Loading [MathJax]/jax/output/SVG/jax.js
Research article Special Issues

Empirical coordination of separable quantum correlations

  • We introduce the notion of empirical coordination for quantum correlations. Quantum mechanics enables the calculation of probabilities for experimental outcomes, emphasizing statistical averages rather than detailed descriptions of individual events. Empirical coordination is thus a natural framework for quantum systems. Focusing on the cascade network, the optimal coordination rates are established, indicating the minimal resources required to simulate, on average, a quantum state. As we consider a network with classical communication links, superposition cannot be maintained, hence the quantum correlations are separable (i.e., a convex combination of product states). This precludes entanglement. Providing the users with shared randomness, before communication begins, does not affect the optimal rates for empirical coordination. We begin with a rate characterization for a basic two-node network, and then generalize to a cascade network. The special case of a network with an isolated node is considered as well. The results can be further generalized to other networks as our analysis includes a generic achievability scheme. The optimal rate formula involves optimization over a collection of state extensions. This is a unique feature of the quantum setting, as the classical parallel does not include optimization. As demonstrated through examples, the performance depends heavily on the choice of decomposition. We further discuss the consequences of our results for quantum cooperative games.

    Citation: Husein Natur, Uzi Pereg. Empirical coordination of separable quantum correlations[J]. AIMS Mathematics, 2025, 10(4): 10028-10061. doi: 10.3934/math.2025458

    Related Papers:

    [1] İsmet Gölgeleyen, Özlem Kaytmaz . Uniqueness for a Cauchy problem for the generalized Schrödinger equation. AIMS Mathematics, 2023, 8(3): 5703-5724. doi: 10.3934/math.2023287
    [2] Luigi Accardi, El Gheted Soueidi, Abdessatar Souissi, Mohamed Rhaima, Farrukh Mukhamedov, Farzona Mukhamedova . Structure of backward quantum Markov chains. AIMS Mathematics, 2024, 9(10): 28044-28057. doi: 10.3934/math.20241360
    [3] Miguel Vivas-Cortez, Muhammad Aamir Ali, Hüseyin Budak, Ifra Bashir Sial . Post-quantum Ostrowski type integral inequalities for functions of two variables. AIMS Mathematics, 2022, 7(5): 8035-8063. doi: 10.3934/math.2022448
    [4] A. M. Mishra . On quantum aspects of general theory of relativity and its detectors. AIMS Mathematics, 2020, 5(1): 332-341. doi: 10.3934/math.2020022
    [5] Nawres A. Alwan, Suzan J. Obaiys, Nadia M. G. Al-Saidi, Nurul Fazmidar Binti Mohd Noor, Yeliz Karaca . A multi-channel quantum image representation model with qubit sequences for quantum-inspired image and image retrieval. AIMS Mathematics, 2025, 10(5): 10994-11035. doi: 10.3934/math.2025499
    [6] Abdessatar Souissi, El Gheteb Soueidy, Mohamed Rhaima . Clustering property for quantum Markov chains on the comb graph. AIMS Mathematics, 2023, 8(4): 7865-7880. doi: 10.3934/math.2023396
    [7] Mohra Zayed, Taghreed Alqurashi, Shahid Ahmad Wani, Cheon Seoung Ryoo, William Ramírez . Several characterizations of bivariate quantum-Hermite-Appell Polynomials and the structure of their zeros. AIMS Mathematics, 2025, 10(5): 11184-11207. doi: 10.3934/math.2025507
    [8] Limin Zhou, Qiuyang Wang . Simple PKE schemes from the TSPEM problem. AIMS Mathematics, 2024, 9(8): 22197-22212. doi: 10.3934/math.20241079
    [9] Changlong Yu, Jing Li, Jufang Wang . Existence and uniqueness criteria for nonlinear quantum difference equations with p-Laplacian. AIMS Mathematics, 2022, 7(6): 10439-10453. doi: 10.3934/math.2022582
    [10] Wael Alosaimi, Abdullah Alharbi, Hashem Alyami, Bader Alouffi, Ahmed Almulihi, Mohd Nadeem, Rajeev Kumar, Alka Agrawal . Analyzing the impact of quantum computing on IoT security using computational based data analytics techniques. AIMS Mathematics, 2024, 9(3): 7017-7039. doi: 10.3934/math.2024342
  • We introduce the notion of empirical coordination for quantum correlations. Quantum mechanics enables the calculation of probabilities for experimental outcomes, emphasizing statistical averages rather than detailed descriptions of individual events. Empirical coordination is thus a natural framework for quantum systems. Focusing on the cascade network, the optimal coordination rates are established, indicating the minimal resources required to simulate, on average, a quantum state. As we consider a network with classical communication links, superposition cannot be maintained, hence the quantum correlations are separable (i.e., a convex combination of product states). This precludes entanglement. Providing the users with shared randomness, before communication begins, does not affect the optimal rates for empirical coordination. We begin with a rate characterization for a basic two-node network, and then generalize to a cascade network. The special case of a network with an isolated node is considered as well. The results can be further generalized to other networks as our analysis includes a generic achievability scheme. The optimal rate formula involves optimization over a collection of state extensions. This is a unique feature of the quantum setting, as the classical parallel does not include optimization. As demonstrated through examples, the performance depends heavily on the choice of decomposition. We further discuss the consequences of our results for quantum cooperative games.



    Shannon theory for point-to-point networks has had a profound influence on communication in the digital age [1,2,3]. However, the simplistic model of a single source-destination pair does not capture many critical aspects of real-world networks [4]. In practice, networked systems often involve multiple sources and destinations, requiring the network to compute functions or make decisions rather than merely transmit data. The Internet of Things (IoT) introduces additional challenges due to its reliance on a shared medium [5]. Furthermore, networks entail intricate tradeoffs between competition for resources [6,7], cooperation for collective gain [8], and security [9]. Network information theory seeks to address fundamental questions of information flow and processing while incorporating these essential characteristics of real-world networks [10,11,12,13]. Recent advances in IoT have drawn attention to the role of coordination in networks with diverse topologies [14].

    Coordination is a fundamental framework in network information theory [15]. Cuff et al. [16] introduced a general information-theoretic model for network coordination where, as opposed to traditional coding tasks, the objective is not to exchange messages between network nodes but rather generate correlation [17]. Two types of coordination tasks were introduced in the classical framework [16]. In strong coordination, the users produce actions in order to simulate a product distribution. That is, the joint distribution resembles that of a particular memoryless source [18]. Empirical coordination imposes a weaker and less stringent condition compared to strong coordination. It requires the type, i.e., the frequency of actions, to converge into a desired distribution [19]. There are many information-theoretic tasks that are closely related to coordination, such as channel/source simulation [21,22,23,24], randomness extraction [25,26], entanglement distribution [27], state transformation [28,29], state merging [30,31], entanglement dilution [32,33,34], and compression [35,36,37,38,39,40].

    Empirical coordination and its variations are widely studied in the classical information theory literature. Le Treust [17] considered joint source-channel empirical coordination. Le Treust and Bloch [41] further used empirical coordination as a unified perspective for masking, amplification, and parameter estimation at the receiver. Cuff and Zhao [42] studied empirical coordination using implicit communication, with information embedding applications, such as digital watermarking, steganography, cooperative communication, and strategic play in team games. Cervia et al. [43] devised a polar coding scheme for empirical coordination. Related models can also be found in [44,45].

    Quantum mechanics enables the calculation of probabilities for experimental outcomes, emphasizing statistical averages rather than detailed descriptions of individual events. For instance, the Heisenberg uncertainty principle states that the standard deviations of position and momentum cannot be minimized simultaneously [46]. Some scholars, such as Fuchs and Peres [47], contend that quantum theory does not describe physical reality at all but is instead confined to represent statistical correlations [48]. Empirical coordination is thus a natural framework for quantum systems.

    Empirical coordination also plays a role in quantum data compression [49]. Barnum et al. [50] addressed a source of commuting density operators, and Kramer and Savari [36] developed a rate-distortion theory that unifies the visible and blind approaches (cf. [51,52]). Khanian and Winter have recently solved the general problem of a quantum source of mixed states (see also [52,53,54,55,56,57,58]).

    Coordination of separable correlations with classical links is described as follows. Consider a network of K nodes, where Node k performs an encoding operation Ek on a system Ak, for k{1,,K}. Some of the nodes are connected by one-way classical links. We denote the rate limit for the link from Node k to Node l by Rk,l. Before the coordination protocol begins, the nodes may also share common randomness (CR). Furthermore, some of the nodes can have access to side information. The objective in the coordination problem is to establish a specific correlation, i.e., to simulate a desired quantum state ωA1AK. Since the links are classical, the correlation is separable. The optimal performance is defined by the communication rates that are necessary and sufficient for simulating the desired correlation on average.

    In analogy to the classical framework, we distinguish between two types of coordination tasks: strong coordination and empirical coordination. In strong coordination, the users encode in order to simulate an n-fold product state, ωnA1AK. That is, the joint state resembles that of a memoryless quantum source. In our previous work, we have considered strong coordination [59,60,61]. In particular, we addressed strong coordination for entanglement generation using quantum links [59] and for classical-quantum (c-q) correlations with classical links [60]. Strong coordination can be viewed as a unified framework for various models. We list a few examples of related protocols:

    1) Channel resolvability: Resolvability aims to approximate the output of a c-q channel using a uniformly distributed codebook [62]. This is equivalent to c-q state simulation [62]. Resolvability is also referred to as c-q soft covering [63]. Quantum soft covering is further studied in [64].

    2) Entanglement dilution and distillation: In the dilution task, Alice and Bob use a maximally entangled state as well as local operations and classical communication (LOCC) in order to prepare a joint state [32,33]. In the other direction, maximal entanglement can be distilled from a bipartite state ωAB using classical communication at a rate R1,2H(A|B)ω (see [65]). A similar rate also appears in the distillation of a secret key [65, Remark 2]. Further work can be found in [65,66,67,68,69,70,71,72,73,74].

    3) State merging and splitting: In state merging, Alice and Bob share ωAB, and Alice would like to send her part to Bob [31,75]. The mother protocol generalizes this task [76,77]. Whereas, state splitting is the reverse task, where Alice holds AB, and would like to send B to Bob [78,79,80].

    4) Channel simulation: A classical channel of capacity C can be simulated at a rate of R1,2 if and only if R1,2C, given sufficient common randomness [81,82]. The quantum analog is not necessarily true [67]. The entanglement cost with LOCC is related to the entanglement of formation [20].

    Multi-user versions of the protocols above have been studied extensively in recent years. The mother protocol can generate distributed compression protocols for correlated quantum sources [76,83,84,85,86,87,88]. Simulation of broadcast and multiple-access channels is considered in [89,90] and [91], respectively. George and Cheng [92] have recently studied multipartite state splitting. Multi-user distillation and manipulation were considered in [93,94,95,96,97,98,99]. Streltsov et al. [100] studied multipartite state merging. A more detailed overview is given in [61].

    Here, we introduce the notion of empirical coordination for separable correlations, imposing a weaker and less stringent condition compared to strong coordination. We require the empirical average state to converge into the desired state ωA1AK. Specifically, let A(1),,A(n) denote the output sequence from all network nodes, where A(i)(A1(i),,AK(i)) is the output, at time i, for i{1,,n}. Then, we would like the nodes to produce an empirical average state 1nni=1ρA(i) that is arbitrarily close to ωA, where A(A1,,AK). That is, we require that the distance,

    1nni=1ρA(i)ωA1 (1.1)

    converges to zero as the block length n tends to infinity. In this work, we focus on the quantum empirical coordination of separable correlations. Our networks consist of nodes possessing quantum systems, and are connected with classical links of limited communication rates.

    After introducing the definition of empirical coordination for quantum states, we discuss the justification for our definition and its physical interpretation. We focus on the 3-node cascade network and determine the optimal coordination rates, which represent the minimal resources for the empirical simulation of a separable state among multiple parties. The cascade network with K=3 users can be viewed as a building block for larger multiuser systems [101,102]. The cascade setting is depicted in Figure 1. Alice, Bob, and Charlie wish to simulate a separable state ωABC. They are provided with rate-limited communication links, R12 from Alice to Bob, and R23 from Bob to Charlie.

    Figure 1.  Cascade network.

    Our results are summarized below. We show that CR between the network users does not affect the optimal rates for empirical coordination. We begin with the rate characterization for the basic two-node network, and then generalize to a cascade network. The special case of a network with an isolated node is considered as well. The results can be further generalized to other networks as our analysis includes a generic achievability scheme. The characterization involves optimization over a collection of state extensions. This is a unique feature of the quantum setting, as the classical parallel does not include optimization [16]. As will be seen in the examples, the performance depends heavily on the choice of decomposition. We further discuss the consequences of our results for cooperative games.

    In Section 2, we set our notation conventions. In Section 3, we present the definitions for our model and their physical interpretation. Section 4 is dedicated to the results, including the statement about CR and the capacity theorems for the two-node, cascade, and isolated node networks. Section 5 and Section 6 provide the achievability and converse analysis, respectively. Section 7 concludes with a discussion about the comparison between strong coordination and empirical coordination, as well as the implications of our results on quantum cooperative games.

    We use standard notation in quantum information theory, as in [103], X,Y,Z, are discrete random variables on finite alphabets X,Y,Z,..., respectively, The distribution of X is specified by a probability mass function (PMF) pX(x) on X. The set of all PMFs over X is denoted by P(X). The normalized total variation distance between two PMFs in P(X) is defined as

    12pXqX1=12aX|pX(a)qX(a)| (2.1)

    for every pX, qX P(X).

    The classical Shannon entropy is then defined as H(pX)=xsupp(pX)pX(x)log(1pX(x)), with logarithm to base 2. We often use the short notation H(X)H(pX) for XpX. Similarly, given a joint PMF pXYP(X×Y), we write H(XY)H(pXY). The mutual information between X and Y is I(X;Y)=H(X)+H(Y)H(XY). A classical channel is defined by a probability kernel {pY|X(y|x):xX,yY}. The conditional entropy with respect to pX×pY|X is defined as H(Y|X)=xsupp(pX)pX(x)H(Y|X=x), where H(Y|X=x)H(pY|X(|x)). According to the entropy chain rule, H(Y|X)=H(XY)H(X).

    We use xn=(xi:i[n]) for a sequence of letters from X, where [n]{1,,n}. We define the type of a sequence xn as the empirical distribution, ˆPxn(a)=1nN(a|xn), where N(a|xn) is the number of occurrences of the letter a in the sequence xn, for aX. The δ-typical set for a PMF pX is defined here as the set of sequences whose type is δ-close to pX in total variation distance. Formally,

    T(n)δ(pX){xnXn:12ˆPxnpX1<δ}. (2.2)

    A quantum system is associated with a Hilbert space, H. The dimensions are assumed to be finite throughout. Denote the set of all linear operators F:HH by L(H). The Hermitian conjugate of F is denoted by F. The extension of a real-valued function to Hermitian operators is defined in the usual manner. Analogously to the total variation distance between classical PMFs, the normalized trace distance between two Hermitian operators satisfies

    12PQ1=12Tr[|PQ|] (2.3)

    for every Hermitian P, QL(H).

    Let System A be associated with HA. The quantum state of A is described by a density operator ρAL(HA), i.e., a unit-trace positive semidefinite operator. Let Δ(HA) denote the set of all such operators. The probability distribution of a measurement outcome is derived from a positive operator-valued measure (POVM). In finite dimensions, this reduces to a finite set of positive semidefinite operators {Dj:j[N]} that satisfy Nj=1Dj=1, where 1 denotes the identity operator. By the Born rule, the probability of a measurement outcome j is given by pJ(j)=Tr(DjρA), for j[N].

    The von Neumann entropy of a quantum state ρAΔ(HA) is defined as H(ρA)Tr[ρAlog(ρA)]. We often denote the quantum entropy by H(A)ρH(ρA). Similarly, given a joint state ρABΔ(HAHB), we write H(AB)ρH(ρAB). A pure state has zero entropy, in which case, there exists |ψHA such that ρ=|ψ×ψ|, where ψ|(|ψ). The conditional quantum entropy is defined by H(A|B)ρ=H(AB)ρH(B)ρ. The conditional mutual information is defined accordingly, as I(A;B|C)ρH(A|C)ρ+H(B|C)ρH(A,B|C)ρ for ρABCΔ(HAHBHC).

    A bipartite state ρAB is said to be separable if a set of product states {ρxσx} in Δ(HAHB) can be found such that

    ρAB=xXpX(x)ρxσx (2.4)

    for some alphabet X and PMF pX on X. Otherwise, ρAB is called entangled. If the state is entangled, then the conditional entropy H(A|B)ρ can be negative. The definition can also be extended to a multipartite system. A state ρA1AK in Δ(HA1HAK) is said to be separable if

    ρA1AK=xXpX(x)ρ(1)xρ(K)x (2.5)

    for some ensemble {pX,ρ(1)xρ(K)x,xX}.

    A quantum channel is defined by a completely positive trace-preserving map, NAB:L(HA)L(HB). In general, the channel maps a state ρΔ(HA) into a state NAB(ρ)Δ(HB). A classical-quantum (c-q) channel NXB is specified by a collection of quantum states {ρ(x)B:xX} in Δ(HB), where ρ(x)BNXB(x) for xX.

    In this subsection, we introduce the basic definitions for empirical coordination. Consider the cascade network shown in Figure 2, which involves three users, Alice, Bob, and Charlie. Let {pX(x),ωxABC,xX} be a given ensemble, with an average

    ωABC=xXpX(x)ωxABC. (3.1)

    Suppose that Alice receives a random sequence Xn, drawn from a memoryless (i.i.d) source pX. This can be viewed as side information that Alice obtains from a local measurement on her environment. Alice sends a classical message m12 to Bob via a noiseless link of limited rate R12, and Bob sends m23 to Charlie at a limited rate R23. Next, Alice, Bob, and Charlie encode their respective quantum outputs An,Bn, and Cn. The objective of the empirical coordination protocol is for the average state to be arbitrarily close to a particular state ωABC.

    Figure 2.  Cascade network with common randomness.

    Remark 1. Achieving empirical coordination allows the network users to perform local measurements such that the outcome statistics follow a desired behavior.

    In other words, the users utilize a coding scheme that simulates, on average, a desired state ωABC. We are interested in the lowest communication rates (R12,R23) that are required in order to achieve this goal.

    In the beginning, we assume that Alice, Bob, and Charlie share unlimited common randomness (CR). That is, a random element S is drawn a priori and distributed to Alice, Bob, and Charlie before the protocol begins. Later, we will show that CR does not affect the achievable rates.

    Definition 1. A (2nR12,2nR23,n) empirical coordination code for the cascade network shown in Figure 2 consists of:

    a CR source pS over a randomization set Sn,

    a pair of classical encoding channels, pM12|XnS and pM23|M12S, over the index sets [2nR12] and [2nR23], respectively, and

    c-q encoding channels,

    EXSA:X×SnΔ(HA), (3.2)
    FM12M23SBn:[2nR12]×[2nR23]×SnΔ(HnB), (3.3)

    and

    DM23SCn:[2nR23]×SnΔ(HCn), (3.4)

    for Alice, Bob, and Charlie, respectively, where Sn is an unbounded set of realizations for the CR resource that is shared between the users a priori.

    The protocol works as follows: Before communication begins, Alice, Bob, and Charlie share a CR element s, drawn from the source pS. Alice receives a sequence xn, generated from a memoryless source pX. That is, the random sequence is distributed according to pnX(xn)ni=1pX(xi). She selects an index

    m12pM12|XnS(|xn,s) (3.5)

    at random, and sends it through a noiseless classical link at rate R12. She then applies the encoding channel EnXSA, to prepare the state of her system An, hence

    ρ(xn,s)An=ni=1EXSA(xi,s). (3.6)

    As Bob receives the message m12 and the CR element s, he selects a random index

    m23pM23|M12S(|m12,s) (3.7)

    and sends it through a noiseless classical link at rate R23 to Charlie. Bob and Charlie encode their systems, Bn and Cn, by

    ρ(m12,m23,s)Bn=FM12M23SBn(m12,m23,s), (3.8)

    and

    ρ(m23,s)Cn=DM23SCn(m23,s) (3.9)

    respectively.

    Given a value s, i.e., a realization of the random element, consider the average state ¯ρABC(s)Δ(HAHBHC) that is induced by the code:

    ¯ρABC(s)1nni=1xnXnm12[2nR12]m23[2nR23]pnX(xn)pM12|XnS(m12|xn,s)pM23|M12S(m23|m12,s)ρ(xi,s)Aiρ(m12,m23,s)Biρ(m23,s)Ci. (3.10)

    We now define achievable rates as rates that are sufficient to encode ¯ρABC(s) that converges to ωABC.

    Definition 2. A rate pair (R12,R23) is achievable for the empirical coordination of a desired separable state ωABC, if for every α,ε,δ>0 and a sufficiently large n, there exists a (2n(R12+α),2n(R23+α),n) coordination code that achieves

    Pr(12¯ρABC(S)ωABC1>ε)δ, (3.11)

    where the probability is computed with respect to the CR element SpS.

    Equivalently, there exists a sequence of empirical coordination codes such that the error converges to zero in probability, i.e.,

    ¯ρABC(S)ωABC10inprobability. (3.12)

    Remark 2. Since the communication links are classical, entanglement cannot be generated. Therefore, we only consider separable states ωABC.

    In this subsection, we discuss the justification and the physical interpretation of our coordination criterion. As mentioned in the Introduction, quantum mechanics enables the calculation of probabilities for experimental outcomes, emphasizing statistical averages rather than detailed descriptions of individual events. For instance, the Heisenberg uncertainty principle states that the standard deviations of position and momentum cannot be minimized simultaneously [46]. Some scholars, such as Fuchs and Peres [47], contend that quantum theory does not describe physical reality at all but is instead confined to represent statistical correlations [48]. Empirical coordination is thus a natural framework for quantum systems. Further justification is provided below.

    Consider an observable represented by an Hermitian operator ˆO on HAHBHC. In practice, statistics are collected by performing measurements on n systems (Ai,Bi,Ci:i[n]). The expected value of the observable in the ith measurement is thus,

    ˆOi=Tr[ˆOρAiBiCi] (3.13)

    for i[n]. Therefore, the empirical average is

    1nni=1ˆOi=Tr[ˆO(1nni=1ρAiBiCi)]=Tr[ˆO¯ρABC]. (3.14)

    Similarly, consider a POVM {D:[L]} on HAHBHC. The probability that we obtain the measurement outcome in the ith measurement is pi()=Tr(DρAiBiCi). Thereby, the average distribution is given by

    ˉp()=Tr(D¯ρABC). (3.15)

    Theorem 1. Any desired state ωABC that can be simulated at rate (R12,R23) through empirical coordination in the cascade network with CR assistance can also be simulated with no CR, i.e., with |Sn|=1.

    We will discuss the interpretation of this result in Subsection 7.4. The proof is provided below.

    Proof. Let (R12,R23) be an achievable rate pair for empirical coordination. Consider the setting in Section 3. Let the CR element S and the classical side information Xn be drawn according to pS and pnX, respectively. Then, Alice, Bob, and Charlie encode by

    m12pM12|XnS(|xn,s),ρ(xn,s)An=ni=1EXSA(xi,s), (4.1)
    m23pM23|M12S(|m12,s),ρ(m12,m23,s)Bn=FM12M23SBn(m12,m23,s), (4.2)
    ρ(m23,s)Cn=DM23SCn(m23,s). (4.3)

    Denote the normalized trace distance by

    d(s)=12¯ρABC(s)ωABC1 (4.4)

    for sSn.

    According to Definition 2, if a rate pair (R12,R23) is achievable, then for every α,ε,δ>0 and sufficiently large n, there exists a sequence of (2n(R12+α),2n(R23+α),n) empirical coordination codes, for which the following holds:

    Pr(d(S)>ε2)δ. (4.5)

    Averaging over the CR element yields the following average state

    ˆρAnBnCn=E[ρ(m12,S)Anρ(m12,S)Bnρ(m23,S)Cn]=sSnpS(s)ρ(m12,s)Anρ(m12,s)Bnρ(m23,s)Cn. (4.6)

    By the total expectation formula,

    E[d(S)]=Pr(d(S)>ε2)E[d(S)|d(S)>ε2]+Pr(d(S)ε2)E[d(S)|d(S)ε2]δ1+1ε2<ε, (4.7)

    where the second line follows from (4.5), and the last inequality holds by choosing δ<ε2. Therefore, there exists sSn for which d(s)ε. We can thus satisfy the coordination requirement with the following encoding maps,

    m12pM12|XnS(|xn,s),ρ(xn)An=ni=1EXSA(xi,s), (4.8)
    m23pM23|M12S(|m12,s),ρ(m12,m23)Bn=FM12M23SBn(m12,m23,s), (4.9)
    ρ(m23)Cn=DM23SCn(m23,s), (4.10)

    which no longer require CR.

    Next, we characterize the achievable rates for empirical coordination. We begin with a basic two-node network, and then generalize to a cascade network. Based on Theorem 1 above, introducing CR does not affect the achievable rates. Therefore, we will focus our definitions on empirical coordination without CR.

    Consider the two-node network. See Figure 3. Alice and Bob would like to simulate a separable state ωAB on average using the following coding scheme. Alice receives classical side information from a memoryless source pX. She encodes An, and then sends an index m12, i.e., a classical message to Bob, at a rate R12.

    Figure 3.  Two-node network.

    Formally, a (2nR12,n) empirical coordination code for a separable state ωAB consists of an input distribution pM12|Xn over an index set [2nR12], and two c-q encoding channels EXA, and FM12Bn. The protocol works as follows. Alice receives xn, drawn according to pnX. She selects a random index

    m12pM12|Xn(|xn), (4.11)

    and sends it through a noiseless link. Furthermore, she encodes An by

    ρ(xn)An=ni=1EXA(xi). (4.12)

    As Bob receives the message m12, he prepares the state

    ρ(m12)Bn=DM12Bn(m12). (4.13)

    Hence, the resulting average (joint) state is

    ¯ρAB=1nni=1xnXnm12[2nR12]pnX(xn)pM12|Xn(m12|xn)ρ(xn)Aiρ(m12)Bi. (4.14)

    Definition 3. A rate R120 is achievable for the empirical coordination of ωAB if for every ε,α>0 and sufficiently large n, there exists a (2n(R12+α),n) code that achieves

    ¯ρABωAB1ε. (4.15)

    Definition 4. The empirical coordination capacity for the simulation of a separable state ωAB over the two-node network is defined as the infimum of achievable rates. We denote the capacity by C2-node(ω).

    The optimal rate for empirical coordination is established below. Consider the extended c-q state,

    ωXAB=xXpX(x)|x×x|XωxAB. (4.16)

    Here, X plays the role of a classical register. Furthermore, let S2-node(ω) be the set of all c-q extensions

    σXYAB=(x,y)X×YpXY(x,y)|x×x||y×y|σxAσyB (4.17a)

    such that

    σXAB=ωXAB. (4.17b)

    Notice that given a classical pair (X,Y)=(x,y), there is no correlation between A and B. We also note that if ωAB is entangled, then S2-node(ω) is an empty set.

    Theorem 2. Let ωAB be a bipartite state in Δ(HAHB). If the set S2-node(ω) is nonempty, then the empirical coordination capacity for the two-node network in Figure 3 is given by

    C2-node(ω)=infσS2-node(ω)I(X;Y)σ. (4.18)

    Otherwise, if S2-node(ω)=, then coordination is impossible.

    The achievability proof for Theorem 2 is given in Subsection 5.2, and the converse in Subsection 6.1.

    Remark 3. The set S2-node(ω) is empty if and only if ωAB is entangled. As mentioned in Remark 2, classical links cannot generate entanglement, hence, coordination is impossible in this case.

    Remark 4. The characterization involves optimization over a collection of separable states, S2-node(ω). This is a unique feature of the quantum setting. In the classical setting, there is no optimization. As will be seen in Examples 1 and 2, the performance depends heavily on the chosen decomposition.

    Remark 5. In the special case of orthonormal sets, {|σxA} and {|σyB}, the coordination capacity satisfies C2-node(ω)=I(A;B)ω. This case is essentially classical.

    Remark 6. One may always find a decomposition of a separable state into a combination of pure states. In particular, consider

    ωAB=(x,y)X×YpXY(x,y)σxAσyB. (4.19)

    By inserting spectral decompositions,

    σxA=v1V1pV1|X(v1|x)|ψx,v1A×ψx,v1A|,σyB=v2V2pV2|Y(v2|y)|ϕy,v2B×ϕy,v2B|, (4.20)

    we obtain

    ωAB=w1,w2pW1W2(w1,w2)|ψw1A×ψw1A||ϕw2B×ϕw2B|, (4.21)

    where W1(X,V1) and W2(Y,V2). If one uses this pure-state decomposition, then the coordination rate would be R12>I(W1;W2)σ. Nevertheless, the theorem shows that this can be suboptimal, since I(XV1;YV2)σI(X;Y)σ.

    Remark 7. Based on our previous result [60], strong coordination can be achieved at the same rate if Alice and Bob share sufficient CR before communication begins. Here, however, we assume that CR is not available to Alice and Bob. Yet, they can perform the coordination task at this rate, since the requirement of empirical coordination is less strict.

    Example 1. Let A and B be a qubit pair, i.e., dim(HA)=dim(HB)=2. Consider the state

    ωAB=12|0×0||0×0|+14|1×1||0×0|+14|1×1||+×+|, (4.22)

    where {|0,|1} and {|+,|} are the computational basis and conjugate basis, respectively. Such decomposition can be associated with a joint distribution pXY, where Y=X with probability 1, an alphabet of size |X|=3, and

    pX=(12,14,14). (4.23)

    Based on Theorem 2, we can achieve the rate R12=I(X;Y)σ=1.5. The coordination rate can be significantly improved by using the decomposition below instead.

    ωAB=12|0×0||0×0|+12|1×1|η, (4.24)

    where η is the BB84 state,

    η=12|0×0|+12|+×+|. (4.25)

    This yields the improved rate of R12=I(X;Y)σ=0.3112.

    Example 2. Consider the following qubit state,

    ωAB=12|0×0|[(1p)|+×+|+p|×|]+12|1×1|[p|+×+|+(1p)|×|] (4.26)

    where the second qubit can be viewed as the output of a phase-flip channel, p(0,1). In this case, we obtain

    I(X;Y)σ=1h(p) (4.27)

    where h(x)=(1x)log(1x)xlog(x) is the binary entropy function on (0,1). For p=12, we have a product state ωAB=1212. Hence, communication is not necessary and the coordination capacity is C2-node(ω)=0.

    Consider the cascade network (see Figure 4).

    Figure 4.  Cascade network without common randomness.

    Alice, Bob, and Charlie wish to simulate a separable state ωABC using the following scheme. Alice receives classical side information from a memoryless source pX. She encodes An, and she sends an index m12, i.e., a classical message to Bob, at a rate R12. Then Bob uses the message m12 to encode his systems Bn, and sends a message m23 to Charlie who uses it to encode his systems Cn.

    Formally, a (2nR12,2nR23,n) empirical coordination code for the simulation of a separable state ωABC in the cascade network consists of two input distributions pM12|Xn and pM23|XnM12 over index sets [2nR12] and [2nR23], and three c-q encoding channels EXA, FM12Bn, and DM23Cn. The protocol works as follows:

    Alice selects a random index,

    m12pM12 (4.28)

    and sends it through a noiseless link. Furthermore, she encodes An by

    ρ(xn)An=ni=1EXA(xi). (4.29)

    As Bob receives the message m12, he generates m23 according to pM23|XnM12(|xn,m12), sends m23 to Charlie, and prepares the state

    ρ(m12)Bn=FM12Bn(m12). (4.30)

    Having received the classical message m23, Charlie applies his c-q encoding map and prepares

    ρ(m23)Cn=DM23Cn(m23). (4.31)

    Hence, the resulting average (joint) state is

    ¯ρABC=xnXnpnX(xn)m12[2nR12]m23[2nR23]pM12|Xn(m12|xn)pM23|M12Xn(m23|m12,xn)1nni=1ρ(xn)Aiρ(m12)Biρ(m23)Ci. (4.32)

    Definition 5. A rate pair (R12,R23) is achievable for the empirical coordination of ωABC if for every ε,δ>0 and a sufficiently large n, there exists a (2n(R12+δ),2n(R23+δ),n) code that achieves

    ¯ρABCωABC1ε. (4.33)

    Definition 6. The empirical coordination capacity region for the simulation of a separable state ωABC over the cascade network is defined as the closure of all the achievable rate pairs (R1,2,R2,3).

    We denote the capacity region by CCascade(ω).

    The main result for the cascade network is established below. Consider the extended c-q state,

    ωXABC=xXpX(x)|x×x|XωxABC. (4.34)

    Furthermore, let SCascade(ω) be the set of all c-q extensions

    σXYZABC=(x,y,z)X×Y×ZpXYZ(x,y,z)|x×x||y×y||z×z|σxAσyBσzC (4.35a)

    such that

    σXABC=ωXABC. (4.35b)

    As before, coordination with classical links is limited to separable states (see Remarks 2 and 3).

    Theorem 3. Let ωABC be a tripartite state in Δ(HAHBHC). If the set SCascade(ω) is nonempty, then the empirical coordination capacity region for the cascade network in Figure 4 is

    CCascade(ω)=SCascade(ω){(R12,R23)SCascade(ω):R12I(X;YZ)σ,R23I(X;Z)σ}. (4.36)

    Otherwise, if SCascade(ω)=, then coordination is impossible.

    The achievability proof for Theorem 3 is provided in Subsection 5.3, and the converse part is provided in Subsection 6.2. We note that based on the Caratheodory's [104], we may limit the union to auxiliary variables of cardinality |Y||X|+|X|2dim(HA)2dim(HB)2dim(HC)21 and |Z||X|+|X|2dim(HA)2dim(HB)2dim(HC)2 (see also [105, App. B]).

    Remark 8. The cascade model has a Markov structure in the sense that given the message m23 from Bob, Charlie's state ρm23Cn has no correlation with Alice. Nevertheless, the correlation that Alice, Bob, and Charlie simulate does not satisfy a Markov chain property. In particular, the auxiliary random variables X, Y, and Z may follow a general Bayesian rule, and do not necessarily form a Markov chain.

    Consider the isolated node network in Figure 5. This is a special case of a cascade network with R23=0. The coordination capacity CIsolated(ω) is defined similarly as in Definition 4, and can be established as a consequence of Theorem 3. Consider the extended c-q state,

    ωXABC=xXpX(x)|x×x|XωxABC. (4.37)
    Figure 5.  Isolated node network.

    Let SIsolated(ω) be the set of all c-q extensions σXYZABC of the form

    σXYZABC=(x,y,z)X×Y×ZpXYZ(x,y,z)|x×x||y×y||z×z|σxAσyBσzC (4.38a)

    such that

    σXABC=ωXABC (4.38b)

    and

    σAC=σAσC. (4.38c)

    Corollary 4. Let ωABC be as in Theorem 3. If the set SIsolated(ω) is nonempty, then the empirical coordination capacity for the isolated node network in Figure 5 is given by

    CIsolated(ω)=infσSIsolated(ω)I(X;Y|Z)σ. (4.39)

    Otherwise, if SIsolated(ω)=, then coordination is impossible.

    In this case, coordination is only possible for a separable state ωABC such that ωAC=ωAωC.

    Remark 9. Notice that B and C can still be correlated, see Example 3. Given unlimited CR, it is clear that we may generate such a correlation. Even in the extreme case of no communication, we can generate Yn from a memoryless source, treat Yn as the CR element, and let Zn=Yn (see discussion in [16, Sec. III-B]). We have seen that CR does not affect the coordination capacity, and thus, the same rates can be achieved without CR. Further intuition is given in the discussion in Subsection 7.4.

    In the following example, we consider empirical coordination in the isolated node network with a tripartite state ωABC, in which B and C are correlated.

    Example 3. Consider the following qubit state,

    ωABC=(1α)|0×0|[(1p)|+×+||+×+|+p|×||×|]+α|1×1|[(1p)|+×+||×|+p|×||+×+|] (4.40)

    with α,p(0,1). In this case, I(X;Y|Z)σ=H(X)=h(α).

    To show the direct part of our coordination capacity theorems, we will use the generic lemma below. Consider the generic two-node network in Figure 6, where Alice receives xn and yn as input to her encoder and encodes a quantum system An. Whereas, Bob receives yn and zn as input and encodes a quantum system Bn. In this case, Alice has encoding maps of the form pM12|XnYnS and EXnYnSAn, and Bob encodes by FM12YnZnSBn. The resulting average state is

    ¯ρAB(xn,yn,zn,s)=1nni=1m12[2nR12]pM12|S(m12|s)ρ(m12,xn,yn,s)Aiρ(m12,yn,zn,s)Bi, (5.1)

    where ρ(xn,yn,s)An=EXnYnSAn(xn,yn,s) and ρ(m12,yn,zn,s)Bn=FM12YnZnSBn(m12,yn,zn,s).

    Figure 6.  Generic two-node network.

    Lemma 5. Consider a state ensemble, {pXYZpU|XY,σx,yAσy,z,uB}. Let

    ηx,y,zB=uUpU|XY(u|x,y)σy,z,uB. (5.2)

    For every δ>0, if

    R12>I(X;U|YZ)σ, (5.3)

    then there exists a sequence of randomized (2nR12,n) empirical coordination codes such that

    limnPr(¯ρAB(xn,yn,zn,S)1nni=1σxi,yiAηxi,yi,ziB1>γ(δ))=0, (5.4)

    uniformly for all (xn,yn,zn)T(n)δ(pXYZ), where the probability is computed with respect to the CR element S, and γ(δ) tends to zero as δ0.

    The proof for Lemma 5 is provided below. Consider the extended c-q state,

    σXYZUAB=(x,y,z,u)X×Y×Z×UpXYZ(x,y,z)pU|XY(u|x,y)|x,y,z,u×x,y,z,u|σx,yAσy,z,uB (5.5)

    where X, Y, Z, and U are classical registers. We note that Z(X,Y)U forms a Markov chain.

    By Theorem 1, we may assume that Alice and Bob share unlimited CR. Therefore, they can generate the codebook jointly using their random element.

    Classical codebook construction Select 2nR0 sequences un(), [2nR0], independently at random, each i.i.d. according to pU, where

    pU(u)=x,y,zpXYZ(x,y,z)pU|XY(u|x,y). (5.6)

    Assign each sequence with a bin index b(un()), where b:Un[2nR12], independently at random. We thus identify the CR element S as the random codebook {un(),b()}.

    Encoding First, consider the classical encoding function M12:Xn×Yn[2nR12]. Given a pair (xn,yn)Xn×Yn, find an index [2nR0] such that (xn,yn,un())T(n)2δ(pXYU). If there is none, set =1. If there is more than one, choose the smallest. Send the corresponding bin index, i.e., m12(xn,yn)=b(un()).

    Then, prepare

    ρxn,ynAnni=1σxi,yiA. (5.7)

    Decoding Given (yn,zn) and m12, find an index ˆ[2nR0] such that

    (yn,zn,un(ˆ))T(n)8δ(pYZU) and b(un(ˆ))=m12. (5.8)

    If there is none, set ˆ=1. If there is more than one, choose the smallest. Prepare the state

    ρyn,zn,un(ˆ)Bnni=1σyi,zi,ui(ˆ)B. (5.9)

    This results in an average state,

    ¯ρAB(un,xn,yn,zn)=1nni=1ρxn,ynAiρyn,zn,unBi=1nni=1σxi,yiAσyi,zi,uiBn, (5.10)

    with unun(ˆ).

    Error analysis Given Un(ˆ)=un, we have

    ¯ρAB(un,xn,yn,zn)=1n(a,b,c,d)X×Y×Z×Ui:(xi,yi,zi,ui)=(a,b,c,d)σxi,yiAσyi,zi,uiB=1n(a,b,c,d)X×Y×Z×Ui:(xi,yi,zi,ui)=(a,b,c,d)σa,bAσb,c,dB=(a,b,c,d)X×Y×Z×UˆPxn,yn,zn,un(a,b,c,d)σa,bAσb,c,dB=(a,b,c)X×Y×ZˆPxn,yn,zn(a,b,c)dUˆPun|xn,yn,zn(d|a,b,c)σa,bAσb,c,dB. (5.11)

    For every un such that (xn,yn,zn,un)T(n)γ(δ)(pXYZU),

    ¯ρAB(un,xn,yn,zn)τAB1γ(δ) (5.12)

    where

    τAB=(a,b,c)X×Y×ZˆPxn,yn,zn(a,b,c)dUpU|XYZ(d|a,b,c)σa,bAσb,c,dB=(a,b,c)X×Y×ZˆPxn,yn,zn(a,b,c)σa,bAdUpU|XY(d|a,b)σb,c,dB=(a,b,c)X×Y×ZˆPxn,yn,zn(a,b,c)σa,bAηa,b,cB=1n(a,b,c)X×Y×ZN(a,b,c|xn,yn,zn)σa,bAηa,b,cB=1n(a,b,c)X×Y×Zi:(xi,yi,zi)=(a,b,c)σa,bAηa,b,cB=1n(a,b,c)X×Y×Zi:(xi,yi,zi)=(a,b,c)σxi,yiAηxi,yi,ziB=1nni=1σxi,yiAηxi,yi,ziB. (5.13)

    Consider the event

    A1{(xn,yn,zn,Un(ˆj))T(n)γ(δ)(pXYZU)}. (5.14)

    Based on the classical result due to Cuff et al. [16],

    Pr(A1)1αn (5.15)

    for all (xn,yn,zn)T(n)δ(pXYZ), where γγ(δ) tends to zero as δ0, and αn tends to zero as n, provided that

    R>I(X;U|YZ)+γ(δ). (5.16)

    Therefore,

    Pr(¯ρAB(Un(ˆ),xn,yn,zn)τAB1>γ(δ))αn. (5.17)

    This completes the proof of Lemma 5.

    We are now in a position to give the achievability proofs for the two-node and cascade networks.

    The proof essentially follows from Lemma 5, with the following addition. If Alice receives a random sequence Xn that is not δ-typical, then she sends an arbitrary transmission. Otherwise, she encodes using the encoder in Lemma 5. Since Pr(XnT(n)δ(pX)) tends to 1 as n, achievability for the two-node network follows.

    We use rate splitting, where Alice's message consists of two components m12 and m12, at rates R12 and R12, respectively, where R12=R12+R12.

    Classical codebook construction Select 22nR0 sequences yn(), zn(), ,[2nR0], independently at random, each i.i.d. according to pY and pZ, where

    pYZ(y,z)=xpX(x)pYZ|X(y,z|x). (5.18)

    Assign each sequence with a bin index b(yn()) and c(zn()), where b:Yn[2nR12] and c:Zn[2nR12], independently at random.

    Alice's encoder As before, if Alice receives xnT(n)δ(pX), she sends an arbitrary transmission. Otherwise, consider the classical encoding function M12:Xn[2nR12]×[2nR12] below. Given xnT(n)δ(pX), find an index pair (,)[2nR0]×[2nR0] such that (xn,yn(),zn())T(n)2δ(pXYZ). If there is none, set (,)=(1,1). If there is more than one, choose the first. Send the corresponding bin indices, i.e., m12(xn)=b(yn()) and m12(xn)=c(zn()).

    Then, prepare

    ρxnAnni=1σxiA. (5.19)

    Bob's encoder Bob receives m12=(m12,m12), and encodes in three stages:

    (i) Given m12, find an index ˆ[2nR0] such that

    zn(ˆ)T(n)8δ(pZ) and c(zn(ˆ))=m12. (5.20)

    If there is none, set ˆ=1. If there is more than one, choose the smallest. Send m23=m12 to Charlie.

    (ii) Now given m12 and ˆ, find an index ˆ[2nR0] such that

    (yn(ˆ),zn(ˆ))T(n)8δ(pYZ) and b(yn(ˆ))=m12. (5.21)

    If there is none, set ˆ=1. If there is more than one, choose the smallest.

    (iii) Prepare the state

    ρyn(ˆ)BnˉZnni=1σyi(ˆ)B|zi(ˆ)×zi(ˆ)|ˉZ (5.22)

    where ˉZn is an auxiliary system for Bob.

    Charlie's encoder Given m23=m12, find an index ˜[2nR0] such that

    zn(˜)T(n)8δ(pZ) and c(zn(˜))=m12. (5.23)

    If there is none, set ˜=1. If there is more than one, choose the smallest.

    Prepare the state

    ρzn(˜)Cnni=1σzi(˜)C. (5.24)

    This results in an average state,

    ¯ρABˉZC(xn,yn,zn)=1nni=1ρxnAiρynBi|ˉzn×ˉzn|ρznCi=1nni=1σxiAσyiB|ˉzi×ˉzi|σziC, (5.25)

    with ynyn(ˆ), ˉznzn(ˆ), and znzn(˜). Based on the analysis in the proof of Lemma 5 (see Section 5.1), Alice, Bob, and Charlie achieve empirical coordination of σABZC, provided that

    R23=R12>I(X;Z), (5.26)
    R12>I(X;Y|Z) (5.27)

    which requires R12=R12+R12>I(X;YZ).

    We now show the converse part of the coordination capacity theorems.

    Consider the two-node network in Figure 3. Let R12 be an achievable rate for empirical coordination with a desired state ωAB. Then, there exists a sequence of (2nR12,n) empirical coordination codes that achieves an error,

    ¯ρXABωXAB1εn, (6.1)

    where εn tends to zero as n. Now, suppose that Bob performs a projective measurement in a particular basis, say, {|y}. This yields a sequence Yn as the measurement outcome, with some distribution pYn|Xn(yn|xn).

    Then, consider the classical variables XJ and YJ, where J is a uniformly distributed random variable, over the index set [n], drawn independently of Xn, Yn. Their joint distribution is

    ˉpXJYJ(x,y)=1nni=1pXiYi(x,y)=(x|y|)¯ρXB(|x|y), (6.2)

    where pXiYi is the marginal distribution of pnX×pYn|Xn. Based on (6.1), we have the following total variation bound:

    ˉpXJYJπXY1εn, (6.3)

    where πXY is defined as

    πXY(x,y)=(x|y|)ωAB(|x|y), (6.4)

    for (x,y)X×Y.

    Next, consider that

    nR12H(M12)I(Xn;M12)I(Xn;Yn)=ni=1I(Xi;Yn|Xi1)=ni=1I(Xi;Xi1Yn)ni=1I(Xi;Yi)=nI(XJ;YJ|J)ˉp (6.5)

    where the third inequality holds by the data processing inequality and the following equalities by the chain rule. Since Xn is i.i.d., it follows that XJ and J are statistically independent, hence,

    I(XJ;YJ|J)ˉp=I(XJ;YJJ)ˉpI(XJ;YJ)ˉp. (6.6)

    Based on entropy continuity [106],

    I(XJ;YJ)ˉpI(X;Y)παn (6.7)

    where αn=3εnlog(εn|X||Y|) [107, Lemm. 2.7], which tends to zero as n. This concludes the converse proof for the two-node network.

    Consider the cascade network in Figure 4. If (R12,R23) is achievable, then there exists a sequence of (2nR12,2nR23,n) codes such

    ¯ρXABCωXABC1εn, (6.8)

    where εn tends to zero as n. Suppose that Bob and Charlie perform projective measurements in a particular basis, say, {|y} and {|z}, respectively. This yields a sequence (Yn,Zn) as the measurement outcomes, with some distribution pYnZn|Xn(yn,zn|xn).

    Then, consider the classical variables XJ, YJ, and ZJ, where J is uniform over [n], independent of Xn, Yn, and Zn. Their joint distribution is

    ˉpXJYJZJ(x,y,z)=1nni=1pXiYiZi(x,y,z)=(x|y|z|)¯ρXBC(|x|y|z), (6.9)

    where pXiYiZi is the marginal distribution of pnX×pYnZn|Xn. By (6.8),

    ˉpXJYJZJπXYZ1εn, (6.10)

    where

    πXYZ(x,y,z)=(x|y|z|)ωABC(|x|y|z), (6.11)

    for (x,y,z)X×Y×Z.

    Consider Alice's communication rate, R12. Now, we may view the overall encoding operation of Bob and Charlie as a "black box" with M12 as input and (Bn,Cn) as output, as shown in Figure 7. Thus,

    nR12H(M12)I(Xn;M12)I(Xn;YnZn)=ni=1I(Xi;YnZn|Xi1)=ni=1I(Xi;Xi1YnZn)ni=1I(Xi;YiZi)=nI(XJ;YJZJ|J)ˉp (6.12)
    Figure 7.  Encoding by Bob and Charlie.

    based on the same arguments as in (6.5). Since XJ and J are statistically independent, we have

    R12I(XJ;YJZJ|J)ˉp=I(XJ;JYJZJ)ˉpI(XJ;YJZJ)ˉp. (6.13)

    Following similar steps, we also have

    R23I(XJ;ZJ)ˉp. (6.14)

    Based on entropy continuity [106],

    I(XJ;YJZJ)I(X;YZ)παn, (6.15)
    I(XJ;ZJ)I(X;Z)παn (6.16)

    where αn=3εnlog(εn|X||Y||Z|) [107, Lemma 2.7], which tends to zero as n.

    We have introduced the notion of empirical coordination for quantum correlations. Quantum mechanics enables the calculation of probabilities for experimental outcomes, emphasizing statistical averages rather than detailed descriptions of individual events. Empirical coordination is thus a natural framework for quantum systems. Focusing on the cascade network, we established the optimal coordination rates, indicating the minimal resources for the empirical simulation of a quantum state. As we consider a network with classical communication links, superposition cannot be maintained, hence the quantum correlations are separable. This precludes entanglement. We have shown that providing the users with shared randomness, before communication begins, does not affect the optimal rates for empirical coordination (see Theorem 1). We began with the rate characterization for the basic two-node network (Theorem 2), and then generalized to a cascade network (Theorem 3). The special case of a network with an isolated node was addressed as well (see Corollary 4). The results generalize to other networks as our analysis includes a generic achievability scheme (see Lemma 5). Nonetheless, we do not claim to have solved all coordination scenarios or network topologies.

    Next, we discuss the consequences of our results for quantum cooperative games.

    In many cooperative games, the payoff is associated with the correlation between the players. In the penny matching game, as introduced by Gossner et al. [108], Alice receives a classical sequence xn from an i.i.d source; thereafter, Alice and Bob produce sequences an and bn that should be close to one another and to xn as well. In other words, Alice and Bob try to guess the source sequence one bit at a time. They gain a point for every bit they both guess correctly. Alice's action an is referred to as a guess, even though she knows the original source sequence xn. As it turns out, an optimal strategy could let Alice guess wrong, i.e., aixi, for some of the time [108]. Cuff and Zhao [42] analyzed a generalized version of the game through the classical two-node network. Here, we introduce a quantum version of the game.

    Suppose that Alice receives a classical sequence xn from an i.i.d source pX, as depicted in the two-node network 3. The quantum encoding of each user is viewed as the actions [109]. The game is specified by a payoff map

    G:Δ(HAHB)[0,). (7.1)

    Given a joint strategy ωAB, the payoff to Alice and Bob is G(ωAB).

    Suppose that Alice uses an empirical coordination code and sends nR12 bits to Bob. Furthermore, let SEP(γ) be the set of all separable strategies ωAB for which Alice and Bob receive a payoff γ=G(ωAB). Alice and Bob can then reach an average payoff γ0 asymptotically, if and only if Alice can send a message to Bob at rate R12>C2-node(ω) for some ωABSEP(γ). The optimal rate C2-node(ω) is characterized by Theorem 2.

    Analogously to the classical framework, we distinguish between two types of coordination tasks: Strong coordination and empirical coordination.

    In the classical setting, strong coordination means that a statistician cannot reliably distinguish between the constructed sequence of actions Xn1,,XnK, and random samples from the desired distribution [16]. This requires the joint distribution pXn1XnK that the code induces to be arbitrarily close to the desired source ππX1XK in total variation distance. That is, strong coordination is achieved if there exists a code sequence such that

    limnpXn1XnKπn1=0, (7.2)

    where πn denotes the i.i.d. distribution corresponding to the desired source.

    Consider a network of K quantum nodes, where the users have access to classical communication links with limited rates Ri,j and may share common randomness (CR) at a limited rate R0. We say that strong coordination is achieved if there exists a code sequence such that the joint state ρAn1AnK that is the code induces converges to the desired state, i.e.,

    limnρAn1AnKωn1=0, (7.3)

    where ωωA1AK is the desired state. In our previous work [60], we have considered strong coordination for classical-quantum (c-q) correlations with classical links.

    In the classical description, empirical coordination uses network communication in order to construct a sequence of actions that have an empirical joint distribution closely matching the desired distribution [16]. In this case, the error criterion sets a weaker requirement, given in terms of the joint type, i.e., the empirical distribution of the actions in the network. Formally, the requirement for empirical coordination is that for every ε>0,

    limnPr(ˆPXn1XnKπ1ε)=0, (7.4)

    where Xn1,,XnK are the encoded actions, and the probability is computed with respect to the CR distribution.

    We say that empirical coordination is achieved in a quantum coordination network if there exists a sequence of coordination codes of length n, such that the time-average state 1nni=1ρA1(i)AK(i) that is induced by the code converges in probability to the desired source ωA1AK, i.e.,

    limnPr(1nni=1ρA1(i)AK(i)ω1ε)=0, (7.5)

    where ωωA1AK is the desired state, and the probability is computed with respect to the CR distribution. We note that the quantum definition differs in nature from the classical one (c.f. (7.4) and (7.5)).

    Remark 10. To see that strong coordination is indeed a stronger condition, note that by trace monotonicity, strong coordination implies ρA1(i)AK(i)ω10 as n, for every i[n]. Hence, by the triangle inequality,

    1nni=1ρA1(i)AK(i)ω11nni=1ρA1(i)AK(i)ω1 (7.6)

    which also tends to zero as n.

    We have discussed the justification and the physical interpretation of our coordination criterion in Subsection 3.2. Consider an observable represented by an Hermitian operator ˆO on HA1HAK. In practice, statistics are collected by performing measurements on n systems (A1(i),,AK(i):i[n]). The expected value of the observable in the ith measurement is thus,

    ˆOi=Tr[ˆOρA1,iAK,i] (7.7)

    for i[n]. Roughly speaking, our coordination criterion guarantees that the empirical average is close to the expected value with respect to a desired state, i.e.,

    1nni=1ˆOi=Tr[ˆO(1nni=1ρA1,iAK,i)]Tr[ˆOωA1AK], (7.8)

    with high probability.

    We have shown that CR does not improve the coordination capacity. That is, if Rj is achievable with CR, it is also achievable without CR. We provide an intuitive explanation below. Suppose we use a coding scheme where the CR element is a sequence Un, drawn from a memoryless source pU over U, and each user encodes by a collection of maps {E(u)}, taking u=Ui at time i. Then, this CR-assisted coding scheme can be replaced with a code based on a fixed agreed-upon sequence ˜un of type ˆP˜unpU.

    Since our coding scheme uses binning and not an encoder of the form E(ui), the description above is only a rough explanation to gain intuition.

    Recent advances in machine-to-machine communication [19] and the Internet of Things (IoT) [14] have raised interest in networks with various topologies [5]. These network topologies are relevant for various applications, such as distributed computing [110], autonomous vehicles [111], embedded sensors [112], players in a cooperative game [42], and quantum-enhanced IoT [114]. Coordination with classical links is motivated by quantum-enhanced IoT networks in which the communication links are classical [113,114,115,116]. The problem at hand is to find the optimal transmission rates required in order to establish a desired correlation. Empirical coordination also plays a role in quantum data compression [49,50,52]. The optimal compression rate for a quantum source of pure states was first established by Schumacher [117] for a quantum source of pure states (see also [118,119]). Empirical coordination is thus a natural framework for quantum systems.

    Empirical coordination also plays a role in quantum data compression [49]. Barnum et al. [50] addressed a source of commuting density operators, and Kramer and Savari [36] developed a rate-distortion theory that unifies the visible and blind approaches (cf. [51] and [52]). Khanian and Winter have recently solved the general problem of a quantum source of mixed states (see also [52,53,54,55,56,57,58]). Rate distortion can be viewed as a special case of empirical coordination.

    In another work by the authors [59], we have also considered strong coordination in a network with quantum links. This allows for the generation of multipartite entanglement and is closely related to tasks such as quantum channel/source simulation [21,22,23,24,89,120], state merging [30,31], state redistribution [77,121], zero-communication state transformation [28,29], entanglement dilution [32,33,34,98], randomness extraction [25,26], source coding [35,36,37,38,39,40], and many others. An interesting avenue for future research is to study empirical coordination in such networks. There are many other coordination scenarios and network topologies that could be studied further, e.g., empirical coordination with entanglement assistance. Other interesting directions include the one-shot setting (n=1) and coordination with two-way communication.

    The authors made equal contributions.

    The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.

    The authors would like to thank Ian George (National University of Singapore), Eric Chitambar (University of Illinois at Urbana-Champaign), and Marius Junge (University of Illinois at Urbana-Champaign) for useful discussions during the conference "Beyond IID in Information Theory, " held at the University of Illinois Urbana-Champaign from July 29 to August 2, 2024, supported by NSF Grant n. 2409823.

    H. Natur and U. Pereg were supported by the Israel Science Foundation (ISF), Grants n. 939/23 and 2691/23, German-Israeli Project Cooperation (DIP) within the Deutsche Forschungsgemeinschaft (DFG) under Grant n. 2032991, Ollendorff Minerva Center (OMC) of the Technion n. 86160946, and Nevet Program of the Helen Diller Quantum Center at the Technion n. 2033613. U. Pereg was also supported by the Junior Faculty Program for Quantum Science and Technology of Israel Planning and Budgeting Committee of the Council for Higher Education (VATAT) under Grant 86636903, and the Chaya Career Advancement Chair through Grant n. 8776026.

    The authors declare no conflict of interest.



    [1] J. Soni, R. Goodman, A mind at play: how Claude Shannon invented the information age, Simon and Schuster, 2017.
    [2] D. J. Costello, G. D. Forney, Channel coding: The road to channel capacity, P. IEEE, 95 (2007), 1150–1177. https://doi.org/10.1109/JPROC.2007.895188 doi: 10.1109/JPROC.2007.895188
    [3] K. Arora, J. Singh, Y. S. Randhawa, A survey on channel coding techniques for 5G wireless networks, Telecommun. Syst., 73 (2020), 637–663. https://doi.org/10.1007/s11235-019-00630-3 doi: 10.1007/s11235-019-00630-3
    [4] A. E. Gamal, Y. H. Kim, Network information theory, Cambridge University Press, 2011. https://doi.org/10.1017/CBO9781139030687
    [5] L. He, M. Xue, B. Gu, Internet-of-things enabled supply chain planning and coordination with big data services: Certain theoretic implications, J. Manage. Sci. Eng., 5 (2020), 1–22. https://doi.org/10.1016/j.jmse.2020.03.002 doi: 10.1016/j.jmse.2020.03.002
    [6] H. Boche, M. Schubert, S. Stanczak, A unifying approach to multiuser receiver design under QoS constraints, In: 2005 IEEE 61st Vehicular Technol. Conf., 2 (2005), 992–996. https://doi.org/10.1016/j.talanta.2005.04.029
    [7] P. Gupta, P. Kumar, The capacity of wireless networks, IEEE T. Inform. Theory, 46 (2000), 388–404. https://doi.org/10.1109/18.825799 doi: 10.1109/18.825799
    [8] U. Pereg, C. Deppe, H. Boche, The quantum multiple-access channel with cribbing encoders, IEEE T. Inform. Theory, 68 (2022), 3965–3988. https://doi.org/10.1109/TIT.2022.3149827 doi: 10.1109/TIT.2022.3149827
    [9] M. Lederman, U. Pereg, Secure communication with unreliable entanglement assistance, In: 2024 IEEE Int. Symp. Inf. Theory (ISIT), 2024, 1017–1022. https://doi.org/10.1109/ISIT57864.2024.10619085
    [10] D. N. C. Tse, S. V. Hanly, Linear multiuser receivers: Effective interference, effective bandwidth and user capacity, IEEE T. Inform. Theory, 45 (1999), 641–657. https://doi.org/10.1109/18.749008 doi: 10.1109/18.749008
    [11] J. Rosenberger, C. Deppe, U. Pereg, Identification over quantum broadcast channels, Quantum Inf. Process., 22 (2023), 361. https://doi.org/10.1109/ISIT50566.2022.9834865 doi: 10.1109/ISIT50566.2022.9834865
    [12] U. Pereg, C. Deppe, H. Boche, The multiple-access channel with entangled transmitters, IEEE T. Inform. Theory, 71 (2025), 1096–1120. https://doi.org/10.1109/TIT.2024.3516507 doi: 10.1109/TIT.2024.3516507
    [13] K. S. K. Arumugam, M. R. Bloch, Covert communication over a k-user multiple-access channel, IEEE T. Inform. Theory, 65 (2019), 7020–7044. https://doi.org/10.1109/TIT.2019.2930484 doi: 10.1109/TIT.2019.2930484
    [14] L. Torres-Figueroa, R. Ferrara, C. Deppe, H. Boche, Message identification for task-oriented communications: Exploiting an exponential increase in the number of connected devices, IEEE Internet Things Mag., 6 (2023), 42–47. https://doi.org/10.1109/IOTM.001.2300166 doi: 10.1109/IOTM.001.2300166
    [15] M. Sudan, H. Tyagi, S. Watanabe, Communication for generating correlation: A unifying survey, IEEE T. Inform. Theory, 66 (2019), 5–37. https://doi.org/10.1109/TIT.2019.2946364 doi: 10.1109/TIT.2019.2946364
    [16] P. W. Cuff, H. H. Permuter, T. M. Cover, Coordination capacity, IEEE T. Inform. Theory, 56 (2010), 4181–4206. https://doi.org/10.1109/TIT.2010.2054651
    [17] M. Le Treust, Joint empirical coordination of source and channel, IEEE T. Inform. Theory, 63 (2017), 5087–5114. https://doi.org/10.1109/TIT.2017.2714682 doi: 10.1109/TIT.2017.2714682
    [18] M. R. Bloch, J. Kliewer, Strong coordination over a line network, In: 2013 IEEE Int. Symp. Inf. Theory (ISIT 2013), 2013, 2319–2323. https://doi.org/10.1109/ISIT.2013.6620640
    [19] M. Mylonakis, P. A. Stavrou, M. Skoglund, Remote empirical coordination, In: 2020 Int. Symp. Inf. Theory Appl. (ISITA 2020), IEEE, 2020, 31–35. https://doi.org/10.1007/s15016-020-7534-6
    [20] M. Berta, F. G. Brandão, M. Christandl, S. Wehner, Entanglement cost of quantum channels, IEEE T. Inform. Theory, 59 (2013), 6779–6795. https://doi.org/10.1109/TIT.2013.2268533 doi: 10.1109/TIT.2013.2268533
    [21] C. H. Bennett, I. Devetak, A. W. Harrow, P. W. Shor, A. Winter, The quantum reverse Shannon theorem and resource tradeoffs for simulating quantum channels, IEEE T. Inform. Theory, 60 (2014), 2926–2959. https://doi.org/10.1109/TIT.2014.2309968 doi: 10.1109/TIT.2014.2309968
    [22] M. M. Wilde, Entanglement cost and quantum channel simulation, Phys. Rev. A, 98 (2018), 042338. https://doi.org/10.1103/PhysRevA.98.042338 doi: 10.1103/PhysRevA.98.042338
    [23] I. George, M. H. Hsieh, E. Chitambar, One-shot bounds on state generation using correlated resources and local encoders, In: IEEE Int. Symp. Inf. Theory (ISIT 2023), 2023, 96–101. https://doi.org/10.1109/ISIT54713.2023.10206997
    [24] H. A. Salehi, F. Shirani, S. S. Pradhan, Quantum advantage in non-interactive source simulation, arXiv preprint, 2024. https://doi.org/10.48550/arXiv.2402.00242
    [25] M. Berta, O. Fawzi, S. Wehner, Quantum to classical randomness extractors, IEEE T. Inform. Theory, 60 (2014), 1168–1192. https://doi.org/10.1109/TIT.2013.2291780 doi: 10.1109/TIT.2013.2291780
    [26] M. Tahmasbi, M. R. Bloch, Steganography protocols for quantum channels, J. Math. Phys., 61 (2020). https://doi.org/10.1063/5.0004731
    [27] G. Vardoyan, E. van Milligen, S. Guha, S. Wehner, D. Towsley, On the bipartite entanglement capacity of quantum networks, IEEE Trans. Quantum Eng., 5 (2024), 1–14. https://doi.org/10.1109/TQE.2024.3443660 doi: 10.1109/TQE.2024.3443660
    [28] I. George, E. Chitambar, Revisiting pure state transformations with zero communication, arXiv preprint, 2023. https://doi.org/10.48550/arXiv.2301.04735
    [29] I. George, E. Chitambar, Reexamination of quantum state transformations with zero communication, Phys. Rev. A, 109 (2024), 062418. https://doi.org/10.1103/PhysRevA.109.062418 doi: 10.1103/PhysRevA.109.062418
    [30] I. Bjelaković, H. Boche, G. Janßen, Universal quantum state merging, J. Math. Phys., 54 (2013). https://doi.org/10.1063/1.4795243
    [31] M. Horodecki, J. Oppenheim, A. Winter, Quantum state merging and negative information, Commun. Math. Phys., 269 (2007), 107–136. https://doi.org/10.1007/s00220-006-0118-x doi: 10.1007/s00220-006-0118-x
    [32] P. Hayden, A. Winter, Communication cost of entanglement transformations, Phys. Rev. A, 67 (2003), 012326. https://doi.org/10.1103/PhysRevA.67.012326 doi: 10.1103/PhysRevA.67.012326
    [33] A. W. Harrow, H. K. Lo, A tight lower bound on the classical communication cost of entanglement dilution, IEEE T. Inform. Theory, 50 (2004), 319–327. https://doi.org/10.1109/TIT.2003.822597 doi: 10.1109/TIT.2003.822597
    [34] W. Kumagai, M. Hayashi, Entanglement concentration is irreversible, Phys. Rev. Lett., 111 (2013), 130407. https://doi.org/10.1103/PhysRevLett.111.130407 doi: 10.1103/PhysRevLett.111.130407
    [35] Z. Goldfeld, H. H. Permuter, G. Kramer, The Ahlswede-Körner coordination problem with one-sided encoder cooperation, In: Proc. IEEE Int. Symp. Inf. Theory (ISIT 2014), IEEE, 2014, 1341–1345. https://doi.org/10.1109/ISIT.2014.6875051
    [36] G. Kramer, S. A. Savari, Quantum data compression of ensembles of mixed states with commuting density operators, arXiv preprint, 2001. https://doi.org/10.48550/arXiv.quant-ph/0101119
    [37] E. Soljanin, Compressing quantum mixed-state sources by sending classical information, IEEE T. Inform. Theory, 48 (2002), 2263–2275. https://doi.org/10.1109/TIT.2002.800500 doi: 10.1109/TIT.2002.800500
    [38] Z. Goldfeld, H. H. Permuter, G. Kramer, Duality of a source coding problem and the semi-deterministic broadcast channel with rate-limited cooperation, IEEE T. Inform. Theory, 62 (2016), 2285–2307. https://doi.org/10.1109/TIT.2016.2533479 doi: 10.1109/TIT.2016.2533479
    [39] M. A. Sohail, T. A. Atif, S. S. Pradhan, A new formulation of lossy quantum-classical and classical source coding based on a posterior channel, In: IEEE Int. Symp. Inf. Theory (ISIT 2023), 2023,743–748. https://doi.org/10.1109/ISIT54713.2023.10206859
    [40] H. M. Garmaroudi, S. S. Pradhan, J. Chen, Rate-limited quantum-to-classical optimal transport: A lossy source coding perspective, In: IEEE Int. Symp. Inf. Theory (ISIT 2023), 2023, 1925–1930. https://doi.org/10.1109/ISIT54713.2023.10206947
    [41] M. Le Treust, M. Bloch, Empirical coordination, state masking and state amplification: Core of the decoder's knowledge, In: 2016 IEEE Int. Symp. Inf. Theory (ISIT 2016), IEEE, 2016,895–899. https://doi.org/10.1109/ISIT.2016.7541428
    [42] P. Cuff, L. Zhao, Coordination using implicit communication, In: 2011 IEEE Inf. Theory Workshop (ITW 2011), IEEE, 2011,467–471. https://doi.org/10.1109/ITW.2011.6089504
    [43] G. Cervia, L. Luzzi, M. R. Bloch, M. Le Treust, Polar coding for empirical coordination of signals and actions over noisy channels, In: 2016 IEEE Inf. Theory Workshop (ITW 2016), IEEE, 2016, 81–85. https://doi.org/10.1109/ITW.2016.7606800
    [44] R. Blasco-Serrano, R. Thobaben, M. Skoglund, Communication and interference coordination, In: 2014 Inf. Theory Appl. Workshop (ITA 2014), IEEE, 2014, 1–8. https://doi.org/10.1109/ITA.2014.6804218
    [45] F. Haddadpour, M. H. Yassaee, A. Gohari, M. R. Aref, Coordination via a relay, In: 2012 IEEE Int. Symp. Inf. Theory Proc. (ISIT 2012), IEEE, 2012, 3048–3052. https://doi.org/10.1109/ISIT.2012.6284121
    [46] C. A. Fuchs, A. Peres, Quantum-state disturbance versus information gain: Uncertainty relations for quantum information, Phys. Rev. A, 53 (1996), 2038. https://doi.org/10.1103/PhysRevA.53.2038 doi: 10.1103/PhysRevA.53.2038
    [47] C. A. Fuchs, A. Peres, Quantum theory needs no 'interpretation', Phys. Today, 53 (2000), 70–71. https://doi.org/10.1063/1.1325194 doi: 10.1063/1.1325194
    [48] J. Bricmont, Making sense of quantum mechanics, Springer, 37 (2016). https://doi.org/10.1007/978-3-319-25889-8
    [49] E. Soljanin, Compressing quantum mixed-state sources by sending classical information, IEEE T. Inform. Theory, 48 (2002), 2263–2275. https://doi.org/10.1109/TIT.2002.800500 doi: 10.1109/TIT.2002.800500
    [50] H. Barnum, C. M. Caves, C. A. Fuchs, R. Jozsa, B. Schumacher, On quantum coding for ensembles of mixed states, J. Phys. A: Math. General, 34 (2001), 35. https://doi.org/10.1023/A:1010336118743 doi: 10.1023/A:1010336118743
    [51] W. Dür, G. Vidal, J. Cirac, Visible compression of commuting mixed states, Phys. Rev. A, 64 (2001), 022308. https://doi.org/10.1103/PhysRevA.64.022308 doi: 10.1103/PhysRevA.64.022308
    [52] M. Horodecki, Limits for compression of quantum information carried by ensembles of mixed states, Phys. Rev. A, 57 (1998), 3364. https://doi.org/10.1103/PhysRevA.57.3364 doi: 10.1103/PhysRevA.57.3364
    [53] M. Horodecki, Towards optimal compression for mixed signal states, Preprint, 1999. https://doi.org/10.1103/PhysRevA.61.052309
    [54] M. Koashi, N. Imoto, Compressibility of quantum mixed-state signals, Phys. Rev. Lett., 87 (2001), 017902. https://doi.org/10.1103/PhysRevLett.87.017902 doi: 10.1103/PhysRevLett.87.017902
    [55] M. Koashi, N. Imoto, Operations that do not disturb partially known quantum states, Phys. Rev. A, 66 (2002), 022318. https://doi.org/10.1103/PhysRevA.66.022318 doi: 10.1103/PhysRevA.66.022318
    [56] M. Hayashi, Optimal visible compression rate for mixed states is determined by entanglement of purification, Phys. Rev. A-At., Mol. Opt. Phys., 73 (2006). https://doi.org/10.1103/PhysRevA.73.060301
    [57] Z. B. Khanian, From quantum source compression to quantum thermodynamics, arXiv preprint, 2020. https://doi.org/10.48550/arXiv.2012.14143
    [58] Z. B. Khanian, Strong converse bounds for compression of mixed states, arXiv preprint, 2022. https://doi.org/10.48550/arXiv.2206.09415
    [59] H. Natur, U. Pereg, Entanglement coordination rates in multi-user networks, In: 2024 IEEE Information Theory Workshop (ITW), 2024. https://doi.org/10.1109/ITW61385.2024.10807005
    [60] H. Nator, U. Pereg, Coordination capacity for classical-quantum correlations, In: 2024 IEEE Information Theory Workshop (ITW), 2024. https://doi.org/10.1109/ITW61385.2024.10807032
    [61] H. Nator, U. Pereg, Quantum coordination rates in multi-user networks, IEEE T. Inform. Theory, 2024, In press. https://doi.org/10.1109/TIT.2025.3554042
    [62] M. Hayashi, Quantum information theory: Mathematical foundation, Springer, 2016. https://doi.org/10.1007/978-3-662-49725-8
    [63] R. A. Chou, M. R. Bloch, J. Kliewer, Empirical and strong coordination via soft covering with polar codes, IEEE T. Inform. Theory, 64 (2018), 5087–5100. https://doi.org/10.1109/TIT.2018.2817519 doi: 10.1109/TIT.2018.2817519
    [64] T. A. Atif, S. S. Pradhan, A. Winter, Quantum soft-covering lemma with applications to rate-distortion coding, resolvability and identification via quantum channels, arXiv preprint, 2023. https://doi.org/10.48550/arXiv.2306.12416
    [65] I. Devetak, A. Winter, Distillation of secret key and entanglement from quantum states, P. Roy. Soc. A-Math. Phys., 461 (2005), 207–235. https://doi.org/10.1098/rspa.2004.1372 doi: 10.1098/rspa.2004.1372
    [66] M. Christandl, A. Ekert, M. Horodecki, P. Horodecki, J. Oppenheim, R. Renner, Unifying classical and quantum key distillation, Theory of Cryptography: 4th Theory of Cryptography Conference, TCC 2007, Amsterdam, The Netherlands, Springer, 2007,456–478. https://doi.org/10.1007/978-3-540-71095-0_8944
    [67] C. H. Bennett, I. Devetak, A. W. Harrow, P. W. Shor, A. Winter, The quantum reverse Shannon theorem and resource tradeoffs for simulating quantum channels, IEEE T. Inform. Theory, 60 (2014), 2926–2959. https://doi.org/10.1109/TIT.2014.2309968 doi: 10.1109/TIT.2014.2309968
    [68] A. K. Ekert, Quantum cryptography based on bell's theorem, Phys. Rev. Lett., 67 (1991), 661. https://doi.org/10.1103/PhysRevLett.67.661 doi: 10.1103/PhysRevLett.67.661
    [69] C. H. Bennett, G. Brassard, J. M. Robert, Privacy amplification by public discussion, SIAM J. Comput., 17 (1988), 210–229. https://doi.org/10.1137/0217014 doi: 10.1137/0217014
    [70] F. Dupuis, Privacy amplification and decoupling without smoothing, IEEE T. Inform. Theory, 69 (2023), 7784–7792. https://doi.org/10.1109/TIT.2023.3301812 doi: 10.1109/TIT.2023.3301812
    [71] Y. C. Shen, L. Gao, H. C. Cheng, Optimal second-order rates for quantum soft covering and privacy amplification, IEEE T. Inform. Theory, 70 (2024), 5077–5091. https://doi.org/10.1109/TIT.2024.3351963 doi: 10.1109/TIT.2024.3351963
    [72] M. Berta, O. Fawzi, S. Wehner, Quantum to classical randomness extractors, IEEE T. Inform. Theory, 60 (2013), 1168–1192. https://doi.org/10.1109/TIT.2013.2291780 doi: 10.1109/TIT.2013.2291780
    [73] K. Cheng, X. Li, Randomness extraction in AC0 and with small locality, arXiv preprint, 2016. https://doi.org/10.48550/arXiv.1602.01530
    [74] K. G. Anco, T. Nemoz, P. Brown, How much secure randomness is in a quantum state? arXiv preprint, 2024. https://doi.org/10.48550/arXiv.2410.16447
    [75] M. Horodecki, J. Oppenheim, A. Winter, Partial quantum information, Nature, 436 (2005), 673–676. https://doi.org/10.1038/nature03909
    [76] A. Abeyesinghe, I. Devetak, P. Hayden, A. Winter, The mother of all protocols: Restructuring quantum information's family tree, P. Roy. Soc. A-Math. Phys., 465 (2018), 2537–2563. https://doi.org/10.1098/rspa.2009.0202 doi: 10.1098/rspa.2009.0202
    [77] M. Berta, M. Christandl, D. Touchette, Smooth entropy bounds on one-shot quantum state redistribution, IEEE T. Inform. Theory, 62 (2016), 1425–1439. https://doi.org/10.1109/TIT.2016.2516006 doi: 10.1109/TIT.2016.2516006
    [78] I. Devetak, A triangle of dualities: reversibly decomposable quantum channels, source-channel duality, and time reversal, arXiv preprint, 2005. https://doi.org/10.48550/arXiv.quant-ph/0505138
    [79] J. Oppenheim, State redistribution as merging: Introducing the coherent relay, arXiv preprint, 2008. https://doi.org/10.48550/arXiv.0805.1065
    [80] M. Berta, M. Christandl, R. Renner, The quantum reverse Shannon theorem based on one-shot information theory, Commun. Math. Phys., 306 (2011), 579–615. https://doi.org/10.1007/s00220-011-1309-7 doi: 10.1007/s00220-011-1309-7
    [81] C. H. Bennett, P. W. Shor, J. A. Smolin, A. V. Thapliyal, Entanglement-assisted capacity of a quantum channel and the reverse Shannon theorem, IEEE T. Inform. Theory, 48 (2002), 2637–2655. https://doi.org/10.1109/TIT.2002.802612 doi: 10.1109/TIT.2002.802612
    [82] P. Cuff, Communication requirements for generating correlated random variables, In: 2008 IEEE Int. Symp. Inf. Theory, 2008, 1393–1397. https://doi.org/10.1109/ISIT.2008.4595216
    [83] C. Ahn, A. C. Doherty, P. Hayden, A. J. Winter, On the distributed compression of quantum information, IEEE T. Inform. Theory, 52 (2006), 4349–4357. https://doi.org/10.1109/TIT.2006.881734 doi: 10.1109/TIT.2006.881734
    [84] Z. B. Khanian, A. Winter, Distributed compression of correlated classical-quantum sources or: The price of ignorance, IEEE T. Inform. Theory, 66 (2020), 5620–5633. https://doi.org/10.1109/TIT.2020.2981322 doi: 10.1109/TIT.2020.2981322
    [85] S. Salek, D. Cadamuro, P. Kammerlander, K. Wiesner, Quantum rate-distortion coding of relevant information, IEEE T. Inform. Theory, 65 (2018), 2603–2613. https://doi.org/10.1109/TIT.2018.2878412 doi: 10.1109/TIT.2018.2878412
    [86] T. A. Atif, M. Heidari, S. S. Pradhan, Faithful simulation of distributed quantum measurements with applications in distributed rate-distortion theory, IEEE T. Inform. Theory, 68 (2022), 1085–1118. https://doi.org/10.1109/TIT.2021.3124976 doi: 10.1109/TIT.2021.3124976
    [87] Z. B. Khanian, K. Kuroiwa, D. Leung, Rate-distortion theory for mixed states, IEEE T. Inform. Theory, 71 (2024), 1077–1095. https://doi.org/10.1109/TIT.2024.3509825 doi: 10.1109/TIT.2024.3509825
    [88] P. Colomer, A. Winter, Decoupling by local random unitaries without simultaneous smoothing, and applications to multi-user quantum information tasks, Commun. Math. Phys., 405 (2024), 281. https://doi.org/10.1007/s00220-024-05191-4 doi: 10.1007/s00220-024-05191-4
    [89] H. C. Cheng, L. Gao, M. Berta, Quantum broadcast channel simulation via multipartite convex splitting, arXiv preprint, 2023. https://doi.org/10.48550/arXiv.2304.12056
    [90] M. X. Cao, N. Ramakrishnan, M. Berta, M. Tomamichel, Channel simulation: Finite blocklengths and broadcast channels, IEEE T. Inform. Theory, 70 (2024), 6780–6808. https://doi.org/10.1109/TIT.2024.3445998 doi: 10.1109/TIT.2024.3445998
    [91] A. Nema, S. Sreekumar, M. Berta, One-shot multiple access channel simulation, In: 2024 IEEE Int. Symp. Inf. Theory (ISIT), IEEE, 2024, 2981–2986. https://doi.org/10.1109/ISIT57864.2024.10619283
    [92] I. George, H. C. Cheng, Coherent distributed source simulation as multipartite quantum state splitting, In: 2024 IEEE Int. Symp. Inf. Theory (ISIT), IEEE, 2024, 1221–1226. https://doi.org/10.1109/ISIT57864.2024.10619569
    [93] J. A. Smolin, F. Verstraete, A. Winter, Entanglement of assistance and multipartite state distillation, Phys. Rev. A-At., Mol. Opt. Phys., 72 (2005), 052317. https://doi.org/10.1103/PhysRevA.72.052317 doi: 10.1103/PhysRevA.72.052317
    [94] S. Bravyi, D. Fattal, D. Gottesman, GHZ extraction yield for multipartite stabilizer states, J. Math. Phys., 47 (2006). https://doi.org/10.1063/1.2203431
    [95] R. Augusiak, P. Horodecki, Multipartite secret key distillation and bound entanglement, Phys. Rev. A-At., Mol. Opt. Phys., 80 (2009), 042307. https://doi.org/10.1103/PhysRevA.80.042307 doi: 10.1103/PhysRevA.80.042307
    [96] A. Streltsov, C. Meignant, J. Eisert, Rates of multi-partite entanglement transformations and applications in quantum networks, arXiv preprint, 2017. https://doi.org/10.48550/arXiv.1709.09693
    [97] G. Murta, F. Grasselli, H. Kampermann, D. Bruß, Quantum conference key agreement: A review, Adv. Quantum Technol., 3 (2020), 2000025. https://doi.org/10.1002/qute.202070032 doi: 10.1002/qute.202070032
    [98] F. Salek, A. Winter, Multi-user distillation of common randomness and entanglement from quantum states, IEEE T. Inform. Theory, 68 (2022), 976–988. https://doi.org/10.1109/TIT.2021.3124965 doi: 10.1109/TIT.2021.3124965
    [99] F. Salek, A. Winter, New protocols for conference key and multipartite entanglement distillation, arXiv preprint, 2023. https://doi.org/10.48550/arXiv.2308.01134
    [100] A. Streltsov, C. Meignant, J. Eisert, Rates of multipartite entanglement transformations, Phys. Rev. Lett., 125 (2020), 080502. https://doi.org/10.1103/PhysRevLett.125.080502 doi: 10.1103/PhysRevLett.125.080502
    [101] M. Q. Vu, T. V. Pham, N. T. Dang, A. T. Pham, Design and performance of relay-assisted satellite free-space optical quantum key distribution systems, IEEE Access, 8 (2020), 122498–122510. https://doi.org/10.1109/ACCESS.2020.3007461 doi: 10.1109/ACCESS.2020.3007461
    [102] J. L. Jiang, M. X. Luo, S. Y. Ma, Quantum network capacity of entangled quantum internet, IEEE J. Sel. Area. Comm., 2024. https://doi.org/10.1109/JSAC.2024.3380091
    [103] M. M. Wilde, Quantum information theory, 2 Eds., Cambridge Univ. Press, 2017.
    [104] H. G. Eggleston, Convexity, J. Lond. Math. Soc., 1 (1966), 183–186. https://doi.org/10.1112/jlms/s1-41.1.183b
    [105] U. Pereg, Communication over quantum channels with parameter estimation, IEEE T. Inform. Theory, 68 (2022), 359–383. https://doi.org/10.1109/TIT.2021.3123221 doi: 10.1109/TIT.2021.3123221
    [106] C. E. Shannon, A mathematical theory of communication, Bell Syst. Tech. J., 27 (1948), 379–423. https://doi.org/10.1002/j.1538-7305.1948.tb01338.x doi: 10.1002/j.1538-7305.1948.tb01338.x
    [107] I. Csiszár, J. Körner, Information theory: Coding theorems for discrete memoryless systems, Cambridge Univ. Press, 2011. https://doi.org/10.4074/S2113520711012138
    [108] O. Gossner, P. Hernandez, A. Neyman, Online matching pennies, Technical Report, The Federmann Center for the Study of Rationality, Hebrew University of Jerusalem, Discussion Paper Series dp316, 2003.
    [109] A. P. Flitney, D. Abbott, An introduction to quantum game theory, Fluct. Noise Lett., 2 (2002), R175–R187. https://doi.org/10.1142/S0219477502000981 doi: 10.1142/S0219477502000981
    [110] C. Borcea, D. Iyer, P. Kang, A. Saxena, L. Iftode, Cooperative computing for distributed embedded systems, In: Proc. 22nd Int. Conf. Distrib. Comput. Syst., IEEE, 2002,227–236. https://doi.org/10.1109/ICDCS.2002.1022260
    [111] M. N. Ahangar, Q. Z. Ahmed, F. A. Khan, M. Hafeez, A survey of autonomous vehicles: Enabling communication technologies and challenges, Sensors, 21 (2021), 706. https://doi.org/10.3390/s21030706 doi: 10.3390/s21030706
    [112] J. A. Stankovic, T. Abdelzaher, C. Lu, L. Sha, J. C. Hou, Real-time communication and coordination in embedded sensor networks, P. IEEE, 91 (2003), 1002–1022. https://doi.org/10.1109/JPROC.2003.814620 doi: 10.1109/JPROC.2003.814620
    [113] I. Burenkov, M. Jabir, S. Polyakov, Practical quantum-enhanced receivers for classical communication, AVS Quantum Sci., 3 (2021). https://doi.org/10.1116/5.0036959
    [114] F. Granelli, R. Bassoli, J. Nötzel, F. H. Fitzek, H. Boche, N. L. da Fonseca, A novel architecture for future classical-quantum communication networks, Wirel. Commun. Mob. Com., 2022 (2022). https://doi.org/10.1155/2022/3770994
    [115] J. Nötzel, Entanglement-enabled communication, IEEE J. Sel. Area. Inf. Theory, 1 (2020), 401–415. https://doi.org/10.1109/JSAIT.2020.3017121 doi: 10.1109/JSAIT.2020.3017121
    [116] J. Nötzel, S. DiAdamo, Entanglement-enabled communication for the Internet of things, In: 2020 Int. Conf. Comput., Inf. Telecommun. Syst. (CITS 2020), 2020, 1–6. https://doi.org/10.1109/CITS49457.2020.9232550
    [117] B. Schumacher, Quantum coding, Phys. Rev. A, 51 (1995), 2738. https://doi.org/10.1103/PhysRevA.51.2738
    [118] R. Jozsa, B. Schumacher, A new proof of the quantum noiseless coding theorem, J. Mod. Optic., 41 (1994), 2343–2349. https://doi.org/10.1080/09500349414552191 doi: 10.1080/09500349414552191
    [119] H. Barnum, C. A. Fuchs, R. Jozsa, B. Schumacher, General fidelity limit for quantum channels, Phys. Rev. A, 54 (1996), 4707. https://doi.org/10.1103/PhysRevA.54.4707 doi: 10.1103/PhysRevA.54.4707
    [120] S. Pirandola, S. L. Braunstein, R. Laurenza, C. Ottaviani, T. P. Cope, G. Spedalieri, et al., Theory of channel simulation and bounds for private communication, Quantum Sci. Technol., 3 (2018), 035009. https://doi.org/10.1088/2058-9565/aac394 doi: 10.1088/2058-9565/aac394
    [121] J. T. Yard, I. Devetak, Optimal quantum source coding with quantum side information at the encoder and decoder, IEEE T. Inform. Theory, 55 (2009), 5339–5351. https://doi.org/10.1109/TIT.2009.2030494 doi: 10.1109/TIT.2009.2030494
  • Reader Comments
  • © 2025 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(181) PDF downloads(19) Cited by(0)

Figures and Tables

Figures(7)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog