ CP[Θ], we say that:

J LOCC if J is one-way local with respect to some party K, followed by a coarse-graining • ∈ 1 map.

J LOCCr (r 2) if it is LOCC linked to some J LOCCr 1. • ∈ ≥ ∈ − J LOCCN if J LOCC for some r N = 1,2,... . • ∈ ∈ r ∈ { } J LOCC if there exists a sequence J , J ,... in which (i) J LOCCN, (ii) J is LOCC • ∈ { 1 2 } ν ∈ ν linked to Jν 1, and (iii) each Jν has a coarse-graining Jν′ such that Jν′ converges to J. −

5 J =( , ) A1 A2 A1 A2 Coarse-grained instrument: 1 1 2 1 1 2 2 2 B | B | B | B | J′ =( 1 1 1 + 1 2 2, B | ◦A B | ◦A 2 1 1 + 2 2 2) (coarse-grain over matching branches) B | ◦A B | ◦A

Figure 1: The instrument J′ is LOCC linked to the instrument J. Conditional instruments ( 1 1, 2 1) and ( 1 2, 2 2) can be composed with the two elements of J so that after coarse- B | B | B | B | graining, the resulting instrument is J′.

J LOCCN if there exists a sequence J , J , . . . in which (i) J LOCCN and (ii) the se- • ∈ 1 2 j ∈ quence converges to J.

Operationally, LOCCr is the set of all instruments that can be implemented by some r-round LOCC protocol. Here, one round of communication involves one party communicating to all the others, and the sequence of communicating parties can depend on the intermediate measurement outcomes. The set of instruments that can be implemented by some finite round protocol is then LOCCN. On the other hand, so-called infinite round protocols, or those having an unbounded number of non-trivial communication rounds, correspond to instruments in LOCC LOCCN. The \ full set of LOCC then consists of both bounded-round protocols as well as the unbounded ones + [BDF 99, KKB11, Chi11]. Theset LOCCN is the topological closure of LOCCN, and from the chain of inclusions LOCCN LOCC LOCCN, we obtain LOCCN = LOCC. ⊂ ⊂ Both LOCC and LOCCN consist of all instruments that can be approximated better and better with more LOCC rounds. The two sets are distinguished by noting that for any instrument in LOCC, its approximation in finite rounds can be made tighter by just continuing for more rounds within a fixed LOCC protocol; whereas for instruments in LOCCN LOCC, different protocols will \ be needed for different degrees of approximation. Note that according to our definitions, every LOCC instrument is defined with respect to some fixed index set Θ. However, the instruments implemented during intermediate rounds of a pro- tocol might range over different index sets. The requirement is that the intermediate instruments can each be coarse-grained into Θ so to form a convergent sequence of instruments. This coarse- graining need not correspond to an actual discarding of information. Indeed, discarding the mea- surement record midway through the protocol will typically prohibit the parties from completing the final LOCC instrument since the choice of measurement in each round depends on the full measurement history. On the other hand, often there will be an accumulation of classical data su- perflous to the task at hand, and the parties will physically perform some sort of coarse-graining (discarding of information), especially at the very end of the protocol. To complete the picture, we also provide definitions for the related classes of separable (SEP) A :A : :A and positive partial transpose preserving (PPT) instruments. A multipartite state ρ 1 2 ··· N is called (fully) separable if it can be expressed as a convex combination of product states with respect to the partition A : A : : A . Likewise, ρA1:A2: :AN is said to have positive partial- 1 2 · · · N ··· transpose (PPT) if the operator obtained by taking a partial transpose with respect to any subset of parties is positive semi-definite. Now, for the space := A1 A2 AN , we introduce A A A H H ⊗H ⊗···⊗H the auxiliary state space := 1′ 2′ ′N and let denote the identity in ( ( )). H′ H ⊗H ⊗···⊗H I ′ L B H′

6 Let J =( , ,... ) be a quantum instrument acting on ( ) and consider the state E1 E2 B H A A :A A : :A A ( ′ )[ρ 1′ 1 2′ 2 ··· ′N N ]. (2) I ⊗Ej We say that

J =( 1, 2,... ) SEP if each j is a separable map [VPRK97, Rai97], meaning that the state • E E ∈ E A A :A A : :A A in Eq. (2) is separable whenever ρ 1′ 1 2′ 2 ··· ′N N is separable;

J = ( 1, 2,... ) PPT if each j is a PPT map [Rai01, DLT02], meaning that the state in • E E ∈ A A :A AE : :A A Eq. (2) is PPT whenever ρ 1′ 1 2′ 2 ··· ′N N is PPT. Operationally, these classes are more powerful than LOCC, yet they are still more restrictive than the most general quantum operations. At the same time, they admit a simpler mathematical char- acterization than LOCC, and this can be used to derive many limitations on LOCC, such as entan- glement distillation [Rai97, Rai01, HHH98] and state discrimination [DLT02, Che04]. Every instrument belonging to LOCCr, LOCC, LOCC, SEP or PPT has an associated index set Θ. Thus for each Θ, these operational classes naturally become subsets of CP[Θ], and we denote them as LOCCr[Θ], LOCC[Θ], LOCC[Θ], SEP[Θ] and PPT[Θ] respectively. In the following sections we shall consider these different classes in more detail.

2.3 Relationships Between the Classes For a fixed number of parties N 2, and over a sufficiently large (but universal) index set Θ, the ≥ different LOCC classes are related to each other as follows:

LOCC1 ( LOCCr ( LOCCr+1 ( LOCCN ( LOCC ( LOCC ( SEP ( PPT (3) for any r 2. We note that LOCC ( LOCC for some k N implies LOCC ( LOCC . To ≥ r r+k ∈ r r+1 see this, suppose that LOCC = LOCC , and let J be some instrument in LOCC LOCC . r r+1 r+k \ r Then there exists an implementation of J consuming r + k rounds with Jr+1 being the instrument performed during the first r + 1 rounds of this particular implementation. But since LOCCr = LOCCr+1, wehave Jr+1 LOCCr and so J LOCCr+k 1. Here, we have considered Θ sufficiently ∈ ∈ − large such that both J and Jr+1 are instruments over the same index set (this can always be done by Theorem 2). Repeating this argument for k total times gives that J LOCC , which is a ∈ r contradiction. We now explain why all inclusions are proper. The operational advantage of LOCC2 over LOCC1 is well-known, having been observed in entanglement distillation [BDSW96], [GL03], and state discrimination [Coh07, OH08]. On the other hand, only a few examples have been proven to demonstrate the separation between LOCCr and LOCCr+1. For N = 2, Xin and Duan have constructed sets containing O(n2) pure states in two n-dimensional systems that require O(n) rounds of LOCC to distinguish perfectly [XD08]. For N 3, a stronger ≥ separation is shown for random distillation of bipartite entanglement from a three-qubit state. Here the dimension is fixed, and two extra LOCC rounds can always increase the probability of success by a quantifiable amount; moreover, certain distillations only become possible by infinite- round LOCC, thus demonstrating LOCCN = LOCC [Chi11]. By studying the same random distil- 6 lation problem, one can show that LOCC = LOCC [CCL12]. Thus, there are instruments that re- 6 quire different protocols to achieve better and better approximations when more and more LOCC rounds are available, and neither LOCCN nor LOCC is closed. For N = 2, we will demonstrate in Sect. 4 that LOCCN is also not closed for two-qubit systems. The difference between LOCCN and

7 SEP has emerged in various problems such as state discrimination [BDM+99, Coh07, DFXY09] and entanglement transformations [CDS08]. Proving that LOCC = SEP is more difficult, but it 6 indeed has been demonstrated in Refs. [BDF+99, KTYI07, CLMO12] for the task of state discrimi- nation, as well as Ref. [CCL11, CCL12] for random distillation. In Sect. 4 we will provide another example of this. Finally, the strict inclusion between PPT and SEP follows from the existence of non-separable states that possess a positive partial transpose [Hor97]. While LOCC ( SEP, it is possible to stochasitcally perform any instrument from SEP by LOCC. More precisely, let Θ = [m] and J = ( ,... , ) SEP be an arbitrary separable instrument. E1 Em ∈ Then we say that J can be performed by Stochastic LOCC (SLOCC) if there is some nonzero probability p such that the instrument J =(p ,... p , (1 p) ) CP[m + 1] is implementable ′ E1 Em − D ∈ by LOCC. Here, ( ( )) is the completely depolarizing TCP map which acts by (ρ) = D ∈ L B H D Tr(ρ) I , where I is the identity operator in ( ) and d is the dimension of . The outcome d d d B H H D indicates a failure in implementing J and it occurs with probability 1 p. It was shown in Ref. − [DVC00] that every separable CP map has an SLOCC implementation. Here we provide a lower bound on the success probability p. Lemma 1. If J = ( ,... , ) is some N-partite separable instrument with d being the total dimension E1 Em of . Then J can be implemented by SLOCC with a success probability at least 1 . H md2 1 N Proof. For all i [m] let Aij ... Aij j [d2] be the set of operators in some Kraus representation ∈ { ⊗ ⊗ } ∈ of with no more than d2 elements. The existence of such a representation can be deduced using Ei Carath´eodory’s theorem (see Theorem 2 and its proof). For each Kraus operator A1 ... AN, we ij ⊗ ⊗ ij choose the individual matrices so that A1 = ... = Ap . This ensures that I (Ak )† Ak , k ijk∞ k ijk∞ dk ≥ ij ij where I is the identity acting on subspace Ak . Therefore, dk H k := Ak , I (Ak )† AN (4) Mij ij dk − ij ij q is a valid local measurement for each party k. The protocol now consists of the parties first collec- tively choosing a pair (i, j) [m] [d2] uniformly at random. They then take turns performing ∈ × their respective local measurements k and broadcasting their result. If all parties obtain the first Mij outcome, their implementation is a success and they fully coarse-grain the classical data (i, j) over the index j (hence recovering the maps ). If at least one party obtains the second outcome, all the Ei parties locally depolarize and this is a failure outcome. Coarse-graining over all failure outcomes 1 1 md2 1 generates the (N-round) LOCC instrument ( ,... , , − ). md2 E1 md2 Em md2 D 3 What is the shape of LOCC?

In this section we describe some topological properties of the set LOCC. Proposition 1. LOCC forms a convex subset of the set of all quantum instruments. Proof. Given two LOCC instruments J and J , the convex combination λJ +(1 λ)J can be 1 2 1 − 2 implemented by LOCC by introducing some globally accessible randomness into the first round of the protocol that determines whether the parties perform J1 or J2. Next we turnto the questionof LOCC interior. In what follows, let d denote the total dimension of the N-party state space = A1 ... AN . H H ⊗ ⊗H Theorem 1. For a finite index set Θ = [m], LOCC has a non-empty interior.

8 To prove this, we will show the existence of a non-empty ball that consists entirely of LOCC instruments around the completely depolarizing instrument D := 1 ,..., 1 .2 Our argument m D m D will rely on the following result of Gurvits and Barnum.  Proposition 2 ([GB03]). If A ( ) and A R := 21 N/2, then I + A is a separable operator ∈ B H k k2 ≤ N − d with respect to the partition A : A : : A . 1 2 · · · N While this is a statement about separable operators, we can easily translate it into a statement about separable maps. Recall that for a CP map ( ( A1 ... AN )), with d being the E ∈ L B H ⊗ ⊗H i dimension of Ai , the N-partite Choi matrix is given by H A′ A : :A′ A A A A A A A A A A A Ω 1 1 ··· N N :=( 1 2... N 1 2... N )[Φ 1′ 1 ... Φ ′N N ], E I ⊗E ⊗ ⊗ A A di where Φ i′ i = ∑ ii jj [Cho75]. In short, the Choi matrix is obtained by distributing maxi- i,j=1 | ih | mally entangled states across two copies of the original N-partite system and applying the map A A : :A A E to just half of it. It is known that is a separable CP map if and only if Ω 1′ 1 ′N N is a separable E ··· operator with respect to the partition A A : : A A [CDKL01]. Then as a first corollary to 1′ 1 · · · ′N N Proposition 2, we have: Corollary 1. Let J =( ,..., ) CP[m]. Then the instrument E1 Em ∈ (1 η)D + ηI SEP, (5) − ∈ 2 2 N 1 1 for all η R := R /(md + R )=(md 2 2 + 1) . ≤ SEP N N − − Proof. By the above discussion, we need to show that for each CP map in Eq. (5), its corresponding 1 I Choi matrix is separable. For the depolarizing map , we have Ω = d2 , and therefore, the Choi D d 1 1 D ηmd I 2 Ω I 2 Ω matrix for (1 η) m + η i is (1 η) md d + η i . From Proposition 2 we get that d + 1 η i − D E − E − E ηmd Ω Ω Ω is separable whenever 1 η i 2 RN. Note that i 2 i 1 = d. Therefore, choosing − k E k ≤ k E k ≤ k E k RN η 2 yields the desired statement. ≤ md +RN Now we are in a position to take Proposition 2 one step further and prove Theorem 1. Corollary 2. Any instrument J =( ,..., ) CP[m] such that E1 Em ∈

D (D, J) RLOCC (6) ⋄ ≤ R N SEP 3 6 2 1 2 4 1 can be implemented by an N-round LOCC protocol, where RLOCC := m2d4 =(m d 2 − + m d )− . J 1 1 Proof. We begin by decomposing each i in as i = (1 mdδ) m + mdδ i where i = m 1 1 E E Ω − D F F D− mdδ ( m i). By considering the Choi matrices i , we see that each i is CP when taking D−E 1 F F I 2 Ω δ = maxi [m] md d i ∞. From the definition of the diamond norm, it immediately follows ∈ k − E k that 1 1 1 D J I 2 Ω D ( , )= max ∑ ( i)[ρ] ∑ d i ⋄ 0 ρ I I⊗ m D−I⊗E 1 ≥ d md − E 1 ≤ ≤ i [m] i [m] ∈ ∈ 1 1 1 1 δ I 2 Ω I 2 Ω ∑ d i max d i = . (7) ≥ d md − E ∞ ≥ d i [m] md − E ∞ d i [m] ∈ ∈ 2Via private communication, we have learned that Marco Piani has independently obtained a similar result for TCP maps.

9 Therefore, we have J = (1 mdδ)D + mdδJ′ where J′ = ( 1,... , m) CP[m] and mdδ 2 − F F ∈ ≤ md D (D, J). Now Corollary 1 guarantees that (1 η)D + ηJ′ SEP when η RSEP. From ⋄ − ∈ ≤ Lemma 1 and the convexity of LOCC, there exists an N-round LOCC protocol that implements this instrument with probability p and depolarizes (fails) with probability 1 p, for any p 1 . − ≤ md2 By coarse-graining the failure outcome into each of the m success outcomes by equal amounts, we have that the instrument

p (1 η)D + ηJ′ +(1 p)D =(1 pη)D + pηJ′ LOCC (8) − − − ∈ N 2 2 whenever pη RSEP /(md ). Taking pη = mdδ md D (D, J), it follows that J = (1 mdδ)D + ≤ ≤ 2 2 ⋄ − mdδJ′ LOCCN whenever D (D, J) RSEP/(md ) . ∈ ⋄ ≤ We next turn to the question of compactness. As LOCC itself is not closed (see Sect. 2.3), clearly it is not a compact set. However, we will prove that when restricted to finite round protocols with finite number of outcomes, compactness indeed holds. Such a result might not be entirely obvious; it is conceivable that, before coarse-graining to a finite number of outcomes in the final round, the protocol requires intermediate measurements with an unbounded number of outcomes. We will show that no such requirement can exist for a finite roundLOCC protocol. Our first stepis to apply Carath´eodory’s Theorem [Roc96] repeatedly to bound the number of measurement outcomes in each step. The following theorem is proven in Appendix A.

Theorem 2. For an N-partite system, suppose that J = ( 1,..., m) is an instrument in LOCCr. Let N E E dK denote the local dimension of party AK and D = ∏K=1 dK the dimension of the underlying global state space. Then there exists an r-round protocol that implements ( ,..., ) such that each instrument in E1 Em round l [r] consists of no more than mD4(r l+1) CP maps of the form M( )M† for some D D matrix ∈ − · × M, with some overall coarse-graining performed at the end of the protocol. Corollary 3. Consider an N-partite system of total dimension D. For any r, m N, the subset of instru- ∈ ments in LOCCr with m outcomes is compact in the set of all quantum instruments. Proof. By the previous theorem, any such LOCC instrument belonging to this set can be charac- terized by matrices:

(i1) (i1...ir 1) M , M ,..., M − : i1 [n1], i2 [n2],..., ir [nr] (9) i1 i2 ir ∈ ∈ ∈  (i1...il 1) 4(r l 1) where (i) for each i the M − are square matrices of some fixed size < D, (ii) n mD l il l − − (~) † (~) ≤ for 1 l r and (iii) ∑nl M s M s = 1 for all 1 l r and outcome strings ~s. As this ≤ ≤ il=1 il il ≤ ≤ is a finite collection of algebraic constraints, the set of feasible instruements is both closed and  bounded.

4 Is LOCC closed?

Bipartite entanglement is known to behave differently than its multipartite counterpart in many circumstances. One such example is the LOCC transformation of some pure state from one form to another. While in the bipartite case such a transformation can always be completed in one round of LOCC if it is possible at all [LP01], the same is not true for tripartite pure state manipulations. Among the separation results listed in Sect. 2.3, whether LOCC = LOCC in the bipartite case is unknown prior to this work. In this section, we exhibit a bipartite instrument J which is in LOCC but not in LOCC. This instrument acts only on two , and is given by J =( , , ) (10) E00 E01 E10

10 where

(ρ) := 11 11 ρ 11 11 , E00 | ih | | ih | 2 † 01(ρ) := ∑ Ti 0 0 ρ Ti 0 0 , (11) E i=1 ⊗ | ih | ⊗ | ih | 2   † 10(ρ) := ∑ 0 0 Ti ρ 0 0 Ti , E i=1 | ih |⊗ | ih |⊗   with

1 1 0 √ 0 T := √3 , T := 6 . (12) 1 0 1 2  2  √3 ! 0 3  q  We construct in Sect. 4.1 a sequence of LOCC instruments that converges to J, thereby showing J LOCC. In Sect. 4.3 we prove that J / LOCC, by demonstrating that the induced instrument ∈ ∈ J I on three qubits, where I denotes the identity instrument, is not in LOCC. We show that, if ⊗ J I is in LOCC, its effect on a particular three-qubit pure state violates the monotonicity of a ⊗ quantity that we derive in Sect. 4.2. This quantity pertains to a random entanglement distillation task that we also review and discuss.

4.1 Proofof J LOCC ∈ We construct a sequence of LOCC instruments J , J ,... taken directly from the Fortescue-Lo { 1 2 } random distillation scheme [FL07], except here we just consider bipartite systems. Let ǫ > 0 be fixed. Consider the measurement (ρ)= M ρ M 0 0 + M ρ M 1 1 where M 0 0 ⊗ | ih | 1 1 ⊗ | ih | M := √1 ǫ 0 0 + 1 1 , M := √ǫ 0 0 (13) 0 − | ih | | ih | 1 | ih | are diagonal. For each ν N, J is implemented by the following protocol. Alice and Bob each ∈ ν perform the measurement locally and share the measurement outcome. If the joint outcome is M one of 01, 10, or 11, they stop. If the joint outcome is 00, then they repeat the same measurement procedure again. After a maximum of ν iterations they stop. Then, coarse-graining is applied to obtain J = ( , , , ) where includes all the cases when Alice and Bob stop upon ν Eν00 Eν01 Eν10 Eν11 Eνij obtaining the joint outcome ij. More specifically, these four CP maps are respectively generated by ν ν µ µ 1 µ 1 µ the following sets of Kraus operators: M0 M0 , M0 M1 M0 − : µ [ν] , M1 M0 − M0 : µ 1 µ 1 { ⊗ } { ⊗ ∈ } { ⊗ µ [ν] , and M M − M M − : µ [ν] . ∈ } { 1 0 ⊗ 1 0 ∈ } To see what this instrument looks like, we consider the Choi matrices of its CP maps. The four Choi matrices corresponding to the four elements in Jν (up to the ordering of spaces in the tensor

11 product) are:

1 A′ B′ ν ν AB ν ν Ων00 = ∑ wx yz (M0 M0) wx yz (M0 M0), w,x,y,z=0 | ih | ⊗ ⊗ | ih | ⊗ 1 ν µ µ 1 µ µ 1 Ω A′ B′ AB ν01 = ∑ wx yz ∑ (M0 M1 M0 − ) wx yz (M0 M0 − M1), w,x,y,z=0 | ih | ⊗ µ=1 ⊗ | ih | ⊗ 1 ν µ 1 µ µ 1 µ Ω A′ B′ AB ν10 = ∑ wx yz ∑ (M1 M0 − M0 ) wx yz (M0 − M1 M0 ), w,x,y,z=0 | ih | ⊗ µ=1 ⊗ | ih | ⊗ 1 ν µ 1 µ 1 µ 1 µ 1 Ω A′ B′ AB ν11 = ∑ wx yz ∑ (M1 M0 − M1 M0 − ) wx yz (M0 − M1 M0 − M1). (14) w,x,y,z=0 | ih | ⊗ µ=1 ⊗ | ih | ⊗

We want to show that limǫ 0 limν ∞ Jν = J where the target instrument has been padded → → to become J = ( , , , 0) where the CP maps are defined in Eq. (11). We will do this by E00 E01 E10 Eij considering the limiting matrices Ωij := limǫ 0 limν ∞ Ωνij. Note that formally we have to choose → → ǫ to be some function of ν for specifying a sequence of instruments J1, J2,... that converges to J. However, we have verified that in all cases the double limit agrees with the single limit ν ∞ if c → we choose ǫ := ν− for some 0 < c < 1. Thus, for the sake of simplicity of the argument we will compute the double limit. µ 1 µ 1 For ij = 11, note that M M − = M − M = ǫ(1 ǫ)µ 1 0 0 , so we have 1 0 0 1 − − | ih | ν p 2 2µ 2 A′ A B′ B Ων11 = ∑ ǫ (1 ǫ) − 00 00 00 00 . (15) µ=1 − | ih | ⊗ | ih |

∞ 2 ∑∞ 2µ ǫ2 ǫ Ω Ω In the ν limit we get ǫ µ=0(1 ǫ) = 1 (1 ǫ)2 = 2 ǫ , so 11 = limǫ 0 limν ∞ ν11 = 0 → − − → → − − A A B B as desired. Similarly, for ij = 00, it is easy to see that Ω00 = 11 11 ′ 11 11 ′ . µ | ih | ⊗ | ih | For ij = 01, note that M = (1 ǫ)µ 0 0 + 1 1 , so we have 0 − | ih | | ih | ν 1 p 1 µ µ µ 1 µ 1 Ω A′ A B′ B ν01 = ∑ ∑ w y M0 w y M0 ∑ x z M1 M0 − x z M0 − M1 (16) µ=1 w,y=0 | ih | ⊗ | ih | !⊗ x,z=0 | ih | ⊗ | ih | !

ν µ µ A′ A (1 ǫ) (1 ǫ) µ 1 B B = ∑ − − ǫ(1 ǫ) − 00 00 ′ (17) (1 ǫ)µ 1 ⊗ − | ih | µ=1  − p  A A ǫ pν (1 ǫ)2µ (1 ǫ)3µ/2 ′ = ∑ − − 00 00 B′ B (18) 1 ǫ (1 ǫ)3µ/2 (1 ǫ)µ ⊗ | ih | − µ=1  − −  where the first tensor factor in the last two lines is written in the basis 00 , 11 . Evaluat- {| iA′ A | iA′ A} ing the geometric series as ν ∞ and then ǫ 0, we see that → → A A 1/2 2/3 ′ Ω = 00 00 B′ B. (19) 01 2/3 1 ⊗ | ih |   By permuting the parties, we obtain Ω10. Finally, it is easy to verify that Ω is indeed the Choi matrix of the map given by Eq. (11), 01 E01 and similarly for all other Ωij. Since the Hilbert space dimension is 4 (a small constant), entry- wise convergence of the Choi matrix is equivalence to the convergence of instruments defined in Eq. (1).

12 4.2 Digression: Random Concurrence Distillation As mentioned in Sect. 4, we will prove LOCC infeasibility of the bipartite instrument J in Eq. (10) by showing infeasibility of the induced tripartite transformation J I. Specifically, this is shown ⊗ by considering the action of the instrument J I on the state W = √1/3( 100 + 010 + 001 ). ⊗ | i | i | i | i The state W is the canonical representative of the so-called W-class of states which con- | i sists of all three-qubit pure states that are locally unitarily (LU) equivalent to ~x = √x 000 + | i 0| i √x 100 + √x 010 + √x 001 [DVC00] for some ~x = (x , x , x ) (with x = 1 x x A| i B| i C| i A B C 0 − A − B − x ). For any normalized pure W-class state ψ , we can unambiguously represent it by an ~x C | i ψ [KT10] as follows. First, it is not difficult to show that, if ~x is LU equivalent to ~x , and all entries | i | ′i of ~x are nonzero, then ~x = ~x′. If xK = 0 for exactly one party K, signifying that only bipartite entanglement exists, then by Schmidt decomposing the two-qubit state of the entangled parties, we can take x = 0 and can choose x max x , x if x = 0, and x x otherwise. Fi- 0 A ≥ { B C} A 6 B ≥ C nally, if xK = 0 for two or more parties, the state is fully separable, and we set xA = 1. Thus, LOCC transformations on W-class states can be viewed as transformations between their unique representations, and can be understood by tracking the changes to ~x. Now, consider the following transformation task. Suppose that Alice, Bob and Charlie are in possession of some W-class state, and they wish to obtain a bipartite pure state held by either Charlie-Alice or Charlie-Bob such that the expected concurrence is maximized. We call this task random concurrence distillation. Recall that for a two-qubit pure state of the form ψ = α 01 + | i | i β 10 , its concurrence C(ψ) is given by 2 αβ [Woo98]. | i | | The described problem generalizes two different tasks in quantum information processing: fixed-pair concurrence distillation and random EPR combing. In fixed-pair concurrence distillation, one target pair of parties is a priori specified, and the goal is for the third party to measure in such a way that maximizes the average post-measurement concurrence shared between the target pair. This optimal value is known as the Concurrence of Assistance (COA). In Ref. [LVvE03] it is shown that for a W-class state with x0 = 0, its COA is 2√xJ xK when parties J and K constitute the target pair. In contrast, in random EPR combing, two target pairs of parties are a priori specified, and the goal is to maximize the probability that either of these pairs obtains an EPR state. For example, if Alice is the common party to both target pairs, it has recently been shown that for W- class states with x = 0, the optimal probability is 2(x + x x x /x ) when x max x , x 0 B C − B C A A ≥ { B C} and 2xA otherwise [CCL12]. Random concurrence distillation is a hybrid of these two problems in that the goal is to randomly distill entanglement to two different pairs, but instead of using the probability of obtaining an EPR state as a success measure, we consider the average concurrence distilled. The following lemma, proved in Appendix B, provides an upper bound of the random concurrence distillable between a fixed party ⋆ A, B, C and the remaining two parties. ∈ { } Lemma 2. Let ⋆ A, B, C denote any fixed party. For a W-class state √x 100 + √x 010 + ∈ { } A| i B| i √x 001 take n , n = A, B, C ⋆ such that x x , and consider the function C| i { 1 2} { } \ { } n1 ≥ n2 x⋆ ⋆ 2 2√x xn1 + 3 xn2 x if xn1 = 0, (~x) := n1 6 (20) C (0 q otherwise.

The function is non-increasing on average under LOCC. Furthermore, it is strictly decreasing on average C when either party “⋆”orn1 performs a non-trivial measurement.

Remark. Note that if x⋆ = 0, then x > 0 and so (~x) = 0 by Eq. (20). If x = 0, then (~x) n1 C n1 C is defined to be 0 since xn2 = 0 and so the state is fully separable. When xn2 = 0, the function reduces to the bipartite concurrence measure between parties n and ⋆. Consequently, what C 1

13 Lemma 2 says is that for any tripartite to bipartite entanglement conversion of a W-class state, the maximum average concurrence that some fixed party “⋆” can share with any of the other two parties is no greater than (~x). Finally, observe that for the state W = √1/3( 100 + 010 + C | i | i | i 001 ), we have (W)= 8/9. | i C 4.3 LOCC Impossibility We now put the pieces together to prove that J / LOCC where J is the bipartite instrument ∈ defined in Eq. (10). Suppose J I is applied to W = √1/3( 100 + 010 + 001 ). This induces ⊗ | i | i | i | i the following state transformation: ω(AC) 0 0 (B) with probability 1/2, W W ⊗ | ih | (21) | ih | 7→ 0 0 (A) ω(BC) with probability 1/2, (| ih | ⊗ where 00 00 0 1/3 4/9 0 ω :=   . 0 4/9 2/3 0 00 00   Since Eq. (21) describes a mixed state transformation, Lemma 2 cannot be used directly. However, using the convex roof construction [Hor01], we can extend the monotone to mixed W-class C states. Recall that the set of mixed W-class states consists of all convex combinations of pure W- class states, biseparable states, and product states [ABLS01]. It represents a convex compact set in state space that is closed under LOCC operations. Therefore, the following is well-defined. Corollary 4. Let ⋆ A, B, C denote any fixed party and suppose ρ is some mixed W-class state. Define ∈ { } the function ˆ(ρ)= min ∑ pi (φi) (22) C p , φ C i | ii pi where the minimization is taken over all pure state decompositions of ρ = ∑ p φ φ , and is the pure i i| iih i| C state entanglement monotone defined by Eq. (20). Then ˆ is an entanglement monotone on the class of C mixed W-class states. For transformation (21), we see that ˆ(W) = 8/9 and the final value of ˆ in both outcomes C C reduces to (ω), the concurrence of ω (i.e., it is the convex-roof extension of the bipartite pure C state concurrence). Using Wootters’ formula for the concurrence [Woo98], we can compute that (ω) = 8/9. Thus, ˆ must remain invariant on average during each measurement in a proto- C C col that performs transformation (21). However, we can always decompose the first non-unitary measurement into a local pure state transformation on W , which by Lemma 2 will cause to | i C strictly decrease (choose n as the first measuring party). And because ˆ reduces to on pure 1 C C states, we therefore have that ˆ will necessarily decrease on average during the first non-trivial C measurement. Hence, transformation (21) cannot be performed by LOCC, and so likewise, the original bipartite instrument is infeasible by LOCC. This argument can also be used to prove the LOCC impossibility of other separable bipartite instruments. For instance, consider the three CP maps (ρ) := Π ρΠ† where Ei i i Π = 1 0 0 + 1 1 0 0 , 1 √2 | ih | | ih | ⊗ | ih |   Π = 0 0 1 0 0 + 1 1 , (23) 2 | ih |⊗ √2 | ih | | ih | Π = 1 1 1 1 .  3 | ih | ⊗ | ih |

14 As ∑3 Π†Π = I we have that ( , , ) is a separable quantum instrument, and its action on i=1 i i E1 E2 E3 W will yield bipartite mixed states with an average concurrence of 2√2/3 > 8/9. Hence, this | i transformation cannot even be approximated asymptotically, which provides another example of the operational gap between LOCC and SEP.

5 Whatdidwelearn?

In this article, we have closely studied the structure of LOCC operations. In light of recent findings concerning the of asymptotic LOCC processes, we have adopted the formalism of quan- tum instruments to precisely characterize the topological closure of LOCC. Additionally, we have proven that the class of LOCC instruments acting on two qubits is not closed. This resolves an open problem and reveals the complexity of LOCC even when dealing with the smallest bipartite systems. There are a few interesting questions related to our work that deserve additional investigation. First, all known examples that separate LOCC from LOCC or LOCC from SEP make use of the classical information obtained from a quantum measurement. For quantum channels with no classical register, it is unknown whether the same separation results hold (although, LOCCN can be separated from SEP by such channels [CDS08]). For instance, if we coarse grain over the three different maps given by Eq. (11), is the resulting channel feasible by LOCC? We conjecture that it is not, however our current proof techniques are unable to show this. On the other hand, one can ask how the operational classes compare if one is only interested in the classical information extracted from a quantum measurement, i.e., if attention is restricted only to POVMs. While the state discrimination results take such an approach to separate LOCC from SEP, the random distillation examples demonstrating LOCCN = LOCC = LOCC depend 6 6 crucially on the quantum outputs of the measurement. Thus, it may be possible that LOCC = LOCC for POVMs. We draw the reader’s attention to Ref. [KKB11] in which a particular state discrimination problem is presented that directly questions this possibility. Finally, by using the distance measure between instruments described in Sect. 2, one can mean- ingfully inquire about the size in separation between operational classes. When given some in- strument in SEP, what is the closet LOCC instrument? Furthermore, is this distance related to the nonlocal resources needed to implement the separable instrument? We hope this paper stimulates further research into such questions concerning the structure of LOCC.

Acknowledgments

We thank Matthias Kleinmann, Hermann Kampermann, and Dagmar Bruß for a helpful discus- sion on asymptotic LOCC. Additionally, much appreciation is extended to Daniel Gottesman for spotting an error in an earlier version of Sect. 4. DL, LM, and MO were supported by CFI, ORF, CIFAR, CRC, NSERC Discovery grant, NSERC QuantumWorks grant, MITACS, the Ontario Min- istry of Research and Innovation, and the US ARO/DTO. MO acknowledges additional support from the DARPA QUEST program under contract number HR0011-09-C-0047. AW was supported by the European Commission (STREP “QCS” and Integrated Project “QESSENCE”), the ERC (Ad- vanced Grant “IRQUAT”), a Royal Society Wolfson Merit Award and a Philip Leverhulme Prize. The Centre for Quantum Technologies is funded by the Singapore Ministry of Education and the National Research Foundation as part of the Research Centres of Excellence programme.

15 Appendices

A Proof of Theorem 2

Theorem 2. For an N-partite system, suppose that J = ( 1,..., m) is an instrument in LOCCr. Let N E E dK denote the local dimension of party AK and D = ∏K=1 dK the dimension of the underlying global state space. Then there exists an r-round protocol that implements ( ,..., ) such that each instrument in E1 Em round l [r] consists of no more than mD4(r l+1) CP maps of the form M( )M† for some D D matrix ∈ − · × M, with some overall coarse-graining performed at the end of the protocol.

Proof. We will prove Theorem 2 in multiple steps. A general r-round LOCC instrument can be represented by a tree partitioned into r levels. Within each level are nodes that correspond to the different one-way local LOCC instruments performed in that round. The nodes are specified by their respective measurement histories (i1i2 ... il ). Many parties may perform a quantum opera- tion at each node, but only one party can perform a non-trace-preserving operation, and this party may vary across different nodes at each level. Our first task is to convert a general protocol into one for which (i) except for the final round, the dimensions of the input and output spaces are the same for each map, and (ii) the party acting non-trivially at any node along the same level is fixed. To obtain (i) we first fine-grain each local map into the form specified by the Theorem, i.e., ρ MρM†. This modification will only increase 7→ the number of edges coming out from a node, but will not change the total number of rounds, and the original instrument can be recovered by suitable coarse-graining at the end. Then, we apply the polar decomposition M = UA where A is a square matrix and U is an isometry. Thus whenever M( )M† is performed within the protocol, it can be replaced with A( )A† combined · · with a pre-application of U in the next level. In other words, if the same party applies Kraus operator N next in the original protocol, then NU is applied instead. Polar decomposing NU, we obtain U′ A′ where A′ again maps the initial system to itself, and U′ is some other isometry to be moved to yet the next level. Doing this inductively for all levels yields condition (i) in which the state lives in the same input system throughout, except for the last step. To describe simplification (ii), we introduce some terminology. For an r-round LOCC protocol, we say its measurement-ordered expansion is a protocol obtained by increasing the number of rounds through the addition of trivial maps so that there is a unique predetermined party who can act non-trivially along all nodes in a given level. This new protocol will consist of no more than Nr levels. A measurement-ordered compression is a reversal of the expansion. Specifically, for a given LOCC protocol, consider every node (i1 ... il 2) in which a trivial map is performed at that node (this will be a measurement in round l 1), and− suppose that T is the subsequent one- − way local instrument implemented in round l. We modify the original protocol by performing T instead of the trivial map at node (i1 ... il 2). Doing this for each trivial operation and compressing recursively generates the measurement-ordered− compression of the original protocol. Now we return to the specific instrument J = ( ,..., ). For simplicity we will assume E1 Em that N = 2, and more general cases follow by analogous constructions. Suppose we are given some measurement-ordered protocol that can be compressed to r rounds and which implements J. Again for simplicity we will assume that the protocol just consists of 4 rounds in which the order of measurement is Alice, Bob, Alice, Bob; the case of more rounds can be proved by induction. For (i1...il 1) (i1...il 1) this protocol, the CP maps performed at node (i1 ... il 1) will be denoted by − ( − ) − Ai1 Bi1 when Alice (Bob) is the acting party in round l. Thus the entire instrument J can be expressed

16 A0 ρ ( B0 Ω A1′

A 1 Ai1

B2′

(i1) B2 Bi2

A3′

(i1i2) A3 Ai3

B4′

(i1i2i3) B4 Bi4

Figure 2: Construction of Ω with four rounds of LOCC according to Eq. (25). From the state Ω the action of [J] on ρ can be obtained by inputting ρ into wires A B and applying the projector Φ E 0 0 | i onto each of the the pairs A0 A1′ , B0B2′ , A1 A3′ , and B2B4′ as in Eq. (26). through the TCP map

m m ~ (i1i2i3) (i1i2) (i1) [J](ρ) = ∑ λ(ρ) λ λ = ∑ ∑ δλ(i) i (ρ) λ λ , (24) E E ⊗ | ih | Bi4 ◦Ai3 ◦ Bi2 ◦A 1 ⊗ | ih | λ=1 λ=1 ~i where δ (~i) = 1 if~i = (i , i , i , i ) is included in the coarse-graining of and 0 otherwise. Our λ 1 2 3 4 Eλ goal is to obtain a bound on the number of different~i needed to implement J. We construct an operator representation of [J] by introducing a copy of the local Hilbert space for each measure- E ment performed in each round. Then, we have the operator

m A′ A : :B′ B ~ A′ A′ A B′ (i1...i3) B′ B Ω 1 1 ··· 4 4 = ∑ ∑ δλ(i)(I 1 i )[Φ 1 1 ] (I 4 )[Φ 4 4 ] λ λ . (25) ⊗A 1 ⊗···⊗ ⊗ Bi4 ⊗ | ih | λ=1 ~i

A A di Here, Φ i′ i = ∑ ll mm is a maximally entangled state on two copies of the space possessed l,m=1 | ih | B′ B (i1...il 1) (i1...il 1) by Ai (likewise for Φ i i ), and the − and − act on the “unprimed” copies in their Ail Bil respective rounds. The action of [J] is recovered by first supposing that the input state ρ is shared E across local registers A and B . Then [J](ρ) is given by the following equality (see Fig. 2): 0 0 E

A0 B0 A A′ : :B B′ A A′ B B′ A A′ B B′ [J](ρ) = TrA B A A B B A B ρ Ω 1 1 ··· 4 4 Φ 0 1 Φ 0 2 Φ 1 3 Φ 2 4 . (26) E 0 0 1 1′ 2 2′ 3′ 4′ ⊗ h  i We next use Carath´eodory’s Theorem to rewrite the sum over~i in Eq. (25). Lemma 3 (Carath´eodory’s Theorem [Roc96]). Let S be a subset of Rn and conv(S) its convex hull. Then any x conv(S) can be expressed as a convex combination of at most n + 1 elements of S. ∈

17 As a Hermitian operator, every (unnormalized) mixed can be expressed by (dim)2 real numbers, where dim is the dimension of the underlying space on which the operator acts. We first consider the sum over λ and i4 in Eq. (25) with i1, i2, i3 fixed. This sum can be expressed as

m (i1i2i3) ~ B′ (i1i2i3) B′ B ρ := ∑ ∑ δλ(i)(I 4 )[Φ 4 4 ] λ λ ⊗ Bi4 ⊗ | ih | λ=1 i4 m (i1i2i3) B (i1i2i3) B B = ∑ ∑ q (I 4′ )[Φ 4′ 4 ] λ λ (27) λ,i4 ⊗ Bi4 ⊗ | ih | λ=1 i4

( ) where (i) for each~i, q i1i2i3 0 is nonzero for only one value of λ (the one for which δ (~i) = 1), λ,i4 λ ( ) ≥ (ii) for each λ, q i1i2i3 is nonzero for at most d4 values of ~i (by Carath´eodory’s Theorem), and λ,i4 B ( ) ( ) (iii) ∑m ∑ q i1i2i3 i1i2i3 is trace-preserving. Thus by (i) and (ii) the sum in Eq. (27) contains no λ=1 i4 λ,i4 i4 4 B more than mdB terms. From these operators ρ(i1i2i3), we construct the set

(i1i2) A′ (i1i2) A′ A (i1i2i3) T := (I 3 )[Φ 3 3 ] ρ : i3 (28) ⊗Ai3 ⊗

(i i2)  and its convex hull conv(T 1 ). For each fixed pair i1i2, we can apply Carath´eodory’s Theorem again when summing over i3 in Eq. (25) to obtain

m ~ A′ (i1i2) A′ A B′ (i1i2i3) B′ B ∑ ∑ ∑δλ(i)(I 3 )[Φ 3 3 ] (I 4 )[Φ 4 4 ] λ λ ⊗Ai3 ⊗ ⊗ Bi4 ⊗ | ih | i3 λ=1 i4

A (i1i2) A A (i i i ) = ∑(I 3′ )[Φ 3′ 3 ] ρ 1 2 3 ⊗Ai3 ⊗ i3

(i1i2) (i1i2) ( ) = ∑ q (IA3′ )[ΦA3′ A3 ] ρ i1i2i3 i3 ⊗Ai3 ⊗ i3 m (i1i2) (i1i2i3) A (i1i2) A A B (i1i2i3) B B = ∑ ∑ q q (I 3′ )[Φ 3′ 3 ] (I 4′ )[Φ 4′ 4 ] λ λ , (29) i3 λ,i4 ⊗Ai3 ⊗ ⊗ Bi4 ⊗ | ih | i3,i4 λ=1

(i1i2) 4 (i1i2) (i1i2) where q 0is nonzerofor at most m(dAdB) values of i3 and ∑ q is trace-preserving. i3 ≥ i3 i3 Ai3 Repeating this argument on the sums over i1 and i2 respectively, Eq. (25) becomes

m A′ A : :B′ B (i1) (i1i2) (i1i2i3) A′ A′ A B′ (i1i2i3) B′ B Ω 1 1 ··· 4 4 = ∑ ∑ q q q q (I 1 i )[Φ 1 1 ] (I 4 )[Φ 4 4 ] λ λ i1 i2 i3 λ,i4 ⊗A 1 ⊗···⊗ ⊗ Bi4 ⊗ | ih | λ=1 ~i

(i1) 4 where again, q 0 is nonzero for at most m(dBdAdB) values of i2 while qi 0 is nonzero for i2 ≥ 1 ≥ 4 (i1) (i1) at most m(dAdBdAdB) values of i1. Both ∑ qi i and ∑ q are trace-preserving. i1 1 A 1 i2 i2 Bi2 Thus the modified protocol consists of the instrument sequence 1. Alice: q : q > 0 , { i1 Ai1 i1 } ( ) ( ) ( ) 2. Bob: q i1 i1 : q i1 > 0 , { i2 Bi2 i2 } ( ) ( ) ( ) 3. Alice: q i1i2 i1i2 : q i1i2 > 0 , { i3 Ai3 i3 } ( ) ( ) ( ) 4. Bob: q i1i2i3 i1i2i3 : q i1i2i3 > 0 . { λ,i4 Bi4 λ,i4 }

18 An overall coarse-graining for each value of λ is applied after the final measurement to obtain J. Finally, observe that this modified protocol only differs from the original in the one-way local instruments performed at each node. Furthermore, if a trivial operation is performed at any node in the original protocol, it will likewise be performed at the same node in this modified proto- col. Therefore, the ability to compress the original protocol to r rounds of LOCC implies that the modified protocol can also be compressed to r rounds.

B Proof of Lemma 2

Lemma 2. Let ⋆ A, B, C denote any fixed party. For a W-class state √x 100 + √x 010 + ∈ { } A| i B| i √x 001 take n , n = A, B, C ⋆ such that x x , and consider the function C| i { 1 2} { } \ { } n1 ≥ n2 x⋆ ⋆ 2 2√x xn1 + 3 xn2 x if xn1 = 0, (~x) := n1 6 (20) C (0 q otherwise.

The function is non-increasing on average under LOCC. Furthermore, it is strictly decreasing on average C when either party “⋆”orn1 performs a non-trivial measurement.

Proof. Note that if any of the xA, xB, xC is zero, the lemma is already true, so, we can focus on the case when all three parameters are nonzero. The proof is analogous to the ones given in Refs. [CCL12, CCL11] where similar W-class entanglement monotones were derived. We can de- compose each local measurement into a sequence of binary measurements [AO08, OB05]. By continuity of , it is then sufficient to show that is non-increasing on average under such a mea- C C surement. Each such measurement is specified by two Kraus operators M : λ = 1, 2 , which { λ } can always be expressed in the form

√a b M := U λ λ , (30) λ λ · 0 √c  λ where aλ, cλ 0, bλ C and Uλ is unitary (which acts trivially on the representation ~x). By the ≥ ∈ 2 † completeness relation ∑λ=1 Mλ Mλ = I we have

2 a √a b 1 0 ∑ λ λ λ = , (31) √a b b 2 + c 0 1 λ=1  λ λ∗ | λ| λ   thus a + a = 1 and c + c 1 with equality if and only if b = b = 0. 1 2 1 2 ≤ 1 2 The state change induced by any such measurement can be understood readily by considering a concrete example, say when the first party measures:

√x 000 + √x 100 + √x 010 + √x 001 0| i A| i B| i C| i 1 (√aλ x0 + bλ√xA) 000 + √cλxA 100 + √aλ xB 010 + √aλ xC 001 , (32) 7→ √pλ | i | i | i | i   where the final state is normalized with pλ being the probability of obtaining outcome λ. The post-measurement state in Eq. (32) is represented by vector ~x = ( cλ x , aλ x , aλ x ). More gen- λ pλ A pλ B pλ C erally, when party K A, B, C measures and outcome λ occurs, then the representing vector ∈ { } components transform as

cλ aλ xK xK and xJ xJ for J A, B, C K . (33) 7→ pλ 7→ pλ ∈ { } \ { }

19 We will consider three separate cases: (i) when K = ⋆, (ii) when K n , n with x > x , ∈ { 1 2} n1 n2 and (iii) when K n , n and x = x . Note from Eq. (33) that in case (i) the assignments ∈ { 1 2} n1 n2 of parties n1 and n2 will remain unchanged in the pre and post-measurement states. To simplify analysis, we can impose this also in case (ii) by assuming a sufficiently weak measurement (i.e., one for which a c and a c ). This is without loss of generality since any measurement can 1 ≈ 1 2 ≈ 2 be decomposed as a sequence of weak measurements. However, in case (iii), the assignment of parties n1 and n2 in the post-measurement states can differ from the pre-measurement state and might even depend on the outcome. Define the average change in incurred by the measurement as C ∆ := p (~x )+ p (~x ) (~x). (34) C 1C 1 2C 2 −C where pλ is the probability to obtain outcome λ and ~xλ is the representation of the corresponding post-measurement state. Case (i) If party “⋆” makes a measurement,

2 x⋆ ∆ = 2√x⋆x + x √c a + √c a 1 0, (35) C n1 3 n2 x 1 1 2 2 − ≤  r n1   with equality obtained only under the trivial measurement a = c = d and a = c = 1 d for 1 1 2 2 − some d [0, 1]. ∈ > Case (ii) Suppose either party n1 or n2 measures and xn1 xn2 . If party n2 measures, the > assignment of parties in all post-measurement states remains the same if aλ xn1 cλxn2 for all λ 1, 2 . Subject to constraints imposed by Eq. (31), we can guarantee this by assuming that the ∈ { } measurement is sufficiently weak (i.e., a c and a c ). Then 1 ≈ 1 2 ≈ 2 2 2 x⋆ 2 x⋆ ∆ = ∑ aλ2√x⋆xn + cλ xn 2√x⋆xn + xn C 1 3 2 x − 1 3 2 x λ=1  r n1   r n1  2 x⋆ =(c + c 1) x 0, (36) 1 2 − 3 n2 x ≤ r n1 with equality attained only when c1 + c2 = 1 (and thus b1 = b2 = 0). Similarly, if party n1 performs a sufficiently weak measurement (i.e., c x > a x for all λ 1, 2 ), λ n1 λ n2 ∈ { }

2 x⋆ a1 a2 ∆ = 2√x⋆x √a c + √a c 1 + x a + a 1 . (37) C n1 1 1 2 2 − 3 n2 x 1 c 2 c − r n1  r 1 r 2   Moreover, under the weakness constraint,

∂ a2 1 xn a2 ∆ = √x⋆x 1 2 > 0. (38) ∂c C n1 c − 3 x c 2 r 2  n1 2  Thus, for fixed a (thus also a ) and c , the value of ∆ is maximized when c = 1 c . Substituting 1 2 1 C 2 − 1 this into Eq. (37) and evaluating around the point (a1, c1)=(1/2, 1/2), we seethat to secondorder in (a 1/2) and (c 1/2): 1 − 1 − xn ∆ √x⋆x 2 1 (a c )2 0, (39) C ≈ n1 x − 1 − 1 ≤  n1  with equality obtained again only by the trivial measurement. Thus, ∆ is non-positive in some C neighborhood of (a1, c1)=(1/2, 1/2).

20 Case (iii) Finally, consider when xn1 = xn2 and either party n1 or n2 measures. Parties n1 and n2 are symmetric prior to measurement, but their ordering can possibly change across the two post-measurement states. Specifically,

2 x⋆x (a + 1 c ) if a c , ~ √ n1 λ 3 λ λ λ pλ (xλ)= 3/2 1/2 ≥ (40) C ⋆ 1 > (2√x xn1 √aλcλ + 3 aλ cλ− if cλ aλ. In both intervals, p (~x ) is an increasing function of c . Therefore, like before, the total average λC λ λ change in is maximized when c and c are both large, i.e., c + c = 1. Thus, taking a such that C 1 2 1 2 1 a c (otherwise take a = 1 a such that a c ), and noting that the function from Eq. (20) 1 ≥ 1 2 − 1 2 ≥ 2 C 8 ⋆ reduces to 3 √x xn1 prior to measurement, we get

1 1 3/2 1/2 4 ∆ = 2√x⋆x a + c + (1 a )(1 c )+ (1 a ) (1 c )− C n1 1 3 1 − 1 − 1 3 − 1 − 1 − 3  q  1     √x⋆x (a c )3 0. (41) ≈−3 n1 1 − 1 ≤ Equality holds only with a trivial measurement.

References

[ABLS01] Antonio Ac´ın, Dagmar Bruß, Maciej Lewenstein, and Anna Sanpera. Clas- sification of mixed three-qubit states. Phys. Rev. Lett., 87(4):040401, Jul 2001. arXiv:quant-ph/0103025, doi:10.1103/PhysRevLett.87.040401. [AO08] Erika Anderson and Daniel K. L. Oi. Binary search trees for generalized measurements. Phys. Rev. A, 77(5):052104, May 2008. arXiv:0712.2665, doi:10.1103/PhysRevA.77.052104. [BBC+93] Charles H. Bennett, Gilles Brassard, Claude Cr´epeau, Richard Jozsa, Asher Peres, and William K. Wootters. Teleporting an unknown quantum state via dual classical and Einstein-Podolsky-Rosen channels. Phys. Rev. Lett., 70(13):1895–1899, Mar 1993. doi:10.1103/PhysRevLett.70.1895. [BBPS96] Charles H. Bennett, Herbert Bernstein, , and . Concentrating partial entanglement by local operations. Phys. Rev. A, 53(4):2046–2052, Apr 1996. arXiv:quant-ph/9511030, doi:10.1103/PhysRevA.53.2046. [BCH+02] Dagmar Bruß, J. Ignacio Cirac, Paweł Horodecki, Florian Hulpke, Barbara Kraus, Maciej Lewenstein, and Anna Sanpera. Reflections upon separability and dis- tillability. J. Mod. Optics, 49(8):1399–1418, 2002. arXiv:quant-ph/0110081, doi:10.1080/09500340110105975. [BDF+99] Charles H. Bennett, David P. DiVincenzo, Christopher A. Fuchs, Tal Mor, Eric Rains, Peter W. Shor, John A. Smolin, and William K. Wootters. without entanglement. Phys. Rev. A, 59(2):1070–1091, Feb 1999. arXiv:quant-ph/9804053, doi:10.1103/PhysRevA.59.1070. [BDM+99] Charles H. Bennett, David P. DiVincenzo, Tal Mor, Peter W. Shor, John A. Smolin, and Barbara M. Terhal. Unextendible product bases and bound entangle- ment. Phys. Rev. Lett., 82(26):5385–5388, Jun 1999. arXiv:quant-ph/9808030, doi:10.1103/PhysRevLett.82.5385.

21 [BDSW96] Charles H. Bennett, David P. DiVincenzo, John A. Smolin, and William K. Wootters. Mixed-state entanglement and . Phys. Rev. A, 54(5):3824–3851, Nov 1996. arXiv:quant-ph/9604024, doi:10.1103/PhysRevA.54.3824.

[BPR+00] Charles H. Bennett, Sandu Popescu, Daniel Rohrlich, John A. Smolin, and Ashish V. Thapliyal. Exact and asymptotic measures of multipartite pure state entan- glement. Phys. Rev. A, 63(1):012307, Dec 2000. arXiv:quant-ph/9908073, doi:10.1103/PhysRevA.63.012307.

[CCL11] Wei Cui, Eric Chitambar, and Hoi-Kwong Lo. Randomly distilling W-class states into general configurations of two-party entanglement. Phys. Rev. A, 84(5):052301, Nov 2011. arXiv:1106.1209, doi:10.1103/PhysRevA.84.052301.

[CCL12] Eric Chitambar, Wei Cui, and Hoi-Kwong Lo. Increasing entanglement mono- tones by separable operations. Phys. Rev. Lett., 108(24):240504, Jun 2012. doi:10.1103/PhysRevLett.108.240504.

[CDKL01] J. Ignacio Cirac, Wolfgang D¨ur, Barbara Kraus, and Maciej Lewenstein. En- tangling operations and their implementation using a small amount of entangle- ment. Phys. Rev. Lett., 86(3):544–547, Jan 2001. arXiv:quant-ph/0007057, doi:10.1103/PhysRevLett.86.544.

[CDS08] Eric Chitambar, Runyao Duan, and Yaoyun Shi. Tripartite entanglement transforma- tions and tensor rank. Phys. Rev. Lett., 101(14):140502, Oct 2008. arXiv:0805.2977, doi:10.1103/PhysRevLett.101.140502.

[Che04] Anthony Chefles. Condition for unambiguous state discrimination using local operations and classical communication. Phys. Rev. A, 69(5):050307, May 2004. doi:10.1103/PhysRevA.69.050307.

[Chi11] Eric Chitambar. Local quantum transformations requiring infinite rounds of classi- cal communication. Phys. Rev. Lett., 107(19):190502, Nov 2011. arXiv:1105.3451, doi:10.1103/PhysRevLett.107.190502.

[Cho75] Man-Duen Choi. Completely positive linear maps on complex matrices. Liner Alg. Appl., 10(3):285–290, Jun 1975. doi:10.1016/0024-3795(75)90075-0.

[CLMO12] Andrew M. Childs, Debbie Leung, Laura Manˇcinska, and Maris Ozols. A framework for bounding nonlocality of state discrimination, 2012. arXiv:1206.5822.

[Coh07] Scott M. Cohen. Local distinguishability with preservation of entangle- ment. Phys. Rev. A, 75(5):052313, May 2007. arXiv:quant-ph/0602026, doi:10.1103/PhysRevA.75.052313.

[DFXY09] Runyao Duan, Yuan Feng, Yu Xin, and Mingsheng Ying. Distinguishability of quan- tum states by separable operations. IEEE Trans. Inf. Theory, 55(3):1320–1330, Mar 2009. arXiv:0705.0795, doi:10.1109/TIT.2008.2011524.

[DHR02] Matthew J. Donald, Michał Horodecki, and Oliver Rudolph. The unique- ness theorem for entanglement measures. J. Math. Phys., 43:4252–4272, 2002. arXiv:quant-ph/0105017, doi:10.1063/1.1495917.

22 [DL70] E. Brian Davies and John T. Lewis. An operational approach to quantum probability. Comm. Math. Phys., 17(3):239–260, 1970. URL: http://projecteuclid.org/euclid.cmp/1103842336, doi:10.1007/BF01647093.

[DLT02] David P. DiVincenzo, Debbie W. Leung, and Barbara M. Terhal. Quantum data hid- ing. IEEE Trans. Inf. Theory, 48(3):580–598, Mar 2002. arXiv:quant-ph/0103098, doi:10.1109/18.985948.

[DVC00] Wolfgang D¨ur, Guifre Vidal, and J. Ignacio Cirac. Three qubits can be en- tangled in two inequivalent ways. Phys. Rev. A, 62(6):062314, Nov 2000. arXiv:quant-ph/0005115, doi:10.1103/PhysRevA.62.062314.

[EW02] Tilo Eggeling and Reinhard F. Werner. Hiding classical data in multipartite quan- tum states. Phys. Rev. Lett., 89(9):097905, Aug 2002. arXiv:quant-ph/0203004, doi:10.1103/PhysRevLett.89.097905.

[FL07] Ben Fortescue and Hoi-Kwong Lo. Random bipartite entanglement from W and W- like states. Phys. Rev. Lett., 98(26):260501, Jun 2007. arXiv:quant-ph/0607126, doi:10.1103/PhysRevLett.98.260501.

[GB03] Leonid Gurvits and Howard Barnum. Separable balls around the maximally mixed multipartite quantum states. Phys. Rev. A, 68(4):042312, Oct 2003. arXiv:quant-ph/0302102, doi:10.1103/PhysRevA.68.042312.

[GL03] Daniel Gottesman and Hoi-Kwong Lo. Proof of security of with two-way classical communications. IEEE Trans. Inf. Theory, 49(2):457–475, Feb 2003. arXiv:quant-ph/0105121, doi:10.1109/TIT.2002.807289.

[HHH96] Michał Horodecki, Paweł Horodecki, and Ryszard Horodecki. Separability of mixed states: necessary and sufficient conditions. Phys. Lett. A, 223(1–2):1–8, 1996. arXiv:quant-ph/9605038, doi:10.1016/S0375-9601(96)00706-2.

[HHH98] Michał Horodecki, Paweł Horodecki, and Ryszard Horodecki. Mixed-state entanglement and distillation: is there a “bound” entanglement in nature? Phys. Rev. Lett., 80(24):5239–5242, Jun 1998. arXiv:quant-ph/9801069, doi:10.1103/PhysRevLett.80.5239.

[HHH00] Michał Horodecki, Paweł Horodecki, and Ryszard Horodecki. Limits for entanglement measures. Phys. Rev. Lett., 84(9):2014–2017, Feb 2000. arXiv:quant-ph/9908065, doi:10.1103/PhysRevLett.84.2014.

[Hor97] Paweł Horodecki. Separability criterion and inseparable mixed states with positive partial transposition. Phys. Lett. A, 232(5):333–339, 1997. arXiv:quant-ph/9703004, doi:10.1016/S0375-9601(97)00416-7.

[Hor01] Michał Horodecki. Entanglement measures. Quant. Inf. Comput., 1(1):3–26, Jun 2001. URL: http://www.rintonpress.com/journals/qic-1-1/miarprz3.pdf.

[Jam72] Andrzej Jamiołkowski. Linear transformations which preserve trace and pos- itive semidefiniteness of operators. Rep. Math. Phys., 3(4):275–278, Dec 1972. doi:10.1016/0034-4877(72)90011-0.

23 [KKB11] Matthias Kleinmann, Hermann Kampermann, and Dagmar Bruß. Asymp- totically perfect discrimination in the local-operation-and-classical-communication paradigm. Phys. Rev. A, 84(4):042326, Oct 2011. arXiv:1105.5132, doi:10.1103/PhysRevA.84.042326.

[KSV02] Alexei Yu. Kitaev, Alexander Shen, and Mikhail N. Vyalyi. Classical and Quantum Com- putation, volume 47 of Graduate Studies in Mathematics. American Mathematical Society, 2002. URL: http://books.google.com/books?id=qYHTvHPvmG8C.

[KT10] Seckin Kıntas¸and Sadi Turgut. Transformations of W-type entangled states. J. Math. Phys., 51:092202, 2010. arXiv:1003.2118, doi:10.1063/1.3481573.

[KTYI07] Masato Koashi, Fumitaka Takenaga, Takashi Yamamoto, and Nobuyuki Imoto. Quan- tum nonlocality without entanglement in a pair of qubits, 2007. arXiv:0709.3196.

[LP01] Hoi-Kwong Lo and Sandu Popescu. Concentrating entanglement by local actions: Be- yond mean values. Phys. Rev. A, 63(2):022301, Jan 2001. arXiv:quant-ph/9707038, doi:10.1103/PhysRevA.63.022301.

[LVvE03] Thomas Laustsen, Frank Verstraete, and Steven J. van Enk. Local vs. joint mea- surements for the entanglement of assistance. Quant. Inf. Comput., 3(1):64–83, Jan 2003. URL: http://www.rintonpress.com/xqic3/qic-3-1/064-083.pdf, arXiv:quant-ph/0206192.

[MP95] Serge Massar and Sandu Popescu. Optimal extraction of information from finite quantum ensembles. Phys. Rev. Lett., 74(8):1259–1263, Feb 1995. doi:10.1103/PhysRevLett.74.1259.

[Nie99] Michael A. Nielsen. Conditions for a class of entanglement transforma- tions. Phys. Rev. Lett., 83(2):436–439, Jul 1999. arXiv:quant-ph/9811053, doi:10.1103/PhysRevLett.83.436.

[OB05] Ognyan Oreshkov and Todd A. Brun. Weak measurements are univer- sal. Phys. Rev. Lett., 95(11):110409, Sep 2005. arXiv:quant-ph/0503017, doi:10.1103/PhysRevLett.95.110409.

[OH08] Masaki Owari and Masahito Hayashi. Two-way classical communication re- markably improves local distinguishability. New J. Phys., 10(1):013006, 2008. arXiv:0708.3154, doi:10.1088/1367-2630/10/1/013006.

[PV07] Martin B. Plenio and Shashank Virmani. An introduction to entan- glement measures. Quant. Inf. Comput., 7(1&2):1–51, Jan 2007. URL: http://www.rintonpress.com/xqic7/qic-7-12/001-051.pdf, arXiv:quant-ph/0504163.

[PW91] Asher Peres and William K. Wootters. Optimal detection of quantum information. Phys. Rev. Lett., 66(9):1119–1122, Mar 1991. doi:10.1103/PhysRevLett.66.1119.

[Rai97] Eric M. Rains. Entanglement purification via separable superoperators, 1997. arXiv:quant-ph/9707002.

24 [Rai01] Eric M. Rains. A semidefinite program for distillable entanglement. IEEE Trans. Inf. Theory, 47(7):2921–2933, Nov 2001. arXiv:quant-ph/0008047, doi:10.1109/18.959270.

[Rin04] Sergio De Rinaldis. Distinguishability of complete and unextendible product bases. Phys. Rev. A, 70(2):022309, Aug 2004. arXiv:quant-ph/0304027, doi:10.1103/PhysRevA.70.022309.

[Roc96] R. Tyrell Rockafellar. Convex analysis. Princeton Math- ematical Series. Princeton University Press, 1996. URL: http://books.google.com/books?id=1TiOka9bx3sC.

[TDL01] Barbara M. Terhal, David P. DiVincenzo, and Debbie W. Leung. Hiding bits in Bell states. Phys. Rev. Lett., 86(25):5807–5810, Jun 2001. arXiv:quant-ph/0011042, doi:10.1103/PhysRevLett.86.5807.

[VPRK97] Vlatko Vedral, Martin B. Plenio, Michael A. Rippin, and Peter L. Knight. Quantifying entanglement. Phys. Rev. Lett., 78(12):2275–2279, Mar 1997. arXiv:quant-ph/9702027, doi:10.1103/PhysRevLett.78.2275.

[Wat05] John Watrous. Notes on super-operator norms induced by Schat- ten norms. Quant. Inf. Comput., 5(1):58–68, Jan 2005. URL: http://www.rintonpress.com/xqic5/qic-5-1/058-068.pdf, arXiv:quant-ph/0411077.

[Wer89] Reinhard F. Werner. Quantum states with Einstein-Podolsky-Rosen correlations admitting a hidden-variable model. Phys. Rev. A, 40(8):4277–4281, Oct 1989. doi:10.1103/PhysRevA.40.4277.

[Woo98] William K. Wootters. Entanglement of formation of an arbitrary state of two qubits. Phys. Rev. Lett., 80(10):2245–2248, Mar 1998. arXiv:quant-ph/9709029, doi:10.1103/PhysRevLett.80.2245.

[XD08] Yu Xin and Runyao Duan. Local distinguishability of orthogonal 2 3 ⊗ pure states. Phys. Rev. A, 77(1):012315, Jan 2008. arXiv:0709.1651, doi:10.1103/PhysRevA.77.012315.

25