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

thanks: This work was presented in part at IEEE International Symposium on Information Theory (ISIT) 2021.

Distributed Quantum Faithful Simulation and Function Computation Using Algebraic Structured Measurements

Touheed Anwar Atif and S. Sandeep Pradhan University of Michigan, Ann Arbor
touheed@umich.edu, pradhanv@umich.edu
Abstract

In this work, we consider the task of faithfully simulating a quantum measurement, acting on a joint bipartite quantum state, in a distributed manner. In the distributed setup, the constituent sub-systems of the joint quantum state are measured by two agents, Alice and Bob. A third agent, Charlie receives the measurement outcomes sent by Alice and Bob. Charlie uses local and pairwise shared randomness to compute a bivariate function of the measurement outcomes. The objective of three agents is to faithfully simulate the given distributed quantum measurement acting on the given quantum state while minimizing the communication and shared randomness rates. We demonstrate a new achievable quantum information-theoretic rate-region that exploits the bivariate function using random structured POVMs based on asymptotically good algebraic codes. The algebraic structure of these codes is matched to that of the bivariate function that models the action of Charlie. The conventional approach for this class of problems has been to reconstruct individual measurement outcomes corresponding to Alice and Bob, at Charlie, and then compute the bivariate function. This is achieved using mutually independent approximating POVMs based on random unstructured codes. In the present approach, using algebraic structured POVMs, the computation is performed on the fly, thus obviating the need to reconstruct individual measurement outcomes at Charlie. Using this, we show that a strictly larger rate region can be achieved. The performance limit is characterized using single-letter quantum mutual information quantities. We provide examples to illustrate the information-theoretic gains attained by endowing POVMs with algebraic structure. One of the challenges in analyzing these structured POVMs is that they exhibit only pairwise independence and induce only uniform single-letter distributions. To address this, we use nesting of algebraic codes and develop a covering lemma applicable to pairwise-independent POVM ensembles. Combining these techniques, we provide a multi-party distributed faithful simulation and function computation protocol.

preprint: APS/123-QED

I Introduction

Measurement compression is one of the foremost and fundamental quantum information processing techniques which form the basis of many quantum protocols [1]. One of the seminal works in this regard was by Winter [2], where he performed a novel information theoretic analysis to compress measurements in an asymptotic sense. The measurement compression problem formulated in [2] is as follows. Consider an agent (Alice) who performs a measurement MM on a quantum state ρ\rho, and sends a set of classical bits to another agent (Bob). Bob intends to faithfully recover the outcomes of Alice’s measurements without having access to ρ\rho, while preserving the correlation with the post-measured state of Alice’s reference. The major contribution of this work (as elaborated in [3]) was in specifying an optimal rate region in terms of classical communication and common randomness needed to faithfully simulate the action of repeated independent measurements performed on many independent copies of the given quantum state.

Wilde et al. [3] extended the measurement compression problem by considering additional resources available to each of the participating parties. One such formulation allows Bob to process the information received from Alice using local private randomness. The authors here also combined the ideas from [2] and [4] to simulate a measurement in presence of quantum side information. In the above problem formulations, authors have derived the results using the prevalent random coding techniques analogous to Shannon’s unstructured random codes [5] involving mutually independent codewords. The point-to-point setup [2, 3] requires randomly generating approximating POVMs and analyzing the error associated with these approximating POVMs, also termed as “covering error”. The key analytical tool that facilitates this is the operator Chernoff bound [6], which crucially exploits the mutual independence of codewords, yielding the quantum covering lemma [7, Lemma 17.2.1].

The measurement compression problem has been studied extensively. Early works on quantifying the information gain of a measurement include [8, 9, 10]. Buscemi et al. [11, 12, 13] later advocated quantum mutual information with respect to a classical-quantum state as the measure to characterize the corresponding information gain. Berta et al. [14] generalized the Winter’s measurement compression theorem by developing a universal measurement compression theorem for arbitrary inputs, and identified the quantum mutual information of a measurement as the information gained by performing the measurement, independent of the input state on which it is performed. They provide a proof based on new “classically coherent state merging protocol” - a variation of the quantum state merging protocol [15, 16], and the post-selection technique for quantum channels [17].

Anshu et al. [18] considered the problem of measurement compression with side information in the one-shot setting. They presented a protocol by proposing a new convex-split lemma for classical-quantum states and employing the position based decoding, and bounded the communication in terms of smooth max and hypothesis testing relative entropies. The original convex-split lemma [19, 20] demanded sub-optimal shared-randomness rate in the one-shot setting, by requiring large amount of additional quantum states in its statement. The authors addressed this by modifying the lemma to only use pairwise independent random variables. This substantially simplified the derandomization required, leading to an exponential reduction in the randomness cost in comparison to [19]. Considering a related problem, Renes and Renner [21] also studied sending of classical messages in the presence of quantum side information in the one-shot setting. For more discussion and results pertaining to one-shot quantum information theory, the reader is directed to [22, 23].

Furthermore, the authors in [24] considered the task of quantifying “relevant information” for the quantum measurements performed in a distributed fashion on bipartite entangled states involving three agents. In this multi-terminal setting, a composite bipartite quantum system ABAB is made available to two agents, Alice and Bob, where they have access to the sub-systems AA and BB, respectively. Two separate measurements, one for each sub-system, are performed in a distributed fashion with no communication taking place between Alice and Bob. A third party, Charlie, is connected to Alice and Bob via two separate classical links. The objective of the three parties is to simulate the action of repeated independent measurements performed on many independent copies of the given composite state. Further, common randomness at rate CC is also shared amidst the three parties. This is achieved using random unstructured code ensembles while still using the operator Chernoff bound.

The measurement compression theorem has found its applications in several quantum information processing protocols. Examples include the quantum reverse Shannon theorem [25, 26, 27], local purity distillation protocols [28, 29, 30, 31], and also in the grandmother protocol [1] which is useful in entanglement distillation from noisy quantum states.

An ubiquitous application of distributed systems in current quantum settings arises due to the inherent vulnerability of the large-scale quantum computation systems to noise. The state-of-art systems exhibits technical difficulties in increasing the number of low-noise qubits in a single quantum device. A solution to this is cooperative processing of information separately on spatially segregated units. This necessitates the need for distributed compression protocols to compress efficiently and recover the data. In addition, when one is interested in solely reconstructing functions of the distributively stored quantum data, the rate of communication may be further reduced by employing structured coding techniques. For this, we need to impose further structure on these POVMs. This is to ensure that the joint decoder (Charlie) is able to reconstruct a lower dimensional quantum state with minimal use of the classical communication resource. Hence, structure of the POVM is desired to match with the structure of the function being computed.

The traditional random coding techniques using unstructured code ensembles may not always achieve optimality for distributed multi-terminal settings. For instance, the work by Korner-Marton [32] demonstrated this sub-optimality for the problem of classical distributed lossless compression with the objective of computing the sum of the sources for the binary symmetric case using random linear codes. Traditionally, algebraic-structured codes are used in information coding problems toward achieving computationally efficient (polynomial-time) encoding and decoding algorithms. However, in multi-terminal communication problems, even if computational complexity is a non-issue, random algebraic structured codes outperform random unstructured codes in terms of achieving improved asymptotic rate regions in many cases [33, 34, 35, 36].

Motivated by this, we consider the quantum distributed faithful measurement simulation problem and present a new achievable rate-region using algebraic structured coding techniques. However, there are two main challenges in using these algebraic structured codes toward an asymptotic analysis in quantum information theory. The first challenge is to be able to induce arbitrary empirical single-letter distributions. For example, if we were to send codewords from a linear code with uniform probability, then the induced empirical distribution of codeword symbols (single-letter distribution on the symbols of the codewords) is uniform. To address this challenge, we use a collection of cosets of a linear code called Unionized Coset Codes (UCCs) [37]. The second challenge is that unlike the random unstructured codes, the codewords generated from a random linear code are only pairwise-independent [38]. This renders the above technique of operator Chernoff bound, or even the covering lemma, unusable. Since our approach relies on the use of UCCs for generating the approximating POVMs, the binning of these POVM elements is performed in a correlated fashion as governed by these structured codes. This is in contrast to the common technique of independent binning. Due to the correlated binning, the pairwise-independence issue gets exacerbated.

We address these challenges using three main ideas summarized as follows:

  • Random structured generation of pruned POVMs - We generate a collection of algebraic structured approximating POVMs randomly using the above described UCC technique, and then prune them. This pruning ensures that these POVMs form a positive resolution of identity, and thus eliminates any need for the operator Chernoff inequality. However, such pruning comes at the cost of additional approximating error. To bound the approximating error caused by pruning the POVMs, we develop a new Operator Inequality which provides a handle to convert the pruning error in the form of covering error expression (dealt within the next idea).

  • Covering Lemma for Pairwise-Independent Ensemble - Since the traditional covering lemma is based on the Chernoff inequality, we develop an alternative proof for the aforementioned covering lemma [39, Lemma 17.2.1]. This alternative proof is based on the second-order analysis using the operator trace inequalities and hence requires the operators to be only pairwise-independent.

  • Multi-partite Packing Lemma - We develop a binning technique for performing computation on the fly so as to achieve a low dimensional reconstruction of a function at the location of Charlie. In an effort towards analysing this binning technique, we develop a multi-partite packing Lemma for the structured POVMs.

Combining these techniques, we provide a multi-party distributed faithful simulation and function computation protocol in a quantum information theoretic setting. We provide a characterization of the asymptotic performance limit of this protocol in terms of a computable single-letter achievable rate-region, which is the main result of the paper (see Theorem 1).

The organization of the paper is as follows. In Section II, we set the notation, state requisite definitions and also provide related results. In Section III.1 we state our main result on the distributed measurement compression and provide the theorem (Theorem 1) characterizing the rate-region. In Section III.2 we provide a new Covering Lemma for pairwise-independent ensembles. Section IV provides useful lemmas. In Section V, we consider the point-to-point setup and provide a theorem characterizing the rate-region using algebraic structured codes. We prove the main result (Theorem 1) in Section VI using the point-to-point result as a building block. Finally, we conclude the paper in Section VII.

II Preliminaries

Notation: Given any natural number MM, let the finite set {1,2,,M}\{1,2,\cdots,M\} be denoted by [1,M][1,M]. Let ()\mathcal{B(H)} denote the algebra of all bounded linear operators acting on a finite dimensional Hilbert space \mathcal{H}. Further, let 𝒟()\mathcal{D(H)} denote the set of all unit trace positive operators acting on \mathcal{H}. Let II denote the identity operator. The trace distance between two operators AA and BB is defined as AB1\ensurestackMath\stackon[1pt]=ΔTr|AB|\|A-B\|_{1}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\Tr|A-B|, where for any operator Λ\Lambda we define |Λ|\ensurestackMath\stackon[1pt]=ΔΛΛ|\Lambda|\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\sqrt{\Lambda^{\dagger}\Lambda}. The von Neumann entropy of a density operator ρ𝒟()\rho\in\mathcal{D}(\mathcal{H}) is denoted by S(ρ)S(\rho). The quantum mutual information for a bipartite density operator ρAB𝒟(AB)\rho_{AB}\in\mathcal{D}(\mathcal{H}_{A}\otimes\mathcal{H}_{B}) is defined as

I(A;B)ρ\displaystyle I(A;B)_{\rho} \ensurestackMath\stackon[1pt]=ΔS(ρA)+S(ρB)S(ρAB).\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}S(\rho_{A})+S(\rho_{B})-S(\rho_{AB}).

A positive-operator valued measure (POVM) acting on a Hilbert space \mathcal{H} is a collection M\ensurestackMath\stackon[1pt]=Δ{Λx}x𝒳M\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\{\Lambda_{x}\}_{x\in\mathcal{X}} of positive operators in ()\mathcal{B}(\mathcal{H}) that form a resolution of the identity:

Λx0,x𝒳,x𝒳Λx=I,\displaystyle\Lambda_{x}\geq 0,\forall x\in\mathcal{X},\qquad\sum_{x\in\mathcal{X}}\Lambda_{x}=I,

where 𝒳\mathcal{X} is a finite set. If instead of the equality above, the inequality xΛxI\sum_{x}\Lambda_{x}\leq I holds, then the collection is said to be a sub-POVM. A sub-POVM MM can be completed to form a POVM, denoted by [M][M], by adding the operator Λ0\ensurestackMath\stackon[1pt]=Δ(IxΛx)\Lambda_{0}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}(I-\sum_{x}\Lambda_{x}) to the collection. Let ΨRAρ\Psi^{\rho}_{RA} denote a purification of a density operator ρD(A)\rho\in D(\mathcal{H}_{A}). Given a POVM M\ensurestackMath\stackon[1pt]=Δ{ΛxA}x𝒳M\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\{\Lambda^{A}_{x}\}_{x\in\mathcal{X}} acting on ρ\rho, the post-measurement state of the reference together with the classical outputs is represented by

(idM)(ΨRAρ)\ensurestackMath\stackon[1pt]=Δx𝒳|xx|TrA{(IRΛxA)ΨRAρ}.(\text{id}\otimes M)(\Psi^{\rho}_{RA})\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\sum_{x\in\mathcal{X}}\outerproduct{x}{x}\otimes\Tr_{A}\{(I^{R}\otimes\Lambda_{x}^{A})\Psi^{\rho}_{RA}\}. (1)

Consider two POVMs MA={ΛxA}x𝒳M_{A}=\{\Lambda^{A}_{x}\}_{x\in\mathcal{X}} and MB={ΛyB}y𝒴M_{B}=\{\Lambda^{B}_{y}\}_{y\in\mathcal{Y}} acting on A\mathcal{H}_{A} and B\mathcal{H}_{B}, respectively. Define MAMB\ensurestackMath\stackon[1pt]=Δ{ΛxAΛyB}x𝒳,y𝒴M_{A}\otimes M_{B}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\{\Lambda^{A}_{x}\otimes\Lambda^{B}_{y}\}_{x\in\mathcal{X},y\in\mathcal{Y}} With this definition, MAMBM_{A}\otimes M_{B} is a POVM acting on AB\mathcal{H}_{A}\otimes\mathcal{H}_{B}. By MnM^{\otimes n} denote the nn-fold tensor product of the POVM MM with itself. For a prime pp, we denote the unique finite field of size pp by 𝔽p\mathbb{F}_{p}, and denote the addition operation over the field by ++.

Definition 1 (Faithful simulation [3]).

Given a POVM M\ensurestackMath\stackon[1pt]=Δ{Λx}x𝒳{M}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\{\Lambda_{x}\}_{x\in\mathcal{X}} acting on a Hilbert space \mathcal{H} and a density operator ρ𝒟()\rho\in\mathcal{D}(\mathcal{H}), a sub-POVM M~\ensurestackMath\stackon[1pt]=Δ{Λ~x}x𝒳\tilde{M}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\{\tilde{\Lambda}_{x}\}_{x\in\mathcal{X}} acting on \mathcal{H} is said to be ϵ\epsilon-faithful to MM with respect to ρ\rho, for ϵ>0\epsilon>0, if the following holds:

x𝒳ρ(ΛxΛ~x)ρ1+Tr{(IxΛ~x)ρ}ϵ.\sum_{x\in\mathcal{X}}\Big{\|}\sqrt{\rho}(\Lambda_{x}-\tilde{\Lambda}_{x})\sqrt{\rho}\Big{\|}_{1}+\Tr\left\{(I-\sum_{x}\tilde{\Lambda}_{x})\rho\right\}\leq\epsilon. (2)
Lemma 1.

Given a density operator ρAB𝒟(AB)\rho_{AB}\in\mathcal{D}(\mathcal{H}_{A}\otimes\mathcal{H}_{B}), a sub-POVM MY\ensurestackMath\stackon[1pt]=Δ{ΛyB:y𝒴}M_{Y}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\left\{\Lambda_{y}^{B}:y\in\mathcal{Y}\right\} acting on B,\mathcal{H}_{B}, for some set 𝒴\mathcal{Y}, and any Hermitian operator ΓA\Gamma^{A} acting on A\mathcal{H}_{A}, we have

y𝒴ρAB(ΓAΛyB)ρAB1ρAΓAρA1,\displaystyle\sum_{y\in\mathcal{Y}}\left\|\sqrt{\rho_{AB}}\left(\Gamma^{A}\otimes\Lambda_{y}^{B}\right)\sqrt{\rho_{AB}}\right\|_{1}\leq\left\|\sqrt{\rho_{A}}\Gamma^{A}\sqrt{\rho_{A}}\right\|_{1}, (3)

with equality if y𝒴ΛyB=I\displaystyle\sum_{y\in\mathcal{Y}}\Lambda_{y}^{B}=I, where ρA=TrB{ρAB}\rho_{A}=\Tr_{B}\{\rho_{AB}\}.

Proof.

The proof is provided in Lemma 3 of [24]. ∎

III Main Results

In this section we present the main results of this paper.

III.1 Simulation of Distributed POVMs using Algebraic-Structured POVMs

Let ρAB\rho_{AB} be a density operator acting on a composite Hilbert Space AB\mathcal{H}_{A}\otimes\mathcal{H}_{B}. Consider two measurements MAM_{A} and MBM_{B} on sub-systems AA and BB, respectively. Imagine again that we have three parties, named Alice, Bob and Charlie, that are trying to collectively simulate the action of a given measurement MABM_{AB} performed on the state ρAB\rho_{AB}, as shown in Fig. 1. Charlie additionally has access to unlimited private randomness. The problem is defined in the following.

Refer to caption
Figure 1: The diagram depicting the distributed POVM simulation problem with stochastic processing. In this setting, Charlie additionally has access to unlimited private randomness.
Definition 2.

For a given finite set 𝒵\mathcal{Z}, and a Hilbert space AB\mathcal{H}_{A}\otimes\mathcal{H}_{B}, a distributed protocol with stochastic processing with parameters (n,Θ1,Θ2,N1,N2)(n,\Theta_{1},\Theta_{2},N_{1},N_{2}) is characterized by
1) a collections of Alice’s sub-POVMs M~A(μ1),μ1[1,N1]\tilde{M}_{A}^{(\mu_{1})},\mu_{1}\in[1,N_{1}] each acting on An\mathcal{H}_{A}^{\otimes n} and with outcomes in a subset 1\mathcal{L}_{1} satisfying |1|Θ1|\mathcal{L}_{1}|\leq\Theta_{1}.
2) a collections of Bob’s sub-POVMs M~B(μ2),μ2[1,N2]\tilde{M}_{B}^{(\mu_{2})},\mu_{2}\in[1,N_{2}] each acting on Bn\mathcal{H}_{B}^{\otimes n} and with outcomes in a subset 2\mathcal{L}_{2}, satisfying |2|Θ2|\mathcal{L}_{2}|\leq\Theta_{2}.
3) a collection of Charlie’s classical stochastic maps P(μ1,μ2)(zn|l1,l2)P^{(\mu_{1},\mu_{2})}(z^{n}|l_{1},l_{2}) for all l11,l22,zn𝒵nl_{1}\in\mathcal{L}_{1},l_{2}\in\mathcal{L}_{2},z^{n}\in\mathcal{Z}^{n}, μ1[1,N1]\mu_{1}\in[1,N_{1}] and μ2[1,N2]\mu_{2}\in[1,N_{2}].
The overall sub-POVM of this distributed protocol, given by M~AB\tilde{M}_{AB}, is characterized by the following operators:

Λ~zn\ensurestackMath\stackon[1pt]=Δ1N11N2μ1,μ2\displaystyle\tilde{\Lambda}_{z^{n}}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\frac{1}{N_{1}}\frac{1}{N_{2}}\sum_{\mu_{1},\mu_{2}} l1,l2P(μ1,μ2)(zn|l1,l2)\displaystyle\sum_{l_{1},l_{2}}P^{(\mu_{1},\mu_{2})}(z^{n}|l_{1},l_{2})
Λl1A,(μ1)Λl2B,(μ2),zn𝒵n,\displaystyle\Lambda^{A,(\mu_{1})}_{l_{1}}\otimes\Lambda^{B,(\mu_{2})}_{l_{2}},\quad\forall z^{n}\in\mathcal{Z}^{n},

where Λl1A,(μ1)\Lambda^{A,(\mu_{1})}_{l_{1}} and Λl2B,(μ2)\Lambda^{B,(\mu_{2})}_{l_{2}} are the operators corresponding to the sub-POVMs M~A(μ1)\tilde{M}_{A}^{(\mu_{1})} and M~B(μ2)\tilde{M}_{B}^{(\mu_{2})}, respectively.

In the above definition, (Θ1,Θ2)(\Theta_{1},\Theta_{2}) determines the amount of classical bits communicated from Alice and Bob to Charlie. The amount of pairwise shared randomness is determined by N1N_{1} and N2N_{2}. The classical stochastic maps P(μ1,μ2)(zn|l1,l2)P^{(\mu_{1},\mu_{2})}(z^{n}|l_{1},l_{2}) represent the action of Charlie on the received classical bits.

Definition 3.

Given a POVM MABM_{AB} acting on AB\mathcal{H}_{A}\otimes\mathcal{H}_{B}, and a density operator ρAB𝒟(AB)\rho_{AB}\in\mathcal{D}(\mathcal{H}_{A}\otimes\mathcal{H}_{B}), a quadruple (R1,R2,C1,C2)(R_{1},R_{2},C_{1},C_{2}) is said to be achievable, if for all ϵ>0\epsilon>0 and for all sufficiently large nn, there exists a distributed protocol with stochastic processing with parameters (n,Θ1,Θ2,N1,N2)(n,\Theta_{1},\Theta_{2},N_{1},N_{2}) such that its overall sub-POVM M~AB\tilde{M}_{AB} is ϵ\epsilon-faithful to MABnM_{AB}^{\otimes n} with respect to ρABn\rho_{AB}^{\otimes n} (see Definition 1), and

1nlog2ΘiRi+ϵ,and1nlog2NiCi+ϵ,i=1,2.\displaystyle\frac{1}{n}\log_{2}\Theta_{i}\leq R_{i}+\epsilon,\!\quad\mbox{and}\!\quad\!\frac{1}{n}\log_{2}N_{i}\leq C_{i}+\epsilon,\quad i=1,2.

The set of all achievable quadruples (R1,R2,C1,C2)(R_{1},R_{2},C_{1},C_{2}) is called the achievable rate region.

Definition 4 (Joint Measurements).

A POVM MAB={ΛzAB}z𝒵M_{AB}=\{\Lambda^{AB}_{z}\}_{z\in\mathcal{Z}}, acting on a Hilbert space AB\mathcal{H}_{A}\otimes\mathcal{H}_{B}, is said to have a separable decomposition with stochastic integration given by (M¯A,M¯B,PZ|S,T)(\bar{M}_{A},\bar{M}_{B},P_{Z|S,T}) if there exist POVMs M¯A={Λ¯sA}s𝒮\bar{M}_{A}=\{\bar{\Lambda}^{A}_{s}\}_{s\in\mathcal{S}} and M¯B={Λ¯tB}t𝒯\bar{M}_{B}=\{\bar{\Lambda}^{B}_{t}\}_{t\in\mathcal{T}} and a stochastic mapping PZ|S,T:𝒮×𝒯𝒵P_{Z|S,T}:\mathcal{S}\times\mathcal{T}\rightarrow\mathcal{Z} such that

ΛzAB=s,tPZ|S,T(z|s,t)Λ¯sAΛ¯tB,z𝒵,\Lambda^{AB}_{z}=\sum_{s,t}P_{Z|S,T}(z|s,t)\bar{\Lambda}^{A}_{s}\otimes\bar{\Lambda}^{B}_{t},\quad\forall z\in\mathcal{Z}, (4)

where 𝒮,𝒯\mathcal{S},\mathcal{T}, and 𝒵\mathcal{Z} are finite sets.

The following theorem provides an inner bound to the achievable rate region, which is proved in Section VI. This is one of the main results of this paper.

Theorem 1.

Consider a density operator ρAB𝒟(AB)\rho_{AB}\in\mathcal{D}(\mathcal{H}_{A}\otimes\mathcal{H}_{B}), and a POVM MAB={ΛzAB}z𝒵M_{AB}=\{\Lambda^{AB}_{z}\}_{z\in\mathcal{Z}} acting on AB\mathcal{H}_{A}\otimes\mathcal{H}_{B} having a separable decomposition with stochastic integration (as in Definition 4), yielding POVMs M¯A={Λ¯sA}s𝒮\bar{M}_{A}=\{\bar{\Lambda}^{A}_{s}\}_{s\in\mathcal{S}} and M¯B={Λ¯tB}t𝒯\bar{M}_{B}=\{\bar{\Lambda}^{B}_{t}\}_{t\in\mathcal{T}} and a stochastic map PZ|S,T:𝒮×𝒯𝒵P_{Z|S,T}:\mathcal{S}\times\mathcal{T}\rightarrow\mathcal{Z}. Define the auxiliary states

σ1RSB\displaystyle\sigma_{1}^{RSB} \ensurestackMath\stackon[1pt]=Δ(idRM¯AidB)(ΨRABρAB),\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}(\emph{id}_{R}\otimes\bar{M}_{A}\otimes\emph{id}_{B})(\Psi^{\rho_{AB}}_{RAB}),
σ2RTV\displaystyle\sigma_{2}^{RTV} \ensurestackMath\stackon[1pt]=Δ(idRidAM¯B)(ΨRABρAB),and\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}(\emph{id}_{R}\otimes\emph{\text{id}}_{A}\otimes\bar{M}_{B})(\Psi^{\rho_{AB}}_{RAB}),\quad\text{and}
σ3RSTZ\displaystyle\sigma_{3}^{RSTZ} \ensurestackMath\stackon[1pt]=Δs,t,zρAB(Λ¯sAΛ¯tB)ρAB\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\sum_{s,t,z}\sqrt{\rho_{AB}}\left(\bar{\Lambda}^{A}_{s}\otimes\bar{\Lambda}^{B}_{t}\right)\sqrt{\rho_{AB}}
PZ|S,T(z|s,t)|ss||tt||zz|,\displaystyle\hskip 30.0pt\otimes P_{Z|S,T}(z|s,t)\outerproduct{s}{s}\otimes\outerproduct{t}{t}\otimes\outerproduct{z}{z},

for some orthonormal sets {|s}s𝒮,{|t}t𝒯\{\ket{s}\}_{s\in\mathcal{S}},\{\ket{t}\}_{t\in\mathcal{T}}, and {|z}z𝒵\{\ket{z}\}_{z\in\mathcal{Z}}, where ΨRABρAB\Psi^{\rho_{AB}}_{RAB} is a purification of ρAB\rho_{AB}. A quadruple (R1,R2,C1,C2)(R_{1},R_{2},C_{1},C_{2}) is achievable if there exists a finite field 𝔽p\mathbb{F}_{p}, for a prime pp, a pair of mappings fS:𝒮𝔽pf_{S}:\mathcal{S}\rightarrow\mathbb{F}_{p} and fT:𝒯𝔽pf_{T}:\mathcal{T}\rightarrow\mathbb{F}_{p}, and a stochastic mapping PZ|W:𝔽p𝒵P_{Z|W}:\mathbb{F}_{p}\rightarrow\mathcal{Z} such that

PZ|S,T(z|s,t)=PZ|W(z|fS(s)+fT(t)),s𝒮,t𝒯,z𝒵,P_{Z|S,T}(z|s,t)\!=\!P_{Z|W}(z|f_{S}(s)+f_{T}(t)),\forall s\!\in\!\mathcal{S},t\!\in\!\mathcal{T},z\!\in\!\mathcal{Z},

yielding U=fS(S)U=f_{S}(S), V=fT(T)V=f_{T}(T), and W=U+VW=U+V, and the following inequalities are satisfied:

R1I(U;R,B)σ1+I(W;V)σ3I(U;V)σ3,\displaystyle R_{1}\geq I(U;R,B)_{\sigma_{1}}+I(W;V)_{\sigma_{3}}-I(U;V)_{\sigma_{3}}, (5a)
R2I(V;R,A)σ2+I(W;U)σ3I(U;V)σ3,\displaystyle R_{2}\geq I(V;R,A)_{\sigma_{2}}+I(W;U)_{\sigma_{3}}-I(U;V)_{\sigma_{3}}, (5b)
R1+C1I(U;R,Z)σ3+I(W;V)σ3I(U;V)σ3,\displaystyle R_{1}+C_{1}\geq I(U;\!R,Z)_{\sigma_{3}}+I(W;V)_{\sigma_{3}}-I(U;V)_{\sigma_{3}}, (5c)
R2+C2I(V;R,Z)σ3+I(W;U)σ3I(U;V)σ3,\displaystyle R_{2}+C_{2}\geq I(V;\!R,Z)_{\sigma_{3}}+I(W;U)_{\sigma_{3}}-I(U;V)_{\sigma_{3}}, (5d)
R1+R2+C1+C2I(U,V;R,Z)σ3+I(W;U)σ3\displaystyle R_{1}+R_{2}+C_{1}+C_{2}\geq I(U,V;R,Z)_{\sigma_{3}}+I(W;U)_{\sigma_{3}}
+I(W;V)σ3I(U;V)σ3.\displaystyle\hskip 101.17755pt+I(W;V)_{\sigma_{3}}-I(U;V)_{\sigma_{3}}. (5e)
Proof.

A proof is provided in Section VI. ∎

Remark 1.

Note that the rate-region obtained in Theorem 6 of [24] using unstructured random code ensembles, contains the constraint R1+R2+C1+C2I(U,V;R,Z)σ3R_{1}+R_{2}+C_{1}+C_{2}\geq I(U,V;R,Z)_{\sigma_{3}}. Hence when

I(W;U)σ3\displaystyle I(W;U)_{\sigma_{3}} +I(W;V)σ3I(U;V)σ3\displaystyle+I(W;V)_{\sigma_{3}}-I(U;V)_{\sigma_{3}}
=2S(U+V)σ3S(U,V)σ3<0,\displaystyle=2S(U+V)_{\sigma_{3}}-S(U,V)_{\sigma_{3}}<0,

the above theorem gives a lower sum rate constraint. As a result, the rate-region above contains points that are not contained within the rate-region provided in [24]. To illustrate this fact further, consider the following example.

Remark 2.

In the above theorem, we restrict our attention to prime finite fields for ease of exposition. The results can be generalized to arbitrary finite fields in a straight-forward manner.

Example 1.

Suppose the composite state ρAB\rho_{AB} is described using one of the Bell states on AB\mathcal{H}_{A}\otimes\mathcal{H}_{B} as

ρAB=12(|00AB+|11AB)(00|AB+11|AB).\displaystyle\rho_{AB}=\cfrac{1}{2}\left(\ket{00}_{AB}+\ket{11}_{AB}\right)\left(\bra{00}_{AB}+\bra{11}_{AB}\right).

Since πA=TrBρAB\pi^{A}=\Tr_{B}{\rho^{AB}} and πB=TrAρAB\pi^{B}=\Tr_{A}{\rho^{AB}}, Alice and Bob would perceive each of their particles in maximally mixed states πA=IA2\pi^{A}=\frac{I_{A}}{2} and πB=IB2\pi^{B}=\frac{I_{B}}{2}, respectively. Upon receiving the quantum state, the two parties wish to independently measure their states, using identical POVMs M¯A\bar{M}_{A} and M¯B\bar{M}_{B}, given by M¯A\ensurestackMath\stackon[1pt]=Δ{Λ¯sA}s𝒮,M¯B\ensurestackMath\stackon[1pt]=Δ{Λ¯vB}t𝒯\bar{M}_{A}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\left\{\bar{\Lambda}_{s}^{A}\right\}_{s\in\mathcal{S}},\bar{M}_{B}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\left\{\bar{\Lambda}_{v}^{B}\right\}_{t\in\mathcal{T}}, where 𝒮=𝒯={0,1}\mathcal{S}=\mathcal{T}=\{0,1\}, and

Λ0A\displaystyle\Lambda_{0}^{A} =Λ0B\ensurestackMath\stackon[1pt]=Δ[0.95010.0826+i0.10890.0826i0.10890.0615],\displaystyle=\Lambda_{0}^{B}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\begin{bmatrix}0.9501&0.0826+i0.1089\\ 0.0826-i0.1089&0.0615\end{bmatrix},
Λ1A\displaystyle\Lambda_{1}^{A} =Λ1B\ensurestackMath\stackon[1pt]=Δ[0.04990.0826i0.10890.0826+i0.10890.9385].\displaystyle=\Lambda_{1}^{B}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\begin{bmatrix}0.0499&-0.0826-i0.1089\\ -0.0826+i0.1089&0.9385\end{bmatrix}.

Alice and Bob together with Charlie are trying to simulate the action of MAB\ensurestackMath\stackon[1pt]=Δ{ΓzAB}z𝒵M_{AB}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\left\{\Gamma_{z}^{AB}\right\}_{z\in\mathcal{Z}}, using the classical communication and common randomness as the resources available to them, where 𝒵={0,1}\mathcal{Z}=\{0,1\}, and

ΓzAB\ensurestackMath\stackon[1pt]=Δs{0,1}t{0,1}PZ|S,T(z|s,t)(ΛsAΛtB),\displaystyle\Gamma_{z}^{AB}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\sum_{s\in\{0,1\}}\sum_{t\in\{0,1\}}P_{Z|S,T}(z|s,t)\left(\Lambda_{s}^{A}\otimes\Lambda_{t}^{B}\right), (6)

for z{0,1},z\in\{0,1\}, and PZ|S,T(0|0,0)=PZ|S,T(0|1,1)=1PZ|S,T(0|0,1)=1PZ|S,T(0|1,0)=λP_{Z|S,T}(0|0,0)=P_{Z|S,T}(0|1,1)=1-P_{Z|S,T}(0|0,1)=1-P_{Z|S,T}(0|1,0)=\lambda, with λ(0,1)\lambda\in(0,1). Note that the above POVM MABM_{AB} admits a separable decomposition as defined in the statement of Theorem 1 with respect to the prime finite field 𝔽2\mathbb{F}_{2}, with U=SU=S and V=TV=T, and

PZ|W(0|0)=1PZ|W(0|1)=λ.P_{Z|W}(0|0)=1-P_{Z|W}(0|1)=\lambda.

Hence the above theorem can be employed. This gives

S(U+V)σ3\displaystyle S(U+V)_{\sigma_{3}} =0.5155,S(U)σ3=S(V)σ3=0.9999,\displaystyle=0.5155,\quad S(U)_{\sigma_{3}}=S(V)_{\sigma_{3}}=0.9999,
S(U,V)σ3\displaystyle S(U,V)_{\sigma_{3}} =1.5154,I(U,V)σ3=0.4844,\displaystyle=1.5154,\quad I(U,V)_{\sigma_{3}}=0.4844,

where σ3\sigma_{3} is as defined in the statement of Theorem 1. Since S(U)σ3S(U+V)σ3=S(V)σ3S(U+V)σ3=I(U,V)σ3,S(U)_{\sigma_{3}}-S(U+V)_{\sigma_{3}}=S(V)_{\sigma_{3}}-S(U+V)_{\sigma_{3}}=I(U,V)_{\sigma_{3}}, the constraints on R1R_{1}, R2R_{2}, R1+CR_{1}+C and R2+CR_{2}+C are the same as obtained in Theorem 6 of [24]. However, with 2S(U+V)σ3S(U,V)σ3=0.4844<02S(U+V)_{\sigma_{3}}-S(U,V)_{\sigma_{3}}=-0.4844<0, the constraint on R1+R2+C1+C2R_{1}+R_{2}+C_{1}+C_{2} in the above theorem (5e) is strictly weaker than the constraint obtained using random unstructured codes in Theorem 6 of [24]. Therefore, the rate-region obtained above using random structured codes in Theorem 1 is strictly larger than the rate-region in Theorem 6 of [24].

Example 2.

For the same state ρAB\rho_{AB} as in the above example, consider the following identical POVMs MA\ensurestackMath\stackon[1pt]=Δ{Λ¯sA}s𝒮M_{A}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\left\{\bar{\Lambda}_{s}^{A}\right\}_{s\in\mathcal{S}} and MB\ensurestackMath\stackon[1pt]=Δ{Λ¯tB}t𝒯M_{B}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\left\{\bar{\Lambda}_{t}^{B}\right\}_{t\in\mathcal{T}}, where 𝒮=𝒯={0,1}\mathcal{S}=\mathcal{T}=\{0,1\}, and

Λ0A\displaystyle\Lambda_{0}^{A} =Λ0B=[0.49740.0471+i0.49750.0471i0.49750.5026],\displaystyle=\Lambda_{0}^{B}=\begin{bmatrix}0.4974&0.0471+i0.4975\\ 0.0471-i0.4975&0.5026\end{bmatrix},
Λ1A\displaystyle\Lambda_{1}^{A} =Λ1B=[0.50260.0471i0.49750.0471+i0.49750.4974].\displaystyle=\Lambda_{1}^{B}=\begin{bmatrix}0.5026&-0.0471-i0.4975\\ -0.0471+i0.4975&0.4974\end{bmatrix}.

Let the joint measurement that Alice and Bob are trying to simulate be given by

ΓzAB\ensurestackMath\stackon[1pt]=Δs{0,1}t{0,1}PZ|S,T(z|s,t)(ΛsAΛtB),\displaystyle\Gamma_{z}^{AB}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\sum_{s\in\{0,1\}}\sum_{t\in\{0,1\}}P_{Z|S,T}(z|s,t)\left(\Lambda_{s}^{A}\otimes\Lambda_{t}^{B}\right), (7)

for z{0,1}z\in\{0,1\} where PZ|S,T:{0,1}[0,1]P_{Z|S,T}:\{0,1\}\rightarrow[0,1] is a conditional PMF on 𝒵×𝒮×𝒯\mathcal{Z}\times\mathcal{S}\times\mathcal{T} with PZ|S,T(0|0,0)=δ0(0,1)P_{Z|S,T}(0|0,0)=\delta_{0}\in(0,1) and PZ|S,T(0|0,1)=PZ|S,T(0|1,0)=PZ|S,T(0|1,1)=δ1(0,1)P_{Z|S,T}(0|0,1)=P_{Z|S,T}(0|1,0)=P_{Z|S,T}(0|1,1)=\delta_{1}\in(0,1). Note that PZ|S,TP_{Z|S,T} depends on the variables (s,t)(s,t) only through sts\lor t, the logical OR function. Now, we define the random variables UU and VV on the prime finite field 𝔽3\mathbb{F}_{3} with the identity mappings U=SU=S and V=TV=T, while noting that UU and VV take values in 𝔽3\mathbb{F}_{3} with P(U=2)=P(V=2)=0P(U=2)=P(V=2)=0. Now with W=U+VW=U+V, we identify the mapping PZ|WP_{Z|W} as

PZ|W(0|0)=δ0,PZ|W(0|1)=PZ|W(0|2)=δ1.\displaystyle P_{Z|W}(0|0)=\delta_{0},\quad P_{Z|W}(0|1)=P_{Z|W}(0|2)=\delta_{1}. (8)

For this identification, we obtain 2S(U+V)S(U,V)=0.9039<02S(U+V)-S(U,V)=-0.9039<0, which gives the constraint on R1+R2+C1+C2R_{1}+R_{2}+C_{1}+C_{2} in the above theorem (5e) strictly weaker than the corresponding constraint obtained using random unstructured codes in Theorem 6 of [24]. Since this is a biting constraint, the above rate-region is strictly larger than the former for this example.

Example 3.

Building upon Example 2, we explore more points in the POVM space such that the above theorem provides constraints (5e) that are strictly weaker than the corresponding constraint obtained in Theorem 6 of [24]. For this, we consider the same state ρAB,\rho_{AB}, as above and the following identical POVMs MA\ensurestackMath\stackon[1pt]=Δ{Λ¯sA}s𝒮M_{A}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\left\{\bar{\Lambda}_{s}^{A}\right\}_{s\in\mathcal{S}} and MB\ensurestackMath\stackon[1pt]=Δ{Λ¯tB}t𝒯M_{B}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\left\{\bar{\Lambda}_{t}^{B}\right\}_{t\in\mathcal{T}}, where 𝒮=𝒯={0,1}\mathcal{S}=\mathcal{T}=\{0,1\}, and

Λ0A=Λ0B=[θ1θ2+iθ3θ2iθ31θ1],Λ1A=Λ1B=IΛ0A\displaystyle\Lambda_{0}^{A}=\Lambda_{0}^{B}=\begin{bmatrix}\theta_{1}&\theta_{2}+i\theta_{3}\\ \theta_{2}-i\theta_{3}&1-\theta_{1}\end{bmatrix},\quad\Lambda_{1}^{A}=\Lambda_{1}^{B}=I-\Lambda_{0}^{A}

for θi[1,1]\theta_{i}\in[-1,1]111The above parametrization is only for illustrative purposes and do not constitute all the two dimensional POVMs.. Figure 2 illustrates the surface where 2S(U+V)=S(U,V)2S(U+V)=S(U,V) and therefore the region inside the surface has 2S(U+V)S(U,V)<02S(U+V)-S(U,V)<0, where the POVMs obtained provides the constraint on R1+R2+C1+C2R_{1}+R_{2}+C_{1}+C_{2} in the above theorem (5e) strictly weaker than the corresponding constraint obtained in Theorem 6 of [24].

Refer to caption
Figure 2: Shown above is a (θ1,θ2,θ3)(\theta_{1},\theta_{2},\theta_{3})-surface with POVMs satisfying 2S(U+V)=S(U,V)2S(U+V)=S(U,V). Although the surface is symmetric in θ3\theta_{3}, but for the ease of illustration only the upper half of the surface is shown.
Remark 3.

Note that for POVMs contained in the above (θ1,θ2,θ3)(\theta_{1},\theta_{2},\theta_{3})-surface of Example 2, the sum rate constraint R1+R2+C1+C2R_{1}+R_{2}+C_{1}+C_{2} is strictly weaker than the corresponding constraint in [24, Theorem 6], and vice-versa outside. One can employ a strategy based on superposition and successive encoding that combines the two coding techniques to yield a unified rate-region.

III.2 Covering Lemma with Change of Measure for Pairwise-Independent Ensemble

The proof of the theorem is based on a construction of algebraic-structured POVM ensemble where the elements are only pairwise independent and not mutually independent. To analyze these POVMs we retreat back to first principles and develop a new one-shot Covering Lemma based on a change of measure technique and a second order analysis. This lemma, which can be of independent interest, is one of the main contributions of this work.

Lemma 2 (Covering Lemma).

Let {λx,σx}x𝒳\{\lambda_{x},\sigma_{x}\}_{x\in\mathcal{X}} be an ensemble, with σx𝒟()\sigma_{x}\in\mathcal{D}(\mathcal{H}) for all x𝒳x\in\mathcal{X}, 𝒳\mathcal{X} being a finite set, and σ=x𝒳λxσx\sigma=\sum_{x\in\mathcal{X}}\lambda_{x}\sigma_{x}. Further, suppose we are given a total subspace projector Π\Pi and a collection of codeword subspace projectors {Πx}x𝒳\{\Pi_{x}\}_{x\in\mathcal{X}} which satisfy the following hypotheses

Tr(Πσx)\displaystyle\Tr{\Pi\sigma_{x}} 1ϵ,\displaystyle\geq 1-\epsilon, (9a)
Tr(Πxσx)\displaystyle\Tr{\Pi_{x}\sigma_{x}} 1ϵ,\displaystyle\geq 1-\epsilon, (9b)
Πσ12\displaystyle\|\Pi\sqrt{\sigma}\|^{2}_{1} D,\displaystyle\leq D, (9c)
ΠxσxΠx\displaystyle\Pi_{x}\sigma_{x}\Pi_{x} 1dΠx,and\displaystyle\leq\frac{1}{d}\Pi_{x},\quad\text{and} (9d)
ΠxσxΠx\displaystyle\Pi_{x}\sigma_{x}\Pi_{x} σx.\displaystyle\leq\sigma_{x}. (9e)

for some ϵ(0,1)\epsilon\in(0,1) and d<Dd<D. Let MM be a finite non-negative integer. Additionally, assume that there exists some set 𝒳¯\bar{\mathcal{X}} containing 𝒳\mathcal{X}, with σx\ensurestackMath\stackon[1pt]=Δ0\sigma_{x}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}0 (null operator) and λx\ensurestackMath\stackon[1pt]=Δ0\lambda_{x}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}0 for x𝒳¯\𝒳x\in\bar{\mathcal{X}}\backslash\mathcal{X}. Suppose {μx¯}x¯𝒳¯\{\mu_{\bar{x}}\}_{\bar{x}\in\bar{\mathcal{X}}} be any distribution on the set 𝒳¯\bar{\mathcal{X}} such that the distribution is {λx}x𝒳\{\lambda_{x}\}_{{x}\in\mathcal{X}} is absolutely continuous with respect to the distribution {μx¯}x¯𝒳¯\{\mu_{\bar{x}}\}_{\bar{x}\in\bar{\mathcal{X}}}. Further, assume that λx/μxκ\lambda_{x}/\mu_{x}\leq\kappa for all x𝒳.x\in\mathcal{X}. Let a random covering code \ensurestackMath\stackon[1pt]=Δ{Cm}m[1,M]\mathbbm{C}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\{C_{m}\}_{m\in[1,M]} be defined as a collection of codewords CmC_{m} that are chosen pairwise independently according to the distribution {μx¯}x¯𝒳¯\{\mu_{\bar{x}}\}_{\bar{x}\in\bar{\mathcal{X}}}. Then we have

𝔼[x𝒳¯λxσx1Mm=1MλCmμCmσCm1]κDMd+2δ(ϵ),\displaystyle\mathbb{E}_{\mathbbm{C}}\left[\Big{\|}\sum_{x\in\bar{\mathcal{X}}}\lambda_{x}\sigma_{x}-\frac{1}{M}\sum_{m=1}^{M}\frac{\lambda_{C_{m}}}{\mu_{C_{m}}}\sigma_{C_{m}}\Big{\|}_{1}\right]\!\leq\!\sqrt{\frac{\kappa D}{Md}}+2\delta(\epsilon), (10)

where δ(ϵ)=4ϵ\delta(\epsilon)=4\sqrt{\epsilon}. Futhermore, for σ~x\tilde{\sigma}_{x} defined as σ~x\ensurestackMath\stackon[1pt]=ΔΠΠxσxΠxΠ\tilde{\sigma}_{x}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\Pi\Pi_{x}{\sigma}_{x}\Pi_{x}\Pi, we have

𝔼[x𝒳¯λxσ~x1Mm=1MλCmμCmσ~Cm1]κDMd.\displaystyle\mathbb{E}_{\mathbbm{C}}\left[\Big{\|}\sum_{x\in\bar{\mathcal{X}}}\lambda_{x}\tilde{\sigma}_{x}-\frac{1}{M}\sum_{m=1}^{M}\frac{\lambda_{C_{m}}}{\mu_{C_{m}}}\tilde{\sigma}_{C_{m}}\Big{\|}_{1}\right]\leq\sqrt{\frac{\kappa D}{Md}}. (11)
Proof.

The proof is provided in Appendix A.1

IV Useful Lemmas

In this section we present a few lemmas which will be used extensively in the sequel.

Definition 5 (Pruning Operators).

Consider an operator A0A\geq 0 acting on Hilbert space A.\mathcal{H}_{A}. We say that a projector PP prunes AA with respect to Identity IAI_{A} on A,\mathcal{H}_{A}, if PP is a projector on to the non-negative eigenspace of IAAI_{A}-A.

IV.1 Pruning Trace Inequality

Lemma 3.

Consider a random operator X0X\geq 0 acting on a Hilbert space A.\mathcal{H}_{A}. Let PP be a pruning operator for XX with respect to IAI_{A}, as in Definition 5. Then we have

𝔼[Tr(IAP)]𝔼[Tr(X)].\mathbb{E}[\Tr{I_{A}-P}]\leq\mathbb{E}[\Tr{X}].
Proof.

The proof follows by noting that Tr(IAP)Tr(X)\Tr{I_{A}-P}\leq\Tr{X}. ∎

Remark 4.

To demonstrate the significance of this inequality, we compare it with the popular Operator Markov Inequality [39]. We know from Operator Markov inequality

(XIA)𝔼[Tr(X)].\displaystyle\mathbb{P}\left(X\nleq I_{A}\right)\leq\mathbb{E}[\Tr{X}].

One can observe that 𝟙{XIA}Tr(IAP)\mathbbm{1}_{\{X\nleq I_{A}\}}\leq\Tr{I_{A}-P}. Taking expectation, we obtain

(XIA)𝔼[Tr(IAP)].\displaystyle\mathbb{P}\left(X\nleq I_{A}\right)\leq\mathbb{E}[\Tr{I_{A}-P}].

Moreover, one can also note that Tr(IAP)Tr(X)\Tr{I_{A}-P}\leq\Tr{X}, and expectation gives

𝔼[Tr(IAP)]𝔼[Tr(X)].\displaystyle\mathbb{E}[\Tr{I_{A}-P}]\leq\mathbb{E}[\Tr{X}].

Hence we conclude that the new inequality is indeed tighter than the operator Markov inequality.

Lemma 4.

(Pruning Trace Inequality) Consider the above random operator X0X\geq 0 acting on a Hilbert space A.\mathcal{H}_{A}. Further, suppose 𝔼[X](1η)IA\mathbb{E}[X]\leq(1-\eta){I_{A}} for η(0,1)\eta\in(0,1). Let PP be a pruning operator for XX with respect to IAI_{A}, as in Definition 5. Then, we have

𝔼[Tr(IAP)]1η𝔼[X𝔼[X]1].\displaystyle\mathbb{E}[\Tr{I_{A}-P}]\leq\frac{1}{\eta}\mathbb{E}\left[\|X-\mathbb{E}[X]\|_{1}\right].
Proof.

The proof is provided in Appendix A.2

V Point-to-point Measurement Compression using Structured Random POVMs

Before presenting the proof of Theorem 1, as a pedagogical first step, we consider the measurement compression problem in the point-to-point setup. This problem was addressed in [2], where the performance limits were derived using unstructured random POVM ensembles. Here, we redrive the performance limit using random algebraic structured POVM ensembles. Since the algebraic structured codes can only induce a uniform distribution, we consider a collection of cosets of a random linear code for this task. The problem setup is described as follows. An agent (Alice) performs a measurement MM on a quantum state ρ\rho, and sends a set of classical bits to a receiver (Bob). Bob has access to additional private randomness, and he is allowed to use this additional resource to perform any stochastic mapping of the received classical bits. The overall effect on the quantum state can be assumed to be a measurement which is a concatenation of the POVM Alice performs and the stochastic map Bob implements. This problem serves as a building block toward the proof of Theorem 1. Formally, the problem is stated as follows.

V.1 Problem Formulation and Main Result

Definition 6.

For a given finite set 𝒵\mathcal{Z}, and a Hilbert space \mathcal{H}, a measurement simulation protocol with parameters (n,Θ,N)(n,\Theta,N) is characterized by
1) a collection of codes 𝒞(μ)𝒲n\mathcal{C}^{(\mu)}\subseteq\mathcal{W}^{n}, for μ[1,N]\mu\in[1,N], such that |𝒞(μ)|Θ|\mathcal{C}^{(\mu)}|\leq\Theta, and 𝒲\mathcal{W}, a finite set, is called the code alphabet,
2) a collection of Alice’s sub-POVMs M~(μ),μ[1,N]\tilde{M}^{(\mu)},\mu\in[1,N] each acting on n\mathcal{H}^{\otimes n} and with outcomes in 𝒞(μ)\mathcal{C}^{(\mu)}.
3) a collection of Bob’s classical stochastic maps P(μ)(zn|wn)P^{(\mu)}(z^{n}|w^{n}) for all wn𝒞(μ)w^{n}\in\mathcal{C}^{(\mu)}, zn𝒵nz^{n}\in\mathcal{Z}^{n} and μ[1,N]\mu\in[1,N].
The overall sub-POVM of this protocol, given by M~\tilde{M}, is characterized by the following operators:

Λ~zn\ensurestackMath\stackon[1pt]=Δ1Nμ=1Nwn𝒞(μ)P(μ)(zn|wn)Λwn(μ),zn𝒵n,\tilde{\Lambda}_{z^{n}}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\frac{1}{N}\sum_{\mu=1}^{N}\sum_{w^{n}\in\mathcal{C}^{(\mu)}}P^{(\mu)}(z^{n}|w^{n})~{}\Lambda^{(\mu)}_{w^{n}},\quad\forall z^{n}\in\mathcal{Z}^{n}, (12)

where {Λwn(μ):wn𝒞(μ)}\{\Lambda^{(\mu)}_{w^{n}}:w^{n}\in\mathcal{C}^{(\mu)}\} is the set of operators corresponding to the sub-POVM M~(μ)\tilde{M}^{(\mu)}. Let 𝒞(μ)(i)\mathcal{C}^{(\mu)}(i) denote the iith codeword of 𝒞(μ)\mathcal{C}^{(\mu)}.

In the above definition, Θ\Theta characterizes the amount of classical bits communicated from Alice to Bob, and the amount of common randomness is determined by NN, with μ\mu being the common randomness bits distributed among the parties. The classical stochastic mappings induced by P(μ)P^{(\mu)} represents the action of Bob on the received classical bits. In building the code, we use the Unionized Coset Code (UCC) [37] defined below. These codes involve two layers of codes (i) a coarse code and (ii) a fine code. The coarse code is a coset of the linear code and the fine code is the union of several cosets of the linear code.

For a fixed k×nk\times n matrix G𝔽pk×nG\in\mathbb{F}_{p}^{k\times n} with knk\leq n, and pp being a prime number, and a 1×n1\times n vector B𝔽pnB\in\mathbb{F}_{p}^{n}, define the coset code as

(G,B)\ensurestackMath\stackon[1pt]=Δ{xn:xn=akG+B, for some ak𝔽pk}.\displaystyle\mathbb{C}(G,B)\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\{x^{n}:x^{n}=a^{k}G+B,\mbox{ for some }a^{k}\in\mathbb{F}_{p}^{k}\}. (13)

In other words, (G,B)\mathbb{C}(G,B) is a shift of the row space of the matrix GG. The row space of GG is a linear code. If the rank of GG is kk, then there are pkp^{k} codewords in the coset code.

Definition 7.

An (n,k,l,p)(n,k,l,p) UCC is characterized by a pair (G,h)(G,h) consisting of a k×nk\times n matrix G𝔽pk×nG\in\mathbb{F}_{p}^{k\times n}, and a mapping h:𝔽pl𝔽pnh:\mathbb{F}_{p}^{l}\rightarrow\mathbb{F}_{p}^{n}, and the code is the following union: m𝔽pl(G,h(m))\bigcup_{m\in\mathbb{F}_{p}^{l}}\mathbb{C}(G,h(m)), where (,)\mathbb{C}(\cdot,\cdot) is defined in (13).

Definition 8.

Given a finite set 𝒵\mathcal{Z}, and a Hilbert space \mathcal{H}, an (n,Θ,κ,N,p)(n,\Theta,\kappa,N,p) UCC-based measurement simulation protocol is a pair of (n,Θ,N)(n,\Theta,N) measurement simulation protocol and a collection of NN UCCs with parameters (n,k,l,p)(n,k,l,p) characterized by {(G,h(μ))}μ[1,N]\{(G,h^{(\mu)})\}_{\mu\in[1,N]} such that (i) the code alphabet of the protocol 𝒲𝔽p\mathcal{W}\subseteq\mathbb{F}_{p} (with suitable relabeling), (ii) κ=pk\kappa=p^{k}, Θ=pl\Theta=p^{l}, and (iii) for all m𝔽plm\in\mathbb{F}_{p}^{l}, we have 𝒞(μ)(m){akG+h(μ)(m):ak𝔽pk}\mathcal{C}^{(\mu)}(m)\in\{a^{k}G+h^{(\mu)}(m):a^{k}\in\mathbb{F}_{p}^{k}\}.

Definition 9.

The UCC grand ensemble is the ensemble of NN UCCs where GG, and {h(μ)}μ[1,N]\{h^{(\mu)}\}_{\mu\in[1,N]} are chosen randomly, independently and uniformly, where the latter is chosen from the set of all mappings with replacement.

Definition 10.

Given a POVM MM acting on \mathcal{H}, and a density operator ρ𝒟()\rho\in\mathcal{D}(\mathcal{H}), a tuple (R,R1,C,p)(R,R_{1},C,p) is said to be achievable using the grand UCC ensemble, if for all ϵ>0\epsilon>0 and for all sufficiently large nn, there exists an ensemble of UCC-based measurement simulation protocols with parameters (n,Θ,κ,N,p)(n,\Theta,\kappa,N,p) (based on the UCC grand ensemble) such that their overall sub-POVM M~\tilde{M} is ϵ\epsilon-faithful to MnM^{\otimes n} with respect to ρn\rho^{\otimes n} in the expected sense:

𝔼\displaystyle\mathbb{E}\! [znρn(ΛznΛ~zn)ρn+Tr{IznΛ~zn}]ϵ,\displaystyle\left[\!\sum_{z^{n}}\!\left\|\!\sqrt{\rho^{\otimes n}}(\Lambda_{z^{n}}\!-\!\tilde{\Lambda}_{z^{n}})\sqrt{\rho^{\otimes n}}\right\|\!+\!\Tr\{I-\sum_{z^{n}}\tilde{\Lambda}_{z^{n}}\}\right]\leq\epsilon,

where the expectation is with respect to the ensemble, and

1nlog2ΘR+ϵ,|1nlogκR1|ϵ,;1nlog2NC+ϵ.\displaystyle\frac{1}{n}\log_{2}\Theta\leq R+\epsilon,\;\left|\frac{1}{n}\log\kappa-R_{1}\right|\leq\epsilon,;\;\frac{1}{n}\log_{2}N\leq C+\epsilon.

Define UCC\mathscr{R}_{\mbox{UCC}} as UCC\ensurestackMath\stackon[1pt]=Δ{(R,R1,C,p):(R,R1,C,p)\mathscr{R}_{\mbox{UCC}}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\{(R,R_{1},C,p):(R,R_{1},C,p) is achievable using the UCC grand ensemble}.

Remark 5.

The appearance of the modulus in the second constraint needs justification. Note that RR is the rate of transmission of information from Alice to Bob and CC is the rate of the common information shared between them. So if (R,R1,C,p)(R,R_{1},C,p) is achievable, then it is clear that any (R~,C~)(\tilde{R},\tilde{C}) is also achievable if R~R\tilde{R}\geq R and C~C\tilde{C}\geq C. However R1R_{1} is a parameter of the UCC grand ensemble, and there is no natural order on R1R_{1}, i.e., it does not naturally follows that (R,R~1,C,p)(R,\tilde{R}_{1},C,p) is achievable for all R~1R1\tilde{R}_{1}\geq R_{1}.

The following theorem characterizes the achievable rate region which characterizes the asymptotic performance of the UCC grand ensemble.

Theorem 2.

For any density operator ρ𝒟()\rho\in\mathcal{D}(\mathcal{H}) and any POVM M\ensurestackMath\stackon[1pt]=Δ{Λz}z𝒵{M}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\{{\Lambda}_{z}\}_{z\in\mathcal{Z}} acting on the Hilbert space \mathcal{H}, a tuple (R,R1,C,p)(R,R_{1},C,p) is achievable using the UCC grand ensemble, i.e., (R,R1,C,p)UCC(R,R_{1},C,p)\in\mathscr{R}_{\mbox{UCC}} if there exist a POVM M¯\ensurestackMath\stackon[1pt]=Δ{Λ¯w}w𝒲\bar{M}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\{\bar{\Lambda}_{w}\}_{w\in\mathcal{W}}, with |𝒲|p|\mathcal{W}|\leq p, and a stochastic map PZ|W:𝒲𝒵P_{Z|W}:\mathcal{W}\rightarrow\mathcal{Z} such that

Λz=w𝒲PZ|W(z|w)Λ¯w,z𝒵,\Lambda_{z}=\sum_{w\in\mathcal{W}}P_{Z|W}(z|w)\bar{\Lambda}_{w},\quad\forall z\in\mathcal{Z},

and

R1+R\displaystyle R_{1}+R I(W;R)σS(W)σ+log(p),\displaystyle\geq I(W;R)_{\sigma}-S(W)_{\sigma}+\log{p}, (14)
R1+R+C\displaystyle R_{1}+R+C I(W;RZ)σS(W)σ+log(p),\displaystyle\geq I(W;RZ)_{\sigma}-S(W)_{\sigma}+\log{p}, (15)
0R1\displaystyle 0\leq R_{1} log(p)S(W)σ,\displaystyle\leq\log{p}-S(W)_{\sigma}, (16)
C\displaystyle C 0,\displaystyle\geq 0, (17)

where σRWZ\ensurestackMath\stackon[1pt]=Δw,zρΛ¯wρPZ|W(z|w)|ww||zz|,\sigma^{RWZ}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\sum_{w,z}\sqrt{\rho}\bar{\Lambda}_{w}\sqrt{\rho}\otimes P_{Z|W}(z|w)\outerproduct{w}{w}\otimes\outerproduct{z}{z}, for some orthogonal sets {|w}w𝒲\{\ket{w}\}_{w\in\mathcal{W}} and {|z}z𝒵.\{\ket{z}\}_{z\in\mathcal{Z}}.

Remark 6.

By choosing R1=log(p)S(W)σR_{1}=\log{p}-S(W)_{\sigma}, we recover the rate region of Wilde et. al [3, Theorem 9].

V.2 Proof of Theorem 2 Using UCC Code Ensemble

As stated earlier, the main objective of proving this theorem is to build a framework for the main theorem of the paper (Theorem 1). In doing so, we observe that the structured POVMs constructed below are only pairwise independent. Since the results in [24] are based on the assumption that approximating POVMs are all mutually independent, the proof below becomes significantly different from [24].

Suppose there exist a POVM M¯\ensurestackMath\stackon[1pt]=Δ{Λ¯w}w𝒲\bar{M}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\{\bar{\Lambda}_{w}\}_{w\in\mathcal{W}} and a stochastic map PZ|W:𝒲𝒵P_{Z|W}:\mathcal{W}\rightarrow\mathcal{Z}, such that M\ensurestackMath\stackon[1pt]=Δ{Λz}z𝒵M\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\{\Lambda_{z}\}_{z\in\mathcal{Z}} can be decomposed as

Λz\ensurestackMath\stackon[1pt]=Δw𝒲PZ|W(z|w)Λ¯w,z𝒵.\displaystyle\Lambda_{z}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\sum_{w\in\mathcal{W}}P_{Z|W}(z|w)\bar{\Lambda}_{w},\quad\forall z\in\mathcal{Z}. (18)

We generate the canonical ensemble corresponding to M¯\bar{M} as

λw\displaystyle\lambda_{w} \ensurestackMath\stackon[1pt]=ΔTr{Λ¯wρ},ρ^w\ensurestackMath\stackon[1pt]=Δ1λwρΛ¯wρ.\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\Tr\{\bar{\Lambda}_{w}\rho\},\quad\hat{\rho}_{w}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\frac{1}{\lambda_{w}}\sqrt{\rho}\bar{\Lambda}_{w}\sqrt{\rho}. (19)

Let 𝒯δ(n)(W)\mathcal{T}_{\delta}^{(n)}(W) denote a δ\delta-typical set associated with the probability distribution induced by {λw}w𝒲,\{\lambda_{w}\}_{w\in\mathcal{W}}, corresponding to a random variable WW. Let Πρ\Pi_{\rho} denote the δ\delta-typical projector (as in [7, Def. 15.1.3]) corresponding to the density operator ρ\ensurestackMath\stackon[1pt]=Δw𝒲λwρ^w\rho\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\sum_{w\in\mathcal{W}}\lambda_{w}\hat{\rho}_{w}, and Πwn\Pi_{w^{n}} denote the strong conditional typical projector (as in [7, Def. 15.2.4]) corresponding to the canonical ensemble {λw,ρ^w}w𝒲\{\lambda_{w},\hat{\rho}_{w}\}_{w\in\mathcal{W}}. For each wn𝒯δ(n)(W)w^{n}\in\mathcal{T}_{\delta}^{(n)}(W), define

ρ~wn\ensurestackMath\stackon[1pt]=ΔΠρΠwnρ^wnΠwnΠρ,\tilde{\rho}_{w^{n}}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\Pi_{\rho}\Pi_{w^{n}}\hat{\rho}_{w^{n}}\Pi_{w^{n}}\Pi_{\rho}, (20)

and ρ~wn=0,\tilde{\rho}_{w^{n}}=0, for wn𝒯δ(n)(W)w^{n}\notin\mathcal{T}_{\delta}^{(n)}(W), with ρ^wn\ensurestackMath\stackon[1pt]=Δiρ^wi\hat{\rho}_{w^{n}}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\bigotimes_{i}\hat{\rho}_{w_{i}}.

V.2.1 Construction of Structured POVMs

We now construct random structured POVM elements. Fix a block length n>0n>0, a positive integer N,N, and a finite field 𝔽p\mathbb{F}_{p} with p|𝒲|p\geq|\mathcal{W}|. Without loss of generality, we assume 𝒲\ensurestackMath\stackon[1pt]=Δ{0,1,,|𝒲|1}\mathcal{W}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\{0,1,\cdots,|\mathcal{W}|-1\}. Furthermore, we assume λw=0\lambda_{w}=0 for all |𝒲|1<w<p|\mathcal{W}|-1<w<p. From now on, we assume that WW takes values in 𝔽p\mathbb{F}_{p} with this distribution. Let μ[1,N]\mu\in[1,N] denote the common randomness shared between the encoder and decoder. In building the code, we use the UCCs [37] as defined in Definition 7 .

For every μ[1,N]\mu\in[1,N], consider a UCC (G,h(μ))(G,h^{(\mu)}) with parameters (n,k,l,p)(n,k,l,p). For each μ\mu, the generator matrix GG along with the function h(μ)h^{(\mu)} generates pk+lp^{k+l} codewords. Each of these codewords are characterized by a triple (a,i,μ)(a,i,\mu), where a𝔽pka\in\mathbb{F}^{k}_{p} and i𝔽pli\in\mathbb{F}^{l}_{p} correspond to the coarse code and the coset indices, respectively. Let Wn,(μ)(a,i)W^{n,(\mu)}(a,i) denote the codewords associated with the encoder (Alice), generated using the above procedure, where

Wn,(μ)(a,i)=aG+h(μ)(i).\displaystyle W^{n,(\mu)}(a,i)=aG+h^{(\mu)}(i). (21)

Now, construct the operators

A¯wn(μ)\displaystyle\bar{A}^{(\mu)}_{w^{n}} \ensurestackMath\stackon[1pt]=Δαwn(ρn1ρ~wnρn1)\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\alpha_{w^{n}}\bigg{(}\sqrt{\rho^{\otimes n}}^{-1}\tilde{\rho}_{w^{n}}\sqrt{\rho^{\otimes n}}^{-1}\bigg{)}\quad
αwn\displaystyle\quad\alpha_{w^{n}} \ensurestackMath\stackon[1pt]=Δ1(1+η)pnλwnpk+l,\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\frac{1}{(1+\eta)}\frac{p^{n}\lambda_{w^{n}}}{p^{k+l}}, (22)

with η(0,1)\eta\in(0,1) being a parameter to be determined. Note that, following the definition of ρ~wn\tilde{\rho}_{w^{n}}, we have A¯wn(μ)=0\bar{A}^{(\mu)}_{w^{n}}=0 for wn𝒯δ(n)(W).w^{n}\notin\mathcal{T}_{\delta}^{(n)}(W). Having constructed the operators A¯wn(μ)\bar{A}^{(\mu)}_{w^{n}}, we normalize these operators, so that they constitute a valid sub-POVM. To do so, we define

Σ(μ)\ensurestackMath\stackon[1pt]=Δwnγwn(μ)A¯wn(μ),γwn(μ)\displaystyle\Sigma^{(\mu)}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\sum_{w^{n}}\gamma_{w^{n}}^{(\mu)}\bar{A}^{(\mu)}_{w^{n}},\;\gamma_{w^{n}}^{(\mu)} \ensurestackMath\stackon[1pt]=Δ|{(a,i):Wn,(μ)(a,i)=wn}|.\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}|\{(a,i):W^{n,(\mu)}(a,i)=w^{n}\}|.

Now, we define Πμ\Pi^{\mu} as the pruning operator for Σ(μ)\Sigma^{(\mu)} with respect to Πρ\Pi_{\rho} using Definition 5. Note that, the pruning operator Πμ\Pi^{\mu} depends on the pair (G,h(μ))(G,h^{(\mu)}). For ease of analysis, the subspace of Πμ\Pi^{\mu} is restricted to Πρ\Pi_{\rho} and hence Πμ\Pi^{\mu} is a projector onto a subspace of Πρ\Pi_{\rho}. Using these pruning operators, for each μ[1,N]\mu\in[1,N], construct the sub-POVM M~(n,μ)\tilde{M}^{(n,\mu)} as

M~(n,μ)\displaystyle\tilde{M}^{(n,\mu)} \ensurestackMath\stackon[1pt]=Δ{γwn(μ)Awn(μ)}wn𝒲n,\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\{\gamma_{w^{n}}^{(\mu)}A^{(\mu)}_{w^{n}}\}_{w^{n}\in\mathcal{W}^{n}},\quad (23)

where Awn(μ)\ensurestackMath\stackon[1pt]=ΔΠμA¯wn(μ)ΠμA^{(\mu)}_{w^{n}}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\Pi^{\mu}\bar{A}^{(\mu)}_{w^{n}}\Pi^{\mu}. Further, using Πμ\Pi^{\mu} we have wnγwn(μ)Awn(μ)=ΠμΣ(μ)ΠμΠρI,\sum_{w^{n}}\gamma_{w^{n}}^{(\mu)}A^{(\mu)}_{w^{n}}=\Pi^{\mu}\Sigma^{(\mu)}\Pi^{\mu}\leq\Pi_{\rho}\leq I, and thus M~(n,μ)\tilde{M}^{(n,\mu)} is a valid sub-POVM for all μ[1,N]\mu\in[1,N]. Moreover, the collection M~(n,μ)\tilde{M}^{(n,\mu)} is completed using the operators Iwn𝒲nγwn(μ)Awn(μ)I-\sum_{w^{n}\in\mathcal{W}^{n}}\gamma_{w^{n}}^{(\mu)}A^{(\mu)}_{w^{n}}.

V.2.2 Binning of POVMs

The next step is to bin the above constructed sub-POVMs. Since, UCC is a union of several cosets, we associate a bin to each coset, and hence place all the codewords of a coset in the same bin. For each i𝔽pli\in\mathbb{F}_{p}^{l}, let (μ)(i)\ensurestackMath\stackon[1pt]=Δ(G,h(μ)(i))\mathcal{B}^{(\mu)}(i)\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\mathbb{C}(G,h^{(\mu)}(i)) denote the iith bin. Further, for all i𝔽pli\in\mathbb{F}_{p}^{l}, we define

ΓiA,(μ)\displaystyle\Gamma^{A,(\mu)}_{i} \ensurestackMath\stackon[1pt]=Δwn𝒲na𝔽pkAwn(μ)𝟙{aG+h(μ)(i)=wn}.\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\sum_{w^{n}\in\mathcal{W}^{n}}\sum_{a\in\mathbb{F}_{p}^{k}}A^{(\mu)}_{w^{n}}\mathbbm{1}_{\{aG+h^{(\mu)}(i)=w^{n}\}}.

Using these operators, we form the following collection:

M(n,μ)\ensurestackMath\stackon[1pt]=Δ{ΓiA,(μ)}i𝔽pl.\displaystyle M^{(n,\mu)}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\{\Gamma^{A,(\mu)}_{i}\}_{i\in\mathbb{F}_{p}^{l}}.

Note that if the collection M~(n,μ)\tilde{M}^{(n,\mu)} is a sub-POVM for each μ[1,N]\mu\in[1,N], then so is the collection M(n,μ)M^{(n,\mu)}, which is due to the relation i𝔽plΓiA,(μ)=wn𝒲nγwn(μ)Awn(μ)I.\sum_{i\in\mathbb{F}_{p}^{l}}\Gamma^{A,(\mu)}_{i}=\sum_{w^{n}\in\mathcal{W}^{n}}\gamma_{w^{n}}^{(\mu)}A^{(\mu)}_{w^{n}}\leq I. To complete M(n,μ)M^{(n,\mu)}, we define Γ0A,(μ)\Gamma^{A,(\mu)}_{0} as Γ0A,(μ)=IiΓiA,(μ)\Gamma^{A,(\mu)}_{0}=I-\sum_{i}\Gamma^{A,(\mu)}_{i} 222Note that Γ0A,(μ)=IiΓiA,(μ)=Iwn𝒯δ(n)(W)γwn(μ)Awn(μ)\Gamma^{A,(\mu)}_{0}=I-\sum_{i}\Gamma^{A,(\mu)}_{i}=I-\sum_{w^{n}\in\mathcal{T}_{\delta}^{(n)}(W)}\gamma_{w^{n}}^{(\mu)}A^{(\mu)}_{w^{n}}.. Now, we intend to use the completions [M(n,μ)][M^{(n,\mu)}] as the POVM for the encoder.

V.2.3 Decoder mapping

We create a decoder which, on receiving the classical bits from the encoder, generates a sequence Wn𝔽pnW^{n}\in\mathbb{F}^{n}_{p} as follows. The decoder first creates a set Di(μ)D^{(\mu)}_{i} and a function F(μ)F^{(\mu)} defined as

Di(μ)\displaystyle D^{(\mu)}_{i} \ensurestackMath\stackon[1pt]=Δ{a~𝔽pk:a~G+h(μ)(i)𝒯δ(n)(W)} and\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\big{\{}\tilde{a}\in\mathbb{F}_{p}^{k}:\tilde{a}G+h^{(\mu)}(i)\in\mathcal{T}_{\delta}^{(n)}(W)\big{\}}\quad\text{ and }
F(μ)(i)\displaystyle F^{(\mu)}(i) \ensurestackMath\stackon[1pt]=Δ{a~G+h(μ)(i) if Di(μ){a~}w0n otherwise ,\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\begin{cases}\tilde{a}G+h^{(\mu)}(i)&\quad\text{ if }D^{(\mu)}_{i}\equiv\{\tilde{a}\}\\ w^{n}_{0}&\quad\text{ otherwise },\end{cases} (24)

where w0nw_{0}^{n} is an arbitrary sequence in 𝔽pn\𝒯δ(n)(W)\mathbb{F}_{p}^{n}\backslash\mathcal{T}_{\delta}^{(n)}(W). Further, F(μ)(i)=w0nF^{(\mu)}(i)=w_{0}^{n} for i=0i=0. Given this and the stochastic processing PZ|WP_{Z|W}, we obtain the approximating sub-POVM M^(n)\hat{M}^{(n)} with the following operators.

Λ^zn\ensurestackMath\stackon[1pt]=Δ1Nμ=1Nwn𝔽pni:F(μ)(i)=wnΓiA,(μ)PZ|Wn(zn|wn),zn𝒵n.\displaystyle\hat{\Lambda}_{z^{n}}\!\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\!\frac{1}{N}\sum_{\mu=1}^{N}\sum_{w^{n}\in\mathbb{F}_{p}^{n}}\sum_{i:F^{(\mu)}(i)=w^{n}}\Gamma^{A,(\mu)}_{i}P^{n}_{Z|W}(z^{n}|w^{n}),~{}\forall z^{n}\in\mathcal{Z}^{n}.

The generator matrix GG and the function h(μ)h^{(\mu)} are chosen randomly uniformly and independently.

V.2.4 Trace Distance

In what follows, we show that M^(n)\hat{M}^{(n)} is ϵ\epsilon-faithful to Mn{M}^{\otimes n} with respect to ρn\rho^{\otimes n} (according to Definition 1), where ϵ>0\epsilon>0 can be made arbitrarily small. More precisely, using (18), we show that, 𝔼[K]ϵ,\mathbb{E}[K]\leq\epsilon, where

K\displaystyle{K} \ensurestackMath\stackon[1pt]=ΔznwnρnΛ¯wnρnPZ|Wn(zn|wn)\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\sum_{z^{n}}\left\|\sum_{w^{n}}\sqrt{\rho^{\otimes n}}\bar{\Lambda}_{w^{n}}\sqrt{\rho^{\otimes n}}P^{n}_{Z|W}(z^{n}|w^{n})\right.
ρnΛ^znρn1,\displaystyle\hskip 115.63243pt\left.-\sqrt{\rho^{\otimes n}}\hat{\Lambda}_{z^{n}}\sqrt{\rho^{\otimes n}}\right\|_{1}, (25)

where the expectation is with respect to the codebook generation.

Step 1: Isolating the effect of error induced by not covering
Consider the second term within K{K}, which can be written as

ρnΛ^znρn\displaystyle\sqrt{\rho^{\otimes n}}\hat{\Lambda}_{z^{n}}\sqrt{\rho^{\otimes n}} =1NμiρnΓiA,(μ)ρn\displaystyle=\frac{1}{N}\sum_{\mu}\sum_{i}\sqrt{\rho^{\otimes n}}\Gamma^{A,(\mu)}_{i}\sqrt{\rho^{\otimes n}}
×PZ|Wn(zn|F(μ)(i))wn𝟙{F(μ)(i)=wn}=1\displaystyle\hskip 14.45377pt\times P^{n}_{Z|W}(z^{n}|F^{(\mu)}(i))\underbrace{\sum_{w^{n}}\!\mathbbm{1}_{\{F^{(\mu)}(i)=w^{n}\}}}_{=1}
=T+T~,\displaystyle=T+\widetilde{T},

where

T\ensurestackMath\stackon[1pt]=Δ\displaystyle T\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}} 1Nμi>0ρnΓiA,(μ)ρnPZ|Wn(zn|F(μ)(i)),\displaystyle\frac{1}{N}\sum_{\mu}\sum_{i>0}\sqrt{\rho^{\otimes n}}\Gamma^{A,(\mu)}_{i}\sqrt{\rho^{\otimes n}}P^{n}_{Z|W}(z^{n}|F^{(\mu)}(i)),
T~\ensurestackMath\stackon[1pt]=Δ\displaystyle\widetilde{T}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}} 1NμρnΓ0A,(μ)ρnPZ|Wn(zn|w0n).\displaystyle\frac{1}{N}\sum_{\mu}\sqrt{\rho^{\otimes n}}\Gamma^{A,(\mu)}_{0}\sqrt{\rho^{\otimes n}}P^{n}_{Z|W}(z^{n}|w^{n}_{0}).

Hence, we have KS+S~,K\leq S+\widetilde{S}, where

S\ensurestackMath\stackon[1pt]=ΔznwnρnΛ¯wnρnPZ|Wn(zn|wn)T1,\displaystyle S\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\sum_{z^{n}}\norm{\sum_{w^{n}}\sqrt{\rho^{\otimes n}}\bar{\Lambda}_{w^{n}}\sqrt{\rho^{\otimes n}}P^{n}_{Z|W}(z^{n}|w^{n})-T}_{1}, (26)

and S~\ensurestackMath\stackon[1pt]=ΔznT~1\widetilde{S}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\sum_{z^{n}}\|\widetilde{T}\|_{1}. Note that S~\widetilde{S} captures the error induced by not covering the state ρn.\rho^{\otimes n}. We further bound S~\widetilde{S} as

S~\displaystyle\widetilde{S} 1NμznPZ|Wn(zn|w0n)ρnΓ0A,(μ)ρn1\displaystyle\leq\frac{1}{N}\sum_{\mu}\sum_{z^{n}}P^{n}_{Z|W}(z^{n}|w^{n}_{0})\left\|\sqrt{\rho^{\otimes n}}\Gamma^{A,(\mu)}_{0}\sqrt{\rho^{\otimes n}}\right\|_{1}
1Nμρn(Iwnγwn(μ)Awn(μ))ρn1\displaystyle\leq\frac{1}{N}\sum_{\mu}\left\|\sqrt{\rho^{\otimes n}}(I-\sum_{w^{n}}\gamma_{w^{n}}^{(\mu)}A_{w^{n}}^{(\mu)})\sqrt{\rho^{\otimes n}}\right\|_{1}
1Nμwnλwnρ^wnwnρnγwn(μ)A¯wn(μ)ρn1\displaystyle\leq\frac{1}{N}\sum_{\mu}\left\|\sum_{w^{n}}\lambda_{w^{n}}\hat{\rho}_{w^{n}}-\sum_{w^{n}}\sqrt{\rho^{\otimes n}}\gamma_{w^{n}}^{(\mu)}\bar{A}^{(\mu)}_{w^{n}}\sqrt{\rho^{\otimes n}}\right\|_{1}
+1Nμwnρnγwn(μ)(A¯wn(μ)Awn(μ))ρn1\displaystyle\hskip 25.0pt+\frac{1}{N}\sum_{\mu}\left\|\sum_{w^{n}}\sqrt{\rho^{\otimes n}}\gamma_{w^{n}}^{(\mu)}\left(\bar{A}^{(\mu)}_{w^{n}}-A^{(\mu)}_{w^{n}}\right)\sqrt{\rho^{\otimes n}}\right\|_{1}
S~1+S~2,\displaystyle\leq\widetilde{S}_{1}+\widetilde{S}_{2},

where

S~1\displaystyle\widetilde{S}_{1} \ensurestackMath\stackon[1pt]=Δ1Nμwnλwnρ^wnwnρnγwn(μ)A¯wn(μ)ρn1,\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\frac{1}{N}\sum_{\mu}\left\|\sum_{w^{n}}\lambda_{w^{n}}\hat{\rho}_{w^{n}}-\sum_{w^{n}}\sqrt{\rho^{\otimes n}}\gamma_{w^{n}}^{(\mu)}\bar{A}^{(\mu)}_{w^{n}}\sqrt{\rho^{\otimes n}}\right\|_{1},
S~2\displaystyle\widetilde{S}_{2} \ensurestackMath\stackon[1pt]=Δ1Nμwnρnγwn(μ)(A¯wn(μ)Awn(μ))ρn1.\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\frac{1}{N}\sum_{\mu}\sum_{w^{n}}\left\|\sqrt{\rho^{\otimes n}}\gamma_{w^{n}}^{(\mu)}\left(\bar{A}^{(\mu)}_{w^{n}}-A^{(\mu)}_{w^{n}}\right)\sqrt{\rho^{\otimes n}}\right\|_{1}.

To provide a bound for the term S~1\widetilde{S}_{1}, we (i) develop a n-letter version of Lemma 2 and (ii) provide a proposition bounding the term corresponding to S~1\widetilde{S}_{1}, using this n-letter lemma.

Lemma 5.

Let {λw,θw}w𝒲\{\lambda_{w},\theta_{w}\}_{w\in\mathcal{W}} be an ensemble, with θw𝒟()\theta_{w}\in\mathcal{D}(\mathcal{H}) for all w𝒲w\in\mathcal{W}, 𝒲𝔽p\mathcal{W}\subseteq\mathbb{F}_{p} for some finite prime pp. Then, for any ϵc(0,1)\epsilon_{c}\in(0,1), and for any η,δ(0,1)\eta,\delta\in(0,1) sufficiently small, and any nn sufficiently large, we have

𝔼\displaystyle\mathbb{E} [wnλwnθwnpnpk+l1Nμ=1Nwna,mλwn(1+η)\displaystyle\bigg{[}\bigg{\|}\sum_{w^{n}}\lambda_{w^{n}}\theta_{w^{n}}-\frac{p^{n}}{p^{k+l}}\frac{1}{N^{\prime}}\sum_{\mu=1}^{N^{\prime}}\sum_{w^{n}}\sum_{a,m}\frac{\lambda_{w^{n}}}{(1+\eta)}
×θwn𝟙{Wn,(μ)(a,m)=wn}1]ϵc,\displaystyle\hskip 83.11005pt\times\theta_{w^{n}}\mathbbm{1}_{\{W^{n,(\mu)}(a,m)=w^{n}\}}\bigg{\|}_{1}\bigg{]}\leq\epsilon_{c}, (27)

if (k+ln)log(p)+1nlog(N)>I(W;R)σθS(W)σθ+log(p)\left(\frac{k+l}{n}\right)\log{p}+\frac{1}{n}\log{N^{\prime}}>I(W;R)_{\sigma_{\theta}}-S(W)_{\sigma_{\theta}}+\log{p}, where θwn\ensurestackMath\stackon[1pt]=Δi=1nθwi\theta_{w^{n}}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\bigotimes_{i=1}^{n}\theta_{w_{i}} and λwn\ensurestackMath\stackon[1pt]=ΔΠi=1nλwi\lambda_{w^{n}}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\Pi_{i=1}^{n}\lambda_{w_{i}}, σθRW\ensurestackMath\stackon[1pt]=Δw𝒲λwθw|ww|\sigma_{\theta}^{RW}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\sum_{w\in\mathcal{W}}\lambda_{w}\theta_{w}\otimes\outerproduct{w}{w}, for some orthogonal set {|w}w𝒲,\{\ket{w}\}_{w\in\mathcal{W}}, and {Wn,(μ)(a,m):a𝔽pk,m𝔽pl,μ[2nC]}\{W^{n,(\mu)}(a,m):a\in\mathbb{F}_{p}^{k},m\in\mathbb{F}_{p}^{l},\mu\in[2^{nC}]\} are as defined in (21), with GG and h(μ)h^{(\mu)} generated randomly uniformly and independently.

Proof.

The proof of the lemma is provided in Appendix A.3

Now we provide the following proposition.

Proposition 1.

For any ϵ(0,1)\epsilon\in(0,1), any η,δ(0,1)\eta,\delta\in(0,1) sufficiently small, and any nn sufficiently large, we have we have 𝔼[S~1]ϵ\mathbb{E}[\widetilde{S}_{1}]\leq\epsilon, if k+lnlog(p)>I(W;R)σS(W)σ+log(p),\frac{k+l}{n}\log{p}>I(W;R)_{\sigma}-S(W)_{\sigma}+\log{p}, where σ\sigma is the auxiliary state defined in the theorem.

Proof.

The proof is provided in Appendix B.1. ∎

Now we provide a bound for S~2.\widetilde{S}_{2}. For that, we first develop another n-letter lemma as follows.

Lemma 6.

For γwn(μ),A¯wn(μ),\gamma_{w^{n}}^{(\mu)},\bar{A}_{w^{n}}^{(\mu)},and Awn(μ)A_{w^{n}}^{(\mu)} as defined above, we have

wnγwn(μ)\displaystyle\sum_{w^{n}}\gamma_{w^{n}}^{(\mu)} ρn(A¯wn(μ)Awn(μ))ρn1\displaystyle\left\|\sqrt{\rho^{\otimes n}}\left(\bar{A}_{w^{n}}^{(\mu)}-A_{w^{n}}^{(\mu)}\right)\sqrt{\rho^{\otimes n}}\right\|_{1}
2 23nδρ(H0+(1ε)(1+η)H1+H2+H3),\displaystyle\leq{2}\;{2^{3n\delta_{\rho}}}\left(H_{0}+\frac{\sqrt{(1-\varepsilon)}}{(1+\eta)}\sqrt{H_{1}+H_{2}+H_{3}}\right),

where

H0\displaystyle H_{0} \ensurestackMath\stackon[1pt]=Δ|Δ(μ)𝔼[Δ(μ)]|,H1\ensurestackMath\stackon[1pt]=ΔTr((ΠρΠμ)wnλwnρ~wn),\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\left|\Delta^{(\mu)}\!-\mathbb{E}[\Delta^{(\mu)}]\right|,H_{1}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\Tr{\!\!(\Pi_{\rho}-\Pi^{\mu})\!\!\sum_{w^{n}}\lambda_{w^{n}}\tilde{\rho}_{w^{n}}\!},
H2\displaystyle H_{2} \ensurestackMath\stackon[1pt]=Δwnλwnρ~wn(1ε)wnαwnγwn(μ)𝔼[Δ(μ)]ρ~wn1,\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\left\|\sum_{w^{n}}\lambda_{w^{n}}\tilde{\rho}_{w^{n}}-(1-\varepsilon)\sum_{w^{n}}\frac{\alpha_{w^{n}}\gamma_{w^{n}}^{(\mu)}}{\mathbb{E}[\Delta^{(\mu)}]}\tilde{\rho}_{w^{n}}\right\|_{1},
H3\displaystyle H_{3} \ensurestackMath\stackon[1pt]=Δ(1ε)wnαwnγwn(μ)Δ(μ)ρ~wnwnαwnγwn(μ)𝔼[Δ(μ)]ρ~wn1,\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}(1-\varepsilon)\left\|\sum_{w^{n}}\frac{\alpha_{w^{n}}\gamma_{w^{n}}^{(\mu)}}{\Delta^{(\mu)}}\tilde{\rho}_{w^{n}}-\sum_{w^{n}}\frac{\alpha_{w^{n}}\gamma_{w^{n}}^{(\mu)}}{\mathbb{E}[\Delta^{(\mu)}]}\tilde{\rho}_{w^{n}}\right\|_{1}, (28)

Δ(μ)=wn𝒯δ(n)(W)αwnγwn(μ),ε\ensurestackMath\stackon[1pt]=Δwn𝒯δ(n)(W)λwn\Delta^{(\mu)}=\sum_{w^{n}\in\mathcal{T}_{\delta}^{(n)}(W)}\alpha_{w^{n}}\gamma_{w^{n}}^{(\mu)},\varepsilon\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\sum_{w^{n}\notin\mathcal{T}_{\delta}^{(n)}(W)}\lambda_{w^{n}} and δρ(δ)0\delta_{\rho}(\delta)\searrow 0 as δ0\delta\searrow 0.

Proof.

The proof is provided in Appendix A.4

Using the above lemma on S~2\widetilde{S}_{2} gives

S~2\displaystyle\widetilde{S}_{2} 2Nμ=1N23nδρ(H0+(1ε)(1+η)H1+H2+H3).\displaystyle\leq\frac{2}{N}\sum_{\mu=1}^{N}{2^{3n\delta_{\rho}}}\left(H_{0}+\frac{\sqrt{(1-\varepsilon)}}{(1+\eta)}\sqrt{H_{1}+H_{2}+H_{3}}\right).

Let us first consider H1H_{1}. By observing wnλwnρ~wnΠρρnΠρ2n(S(ρ)δρ)Πρ\sum_{w^{n}}\lambda_{w^{n}}\tilde{\rho}_{w^{n}}\leq\Pi_{\rho}\rho^{\otimes n}\Pi_{\rho}\leq 2^{-n(S(\rho)-\delta_{\rho})}\Pi_{\rho}, we bound H1H_{1} as

H12n(S(ρ)δρ)Tr((ΠρΠμ)).\displaystyle H_{1}\leq 2^{-n(S(\rho)-\delta_{\rho})}\Tr{(\Pi_{\rho}-\Pi^{\mu})}.

Note that

𝔼[Σ(μ)]\displaystyle\mathbb{E}[\Sigma^{(\mu)}] =𝔼[wnαwnγwn(μ)ρn1ρ~wnρn1]\displaystyle=\mathbb{E}\left[\sum_{w^{n}}\alpha_{w^{n}}\gamma_{w^{n}}^{(\mu)}\sqrt{\rho^{\otimes n}}^{-1}\tilde{\rho}_{w^{n}}\sqrt{\rho^{\otimes n}}^{-1}\right]
=1(1+η)wnλwnρn1ρ~wnρn1Πρ(1+η).\displaystyle=\frac{1}{(1+\eta)}\sum_{w^{n}}\!\lambda_{w^{n}}\!\sqrt{\rho^{\otimes n}}^{-1}\!\tilde{\rho}_{w^{n}}\sqrt{\rho^{\otimes n}}^{-1}\!\!\leq\frac{\Pi_{\rho}}{(1+\eta)}.

Now, we use the Pruning Trace Inequality developed in Lemma 4 on Σ(μ)\Sigma^{(\mu)}, with η(0,1)\eta\in(0,1) to obtain

𝔼[H1]\displaystyle\mathbb{E}[H_{1}] 2n(S(ρ)δρ)(1+η)η𝔼[Σ(μ)𝔼[Σ(μ)]1]\displaystyle\leq 2^{-n(S(\rho)-\delta_{\rho})}\frac{(1+\eta)}{\eta}\mathbb{E}\left[\|\Sigma^{(\mu)}-\mathbb{E}[\Sigma^{(\mu)}]\|_{1}\right]
2n(S(ρ)δρ)(1+η)ηΠρρn1\displaystyle\leq 2^{-n(S(\rho)-\delta_{\rho})}\frac{(1+\eta)}{\eta}\left\|\Pi_{\rho}\sqrt{\rho^{\otimes n}}^{-1}\right\|_{\infty}
×𝔼[wnαwnγwn(μ)ρ~wn𝔼[wnαwnγwn(μ)ρ~wn]1]\displaystyle\hskip 7.0pt\times\mathbb{E}\left[{\left\|\sum_{w^{n}}\alpha_{w^{n}}\gamma_{w^{n}}^{(\mu)}\tilde{\rho}_{w^{n}}-\mathbb{E}\big{[}\sum_{w^{n}}\alpha_{w^{n}}\gamma_{w^{n}}^{(\mu)}\tilde{\rho}_{w^{n}}\big{]}\right\|_{1}}\right]
×Πρρn1\displaystyle\hskip 15.0pt\times\left\|\Pi_{\rho}\sqrt{\rho^{\otimes n}}^{-1}\right\|_{\infty}
22nδρ(1+η)η𝔼[wnλwnρ~wn(1+η)1(1+η)pnpk+l\displaystyle\leq{2^{2n\delta_{\rho}}}\frac{(1+\eta)}{\eta}\mathbb{E}\left[\left\|\sum_{w^{n}}\frac{\lambda_{w^{n}}\tilde{\rho}_{w^{n}}}{(1+\eta)}-\frac{1}{(1+\eta)}\frac{p^{n}}{p^{k+l}}\right.\right.
wna,iλwnρ~wn𝟙{Wn,(μ)(a,i)=wn}1]\displaystyle\hskip 55.0pt\left.\left.\sum_{w^{n}}\sum_{a,i}\lambda_{w^{n}}\tilde{\rho}_{w^{n}}\mathbbm{1}_{\{W^{n,(\mu)}(a,i)=w^{n}\}}\right\|_{1}\right]
=22nδρ(1ε)η𝔼[H~],\displaystyle={2^{2n\delta_{\rho}}}\frac{(1-\varepsilon)}{\eta}\mathbb{E}[\widetilde{H}], (29)

where the second inequality follows from Hólders inequality, and the equality follows by defining H~\widetilde{H} as

H~\ensurestackMath\stackon[1pt]=Δ\displaystyle\widetilde{H}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}} wnλwn(1ε)ρ~wn\displaystyle\left\|\sum_{w^{n}}\frac{\lambda_{w^{n}}}{(1-\varepsilon)}\tilde{\rho}_{w^{n}}\right.
pnpk+lwna,iλwn(1ε)ρ~wn𝟙{Wn,(μ)(a,i)=wn}1.\displaystyle\hskip 10.0pt\left.-\frac{p^{n}}{p^{k+l}}\sum_{w^{n}}\sum_{a,i}\frac{\lambda_{w^{n}}}{(1-\varepsilon)}\tilde{\rho}_{w^{n}}\mathbbm{1}_{\{W^{n,(\mu)}(a,i)=w^{n}\}}\right\|_{1}. (30)

Similarly, using 𝔼[Δ(μ)]=(1ε)(1+η)\mathbb{E}[\Delta^{(\mu)}]=\frac{(1-\varepsilon)}{(1+\eta)}, H2H_{2} can be simplified as

H2\displaystyle H_{2} =wnλwnρ~wn\displaystyle=\left\|\sum_{w^{n}}\lambda_{w^{n}}\tilde{\rho}_{w^{n}}\right.
pnpk+lwna,iλwnρ~wn𝟙{Wn,(μ)(a,i)=wn}1\displaystyle\hskip 20.0pt\left.-\frac{p^{n}}{p^{k+l}}\sum_{w^{n}}\sum_{a,i}\lambda_{w^{n}}\tilde{\rho}_{w^{n}}\mathbbm{1}_{\{W^{n,(\mu)}(a,i)=w^{n}\}}\right\|_{1}
=(1ε)H~.\displaystyle=(1-\varepsilon)\tilde{H}. (31)

Now we consider H3H_{3} and convert it into a similar expression as H0H_{0}.

H3\displaystyle H_{3} (1ε)wn𝒯δ(n)(W)αwnγwn(μ)|1Δ(μ)1𝔼[Δ(μ)]|\displaystyle\leq(1-\varepsilon)\!\!\!\!\sum_{w^{n}\in\mathcal{T}_{\delta}^{(n)}(W)}\!\!\!\!\!\alpha_{w^{n}}\gamma_{w^{n}}^{(\mu)}\left|\frac{1}{\Delta^{(\mu)}}-\frac{1}{\mathbb{E}[\Delta^{(\mu)}]}\right|
=(1+η)|Δ(μ)𝔼[Δ(μ)]|=(1+η)H0.\displaystyle=(1+\eta)\left|{\Delta^{(\mu)}}-{\mathbb{E}[\Delta^{(\mu)}]}\right|=(1+\eta)H_{0}. (32)

Using the above simplification and the concavity of square-root function we obtain:

𝔼[S~2]\displaystyle\mathbb{E}[\widetilde{S}_{2}] 2N23nδρμ=1N(𝔼[H0]+(1ε)(1+η)\displaystyle\leq\frac{2}{N}{2^{3n\delta_{\rho}}}\sum_{\mu=1}^{N}\left(\mathbb{E}[H_{0}]+\frac{\sqrt{(1-\varepsilon)}}{(1+\eta)}\right.
×(1ε)(22nδρη+1)𝔼[H~]+(1+η)𝔼[H0])\displaystyle\hskip 10.0pt\times\left.\sqrt{(1-\varepsilon)\left(\frac{2^{2n\delta_{\rho}}}{\eta}+1\right)\mathbb{E}[\widetilde{H}]+{(1+\eta)}\mathbb{E}[H_{0}]}\right)
2N23nδρμ=1N(𝔼[H0]+(1ε)(1+η)\displaystyle\leq\frac{2}{N}{2^{3n\delta_{\rho}}}\sum_{\mu=1}^{N}\Bigg{(}\mathbb{E}[H_{0}]+\frac{{(1-\varepsilon)}}{(1+\eta)}
×(22nδρη+1)𝔼[H~]+(1ε)(1+η)𝔼[H0]).\displaystyle\hskip 20.0pt\times\!\sqrt{\!\left(\!\frac{2^{2n\delta_{\rho}}}{\eta}\!+\!1\!\right)\!\mathbb{E}[\widetilde{H}]}+\sqrt{\frac{(1-\varepsilon)}{(1+\eta)}}\sqrt{\mathbb{E}[H_{0}]}\Bigg{)}.

The following proposition provides a bound on the above term.

Proposition 2.

For any ϵ(0,1)\epsilon\in(0,1), any η,δ(0,1)\eta,\delta\in(0,1) sufficiently small, and any nn sufficiently large, we have 𝔼[S~2]ϵ\mathbb{E}\left[\widetilde{S}_{2}\right]\leq\epsilon, if k+lnlog(p)>I(W;R)σS(W)σ+log(p)\frac{k+l}{n}\log{p}>I(W;R)_{\sigma}-S(W)_{\sigma}+\log{p}, where σ\sigma is the auxiliary state defined in the theorem.

Proof.

The proof is provided in Appendix B.2

Remark 7.

The term corresponding to the operators that complete the sub-POVMs M(n,μ)M^{(n,\mu)}, i.e., Iwn𝒯δ(n)(W)γwn(μ)Awn(μ)I-\sum_{w^{n}\in\mathcal{T}_{\delta}^{(n)}(W)}\gamma_{w^{n}}^{(\mu)}A_{w^{n}}^{(\mu)} is taken care in T~\widetilde{T}. The expression TT excludes these completing operators.

Step 2: Isolating the effect of error induced by binning
 For this, we simplify TT as

T=\displaystyle T= 1Nμwni>0a𝔽pkρnAwn(μ)ρn\displaystyle\frac{1}{N}\sum_{\mu}\sum_{w^{n}}\sum_{\begin{subarray}{c}i>0\end{subarray}}\sum_{a\in\mathbb{F}_{p}^{k}}\sqrt{\rho^{\otimes n}}A_{w^{n}}^{(\mu)}\sqrt{\rho^{\otimes n}}
×PZ|Wn(zn|F(μ)(i))𝟙{aG+h(μ)(i)=wn}.\displaystyle\hskip 20.0pt\times P^{n}_{Z|W}(z^{n}|F^{(\mu)}(i))\mathbbm{1}_{\{aG+h^{(\mu)}(i)=w^{n}\}}.

We substitute the above expression into SS defined in (26), and isolate the effect of binning by adding and subtracting an appropriate term within SS and applying triangle inequality to obtain SS1+S2,S\leq S_{1}+S_{2}, where

S1\displaystyle S_{1} \ensurestackMath\stackon[1pt]=Δznwnρn(Λ¯wn1Nμγwn(μ)Awn(μ))ρn\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\sum_{z^{n}}\left\|\sum_{w^{n}}\sqrt{\rho^{\otimes n}}\left(\bar{\Lambda}_{w^{n}}-\frac{1}{N}\sum_{\mu}\gamma_{w^{n}}^{(\mu)}A_{w^{n}}^{(\mu)}\right)\sqrt{\rho^{\otimes n}}\right.
×PZ|Wn(zn|wn)1,\displaystyle\hskip 151.76744pt\left.\times P^{n}_{Z|W}(z^{n}|w^{n})\right\|_{1}\!,
S2\displaystyle S_{2} \ensurestackMath\stackon[1pt]=Δzn1Nμa,i>0wnρnAwn(μ)ρn𝟙{aG+h(μ)(i)=wn}\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\sum_{z^{n}}\left\|\frac{1}{N}\!\sum_{\mu}\!\!\sum_{a,i>0}\!\!\sum_{w^{n}}\!\!\!\sqrt{\rho^{\otimes n}}A_{w^{n}}^{(\mu)}\!\sqrt{\rho^{\otimes n}}\mathbbm{1}_{\{aG+h^{(\mu)}(i)=w^{n}\}}\right.
×(PZ|Wn(zn|wn)PZ|Wn(zn|F(μ)(i)))1,\displaystyle\hskip 43.36243pt\left.\times\left(P^{n}_{Z|W}(z^{n}|w^{n})-P^{n}_{Z|W}\left(z^{n}|F^{(\mu)}(i)\right)\right)\right\|_{1}\!,

where F(μ)()F^{(\mu)}(\cdot) is as defined in (24). Note that the term S1S_{1} characterizes the error introduced by approximation of the original POVM with the collection of approximating sub-POVM M~(n,μ)\tilde{M}^{(n,\mu)}, and the term S2S_{2} characterizes the error caused by binning this approximating sub-POVM. In this step, we analyze S2S_{2} and prove the following proposition.

Proposition 3.

For any ϵ(0,1)\epsilon\in(0,1), any η,δ(0,1)\eta,\delta\in(0,1) sufficiently small, and any nn sufficiently large, we have 𝔼[S2]ϵ\mathbb{E}\left[{S}_{2}\right]\leq\epsilon, if k+lnlog(p)R<log(p)S(W)σ\frac{k+l}{n}\log{p}-R<\log{p}-S(W)_{\sigma}, where σ\sigma is the auxiliary state defined in the statement of the theorem.

Proof.

The proof is provided in Appendix B.3

Step 3: Isolating the effect of approximating measurement
In this step, we finally analyze the error induced from employing the approximating measurement, given by the term S1S_{1}. We add and subtract appropriate terms within S1S_{1} and use triangle inequality to obtain S1S11+S12+S13S_{1}\leq S_{11}+S_{12}+S_{13}, where

S11\displaystyle S_{11} \ensurestackMath\stackon[1pt]=Δznwnρn(Λ¯wn1Nμ=1Nαwnγwn(μ)λwnΛ¯wn)\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\sum_{z^{n}}\left\|\sum_{w^{n}}\sqrt{\rho^{\otimes n}}\left(\bar{\Lambda}_{w^{n}}-\frac{1}{N}\sum_{\mu=1}^{N}\frac{\alpha_{w^{n}}\gamma_{w^{n}}^{(\mu)}}{\lambda_{w^{n}}}\bar{\Lambda}_{w^{n}}\right)\right.
×ρnPZ|Wn(zn|wn)1,\displaystyle\hskip 115.63243pt\left.\times\sqrt{\rho^{\otimes n}}P^{n}_{Z|W}(z^{n}|w^{n})\right\|_{1},
S12\displaystyle S_{12} \ensurestackMath\stackon[1pt]=Δzn1Nμ=1Nwnρn(αwnγwn(μ)λwnΛ¯wnγwn(μ)A¯wn(μ))\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\sum_{z^{n}}\left\|\frac{1}{N}\!\sum_{\mu=1}^{N}\sum_{w^{n}}\!\sqrt{\rho^{\otimes n}}\!\left(\!\frac{\alpha_{w^{n}}\gamma_{w^{n}}^{(\mu)}}{\lambda_{w^{n}}}\bar{\Lambda}_{w^{n}}\!\!-\gamma_{w^{n}}^{(\mu)}\bar{A}_{w^{n}}^{(\mu)}\!\right)\right.
×ρnPZ|Wn(zn|wn)1,\displaystyle\hskip 115.63243pt\left.\times\sqrt{\rho^{\otimes n}}P^{n}_{Z|W}(z^{n}|w^{n})\right\|_{1},
S13\displaystyle S_{13} \ensurestackMath\stackon[1pt]=Δzn1Nμ=1Nwnρn(γwn(μ)A¯wn(μ)γwn(μ)Awn(μ))\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\sum_{z^{n}}\left\|\frac{1}{N}\sum_{\mu=1}^{N}\sum_{w^{n}}\sqrt{\rho^{\otimes n}}\left(\gamma_{w^{n}}^{(\mu)}\bar{A}_{w^{n}}^{(\mu)}-\gamma_{w^{n}}^{(\mu)}A_{w^{n}}^{(\mu)}\right)\right.
×ρnPZ|Wn(zn|wn)1.\displaystyle\hskip 115.63243pt\left.\times\sqrt{\rho^{\otimes n}}P^{n}_{Z|W}(z^{n}|w^{n})\right\|_{1}.

Now with the intention of employing Lemma 5, we express S11S_{11} as

S11\displaystyle S_{11} =wnλwnρ^wnϕwn1N1(1+η)pnpk+l\displaystyle=\left\|\sum_{w^{n}}\lambda_{w^{n}}\hat{\rho}_{w^{n}}\otimes\phi_{w^{n}}-\frac{1}{N}\frac{1}{(1+\eta)}\frac{p^{n}}{p^{k+l}}\right.
×μwna,i0𝟙{Wn,(μ)(a,i)=wn}ρ^wnϕwn1,\displaystyle\hskip 28.90755pt\left.\times\sum_{\mu}\sum_{w^{n}}\sum_{a,i\neq 0}\mathbbm{1}_{\{W^{n,(\mu)}(a,i)=w^{n}\}}\hat{\rho}_{w^{n}}\otimes\phi_{w^{n}}\right\|_{1},

where the equality above is obtained by defining ϕwn=znPZ|Wn(zn|wn)|znzn|\phi_{w^{n}}=\sum_{z^{n}}P^{n}_{Z|W}(z^{n}|w^{n})\otimes\outerproduct{z^{n}}{z^{n}} and using the definitions of αwn,γwn(μ)\alpha_{w^{n}},\gamma_{w^{n}}^{(\mu)} and ρ^wn\hat{\rho}_{w^{n}}, followed by using the triangle inequality for the block diagonal operators, Note that the triangle inequality becomes an equality for such block diagonal operators. By identifying θw\theta_{w} with ρ^wϕw\hat{\rho}_{w}\otimes\phi_{w} in Lemma 5 we obtain the following: for all ϵ>0\epsilon>0 and η,δ(0,1)\eta,\delta\in(0,1) sufficiently small, and any nn sufficiently large, 𝔼[S11]ϵ\mathbb{E}\left[{S}_{11}\right]\leq\epsilon, if k+lnlog(p)+1nlog(N)>I(W;R,Z)σ+log(p)S(W)σ\frac{k+l}{n}\log{p}+\frac{1}{n}\log{N}>I(W;R,Z)_{\sigma}+\log{p}-S(W)_{\sigma}, where σ\sigma is the auxiliary state defined in the theorem.

Now we consider the term corresponding to S12S_{12}, and prove that its expectation is small. Recalling S12S_{12}, we get

S12\displaystyle S_{12} 1Nμ=1NwnznPZ|Wn(zn|wn)\displaystyle\leq\frac{1}{N}\sum_{\mu=1}^{N}\sum_{w^{n}}\sum_{z^{n}}P^{n}_{Z|W}(z^{n}|w^{n})
×ρn(αwnγwn(μ)λwnΛ¯wnγwn(μ)A¯wn(μ))ρn1,\displaystyle\hskip 13.0pt\times\left\|\sqrt{\rho^{\otimes n}}\left(\frac{\alpha_{w^{n}}\gamma_{w^{n}}^{(\mu)}}{\lambda_{w^{n}}}\bar{\Lambda}_{w^{n}}-\gamma_{w^{n}}^{(\mu)}\bar{A}_{w^{n}}^{(\mu)}\right)\sqrt{\rho^{\otimes n}}\right\|_{1}\!,
=1Nμ=1Nwnαwnγwn(μ)ρn(1λwnΛ¯wn\displaystyle=\frac{1}{N}\sum_{\mu=1}^{N}\sum_{w^{n}}\alpha_{w^{n}}\gamma_{w^{n}}^{(\mu)}\bigg{\|}\sqrt{\rho^{\otimes n}}\left(\frac{1}{\lambda_{w^{n}}}\bar{\Lambda}_{w^{n}}-\right.
ρn1ρ~wnρn1)ρn1,\displaystyle\hskip 80.0pt\left.\sqrt{\rho^{\otimes n}}^{-1}\tilde{\rho}_{w^{n}}\sqrt{\rho^{\otimes n}}^{-1}\right)\sqrt{\rho^{\otimes n}}\bigg{\|}_{1},

where the inequality above is obtained by using triangle inequality. Applying the expectation, we get

𝔼[S12]\displaystyle\mathbb{E}{\left[S_{12}\right]} 1(1+η)wnλwnρn(1λwnΛ¯wn\displaystyle\leq\frac{1}{(1+\eta)}\sum_{w^{n}}\lambda_{w^{n}}\bigg{\|}\sqrt{\rho^{\otimes n}}\left(\frac{1}{\lambda_{w^{n}}}\bar{\Lambda}_{w^{n}}-\right.
ρn1ρ~wnρn1)ρn1,\displaystyle\hskip 72.26999pt\left.\sqrt{\rho^{\otimes n}}^{-1}\tilde{\rho}_{w^{n}}\sqrt{\rho^{\otimes n}}^{-1}\right)\sqrt{\rho^{\otimes n}}\bigg{\|}_{1},
1(1+η)wn𝒯δ(n)(W)λwn(ρ^wnρ~wn)1\displaystyle\leq\frac{1}{(1+\eta)}\!\!\!\sum_{\begin{subarray}{c}w^{n}\in\mathcal{T}_{\delta}^{(n)}(W)\end{subarray}}\!\!\lambda_{w^{n}}\left\|\left(\hat{\rho}_{w^{n}}-\tilde{\rho}_{w^{n}}\right)\right\|_{1}
+1(1+η)wn𝒯δ(n)(W)λwnρ^wn1\displaystyle\hskip 72.26999pt+\frac{1}{(1+\eta)}\!\!\!\sum_{\begin{subarray}{c}w^{n}\notin\mathcal{T}_{\delta}^{(n)}(W)\end{subarray}}\!\!\lambda_{w^{n}}\left\|\hat{\rho}_{w^{n}}\right\|_{1}
(2ε+2ε′′)+ε(1+η)=ϵS12,\displaystyle\leq\frac{(2\sqrt{\varepsilon^{\prime}}+2\sqrt{\varepsilon^{\prime\prime}})+\varepsilon}{(1+\eta)}=\epsilon_{\scriptscriptstyle S_{12}},

where we have used the fact that 𝔼[αwnγwn(μ)]=λwn(1+η)\mathbb{E}{[\alpha_{w^{n}}\gamma_{w^{n}}^{(\mu)}]}=\frac{\lambda_{w^{n}}}{(1+\eta)}, and the last inequality is obtained by the repeated usage of the Average Gentle Measurement Lemma [7] and setting ϵS12=1(1+η)(2ε+2ε′′+ε)\epsilon_{\scriptscriptstyle{S}_{12}}=\frac{1}{(1+\eta)}(2\sqrt{\varepsilon^{\prime}}+2\sqrt{\varepsilon^{\prime\prime}}+\varepsilon) with ϵS120\epsilon_{\scriptscriptstyle{S}_{12}}\searrow 0 as nn\rightarrow\infty and ε\ensurestackMath\stackon[1pt]=Δεp+2εp\varepsilon^{\prime}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\varepsilon^{\prime}_{p}+2\sqrt{\varepsilon^{\prime}_{p}} and ε′′\ensurestackMath\stackon[1pt]=Δ2εp+2εp\varepsilon^{\prime\prime}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}2\varepsilon^{\prime}_{p}+2\sqrt{\varepsilon^{\prime}_{p}} for εp\ensurestackMath\stackon[1pt]=Δ1min{Tr(Πρρ^wn),Tr(Πwnρ^wn),1ε}\varepsilon^{\prime}_{p}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}1-\min\left\{\Tr{\Pi_{\rho}\hat{\rho}_{w^{n}}},\Tr{\Pi_{w^{n}}\hat{\rho}_{w^{n}}},1-\varepsilon\right\} (see (35) in [3] for details). Now, we move on to bounding the last term within S1S_{1}, i.e., S13.S_{13}. We start by applying triangle inequality to obtain

S13\displaystyle S_{13} znwnPZ|Wn(zn|wn)\displaystyle\leq\sum_{z^{n}}\sum_{w^{n}}P^{n}_{Z|W}(z^{n}|w^{n})
×1Nμ=1Nρn(γwn(μ)A¯wn(μ)γwn(μ)Awn(μ))ρn1\displaystyle\hskip 10.0pt\times\left\|\frac{1}{N}\sum_{\mu=1}^{N}\sqrt{\rho^{\otimes n}}\left(\gamma_{w^{n}}^{(\mu)}\bar{A}_{w^{n}}^{(\mu)}-\gamma_{w^{n}}^{(\mu)}A_{w^{n}}^{(\mu)}\right)\sqrt{\rho^{\otimes n}}\right\|_{1}
1Nμ=1Nwnγwn(μ)ρn(A¯wn(μ)Awn(μ))ρn1\displaystyle\leq\frac{1}{N}\sum_{\mu=1}^{N}\sum_{w^{n}}\gamma_{w^{n}}^{(\mu)}\left\|\sqrt{\rho^{\otimes n}}\left(\bar{A}_{w^{n}}^{(\mu)}-A_{w^{n}}^{(\mu)}\right)\sqrt{\rho^{\otimes n}}\right\|_{1}
=S~2.\displaystyle=\widetilde{S}_{2}. (33)

Since the above term is exactly same as S~2,\widetilde{S}_{2}, we obtain the same rate constraints as in S~2\widetilde{S}_{2} to bound S13,S_{13}, i.e., for all ϵ>0\epsilon>0 and η,δ(0,1)\eta,\delta\in(0,1) sufficiently small, and any nn sufficiently large, 𝔼[S13]ϵ\mathbb{E}[S_{13}]\leq\epsilon if k+lnlog(p)>I(W;R)σ+log(p)S(W)σ\frac{k+l}{n}\log{p}>I(W;R)_{\sigma}+\log{p}-S(W)_{\sigma}.

Since S1S11+S12+S13S_{1}\leq S_{11}+S_{12}+S_{13}, S1S_{1} can be made arbitrarily small for sufficiently large n, if k+lnlogp+1nlogN>I(W;RZ)σS(W)σ+log(p)\frac{k+l}{n}\log p+\frac{1}{n}\log N>I(W;RZ)_{\sigma}-S(W)_{\sigma}+\log{p} and k+lnlog(p)>I(W;R)σS(W)σ+log(p)\frac{k+l}{n}\log{p}>I(W;R)_{\sigma}-S(W)_{\sigma}+\log{p}.

V.2.5 Rate Constraints

To sum-up, we showed 𝔼[K]ϵ\mathbb{E}[K]\leq\epsilon holds for sufficiently large nn if the following bounds hold:

R1+R\displaystyle R_{1}+R >I(W;R)σS(W)σ+log(p),\displaystyle>I(W;R)_{\sigma}-S(W)_{\sigma}+\log{p}, (34a)
R1+R+C\displaystyle R_{1}+R+C >I(W;RZ)σS(W)σ+log(p),\displaystyle>I(W;RZ)_{\sigma}-S(W)_{\sigma}+\log{p}, (34b)
R1\displaystyle R_{1} <log(p)S(W)σ,\displaystyle<\log{p}-S(W)_{\sigma}, (34c)
R1\displaystyle R_{1} 0,C0,\displaystyle\geq 0,\quad C\geq 0, (34d)

where R1\ensurestackMath\stackon[1pt]=Δknlog(p)R_{1}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\frac{k}{n}\log{p} and C\ensurestackMath\stackon[1pt]=Δ1nlog2NC\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\frac{1}{n}\log_{2}N, and R=lnlogpR=\frac{l}{n}\log p. Therefore, there exists a distributed protocol with parameters (n,2nR,2nC)(n,2^{nR},2^{nC}) such that its overall POVM M^\hat{M} is ϵ\epsilon-faithful to MnM^{\otimes n} with respect to ρn\rho^{\otimes n}. This completes the proof of the theorem.

VI Proof of Theorem 1

Suppose there exists a finite field 𝔽p\mathbb{F}_{p}, for a prime pp, a pair of mappings fS:𝒮𝔽pf_{S}:\mathcal{S}\rightarrow\mathbb{F}_{p} and fT:𝒯𝔽pf_{T}:\mathcal{T}\rightarrow\mathbb{F}_{p}, and a stochastic mapping PZ|W:𝔽p𝒵P_{Z|W}:\mathbb{F}_{p}\rightarrow\mathcal{Z} such that

PZ|S,T(z|s,t)=PZ|W(z|fS(s)+fT(t)),P_{Z|S,T}(z|s,t)=P_{Z|W}(z|f_{S}(s)+f_{T}(t)),

s𝒮,t𝒯,z𝒵,\forall s\in\mathcal{S},t\in\mathcal{T},z\in\mathcal{Z}, yielding U=fS(S)U=f_{S}(S), and V=fT(T)V=f_{T}(T). This implies that we have POVMs M¯A\ensurestackMath\stackon[1pt]=Δ\bar{M}_{A}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}} {Λ¯uA}u𝒰\{\bar{\Lambda}^{A}_{u}\}_{u\in\mathcal{U}} and M¯B\ensurestackMath\stackon[1pt]=Δ{Λ¯vB}v𝒱\bar{M}_{B}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\{\bar{\Lambda}^{B}_{v}\}_{v\in\mathcal{V}} with 𝒰=𝒱\ensurestackMath\stackon[1pt]=Δ𝔽p\mathcal{U}=\mathcal{V}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\mathbb{F}_{p} and a stochastic map PZ|W:𝔽p𝒵P_{Z|W}:\mathbb{F}_{p}\rightarrow\mathcal{Z}, such that MAB{M}_{AB} can be decomposed as

ΛzAB=u,vPZ|W(z|u+v)Λ¯uAΛ¯vB,z,\Lambda^{AB}_{z}=\sum_{u,v}P_{Z|W}(z|u+v)\bar{\Lambda}^{A}_{u}\otimes\bar{\Lambda}^{B}_{v},~{}\forall z, (35)

where WW is defined as W=U+VW=U+V. The coding strategy used here is based on Unionized Coset Codes, similar to the one employed in the point-to-point proof (Section V.2), but extended to a distributed setting. Further, the structure in these codes provide a method to exploit the structure present in the stochastic processing applied by Charlie on the classical bits received, i.e., PZ|U+VP_{Z|U+V}. Using this technique, we aim to strictly reduce the rate constraints compared to the ones obtained in Theorem 6 of [24]. Also note that, the results in [24] are based on the assumption that approximating POVMs are all mutually independent. However, since the structured construction of the POVMs only guarantees pairwise independence among the operators of the POVM, the proofs below become significantly different from [24].

We start by generating the canonical ensembles corresponding to M¯A\bar{M}_{A} and M¯B\bar{M}_{B}, defined as

λuA\ensurestackMath\stackon[1pt]=ΔTr{Λ¯uAρA},\displaystyle\lambda^{A}_{u}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\Tr\{\bar{\Lambda}^{A}_{u}\rho_{A}\}, λvB\ensurestackMath\stackon[1pt]=ΔTr{Λ¯vBρB},\displaystyle\quad\lambda^{B}_{v}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\Tr\{\bar{\Lambda}^{B}_{v}\rho_{B}\},
λuvAB\ensurestackMath\stackon[1pt]=ΔTr{(Λ¯uA\displaystyle\lambda^{AB}_{uv}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\Tr\{(\bar{\Lambda}^{A}_{u} Λ¯vB)ρAB},and\displaystyle\otimes\bar{\Lambda}^{B}_{v})\rho_{AB}\},\quad\text{and}
ρ^uA\ensurestackMath\stackon[1pt]=Δ1λuAρAΛ¯uAρA,\displaystyle\hat{\rho}^{A}_{u}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\frac{1}{\lambda^{A}_{u}}\sqrt{\rho_{A}}\bar{\Lambda}^{A}_{u}\sqrt{\rho_{A}}, ρ^vB\ensurestackMath\stackon[1pt]=Δ1λvBρBΛ¯vBρB,\displaystyle\quad\hat{\rho}^{B}_{v}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\frac{1}{\lambda^{B}_{v}}\sqrt{\rho_{B}}\bar{\Lambda}^{B}_{v}\sqrt{\rho_{B}},\quad
ρ^uvAB\ensurestackMath\stackon[1pt]=Δ1λuvABρAB\displaystyle\hat{\rho}^{AB}_{uv}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\frac{1}{\lambda^{AB}_{uv}}\sqrt{\rho_{AB}} (Λ¯uAΛ¯vB)ρAB.\displaystyle(\bar{\Lambda}^{A}_{u}\otimes\bar{\Lambda}^{B}_{v})\sqrt{\rho_{AB}}. (36)

With this notation, corresponding to each of the probability distributions, we can associate a δ\delta-typical set. Let us denote 𝒯δ(n)(U)\mathcal{T}_{\delta}^{(n)}(U), 𝒯δ(n)(V)\mathcal{T}_{\delta}^{(n)}(V) and 𝒯δ(n)(UV)\mathcal{T}_{\delta}^{(n)}(UV) as the δ\delta-typical sets defined for {λuA}\{\lambda^{A}_{u}\}, {λvB}\{\lambda^{B}_{v}\} and {λuvAB}\{\lambda^{AB}_{uv}\}, respectively.

Let ΠρA\Pi_{\rho_{A}} and ΠρB\Pi_{\rho_{B}} denote the δ\delta-typical projectors (as in [7, Def. 15.1.3]) for marginal density operators ρA\rho_{A} and ρB\rho_{B}, respectively. Also, for any un𝒰nu^{n}\in\mathcal{U}^{n} and vn𝒱nv^{n}\in\mathcal{V}^{n}, let ΠunA\Pi_{u^{n}}^{A} and ΠvnB\Pi_{v^{n}}^{B} denote the strong conditional typical projectors (as in [7, Def. 15.2.4]) for the canonical ensembles {λuA,ρ^uA}\{\lambda^{A}_{u},\hat{\rho}^{A}_{u}\} and {λvB,ρ^vB}\{\lambda^{B}_{v},\hat{\rho}^{B}_{v}\}, respectively.

For each un𝒯δ(n)(U)u^{n}\in\mathcal{T}_{\delta}^{(n)}(U) and vn𝒯δ(n)(V)v^{n}\in\mathcal{T}_{\delta}^{(n)}(V) define

ρ~unA\ensurestackMath\stackon[1pt]=ΔΠρAΠunAρ^unAΠunAΠρA,ρ~vnB\ensurestackMath\stackon[1pt]=ΔΠρBΠvnBρ^vnBΠvnBΠρB,\tilde{\rho}_{u^{n}}^{A}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\Pi_{\rho_{A}}\Pi_{u^{n}}^{A}\hat{\rho}^{A}_{u^{n}}\Pi_{u^{n}}^{A}\Pi_{\rho_{A}},\quad\tilde{\rho}_{v^{n}}^{B}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\Pi_{\rho_{B}}\Pi_{v^{n}}^{B}\hat{\rho}^{B}_{v^{n}}\Pi_{v^{n}}^{B}\Pi_{\rho_{B}},

and ρ~unA=0,\tilde{\rho}_{u^{n}}^{A}=0, and ρ~vnB=0\tilde{\rho}_{v^{n}}^{B}=0 for un𝒯δ(n)(U)u^{n}\notin\mathcal{T}_{\delta}^{(n)}(U) and vn𝒯δ(n)(V)v^{n}\notin\mathcal{T}_{\delta}^{(n)}(V), respectively, with ρ^unA\ensurestackMath\stackon[1pt]=Δiρ^uiA\hat{\rho}^{A}_{u^{n}}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\bigotimes_{i}\hat{\rho}^{A}_{u_{i}} and ρ^vnB\ensurestackMath\stackon[1pt]=Δiρ^viB\hat{\rho}^{B}_{v^{n}}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\bigotimes_{i}\hat{\rho}^{B}_{v_{i}}.

VI.1 Construction of Structured POVMs

In what follows, we construct the random structured POVM elements. Fix a block length n>0n>0, positive integers N1N_{1} and N2N_{2}, and a finite field 𝔽p\mathbb{F}_{p}. Let μ1[1,N1]\mu_{1}\in[1,N_{1}] denote the common randomness shared between the first encoder and the decoder, and let μ2[1,N2]\mu_{2}\in[1,N_{2}] denote the common randomness shared between the second encoder and the decoder. Let μ~1[1,N~1]\tilde{\mu}_{1}\in[1,\tilde{N}_{1}] and μ~2[1,N~2]\tilde{\mu}_{2}\in[1,\tilde{N}_{2}] denote additional pairwise shared randomness used for random coding purposes. This randomness is only used to show the existence of a desired distributed protocol (as defined in Definition 2), and is used only for bounding purposes. We denote μ¯i\ensurestackMath\stackon[1pt]=Δ(μi,μ~i)\bar{\mu}_{i}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}(\mu_{i},\tilde{\mu}_{i}), and N¯i\ensurestackMath\stackon[1pt]=ΔNiN~i\bar{N}_{i}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}N_{i}\cdot\tilde{N}_{i} for i=1,2i=1,2. Further, let UU and VV be random variables defined on the alphabets 𝒰\mathcal{U} and 𝒱\mathcal{V}, respectively, where 𝒰=𝒱=𝔽p\mathcal{U}=\mathcal{V}=\mathbb{F}_{p}. In building the code, we use the Unionized Coset Codes (UCCs) [37] as defined above in Definition 7.

For every (μ¯1,μ¯2)(\bar{\mu}_{1},\bar{\mu}_{2}), consider two UCCs (G,h1(μ¯1))(G,h_{1}^{(\bar{\mu}_{1})}) and (G,h2(μ¯2))(G,h_{2}^{(\bar{\mu}_{2})}), each with parameters (n,k,l1,p)(n,k,l_{1},p) and (n,k,l2,p)(n,k,l_{2},p), respectively. Note that, for every (μ¯1,μ¯2),(\bar{\mu}_{1},\bar{\mu}_{2}), they share the same generator matrix G.G.

For each (μ¯1,μ¯2)(\bar{\mu}_{1},\bar{\mu}_{2}), the generator matrix GG along with the function h1(μ¯1)h_{1}^{(\bar{\mu}_{1})} and h2(μ¯2)h_{2}^{(\bar{\mu}_{2})} generates pk+l1p^{k+l_{1}} and pk+l2p^{k+l_{2}} codewords, respectively. Each of these codewords are characterized by a triple (ai,mi,μ¯i)(a_{i},m_{i},\bar{\mu}_{i}), where ai𝔽pka_{i}\in\mathbb{F}^{k}_{p} and mi𝔽plim_{i}\in\mathbb{F}^{l_{i}}_{p} corresponds to the coarse code and the fine code indices, respectively, for i[1,2]i\in[1,2]. Let Un,(μ¯1)(a1,i)U^{n,(\bar{\mu}_{1})}(a_{1},i) and Vn,(μ¯2)(a2,j)V^{n,(\bar{\mu}_{2})}(a_{2},j) denote the codewords associated with Alice and Bob, generated using the above procedure, respectively, where

Un,(μ¯1)(a1,i)\displaystyle U^{n,(\bar{\mu}_{1})}(a_{1},i) \ensurestackMath\stackon[1pt]=Δa1G+h1(μ¯1)(i) and\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}a_{1}G+h_{1}^{(\bar{\mu}_{1})}(i)\quad\text{ and }
Vn,(μ¯2)(a2,j)\displaystyle V^{n,(\bar{\mu}_{2})}(a_{2},j) \ensurestackMath\stackon[1pt]=Δa2G+h2(μ¯2)(j).\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}a_{2}G+h_{2}^{(\bar{\mu}_{2})}(j).

Now, construct the operators

A¯un(μ¯1)\displaystyle\bar{A}^{(\bar{\mu}_{1})}_{u^{n}} \ensurestackMath\stackon[1pt]=Δαun(ρA1ρ~unAρA1) and\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\alpha_{u^{n}}\bigg{(}\sqrt{\rho_{A}}^{-1}\tilde{\rho}_{u^{n}}^{A}\sqrt{\rho_{A}}^{-1}\bigg{)}\quad\text{ and }
B¯vn(μ¯2)\displaystyle\bar{B}^{(\bar{\mu}_{2})}_{v^{n}} \ensurestackMath\stackon[1pt]=Δβvn(ρB1ρ~vnBρB1),\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\beta_{v^{n}}\bigg{(}\sqrt{\rho_{B}}^{-1}\tilde{\rho}_{v^{n}}^{B}\sqrt{\rho_{B}}^{-1}\bigg{)}, (37)

where

αun\displaystyle\alpha_{u^{n}} \ensurestackMath\stackon[1pt]=Δ1(1+η)pnpk+l1λunA,βvn\ensurestackMath\stackon[1pt]=Δ1(1+η)pnpk+l2λvnB,\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\frac{1}{(1+\eta)}\frac{p^{n}}{p^{k+l_{1}}}\lambda_{u^{n}}^{A},\quad\beta_{v^{n}}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\frac{1}{(1+\eta)}\frac{p^{n}}{p^{k+l_{2}}}\lambda_{v^{n}}^{B}, (38)

with η(0,1)\eta\in(0,1) being a parameter to be determined. Having constructed the operators A¯un(μ¯1)\bar{A}^{(\bar{\mu}_{1})}_{u^{n}} and B¯vn(μ¯2)\bar{B}^{(\bar{\mu}_{2})}_{v^{n}}, we normalize these operators, so that they constitute a valid sub-POVM. To do so, we first define

ΣA(μ¯1)\displaystyle\Sigma_{A}^{(\bar{\mu}_{1})} \ensurestackMath\stackon[1pt]=Δunγun(μ¯1)A¯un(μ¯1) and\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\sum_{u^{n}}\gamma_{u^{n}}^{(\bar{\mu}_{1})}\bar{A}^{(\bar{\mu}_{1})}_{u^{n}}\quad\text{ and }
ΣB(μ¯2)\displaystyle\Sigma_{B}^{(\bar{\mu}_{2})} \ensurestackMath\stackon[1pt]=Δvnζvn(μ¯2)B¯vn(μ¯2),\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\sum_{v^{n}}\zeta_{v^{n}}^{(\bar{\mu}_{2})}\bar{B}^{(\bar{\mu}_{2})}_{v^{n}},

where γun(μ¯1)\gamma_{u^{n}}^{(\bar{\mu}_{1})} and ζvn(μ¯2)\zeta_{v^{n}}^{(\bar{\mu}_{2})} are defined as

γun(μ¯1)\displaystyle\gamma_{u^{n}}^{(\bar{\mu}_{1})} \ensurestackMath\stackon[1pt]=Δ|{(a1,i):Un,(μ¯1)(a1,i)=un}| and\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}|\{(a_{1},i):U^{n,(\bar{\mu}_{1})}(a_{1},i)=u^{n}\}|\quad\text{ and }
ζvn(μ¯2)\displaystyle\zeta_{v^{n}}^{(\bar{\mu}_{2})} \ensurestackMath\stackon[1pt]=Δ|{(a2,j):Vn,(μ¯2)(a2,j)=vn}|.\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}|\{(a_{2},j):V^{n,(\bar{\mu}_{2})}(a_{2},j)=v^{n}\}|.

Now, we define ΠAμ¯1\Pi_{A}^{\bar{\mu}_{1}} and ΠBμ¯2\Pi_{B}^{\bar{\mu}_{2}} as pruning operators for ΣA(μ¯1)\Sigma_{A}^{(\bar{\mu}_{1})} and ΣB(μ¯2),\Sigma_{B}^{(\bar{\mu}_{2})}, with respect to ΠρA\Pi_{\rho_{A}} and ΠρB\Pi_{\rho_{B}}, respectively (see Definition 5). Note that, these pruning operators, ΠAμ¯1\Pi_{A}^{\bar{\mu}_{1}} and ΠBμ¯2\Pi_{B}^{\bar{\mu}_{2}}, depend on the triple (G,h1(μ¯1),h2(μ¯2))(G,h_{1}^{(\bar{\mu}_{1})},h_{2}^{(\bar{\mu}_{2})}). Using these pruning operators, for each μ¯1[1,N¯1]\bar{\mu}_{1}\in[1,\bar{N}_{1}] and μ¯2[1,N¯2]\bar{\mu}_{2}\in[1,\bar{N}_{2}], construct the sub-POVMs M1(n,μ¯1)M_{1}^{(n,\bar{\mu}_{1})} and M2(n,μ¯2)M_{2}^{(n,\bar{\mu}_{2})} as

M1(n,μ¯1)\displaystyle M_{1}^{(n,\bar{\mu}_{1})} \ensurestackMath\stackon[1pt]=Δ{γun(μ¯1)Aun(μ¯1):un𝒰n}, and\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\{\gamma_{u^{n}}^{(\bar{\mu}_{1})}A^{(\bar{\mu}_{1})}_{u^{n}}:u^{n}\in\mathcal{U}^{n}\},\quad\text{ and }
M2(n,μ¯2)\displaystyle M_{2}^{(n,\bar{\mu}_{2})} \ensurestackMath\stackon[1pt]=Δ{ζvn(μ¯2)Bvn(μ¯2):vn𝒱n},\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\{\zeta_{v^{n}}^{(\bar{\mu}_{2})}B^{(\bar{\mu}_{2})}_{v^{n}}:v^{n}\in\mathcal{V}^{n}\}, (39)

where Aun(μ¯1)=ΠAμ1A¯un(μ¯1)ΠAμ1A^{(\bar{\mu}_{1})}_{u^{n}}=\Pi_{A}^{\mu_{1}}\bar{A}^{(\bar{\mu}_{1})}_{u^{n}}\Pi_{A}^{\mu_{1}} and Bvn(μ¯2)=ΠBμ2B¯vn(μ¯2)ΠBμ2B^{(\bar{\mu}_{2})}_{v^{n}}=\Pi_{B}^{\mu_{2}}\bar{B}^{(\bar{\mu}_{2})}_{v^{n}}\Pi_{B}^{\mu_{2}}. Further, using these operators ΠAμ¯1\Pi_{A}^{\bar{\mu}_{1}} and ΠBμ¯2,\Pi_{B}^{\bar{\mu}_{2}}, we have unγun(μ¯1)Aun(μ¯1)=ΠAμ¯1ΣA(μ¯1)ΠAμ¯1I\sum_{u^{n}}\gamma_{u^{n}}^{(\bar{\mu}_{1})}A^{(\bar{\mu}_{1})}_{u^{n}}=\Pi_{A}^{\bar{\mu}_{1}}\Sigma_{A}^{(\bar{\mu}_{1})}\Pi_{A}^{\bar{\mu}_{1}}\leq I and vnζvn(μ¯2)Bvn(μ¯2)=ΠBμ¯2ΣB(μ¯2)ΠBμ¯2I,\sum_{v^{n}}\zeta_{v^{n}}^{(\bar{\mu}_{2})}B^{(\bar{\mu}_{2})}_{v^{n}}=\Pi_{B}^{\bar{\mu}_{2}}\Sigma_{B}^{(\bar{\mu}_{2})}\Pi_{B}^{\bar{\mu}_{2}}\leq I, and thus M1(n,μ¯1)M_{1}^{(n,\bar{\mu}_{1})} and M2(n,μ¯2)M_{2}^{(n,\bar{\mu}_{2})} are valid sub-POVMs for all μ¯1[1,N¯1]\bar{\mu}_{1}\in[1,\bar{N}_{1}] and μ¯2[1,N¯2].\bar{\mu}_{2}\in[1,\bar{N}_{2}]. Further, these collections M1(n,μ¯1)M_{1}^{(n,\bar{\mu}_{1})} and M2(n,μ¯2)M_{2}^{(n,\bar{\mu}_{2})} are completed using the operators Iun𝒰nγun(μ¯1)Aun(μ¯1)I-\sum_{u^{n}\in\mathcal{U}^{n}}\gamma_{u^{n}}^{(\bar{\mu}_{1})}A^{(\bar{\mu}_{1})}_{u^{n}} and Ivn𝒱nζvn(μ¯2)Bvn(μ¯2)I-\sum_{v^{n}\in\mathcal{V}^{n}}\zeta_{v^{n}}^{(\bar{\mu}_{2})}B^{(\bar{\mu}_{2})}_{v^{n}}.

VI.2 Binning of POVMs

We next proceed to binning the above constructed collection of sub-POVMs. Since, UCC is already a union of several cosets, we associate a bin to each coset, and hence place all the codewords of a coset in the same bin. For each i𝔽pl1i\in\mathbb{F}_{p}^{l_{1}} and j𝔽pl2j\in\mathbb{F}_{p}^{l_{2}}, let 1(μ¯1)(i)\ensurestackMath\stackon[1pt]=Δ(G,h1(μ¯1)(i))\mathcal{B}^{(\bar{\mu}_{1})}_{1}(i)\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\mathbb{C}(G,h_{1}^{(\bar{\mu}_{1})}(i)) and 2(μ¯2)(j)\ensurestackMath\stackon[1pt]=Δ(G,h2(μ¯2)(j))\mathcal{B}^{(\bar{\mu}_{2})}_{2}(j)\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\mathbb{C}(G,h_{2}^{(\bar{\mu}_{2})}(j)) denote the ithi^{th} and the jthj^{th} bins, respectively. Formally, we define the following operators:

ΓiA,(μ¯1)\displaystyle\Gamma^{A,(\bar{\mu}_{1})}_{i} \ensurestackMath\stackon[1pt]=Δun𝒰na1𝔽pkAun(μ¯1)𝟙{a1G+h1(μ¯1)(i)=un},\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\sum_{u^{n}\in\mathcal{U}^{n}}\sum_{a_{1}\in\mathbb{F}_{p}^{k}}A^{(\bar{\mu}_{1})}_{u^{n}}\mathbbm{1}_{\{a_{1}G+h_{1}^{(\bar{\mu}_{1})}(i)=u^{n}\}},
ΓjB,(μ¯2)\displaystyle\Gamma^{B,(\bar{\mu}_{2})}_{j} \ensurestackMath\stackon[1pt]=Δvn𝒱na2𝔽pk2Bvn(μ¯2)𝟙{a2G+h2(μ¯2)(j)=vn},\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\sum_{v^{n}\in\mathcal{V}^{n}}\sum_{a_{2}\in\mathbb{F}_{p}^{k_{2}}}B^{(\bar{\mu}_{2})}_{v^{n}}\mathbbm{1}_{\{a_{2}G+h_{2}^{(\bar{\mu}_{2})}(j)=v^{n}\}},

for all i𝔽pl1i\in\mathbb{F}_{p}^{l_{1}} and j𝔽pl2j\in\mathbb{F}_{p}^{l_{2}}. Using these operators, we form the following collection:

MA(n,μ¯1)\ensurestackMath\stackon[1pt]=Δ{ΓiA,(μ¯1)}i𝔽pl1,MB(n,μ¯)\ensurestackMath\stackon[1pt]=Δ{ΓjB,(μ¯2)}j𝔽pl2.\displaystyle M_{A}^{(n,\bar{\mu}_{1})}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\{\Gamma^{A,(\bar{\mu}_{1})}_{i}\}_{i\in\mathbb{F}_{p}^{l_{1}}},\quad M_{B}^{(n,\bar{\mu})}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\{\Gamma^{B,(\bar{\mu}_{2})}_{j}\}_{j\in\mathbb{F}_{p}^{l_{2}}}. (40)

Note that if M1(n,μ¯1)M_{1}^{(n,\bar{\mu}_{1})} and M2(n,μ¯2)M_{2}^{(n,\bar{\mu}_{2})} are sub-POVMs, then so are MA(n,μ¯1)M_{A}^{(n,\bar{\mu}_{1})} and MB(n,μ¯2)M_{B}^{(n,\bar{\mu}_{2})}, which is due to the relations

i𝔽pl1ΓiA,(μ¯1)\displaystyle\sum_{i\in\mathbb{F}_{p}^{l_{1}}}\Gamma^{A,(\bar{\mu}_{1})}_{i} =un𝒰nγun(μ1)Aun(μ¯1)I,and\displaystyle=\sum_{u^{n}\in\mathcal{U}^{n}}\gamma_{u^{n}}^{(\mu_{1})}A^{(\bar{\mu}_{1})}_{u^{n}}\leq I,\quad\text{and}
j𝔽pl2ΓjB,(μ¯2)\displaystyle\sum_{j\in\mathbb{F}_{p}^{l_{2}}}\Gamma^{B,(\bar{\mu}_{2})}_{j} =vn𝒱nζvn(μ2)Bvn(μ¯2)I.\displaystyle=\sum_{v^{n}\in\mathcal{V}^{n}}\zeta_{v^{n}}^{(\mu_{2})}B^{(\bar{\mu}_{2})}_{v^{n}}\leq I. (41)

To make MA(n,μ¯1)M_{A}^{(n,\bar{\mu}_{1})} and MB(n,μ¯2)M_{B}^{(n,\bar{\mu}_{2})} complete, we define Γ0A,(μ¯1)\Gamma^{A,(\bar{\mu}_{1})}_{0} and Γ0B,(μ¯2)\Gamma^{B,(\bar{\mu}_{2})}_{0} as Γ0A,(μ¯1)=IiΓiA,(μ¯1)\Gamma^{A,(\bar{\mu}_{1})}_{0}=I-\sum_{i}\Gamma^{A,(\bar{\mu}_{1})}_{i} and Γ0B,(μ¯2)=IjΓjB,(μ¯2)\Gamma^{B,(\bar{\mu}_{2})}_{0}=I-\sum_{j}\Gamma^{B,(\bar{\mu}_{2})}_{j}, respectively333Note that Γ0A,(μ¯1)=IiΓiA,(μ¯1)=IunTδ(n)(U)Aun(μ¯1)\Gamma^{A,(\bar{\mu}_{1})}_{0}=I-\sum_{i}\Gamma^{A,(\bar{\mu}_{1})}_{i}=I-\sum_{u^{n}\in T_{\delta}^{(n)}(U)}A^{(\bar{\mu}_{1})}_{u^{n}} and Γ0B,(μ¯2)=IjΓjB,(μ¯2)=IvnTδ(n)(V)Bvn(μ¯2)\Gamma^{B,(\bar{\mu}_{2})}_{0}=I-\sum_{j}\Gamma^{B,(\bar{\mu}_{2})}_{j}=I-\sum_{v^{n}\in T_{\delta}^{(n)}(V)}B^{(\bar{\mu}_{2})}_{v^{n}}.. Now, we intend to use the completions [MA(n,μ¯1)][M_{A}^{(n,\bar{\mu}_{1})}] and [MB(n,μ¯2)][M_{B}^{(n,\bar{\mu}_{2})}] as the POVMs for encoders associated with Alice and Bob, respectively. Also, note that the effect of the binning is in reducing the communication rates from (k+l1nlog(p),k+l2nlog(p))(\frac{k+l_{1}}{n}\log{p},\frac{k+l_{2}}{n}\log{p}) to (R1,R2)(R_{1},R_{2}), where Ri\ensurestackMath\stackon[1pt]=Δlinlog(p),i{1,2}R_{i}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\frac{l_{i}}{n}\log{p},i\in\{1,2\}. Now, we move on to describing the decoder.

VI.3 Decoder mapping

We create a decoder that takes as an input a pair of bin numbers and produces a sequence Wn𝔽pnW^{n}\in\mathbb{F}^{n}_{p}. More precisely, we define a mapping F(μ¯1,μ¯2)F^{(\bar{\mu}_{1},\bar{\mu}_{2})}, acting on the outputs of [MA(n,μ¯1)][MB(n,μ¯2)][M_{A}^{(n,\bar{\mu}_{1})}]\otimes[M_{B}^{(n,\bar{\mu}_{2})}] as follows. On observing (μ¯1,μ¯2)(\bar{\mu}_{1},\bar{\mu}_{2}) and the classical indices (i,j)𝔽pl1×𝔽pl2(i,j)\in\mathbb{F}_{p}^{l_{1}}\times\mathbb{F}_{p}^{l_{2}} communicated by the encoder, the decoder constructs D(μ¯1,μ¯2)D^{(\bar{\mu}_{1},\bar{\mu}_{2})} and F(μ¯1,μ¯2)(,)F^{(\bar{\mu}_{1},\bar{\mu}_{2})}(\cdot,\cdot) as,

Di,j(μ¯1,μ¯2)\displaystyle D^{(\bar{\mu}_{1},\bar{\mu}_{2})}_{i,j} \ensurestackMath\stackon[1pt]=Δ{a~𝔽pk:a~G+h1(μ¯1)(i)+h2(μ¯2)(j)𝒯δ^(n)(W)},\displaystyle\!\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\!\Big{\{}\!\tilde{a}\in\mathbb{F}_{p}^{k}:\tilde{a}G+h_{1}^{(\bar{\mu}_{1})}(i)+h_{2}^{(\bar{\mu}_{2})}(j)\!\in\!\mathcal{T}_{\hat{\delta}}^{(n)}(W)\!\Big{\}},
F(μ¯1,μ¯2)\displaystyle F^{(\bar{\mu}_{1},\bar{\mu}_{2})} (i,j)\ensurestackMath\stackon[1pt]=Δ\displaystyle(i,j)\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}
{a~G+h1(μ¯1)(i)+h2(μ¯2)(j) if Di,j(μ¯1,μ¯2){a~}w0n otherwise ,\displaystyle\!\!\!\begin{cases}\tilde{a}G+h_{1}^{(\bar{\mu}_{1})}(i)+h_{2}^{(\bar{\mu}_{2})}(j)&\;\text{ if }D^{(\bar{\mu}_{1},\bar{\mu}_{2})}_{i,j}\equiv\{\tilde{a}\}\\ w^{n}_{0}&\;\text{ otherwise },\end{cases} (42)

where δ^=pδ\hat{\delta}=p\delta and w0nw_{0}^{n} is an arbitrary sequence in 𝔽pn\𝒯δ^(n)(W)\mathbb{F}_{p}^{n}\backslash\mathcal{T}_{\hat{\delta}}^{(n)}(W). Further, F(μ¯1,μ¯2)(i,j)=w0nF^{(\bar{\mu}_{1},\bar{\mu}_{2})}(i,j)=w_{0}^{n} for i=0i=0 or j=0j=0. Given this, we obtain the sub-POVM M~AB\tilde{M}_{AB} with the following operators.

Λ~wnAB\ensurestackMath\stackon[1pt]=Δ1N¯1N¯2μ¯1=1N¯1μ¯2=1N¯2(i,j):F(μ¯1,μ¯2)(i,j)=wnΓiA,(μ¯1)ΓjB,(μ¯2),\displaystyle\tilde{\Lambda}_{w^{n}}^{AB}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\frac{1}{\bar{N}_{1}\bar{N}_{2}}\sum_{\bar{\mu}_{1}=1}^{\bar{N}_{1}}\sum_{\bar{\mu}_{2}=1}^{\bar{N}_{2}}\sum_{(i,j):F^{(\bar{\mu}_{1},\bar{\mu}_{2})}(i,j)=w^{n}}\Gamma^{A,(\bar{\mu}_{1})}_{i}\otimes\Gamma^{B,(\bar{\mu}_{2})}_{j},

wn𝔽pn{w0n}.\forall w^{n}\in\mathbb{F}_{p}^{n}\operatorname*{\mathbin{\scalebox{1.0}{$\bigcup$}}}\{w_{0}^{n}\}. Now, we use the stochastic mapping 𝖯Z|W\mathsf{P}_{Z|W} to define the approximating sub-POVM M^AB(n)\ensurestackMath\stackon[1pt]=Δ{Λ^zn}\hat{M}^{(n)}_{AB}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\{\hat{\Lambda}_{z^{n}}\} as

Λ^znAB\ensurestackMath\stackon[1pt]=ΔwnΛ~wnABPZ|Wn(zn|wn),zn𝒵n.\displaystyle\hat{\Lambda}^{AB}_{z^{n}}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\sum_{w^{n}}\tilde{\Lambda}_{w^{n}}^{AB}P^{n}_{Z|W}(z^{n}|w^{n}),~{}\forall z^{n}\in\mathcal{Z}^{n}.

Note that Λ~wnAB=0\tilde{\Lambda}_{w^{n}}^{AB}=0 for wn𝒯δ(n)(W){w0n}.w^{n}\notin\mathcal{T}_{\delta}^{(n)}(W)\operatorname*{\mathbin{\scalebox{1.0}{$\bigcup$}}}\{w^{n}_{0}\}.

UCC Grand Ensemble: The generator matrix GG and the functions h1(μ¯1)h_{1}^{(\bar{\mu}_{1})} and h2(μ¯2)h_{2}^{(\bar{\mu}_{2})} are chosen randomly uniformly and independently, for μ¯1[1,N¯1]\bar{\mu}_{1}\in[1,\bar{N}_{1}] and μ¯2[1,N¯2].\bar{\mu}_{2}\in[1,\bar{N}_{2}].

VI.4 Trace Distance

In what follows, we show that M^AB(n)\hat{M}_{AB}^{(n)} is ϵ\epsilon-faithful to MABnM_{AB}^{\otimes n} with respect to ρABn\rho_{AB}^{\otimes n} (according to Definition 1), where ϵ>0\epsilon>0 can be made arbitrarily small. More precisely, using (35), we show that, 𝔼[K]ϵ,\mathbb{E}[K]\leq\epsilon, where

K\displaystyle{K} \ensurestackMath\stackon[1pt]=Δznun,vnρABn(Λ¯unAΛ¯vnB)ρABn\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\sum_{z^{n}}\bigg{\|}\sum_{u^{n},v^{n}}\sqrt{\rho_{AB}^{\otimes n}}(\bar{\Lambda}^{A}_{u^{n}}\otimes\bar{\Lambda}^{B}_{v^{n}})\sqrt{\rho_{AB}^{\otimes n}}
×PZ|Wn(zn|un+vn)ρABnΛ^znABρABn1,\displaystyle\hskip 28.90755pt\times P^{n}_{Z|W}(z^{n}|u^{n}+v^{n})-\sqrt{\rho_{AB}^{\otimes n}}\hat{\Lambda}_{z^{n}}^{AB}\sqrt{\rho_{AB}^{\otimes n}}\bigg{\|}_{1}, (43)

and the expectation is with respect to the codebook generation.

Step 1: Isolating the effect of error induced by not covering
Consider the second term within K{K}, which can be written as

wn\displaystyle\sum_{w^{n}} ρABnΛ~wnABρABnPZ|Wn(zn|wn)\displaystyle\sqrt{\rho_{AB}^{\otimes n}}\tilde{\Lambda}_{w^{n}}^{AB}\sqrt{\rho_{AB}^{\otimes n}}P^{n}_{Z|W}(z^{n}|w^{n})
=1N¯1N¯2μ¯1,μ¯2i,jρABn(ΓiA,(μ¯1)ΓjB,(μ¯2))ρABn\displaystyle=\frac{1}{\bar{N}_{1}\bar{N}_{2}}\sum_{\bar{\mu}_{1},\bar{\mu}_{2}}\sum_{i,j}\sqrt{\rho_{AB}^{\otimes n}}\left(\Gamma^{A,(\bar{\mu}_{1})}_{i}\otimes\Gamma^{B,(\bar{\mu}_{2})}_{j}\right)\sqrt{\rho_{AB}^{\otimes n}}
×PZ|Wn(zn|F(μ¯1,μ¯2)(i,j))wn𝟙{F(μ¯1,μ¯2)(i,j)=wn}=1\displaystyle\hskip 23.12692pt\times P^{n}_{Z|W}(z^{n}|F^{(\bar{\mu}_{1},\bar{\mu}_{2})}(i,j))\underbrace{\sum_{w^{n}}\mathbbm{1}_{\{F^{(\bar{\mu}_{1},\bar{\mu}_{2})}(i,j)=w^{n}\}}}_{=1}
=T+T~,\displaystyle=T+\widetilde{T},

where

T\ensurestackMath\stackon[1pt]=Δ\displaystyle T\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}} 1N¯1N¯2μ¯1,μ¯2{i>0}{j>0}ρABn(ΓiA,(μ¯1)ΓjB,(μ¯2))\displaystyle\frac{1}{\bar{N}_{1}\bar{N}_{2}}\sum_{\bar{\mu}_{1},\bar{\mu}_{2}}\sum_{\{i>0\}\operatorname*{\mathbin{\scalebox{1.0}{$\bigcap$}}}\{j>0\}}\sqrt{\rho_{AB}^{\otimes n}}\left(\Gamma^{A,(\bar{\mu}_{1})}_{i}\otimes\Gamma^{B,(\bar{\mu}_{2})}_{j}\right)
×ρABnPZ|Wn(zn|F(μ¯1,μ¯2)(i,j)),\displaystyle\hskip 79.49744pt\times\sqrt{\rho_{AB}^{\otimes n}}P^{n}_{Z|W}(z^{n}|F^{(\bar{\mu}_{1},\bar{\mu}_{2})}(i,j)),
T~\ensurestackMath\stackon[1pt]=Δ\displaystyle\widetilde{T}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}} 1N¯1N¯2μ¯1,μ¯2{i=0}{j=0}ρABn(ΓiA,(μ¯1)ΓjB,(μ¯2))\displaystyle\frac{1}{\bar{N}_{1}\bar{N}_{2}}\sum_{\bar{\mu}_{1},\bar{\mu}_{2}}\sum_{\{i=0\}\operatorname*{\mathbin{\scalebox{1.0}{$\bigcup$}}}\{j=0\}}\sqrt{\rho_{AB}^{\otimes n}}\left(\Gamma^{A,(\bar{\mu}_{1})}_{i}\otimes\Gamma^{B,(\bar{\mu}_{2})}_{j}\right)
×ρABnPZ|Wn(zn|w0n).\displaystyle\hskip 122.85876pt\times\sqrt{\rho_{AB}^{\otimes n}}P^{n}_{Z|W}(z^{n}|w^{n}_{0}).

Hence, we have

KS+S~,\displaystyle K\leq S+\widetilde{S}, (44)

where

S\displaystyle S \ensurestackMath\stackon[1pt]=Δznun,vnρABn(Λ¯unAΛ¯vnB\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\sum_{z^{n}}\bigg{\|}\sum_{u^{n},v^{n}}\sqrt{\rho_{AB}^{\otimes n}}\bigg{(}\bar{\Lambda}^{A}_{u^{n}}\otimes\bar{\Lambda}^{B}_{v^{n}}
×PZ|Wn(zn|un+vn))ρABnT1,\displaystyle\hskip 50.58878pt\times P^{n}_{Z|W}(z^{n}|u^{n}+v^{n})\bigg{)}\sqrt{\rho_{AB}^{\otimes n}}-T\bigg{\|}_{1}, (45)

and S~\ensurestackMath\stackon[1pt]=ΔznT~1\widetilde{S}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\sum_{z^{n}}\|\widetilde{T}\|_{1}. Note that S~\widetilde{S} captures the error induced by not covering the state ρABn.\rho_{AB}^{\otimes n}.

Remark 8.

The terms corresponding to the operators that complete the sub-POVMs MA(n,μ¯1)M_{A}^{(n,\bar{\mu}_{1})} and MB(n,μ¯2)M_{B}^{(n,\bar{\mu}_{2})}, i.e., Iun𝒯δ(n)(U)γun(μ¯1)Aun(μ¯1)I-\sum_{u^{n}\in\mathcal{T}_{\delta}^{(n)}(U)}\gamma_{u^{n}}^{(\bar{\mu}_{1})}A_{u^{n}}^{(\bar{\mu}_{1})} and Ivn𝒯δ(n)(V)ζvn(μ¯2)Bvn(μ¯2)I-\sum_{v^{n}\in\mathcal{T}_{\delta}^{(n)}(V)}\zeta_{v^{n}}^{(\bar{\mu}_{2})}B_{v^{n}}^{(\bar{\mu}_{2})} are taken care in T~\widetilde{T}. The expression TT excludes these completing operators.

Step 2: Isolating the effect of error induced by binning
 We begin by simplifying TT as

T=\displaystyle T= 1N¯1N¯2μ¯1,μ¯2un,vni>0,j>0ρABn\displaystyle\frac{1}{\bar{N}_{1}\bar{N}_{2}}\sum_{\bar{\mu}_{1},\bar{\mu}_{2}}\sum_{u^{n},v^{n}}\sum_{\begin{subarray}{c}i>0,\\ j>0\end{subarray}}\sqrt{\rho_{AB}^{\otimes n}}
×(a1𝔽pka2𝔽pkAun(μ¯1)Bvn(μ¯2)𝟙{a1G+h1(μ¯1)(i)=un}\displaystyle\times\!\bigg{(}\sum_{a_{1}\in\mathbb{F}_{p}^{k}}\sum_{a_{2}\in\mathbb{F}_{p}^{k}}A_{u^{n}}^{(\bar{\mu}_{1})}\otimes B_{v^{n}}^{(\bar{\mu}_{2})}\mathbbm{1}_{\{\begin{subarray}{c}a_{1}G+h_{1}^{(\bar{\mu}_{1})}(i)=u^{n}\end{subarray}\}}
×𝟙{a2G+h2(μ¯2)(j)=vn})ρABnPnZ|W(zn|F(μ¯1,μ¯2)(i,j)).\displaystyle\times\!\mathbbm{1}_{\{\begin{subarray}{c}a_{2}G+h_{2}^{(\bar{\mu}_{2})}(j)=v^{n}\end{subarray}\}}\!\bigg{)}\!\sqrt{\rho_{AB}^{\otimes n}}P^{n}_{Z|W}(z^{n}|F^{(\bar{\mu}_{1},\bar{\mu}_{2})}(i,j)).

Note that the (un,vn)(u^{n},v^{n}) that appear in the above summation is confined to (𝒯δ(n)(U)×𝒯δ(n)(V))(\mathcal{T}_{\delta}^{(n)}(U)\times\mathcal{T}_{\delta}^{(n)}(V)), however for ease of notation, we do not make this explicit. We substitute the above expression into SS as in (45), and add and subtract an appropriate term within SS and apply the triangle inequality to isolate the effect of binning as SS1+S2,S\leq S_{1}+S_{2}, where

S1\displaystyle S_{1} \ensurestackMath\stackon[1pt]=Δznun,vnρABn(Λ¯unAΛ¯vnB1N¯1N¯2μ¯1,μ¯2\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\sum_{z^{n}}\bigg{\|}\sum_{u^{n},v^{n}}\sqrt{\rho_{AB}^{\otimes n}}\bigg{(}\bar{\Lambda}^{A}_{u^{n}}\otimes\bar{\Lambda}^{B}_{v^{n}}-\frac{1}{\bar{N}_{1}\bar{N}_{2}}\sum_{\bar{\mu}_{1},\bar{\mu}_{2}}
×γun(μ¯1)Aun(μ¯1)ζvn(μ¯2)Bvn(μ¯2))ρABnPnZ|W(zn|un+vn)1,\displaystyle\hskip 5.0pt\times\gamma_{u^{n}}^{(\bar{\mu}_{1})}\!A_{u^{n}}^{(\bar{\mu}_{1})}\!\otimes\zeta_{v^{n}}^{(\bar{\mu}_{2})}\!B_{v^{n}}^{(\bar{\mu}_{2})}\!\bigg{)}\sqrt{\rho_{AB}^{\otimes n}}P^{n}_{Z|W}(z^{n}|u^{n}+v^{n})\bigg{\|}_{1}\!\!,
S2\displaystyle S_{2} \ensurestackMath\stackon[1pt]=Δzn1N¯1N¯2μ¯1,μ¯2i>0j>0a1,a2un,vnρABn(Aun(μ¯1)\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\sum_{z^{n}}\bigg{\|}\frac{1}{\bar{N}_{1}\bar{N}_{2}}\!\!\sum_{\bar{\mu}_{1},\bar{\mu}_{2}}\sum_{\begin{subarray}{c}i>0\\ j>0\end{subarray}}\sum_{a_{1},a_{2}}\sum_{u^{n},v^{n}}\!\!\sqrt{\rho_{AB}^{\otimes n}}\!\left(\!A_{u^{n}}^{(\bar{\mu}_{1})}\right.
Bvn(μ¯2))ρABn𝟙{a1G+h1(μ¯1)(i)=un,a2G+h2(μ¯2)(j)=vn}\displaystyle\hskip 10.0pt\left.\otimes B_{v^{n}}^{(\bar{\mu}_{2})}\right)\!\sqrt{\rho_{AB}^{\otimes n}}\mathbbm{1}_{\{a_{1}G+h_{1}^{(\bar{\mu}_{1})}(i)=u^{n},a_{2}G+h_{2}^{(\bar{\mu}_{2})}(j)=v^{n}\}}
×(PZ|Wn(zn|un+vn)PZ|Wn(zn|F(μ¯1,μ¯2)(i,j)))1.\displaystyle\hskip 15.0pt\times\left(\!P^{n}_{Z|W}(z^{n}|u^{n}\!+\!v^{n})-P^{n}_{Z|W}\!\left(\!z^{n}|F^{(\bar{\mu}_{1},\bar{\mu}_{2})}(i,j)\!\right)\!\right)\!\!\bigg{\|}_{1}\!.

Note that the term S1S_{1} characterizes the error introduced by approximation of the original POVM with the collection of approximating sub-POVMs M1(n,μ¯1)M_{1}^{(n,\bar{\mu}_{1})} and M2(n,μ¯2)M_{2}^{{(n,\bar{\mu}_{2})}}, and the term S2S_{2} characterizes the error caused by binning of these approximating sub-POVMs. In this step, we analyze S2S_{2} and prove the following proposition.

Proposition 4 (Mutual Packing).

For any ϵ(0,1)\epsilon\in(0,1), any η,δ(0,1)\eta,\delta\in(0,1) sufficiently small, and any nn sufficiently large, we have 𝔼[S2]ϵ\mathbb{E}\left[{S}_{2}\right]\leq\epsilon, if k+l1nlogp>I(U;RB)σ1S(U)σ3+logp\frac{k+l_{1}}{n}\log p>I(U;RB)_{\sigma_{1}}-S(U)_{\sigma_{3}}+\log p, k+l2nlogp>I(V;RA)σ2S(V)σ3+logp\frac{k+l_{2}}{n}\log p>I(V;RA)_{\sigma_{2}}-S(V)_{\sigma_{3}}+\log p, k+l1nlogp+1nlogN¯1>logp\frac{k+l_{1}}{n}\log p+\frac{1}{n}\log\bar{N}_{1}>\log p, k+l2nlogp+1nlogN¯2>logp\frac{k+l_{2}}{n}\log p+\frac{1}{n}\log\bar{N}_{2}>\log p, knlog(p)<log(p)S(W)σ3\frac{k}{n}\log{p}<\log{p}-S(W)_{\sigma_{3}}, where σ1,σ2\sigma_{1},\sigma_{2} and σ3\sigma_{3} are the auxiliary states as defined in the statement of the theorem.

Proof.

The proof is provided in Appendix B.4

Since averaged over μ~1[1,N~1],μ~2[1,N~2]\tilde{\mu}_{1}\in[1,\tilde{N}_{1}],\tilde{\mu}_{2}\in[1,\tilde{N}_{2}], the quantity 𝔼[S2]\mathbb{E}[S_{2}] can be made arbitrarily small, there must exist a pair (μ~1,μ~2)(\tilde{\mu}_{1},\tilde{\mu}_{2}) such that 𝔼[S2]\mathbb{E}[S_{2}] is small for this pair of (μ~1,μ~2)(\tilde{\mu}_{1},\tilde{\mu}_{2}). For the rest of the proof, we fix (μ~1,μ~2)(\tilde{\mu}_{1},\tilde{\mu}_{2}) to be this pair. The dependence of functions defined in the sequel on this pair is not made explicit for ease of notation. For the term corresponding to S~\widetilde{S}, we prove the following result.

Proposition 5.

For any ϵ(0,1)\epsilon\in(0,1), any η,δ(0,1)\eta,\delta\in(0,1) sufficiently small, and any nn sufficiently large, we have 𝔼[S~]ϵ\mathbb{E}[\widetilde{S}]\leq\epsilon, if k+l1nlog(p)>I(U;RB)σ1S(U)σ1+log(p)\frac{k+l_{1}}{n}\log{p}>I(U;RB)_{\sigma_{1}}-S(U)_{\sigma_{1}}+\log{p} and k+l2nlog(p)>I(V;RA)σ2S(V)σ2+log(p),\frac{k+l_{2}}{n}\log{p}>I(V;RA)_{\sigma_{2}}-S(V)_{\sigma_{2}}+\log{p}, where σ1\sigma_{1} and σ2\sigma_{2} are auxiliary states defined in the statement of the theorem.

Proof.

The proof is provided in Appendix B.5. ∎

Step 3: Isolating the effect of Alice’s approximating measurement
In this step, we separately analyze the effect of approximating measurements at the two distributed parties in the term S1S_{1}. For that, we split S1S_{1} as S1Q1+Q2S_{1}\leq Q_{1}+Q_{2}, where

Q1\displaystyle Q_{1} \ensurestackMath\stackon[1pt]=Δznun,vnρABn(Λ¯unAΛ¯vnB1N1μ1=1N1γun(μ1)Aun(μ1)Λ¯vnB)ρABnPZ|Wn(zn|un+vn)1,\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\sum_{z^{n}}\bigg{\|}\!\sum_{u^{n},v^{n}}\!\!\!\sqrt{\rho_{AB}^{\otimes n}}\bigg{(}\!\bar{\Lambda}^{A}_{u^{n}}\otimes\bar{\Lambda}^{B}_{v^{n}}-\frac{1}{N_{1}}\!\sum_{\mu_{1}=1}^{N_{1}}\!\!\gamma_{u^{n}}^{(\mu_{1})}A_{u^{n}}^{(\mu_{1})}\otimes\bar{\Lambda}^{B}_{v^{n}}\bigg{)}\sqrt{\rho_{AB}^{\otimes n}}P^{n}_{Z|W}(z^{n}|u^{n}+v^{n})\bigg{\|}_{1}\!\!,
Q2\displaystyle Q_{2} \ensurestackMath\stackon[1pt]=Δzn1N1μ1=1N1un,vnρABn(γun(μ1)Aun(μ1)Λ¯vnB1N2μ2=1N2γun(μ1)Aun(μ1)ζvn(μ2)Bvn(μ2))ρABnPZ|Wn(zn|un+vn)1.\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\sum_{z^{n}}\left\|\frac{1}{N_{1}}\sum_{\mu_{1}=1}^{N_{1}}\sum_{u^{n},v^{n}}\sqrt{\rho_{AB}^{\otimes n}}\left(\gamma_{u^{n}}^{(\mu_{1})}A_{u^{n}}^{(\mu_{1})}\otimes\bar{\Lambda}^{B}_{v^{n}}-\frac{1}{N_{2}}\sum_{\mu_{2}=1}^{N_{2}}\gamma_{u^{n}}^{(\mu_{1})}A_{u^{n}}^{(\mu_{1})}\otimes\zeta_{v^{n}}^{(\mu_{2})}B_{v^{n}}^{(\mu_{2})}\right)\sqrt{\rho_{AB}^{\otimes n}}P^{n}_{Z|W}(z^{n}|u^{n}+v^{n})\right\|_{1}. (46)

With this partition, the terms within the trace norm of Q1Q_{1} differ only in the action of Alice’s measurement. And similarly, the terms within the norm of Q2Q_{2} differ only in the action of Bob’s measurement. Showing that these two terms are small forms a major portion of the achievability proof.

Analysis of Q1Q_{1}: To prove Q1Q_{1} is small, we characterize the rate constraints which ensure that an upper bound to Q1Q_{1} can be made to vanish in an expected sense. In addition, this upper bound becomes lucrative in obtaining a single-letter characterization for the rate needed to make the term corresponding to Q2Q_{2} vanish. For this, we define JJ as

J\displaystyle J \ensurestackMath\stackon[1pt]=Δzn,vnunρABn(Λ¯unAΛ¯vnB\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\sum_{z^{n},v^{n}}\bigg{\|}\sum_{u^{n}}\sqrt{\rho_{AB}^{\otimes n}}\bigg{(}\bar{\Lambda}^{A}_{u^{n}}\otimes\bar{\Lambda}^{B}_{v^{n}}-
1N1μ1=1N1γun(μ1)Aun(μ1)Λ¯vnB)ρABnPnZ|W(zn|un+vn)1.\displaystyle\hskip 7.0pt\frac{1}{N_{1}}\!\sum_{\mu_{1}=1}^{N_{1}}\!\gamma_{u^{n}}^{(\mu_{1})}A_{u^{n}}^{(\mu_{1})}\otimes\bar{\Lambda}^{B}_{v^{n}}\bigg{)}\sqrt{\rho_{AB}^{\otimes n}}P^{n}_{Z|W}(z^{n}|u^{n}+v^{n})\bigg{\|}_{1}\!\!\!. (47)

By defining JJ and using triangle inequality for block operators (which holds with equality), we add the sub-system VV to RZRZ, resulting in the joint system RZVRZV, corresponding to the state σ3\sigma_{3} as defined in the theorem. Then we approximate the joint system RZVRZV using an approximating sub-POVM MA(n)M_{A}^{(n)} producing outputs on the alphabet 𝒰n\mathcal{U}^{n}. To make JJ small for sufficiently large n, we expect the sum of the rate of the approximating sub-POVM and common randomness, i.e., k+l1nlog(p)+1nlog(N1)\frac{k+l_{1}}{n}\log{p}+\frac{1}{n}\log{N_{1}}, to be larger than I(U;RZV)σ3I(U;RZV)_{\sigma_{3}}. We prove this in the following.

Note that from the triangle inequality, we have Q1J.Q_{1}\leq J. Further, we add and subtract appropriate terms within JJ, and again use the triangle inequality to obtain JJ1+J2J\leq J_{1}+J_{2}, where

J1\displaystyle J_{1} \ensurestackMath\stackon[1pt]=Δzn,vnunρABn(Λ¯unAΛ¯vnB\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\sum_{z^{n},v^{n}}\bigg{\|}\sum_{u^{n}}\sqrt{\rho_{AB}^{\otimes n}}\bigg{(}\bar{\Lambda}^{A}_{u^{n}}\otimes\bar{\Lambda}^{B}_{v^{n}}-
γun(μ1)A¯un(μ1)Λ¯vnB)ρABnPnZ|W(zn|un+vn)1,\displaystyle\hskip 33.0pt\gamma_{u^{n}}^{(\mu_{1})}\bar{A}_{u^{n}}^{(\mu_{1})}\otimes\bar{\Lambda}^{B}_{v^{n}}\bigg{)}\sqrt{\rho_{AB}^{\otimes n}}P^{n}_{Z|W}(z^{n}|u^{n}+v^{n})\bigg{\|}_{1},
J2\displaystyle J_{2} \ensurestackMath\stackon[1pt]=Δzn,vn1N1μ1=1N1unρABn(γun(μ1)A¯un(μ1)Λ¯vnB\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\sum_{z^{n},v^{n}}\bigg{\|}\frac{1}{N_{1}}\sum_{\mu_{1}=1}^{N_{1}}\sum_{u^{n}}\sqrt{\rho_{AB}^{\otimes n}}\bigg{(}\gamma_{u^{n}}^{(\mu_{1})}\bar{A}_{u^{n}}^{(\mu_{1})}\otimes\bar{\Lambda}^{B}_{v^{n}}-
γun(μ1)Aun(μ1)Λ¯vnB)ρABnPnZ|W(zn|un+vn)1.\displaystyle\hskip 33.0pt\gamma_{u^{n}}^{(\mu_{1})}A_{u^{n}}^{(\mu_{1})}\otimes\bar{\Lambda}^{B}_{v^{n}}\bigg{)}\sqrt{\rho_{AB}^{\otimes n}}P^{n}_{Z|W}(z^{n}|u^{n}+v^{n})\bigg{\|}_{1}.

Now we use the following proposition to bound the term corresponding to J1J_{1}.

Proposition 6.

For any ϵ(0,1)\epsilon\in(0,1), any η,δ(0,1)\eta,\delta\in(0,1) sufficiently small, and any nn sufficiently large, we have 𝔼[J1]ϵ\mathbb{E}\left[J_{1}\right]\leq\epsilon if k+l1nlogp+1nlog(N1)>I(U;RZV)σ3+log(p)S(U)σ3\frac{k+l_{1}}{n}\log p+\frac{1}{n}\log{N_{1}}>I(U;RZV)_{\sigma_{3}}+\log{p}-S(U)_{\sigma_{3}}, where σ3\sigma_{3} is the auxiliary state defined in the statement of the theorem.

Proof.

The proof of proposition is provided in Appendix B.6. ∎

Now we move on to bounding the term corresponding to J2.J_{2}. We start by applying triangle inequality followed by Lemma 1 on J2J_{2} to obtain

J2\displaystyle J_{2}\leq znun,vnPZ|Wn(zn|un+vn)1N1μ1=1N1ρAn\displaystyle\sum_{z^{n}}\sum_{u^{n},v^{n}}P^{n}_{Z|W}(z^{n}|u^{n}+v^{n})\bigg{\|}\frac{1}{N_{1}}\sum_{\mu_{1}=1}^{N_{1}}\sqrt{\rho_{A}^{\otimes n}}
×((γun(μ1)A¯un(μ1)γun(μ1)Aun(μ1))Λ¯vnB)ρAn1\displaystyle\hskip 15.0pt\times\left(\left(\gamma_{u^{n}}^{(\mu_{1})}\bar{A}_{u^{n}}^{(\mu_{1})}-\gamma_{u^{n}}^{(\mu_{1})}A_{u^{n}}^{(\mu_{1})}\right)\otimes\bar{\Lambda}^{B}_{v^{n}}\right)\sqrt{\rho_{A}^{\otimes n}}\bigg{\|}_{1}
=\displaystyle= un,vn1N1μ1=1N1ρAn((γun(μ1)A¯un(μ1)γun(μ1)Aun(μ1))\displaystyle\sum_{u^{n},v^{n}}\bigg{\|}\frac{1}{N_{1}}\sum_{\mu_{1}=1}^{N_{1}}\sqrt{\rho_{A}^{\otimes n}}\bigg{(}\left(\gamma_{u^{n}}^{(\mu_{1})}\bar{A}_{u^{n}}^{(\mu_{1})}-\gamma_{u^{n}}^{(\mu_{1})}A_{u^{n}}^{(\mu_{1})}\right)
Λ¯vnB)ρAn1\displaystyle\hskip 151.76744pt\otimes\bar{\Lambda}^{B}_{v^{n}}\bigg{)}\sqrt{\rho_{A}^{\otimes n}}\bigg{\|}_{1}
\displaystyle\leq 1N1μ1=1N1unγun(μ1)ρAn(A¯un(μ1)Aun(μ1))ρAn1.\displaystyle\frac{1}{N_{1}}\sum_{\mu_{1}=1}^{N_{1}}\sum_{u^{n}}\gamma_{u^{n}}^{(\mu_{1})}\left\|\sqrt{\rho_{A}^{\otimes n}}\left(\bar{A}_{u^{n}}^{(\mu_{1})}-A_{u^{n}}^{(\mu_{1})}\right)\sqrt{\rho_{A}^{\otimes n}}\right\|_{1}\!. (48)

Now we use the following proposition to bound the term corresponding to J2.J_{2}.

Proposition 7.

For any ϵ(0,1)\epsilon\in(0,1), any η,δ(0,1)\eta,\delta\in(0,1) sufficiently small, and any nn sufficiently large, we have 𝔼[J2]ϵ\mathbb{E}\left[J_{2}\right]\leq\epsilon if k+l1nlog(p)>I(U;RB)σ1+log(p)S(U)σ3\frac{k+l_{1}}{n}\log{p}>I(U;RB)_{\sigma_{1}}+\log{p}-S(U)_{\sigma_{3}}, where σ1\sigma_{1} and σ3\sigma_{3} are the auxiliary states defined in the statement of the theorem.

Proof.

The proof is provided in Appendix B.7. ∎

Since Q1JJ1+J2Q_{1}\leq J\leq J_{1}+J_{2}, hence 𝔼[J]\mathbb{E}[J], and consequently 𝔼[Q1]\mathbb{E}[Q_{1}], can be made arbitrarily small for sufficiently large n, if k+l1nlog(p)+1nlog(N1)>I(U;RZV)σ3S(U)σ3+log(p)\frac{k+l_{1}}{n}\log{p}+\frac{1}{n}\log{N_{1}}>I(U;RZV)_{\sigma_{3}}-S(U)_{\sigma_{3}}+\log{p} and k+l1nlog(p)>I(U;RB)σ1S(U)σ3+log(p)\frac{k+l_{1}}{n}\log{p}>I(U;RB)_{\sigma_{1}}-S(U)_{\sigma_{3}}+\log{p}. Now we move on to bounding Q2Q_{2}.

Step 4: Analyzing the effect of Bob’s approximating measurement
Step 3 ensured that the sub-system RZVRZV is close to a tensor product state in trace-norm. In this step, we approximate the state corresponding to the sub-system RZRZ using the approximating POVM MB(n)M_{B}^{(n)}, producing outputs on the alphabet 𝒱n\mathcal{V}^{n}. We proceed with the following proposition.

Proposition 8.

For any ϵ(0,1)\epsilon\in(0,1), any η,δ(0,1)\eta,\delta\in(0,1) sufficiently small, and any nn sufficiently large, we have 𝔼[Q2]ϵ\mathbb{E}[{Q}_{2}]\leq\epsilon, if k+l1nlog(p)+1nlog(N1)>I(U;RZV)σ3S(U)σ3+log(p)\frac{k+l_{1}}{n}\log{p}+\frac{1}{n}\log{N_{1}}>I(U;RZV)_{\sigma_{3}}-S(U)_{\sigma_{3}}+\log{p}, k+l2nlog(p)+1nlog(N2)>I(V;RZ)σ3S(V)σ3+log(p)\frac{k+l_{2}}{n}\log{p}+\frac{1}{n}\log{N_{2}}>I(V;RZ)_{\sigma_{3}}-S(V)_{\sigma_{3}}+\log{p}, k+l1nlog(p)>I(U;RB)σ1S(U)σ3+log(p)\frac{k+l_{1}}{n}\log{p}>I(U;RB)_{\sigma_{1}}-S(U)_{\sigma_{3}}+\log{p}, and k+l2nlog(p)>I(V;RA)σ2S(V)σ3+log(p)\frac{k+l_{2}}{n}\log{p}>I(V;RA)_{\sigma_{2}}-S(V)_{\sigma_{3}}+\log{p} where σ1,σ2\sigma_{1},\sigma_{2}, σ3\sigma_{3} are the auxiliary states defined in the statement of the theorem.

Proof.

The proof is provided in Appendix B.8. ∎

VI.5 Rate Constraints

To sum-up, we showed 𝔼[K]ϵ\mathbb{E}[K]\leq\epsilon holds for sufficiently large nn if the following bounds hold:

R~+R1\displaystyle\tilde{R}+R_{1} >I(U;RB)σ1S(U)σ3+log(p),\displaystyle>I(U;RB)_{\sigma_{1}}-S(U)_{\sigma_{3}}+\log{p}, (49a)
R~+R2\displaystyle\tilde{R}+R_{2} >I(V;RA)σ2S(V)σ3+log(p),\displaystyle>I(V;RA)_{\sigma_{2}}-S(V)_{\sigma_{3}}+\log{p}, (49b)
R~+R1+C1\displaystyle\tilde{R}+R_{1}+C_{1} >I(U;RZV)σ3S(U)σ3+log(p),\displaystyle>I(U;RZV)_{\sigma_{3}}-S(U)_{\sigma_{3}}+\log{p}, (49c)
R~+R2+C2\displaystyle\tilde{R}+R_{2}+C_{2} >I(V;RZ)σ3S(V)σ3+log(p),\displaystyle>I(V;RZ)_{\sigma_{3}}-S(V)_{\sigma_{3}}+\log{p}, (49d)
0R~\displaystyle 0\leq\tilde{R} <log(p)S(U+V)σ3,\displaystyle<\log{p}-S(U+V)_{\sigma_{3}}, (49e)
C1\displaystyle C_{1} 0,C20,\displaystyle\geq 0,\quad C_{2}\geq 0, (49f)

where Ci\ensurestackMath\stackon[1pt]=Δ1nlog2Ni,i{1,2}C_{i}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\frac{1}{n}\log_{2}N_{i},i\in\{1,2\} and R~\ensurestackMath\stackon[1pt]=Δknlog(p)\tilde{R}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\frac{k}{n}\log{p}. Therefore, there exists a distributed protocol with parameters (n,2nR1,2nR2,2nC1,2nC2)(n,2^{nR_{1}},2^{nR_{2}},2^{nC_{1}},2^{nC_{2}}) such that its overall POVM M^AB\hat{M}_{AB} is ϵ\epsilon-faithful to MABnM_{AB}^{\otimes n} with respect to ρABn\rho_{AB}^{\otimes n}.

Let us denote the above achievable rate-region by 1\mathcal{R}_{1}. By doing an exact symmetric analysis, but by replacing the first encoder by a product distribution instead of the second encoder in S1S_{1} (as performed in (46)), all the constraints remain the same, except that the constraints on R~+R1+C1\tilde{R}+R_{1}+C_{1} and R~+R2+C2\tilde{R}+R_{2}+C_{2} change as follows

R~+R1+C1\displaystyle\tilde{R}+R_{1}+C_{1} I(U;RZ)σ3S(U)σ3+log(p),\displaystyle\geq I(U;RZ)_{\sigma_{3}}-S(U)_{\sigma_{3}}+\log{p},\quad
R~+R2+C2\displaystyle\tilde{R}+R_{2}+C_{2} I(V;RZU)σ3S(V)σ3+log(p).\displaystyle\geq I(V;RZU)_{\sigma_{3}}-S(V)_{\sigma_{3}}+\log{p}. (50)

Let us denote the above achievable rate-region by 2\mathcal{R}_{2}. By time sharing between the any two points of 1\mathcal{R}_{1} and 2\mathcal{R}_{2} one can achieve any point in the convex closure of (12).(\mathcal{R}_{1}\bigcup\mathcal{R}_{2}). The following lemma gives a symmetric characterization of the closure of convex hull of the union of the above achievable rate-regions.

Lemma 7.

For the above defined rate regions 1\mathcal{R}_{1} and 2\mathcal{R}_{2}, we have 3=Convex Closure(12)\mathcal{R}_{3}=\text{Convex Closure}(\mathcal{R}_{1}\bigcup\mathcal{R}_{2}), where 3\mathcal{R}_{3} is given by the set of all the quintuples (R~,R1,R2,C1,C2)(\tilde{R},R_{1},R_{2},C_{1},C_{2}) satisfying the following constraints:

R~+R1\displaystyle\tilde{R}+R_{1} I(U;RB)σ1S(U)σ3+log(p),\displaystyle\geq I(U;RB)_{\sigma_{1}}-S(U)_{\sigma_{3}}+\log{p},
R~+R2\displaystyle\tilde{R}+R_{2} I(V;RA)σ2S(V)σ3+log(p),\displaystyle\geq I(V;RA)_{\sigma_{2}}-S(V)_{\sigma_{3}}+\log{p},
R~+R1+C1\displaystyle\tilde{R}+R_{1}+C_{1} I(U;RZ)σ3S(U)σ3+log(p),\displaystyle\geq I(U;RZ)_{\sigma_{3}}-S(U)_{\sigma_{3}}+\log{p},
R~+R2+C2\displaystyle\tilde{R}+R_{2}+C_{2} I(V;RZ)σ3S(V)σ3+log(p),\displaystyle\geq I(V;RZ)_{\sigma_{3}}-S(V)_{\sigma_{3}}+\log{p},
2R~+R1+R2+C1+C2\displaystyle 2\tilde{R}\!+\!R_{1}\!+R_{2}+C_{1}+C_{2} I(UV;RZ)σ3S(U,V)σ3\displaystyle\geq I(UV;RZ)_{\sigma_{3}}-S(U,V)_{\sigma_{3}}
+2log(p),\displaystyle\hskip 10.0pt+2\log{p},
0R~\displaystyle 0\leq\tilde{R} log(p)S(U+V)σ3,\displaystyle\leq\log{p}-S(U+V)_{\sigma_{3}}, (51)
R10,R2\displaystyle R_{1}\geq 0,R_{2} 0C10,C20.\displaystyle\geq 0\quad C_{1}\geq 0,C_{2}\geq 0. (52)
Proof.

The proof follows from elementary convex analysis. ∎

Lastly, we complete the proof of the theorem using the following lemma.

Lemma 8.

Let ¯3\bar{\mathcal{R}}_{3} denote the set of all quadruples (R1,R2,C1,C2)(R_{1},R_{2},C_{1},C_{2}) for which there exists R~\tilde{R} such that the quintuple (R1,R2,C1,C2,R~)(R_{1},R_{2},C_{1},C_{2},\tilde{R}) satisfies the inequalities in (7). Let F\mathcal{R}_{F} denote the set of all quadruples (R1,R2,C1,C2)(R_{1},R_{2},C_{1},C_{2}) that satisfy the inequalities in (5) given in the statement of the theorem. Then, ¯3=F\bar{\mathcal{R}}_{3}=\mathcal{R}_{F}.

Proof.

The proof follows from Fourier-Motzkin elimination [40]. ∎

VII Conclusion

We developed a technique of randomly generating structured POVMs using algebraic codes. Using this technique, we demonstrated a new achievable information-theoretic rate-region for the task of faithfully simulating a distributed quantum measurement and function computation. We further devised a Pruning Trace inequality which is a tighter version of the known operator Markov inequality, and a covering lemma which is independent of the operator Chernoff inequality, so as to analyse pairwise-independent POVM elements. Finally, combining these techniques, we demonstrated rate gains for this problem over traditional coding schemes, and provided a multi-party distributed faithful simulation and function computation protocol.

Acknowledgement: We thank Arun Padakandla for his valuable discussion and inputs in developing the proof techniques.

Appendix A Proof of Lemmas

A.1 Proof of Lemma 2

Proof.

We begin by defining the ensemble {λx,σ~x}x𝒳\{\lambda_{x},\tilde{\sigma}_{x}\}_{x\in\mathcal{X}} where σ~x=ΠΠxσxΠxΠ\tilde{\sigma}_{x}=\Pi\Pi_{x}\sigma_{x}\Pi_{x}\Pi for all x𝒳x\in\mathcal{X}. Further, let SS be defined as

S\ensurestackMath\stackon[1pt]=Δx𝒳λxσx1Mx𝒳m=1Mλxμxσx𝟙{Cm=x}1.\displaystyle S\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\Big{\|}\sum_{x\in\mathcal{X}}\lambda_{x}\sigma_{x}-\frac{1}{M}\sum_{x\in\mathcal{X}}\sum_{m=1}^{M}\frac{\lambda_{x}}{\mu_{x}}\sigma_{x}\mathbbm{1}_{\{C_{m}=x\}}\Big{\|}_{1}.

By adding an subtracting appropriate terms within the trace norm of SS and using the triangle inequality we obtain, SS1+S2+S3,S\leq S_{1}+S_{2}+S_{3}, where

S1\displaystyle S_{1} \ensurestackMath\stackon[1pt]=Δx𝒳λxσxx𝒳λxσ~x1,\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\big{\|}\sum_{x\in\mathcal{X}}\lambda_{x}\sigma_{x}-\sum_{x\in\mathcal{X}}\lambda_{x}\tilde{\sigma}_{x}\big{\|}_{1},
S2\displaystyle S_{2} \ensurestackMath\stackon[1pt]=Δ1Mm=1MλCmμCmσ~Cm1Mm=1MλCmμCmσCm1,and\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\big{\|}\frac{1}{M}\sum_{m=1}^{M}\frac{\lambda_{C_{m}}}{\mu_{C_{m}}}\tilde{\sigma}_{C_{m}}-\frac{1}{M}\sum_{m=1}^{M}\frac{\lambda_{C_{m}}}{\mu_{C_{m}}}{\sigma}_{C_{m}}\big{\|}_{1},\quad\text{and}
S3\displaystyle S_{3} \ensurestackMath\stackon[1pt]=Δx𝒳λxσ~x1Mx𝒳m=1Mλxμxσ~x𝟙{Cm=x}1.\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\big{\|}\sum_{x\in\mathcal{X}}\lambda_{x}\tilde{\sigma}_{x}-\frac{1}{M}\sum_{x\in\mathcal{X}}\sum_{m=1}^{M}\frac{\lambda_{x}}{\mu_{x}}\tilde{\sigma}_{x}\mathbbm{1}_{\{C_{m}=x\}}\big{\|}_{1}.

We begin by bounding the term corresponding to S1S_{1} and S2S_{2} as follows:

S1\displaystyle S_{1} x𝒳λxσxΠΠxσxΠxΠ1\displaystyle\leq\sum_{x\in\mathcal{X}}\lambda_{x}\|\sigma_{x}-\Pi\Pi_{x}\sigma_{x}\Pi_{x}\Pi\|_{1}
x𝒳λxσxΠσxΠ1\displaystyle\leq\sum_{x\in\mathcal{X}}\lambda_{x}\|\sigma_{x}-\Pi\sigma_{x}\Pi\big{\|}_{1}
+x𝒳λxΠσxΠΠΠxσxΠxΠ1\displaystyle\hskip 20.0pt+\sum_{x\in\mathcal{X}}\lambda_{x}\|\Pi\sigma_{x}\Pi-\Pi\Pi_{x}\sigma_{x}\Pi_{x}\Pi\|_{1}
2ϵ+x𝒳λxΠσxΠxσxΠx1Π\displaystyle\leq 2\sqrt{\epsilon}+\sum_{x\in\mathcal{X}}\lambda_{x}\|\Pi\|_{\infty}\|\sigma_{x}-\Pi_{x}\sigma_{x}\Pi_{x}\|_{1}\|\Pi\|_{\infty}
4ϵ=δ(ϵ),\displaystyle\leq 4\sqrt{\epsilon}=\delta(\epsilon), (53)

where the first two inequalities use the triangle inequality, the third uses the gentle measurement lemma (given the assumption (9a) from the statement of the Lemma) for the first term, and operator Holder’s inequality (Exercise 12.2.1 in [39]) for the second term. The last inequality follows again from the gentle measurement given the assumption (9b). Similarly, for S2S_{2} we have

𝔼[S2]\displaystyle\mathbb{E}_{\mathbbm{C}}[S_{2}] 𝔼[1Mm=1Mx𝒳λxμx𝟙{Cm=x}σxσ~x1]\displaystyle\leq\mathbb{E}_{\mathbbm{C}}\left[\frac{1}{M}\sum_{m=1}^{M}\sum_{x\in\mathcal{X}}\frac{\lambda_{x}}{\mu_{x}}\mathbbm{1}_{\{C_{m}=x\}}\|\sigma_{x}-\tilde{\sigma}_{x}\|_{1}\right]
=1Mm=1Mx𝒳λxσxσ~x14ϵ=δ(ϵ),\displaystyle=\frac{1}{M}\sum_{m=1}^{M}\sum_{x\in\mathcal{X}}\lambda_{x}\|\sigma_{x}-\tilde{\sigma}_{x}\|_{1}\leq 4\sqrt{\epsilon}=\delta(\epsilon), (54)

where we use the fact that 𝔼[𝟙{cm=x}]=μx\mathbb{E}_{\mathbbm{C}}[\mathbbm{1}_{\{c_{m}=x\}}]=\mu_{x}, and the last inequality uses similar arguments as in (A.1). Finally, we proceed to bound the term corresponding to S3S_{3}. Firstly, note that, 𝔼[1MmλCmμCmσ~Cm]=x𝒳λxσx~\mathbb{E}_{\mathbbm{C}}[\frac{1}{M}\sum_{m}\frac{\lambda_{C_{m}}}{\mu_{C_{m}}}\tilde{\sigma}_{C_{m}}]=\sum_{x\in\mathcal{X}}\lambda_{x}\tilde{\sigma_{x}}. This gives

𝔼[S3]\displaystyle\mathbb{E}_{\mathbbm{C}}[S_{3}] =𝔼[1MmλCmμCmσ~Cm𝔼[1MmλCmμCmσ~Cm]1]\displaystyle=\mathbb{E}_{\mathbbm{C}}\left[\Big{\|}\frac{1}{M}\sum_{m}\frac{\lambda_{C_{m}}}{\mu_{C_{m}}}\tilde{\sigma}_{C_{m}}-\mathbb{E}_{\mathbbm{C}}\bigg{[}\frac{1}{M}\sum_{m}\frac{\lambda_{C_{m}}}{\mu_{C_{m}}}\tilde{\sigma}_{C_{m}}\bigg{]}\Big{\|}_{1}\right]
Tr(𝔼[(1MmλCmμCmσ~Cm𝔼[1MmλCmμCmσ~Cm])2])\displaystyle\leq\Tr{\sqrt{\mathbb{E}_{\mathbbm{C}}\left[\left(\frac{1}{M}\sum_{m}\frac{\lambda_{C_{m}}}{\mu_{C_{m}}}\tilde{\sigma}_{C_{m}}-\mathbb{E}_{\mathbbm{C}}\bigg{[}\frac{1}{M}\sum_{m}\frac{\lambda_{C_{m}}}{\mu_{C_{m}}}\tilde{\sigma}_{C_{m}}\bigg{]}\right)^{2}\right]}}
=Tr(𝔼[(1MmλCmμCmσ~Cm)2](𝔼[1MmλCmμCmσ~Cm])2)\displaystyle=\Tr{\sqrt{\mathbb{E}_{\mathbbm{C}}\left[\left(\frac{1}{M}\sum_{m}\frac{\lambda_{C_{m}}}{\mu_{C_{m}}}\tilde{\sigma}_{C_{m}}\right)^{2}\right]-\left(\mathbb{E}_{\mathbbm{C}}\bigg{[}\frac{1}{M}\sum_{m}\frac{\lambda_{C_{m}}}{\mu_{C_{m}}}\tilde{\sigma}_{C_{m}}\bigg{]}\right)^{2}}}
=Tr(1M2m𝔼[(λCmμCmσ~Cm)2]+1M2m,mmm𝔼[λCmσ~CmμCmλCmσ~CmμCm](1Mm𝔼[λCmσ~CmμCm])2)\displaystyle=\Tr{\sqrt{\frac{1}{M^{2}}\sum_{m}\mathbb{E}_{\mathbbm{C}}\left[\left(\frac{\lambda_{C_{m}}}{\mu_{C_{m}}}\tilde{\sigma}_{C_{m}}\right)^{2}\right]+\frac{1}{M^{2}}\sum_{\begin{subarray}{c}m,m^{\prime}\\ m\neq m^{\prime}\end{subarray}}\mathbb{E}_{\mathbbm{C}}\left[\frac{\lambda_{C_{m}}\tilde{\sigma}_{C_{m}}}{\mu_{C_{m}}}\frac{\lambda_{C_{m^{\prime}}}\tilde{\sigma}_{C_{m^{\prime}}}}{\mu_{C_{m^{\prime}}}}\right]-\left(\frac{1}{M}\sum_{m}\mathbb{E}_{\mathbbm{C}}\bigg{[}\frac{\lambda_{C_{m}}\tilde{\sigma}_{C_{m}}}{\mu_{C_{m}}}\bigg{]}\right)^{2}}}
=Tr(1M𝔼[(λC1σ~C1μC1)2]1M(𝔼[λC1σ~C1μC1])2)\displaystyle=\Tr{\sqrt{\frac{1}{M}\mathbb{E}_{\mathbbm{C}}\left[\left(\frac{\lambda_{C_{1}}\tilde{\sigma}_{C_{1}}}{\mu_{C_{1}}}\right)^{2}\right]-\frac{1}{M}\left(\mathbb{E}_{\mathbbm{C}}\left[\frac{\lambda_{C_{1}}\tilde{\sigma}_{C_{1}}}{\mu_{C_{1}}}\right]\right)^{2}}}
Tr(1M𝔼[(λC1σ~C1μC1)2]),\displaystyle\leq\Tr{\sqrt{\frac{1}{M}\mathbb{E}_{\mathbbm{C}}\left[\left(\frac{\lambda_{C_{1}}\tilde{\sigma}_{C_{1}}}{\mu_{C_{1}}}\right)^{2}\right]}}, (55)

where the first inequality follows from concavity of operator square-root function (Löwner-Heinz theorem, see Theorem 2.62.6 in [41]). The last equality uses the fact that codewords of the random code \mathbbm{C} are pairwise independent, and the last inequality follows from monotonicity of the operator square-root function (Theorem 2.62.6 in [41]).

Moving on, we now bound the operator within the square root of (A.1) as

𝔼[(λC1σ~C1μC1)2]\displaystyle\mathbb{E}_{\mathbbm{C}}\left[\left(\frac{\lambda_{C_{1}}\tilde{\sigma}_{C_{1}}}{\mu_{C_{1}}}\right)^{2}\right] =x𝒳λx2μxσ~x2x𝒳κλxσ~x2\displaystyle=\sum_{x\in\mathcal{X}}\frac{\lambda_{x}^{2}}{\mu_{x}}\tilde{\sigma}_{x}^{2}\leq\sum_{x\in\mathcal{X}}\kappa\lambda_{x}\tilde{\sigma}_{x}^{2}
=κx𝒳λxΠ(ΠxσxΠx)Π(ΠxσxΠx)Π,\displaystyle=\kappa\!\sum_{x\in\mathcal{X}}\!\lambda_{x}\Pi\left(\Pi_{x}\sigma_{x}\Pi_{x}\right)\Pi\left(\Pi_{x}\sigma_{x}\Pi_{x}\right)\Pi,

where we use the assumption λxμxκ\frac{\lambda_{x}}{\mu_{x}}\leq\kappa for all x𝒳x\in\mathcal{X}. Further since, ΠI\Pi\leq I, we have (ΠxσxΠx)Π(ΠxσxΠx)(ΠxσxΠx)2\left(\Pi_{x}\sigma_{x}\Pi_{x}\right)\Pi\left(\Pi_{x}\sigma_{x}\Pi_{x}\right)\leq\left(\Pi_{x}\sigma_{x}\Pi_{x}\right)^{2}, which gives

𝔼[(λC1σ~C1μC1)2]\displaystyle\mathbb{E}_{\mathbbm{C}}\left[\left(\frac{\lambda_{C_{1}}\tilde{\sigma}_{C_{1}}}{\mu_{C_{1}}}\right)^{2}\right] κx𝒳λxΠ(ΠxσxΠx)2Π.\displaystyle\leq\kappa\sum_{x\in\mathcal{X}}\lambda_{x}\Pi\left(\Pi_{x}\sigma_{x}\Pi_{x}\right)^{2}\Pi.

Moreover, using the assumption 9d, i.e., ΠxσxΠx1dΠx1dI\Pi_{x}\sigma_{x}\Pi_{x}\leq\frac{1}{d}\Pi_{x}\leq\frac{1}{d}I, we get

(ΠxσxΠx)2\displaystyle\left(\Pi_{x}\sigma_{x}\Pi_{x}\right)^{2} =ΠxσxΠx(ΠxσxΠx)ΠxσxΠx\displaystyle=\sqrt{\Pi_{x}\sigma_{x}\Pi_{x}}\left(\Pi_{x}\sigma_{x}\Pi_{x}\right)\sqrt{\Pi_{x}\sigma_{x}\Pi_{x}}
1dΠxσxΠx,for all x𝒳.\displaystyle\leq\frac{1}{d}\Pi_{x}\sigma_{x}\Pi_{x},\quad\text{for all }x\in\mathcal{X}.

Thus,

𝔼[(λC1σ~C1μC1)2]\displaystyle\mathbb{E}_{\mathbbm{C}}\left[\left(\frac{\lambda_{C_{1}}\tilde{\sigma}_{C_{1}}}{\mu_{C_{1}}}\right)^{2}\right] κdΠ(x𝒳λxΠxσxΠx)ΠκdΠσΠ,\displaystyle\leq\frac{\kappa}{d}\Pi\left(\sum_{x\in\mathcal{X}}\lambda_{x}\Pi_{x}\sigma_{x}\Pi_{x}\right)\Pi\leq\frac{\kappa}{d}\Pi\sigma\Pi, (56)

where the second inequality uses the assumption (9e) from the statement of the Lemma. Substituting the simplification obtained in (56) into (A.1) and using the monotonicity of square-root operator, we obtain

𝔼[S3]Tr(κMdΠσΠ)κDMd,\displaystyle\mathbb{E}_{\mathbbm{C}}[S_{3}]\leq\Tr{\sqrt{\frac{\kappa}{Md}\Pi\sigma\Pi}}\leq\sqrt{\frac{\kappa D}{Md}}, (57)

where the second inequality uses the assumption (9c). Combining the bounds (A.1), (A.1), and (57) we get the desired result.

A.2 Proof of Lemma 4

Note that if PP prunes XX, then PP also prunes 1η(X(1η)IA)\frac{1}{\eta}(X-(1-\eta)I_{A}) with respect to IA.I_{A}. Using Lemma 3, we obtain

Tr(IAP)1ηTr(X(1η)IA).\displaystyle\Tr{I_{A}-P}\leq\frac{1}{\eta}\Tr{X-(1-\eta)I_{A}}.

Applying expectation and using the assumption on 𝔼[X]\mathbb{E}[X], we get

𝔼[Tr(IAP)]\displaystyle\mathbb{E}[\Tr{\!I_{A}\!-\!P}]\! 1η𝔼[Tr(X𝔼[X])]1η𝔼[X𝔼[X]1].\displaystyle\leq\!\frac{1}{\eta}\mathbb{E}\left[\Tr{X\!-\!\mathbb{E}[X]}\right]\leq\frac{1}{\eta}\mathbb{E}\left[\|X-\mathbb{E}[X]\|_{1}\right].

A.3 Proof of Lemma 5

We begin by defining LL as

L\displaystyle L \ensurestackMath\stackon[1pt]=Δwnλwnθwn\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\bigg{\|}\sum_{w^{n}}\lambda_{w^{n}}\theta_{w^{n}}-
1(1+η)pnpk+lNμ=1Na,mwnλwnθwn𝟙{Wn,(μ)(a,m)=wn}1.\displaystyle\frac{1}{(1+\eta)}\frac{p^{n}}{p^{k+l}N^{\prime}}\sum_{\mu=1}^{N^{\prime}}\sum_{a,m}\sum_{w^{n}}\lambda_{w^{n}}\theta_{w^{n}}\mathbbm{1}_{\{W^{n,(\mu)}(a,m)=w^{n}\}}\bigg{\|}_{1}.

Further, let θ\ensurestackMath\stackon[1pt]=Δw𝒲λwθw\theta\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\sum_{w\in\mathcal{W}}\lambda_{w}\theta_{w} and let Πθ\Pi_{\theta} and Πwnθ\Pi_{w^{n}}^{\theta} denote the δ\delta-typical projector of θ\theta and conditional typical projector of θwn,\theta_{w^{n}}, respectively. Define λ~wn=λwn1ε\tilde{\lambda}_{w^{n}}=\frac{\lambda_{w^{n}}}{1-\varepsilon} for wn𝒯δ(n)(W)w^{n}\in\mathcal{T}_{\delta}^{(n)}(W), and 0 otherwise, where ε=wn𝒯δ(n)(W)λwn.\varepsilon=\sum_{w^{n}\notin\mathcal{T}_{\delta}^{(n)}(W)}\lambda_{w^{n}}. Using the triangle inequality we can bound LL as LL1+L2+L3,L\leq L_{1}+L_{2}+L_{3}, where

L1\displaystyle L_{1} \ensurestackMath\stackon[1pt]=Δwnλwnθwnwnλ~wnθwn1,\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\bigg{\|}\sum_{w^{n}}\lambda_{w^{n}}\theta_{w^{n}}-\sum_{w^{n}}\tilde{\lambda}_{w^{n}}{\theta}_{w^{n}}\bigg{\|}_{1},
L2\displaystyle L_{2} \ensurestackMath\stackon[1pt]=Δwnλ~wnθwn\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\bigg{\|}\sum_{w^{n}}\tilde{\lambda}_{w^{n}}{\theta}_{w^{n}}-
pnpk+lNμ=1Na,mwnλ~wnθwn𝟙{Wn,(μ)(a,m)=wn}1,\displaystyle\hskip 22.0pt\frac{p^{n}}{p^{k+l}N^{\prime}}\sum_{\mu=1}^{N^{\prime}}\sum_{a,m}\sum_{w^{n}}\tilde{\lambda}_{w^{n}}{\theta}_{w^{n}}\mathbbm{1}_{\{W^{n,(\mu)}(a,m)=w^{n}\}}\bigg{\|}_{1},
L3\displaystyle L_{3} \ensurestackMath\stackon[1pt]=Δpnpk+lNμa,mwn(λ~wnλwn(1+η))\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\bigg{\|}\frac{p^{n}}{p^{k+l}N^{\prime}}\sum_{\mu}\sum_{a,m}\sum_{w^{n}}\left(\tilde{\lambda}_{w^{n}}-{\frac{\lambda_{w^{n}}}{(1+\eta)}}\right)
×θwn𝟙{Wn,(μ)(a,m)=wn}1.\displaystyle\hskip 108.405pt\times\theta_{w^{n}}\mathbbm{1}_{\{W^{n,(\mu)}(a,m)=w^{n}\}}\bigg{\|}_{1}.

We begin by bounding the term corresponding to L1L_{1} as

L1\displaystyle L_{1} wn𝒯δ(n)(W)λwnε1εθwn1=1+wn𝒯δ(n)(W)λwnθwn1=1\displaystyle\leq\sum_{w^{n}\in\mathcal{T}_{\delta}^{(n)}(W)}\!\!\!\!\lambda_{w^{n}}\frac{\varepsilon}{1-\varepsilon}\underbrace{\left\|\theta_{w^{n}}\right\|_{1}}_{=1}+\sum_{w^{n}\notin\mathcal{T}_{\delta}^{(n)}(W)}\!\!\!\!\lambda_{w^{n}}\underbrace{\left\|\theta_{w^{n}}\right\|_{1}}_{=1}
=2ε.\displaystyle=2\varepsilon. (58)

Now consider the term corresponding to L2L_{2}, for which we employ Lemma 2. Toward this, we consider the following identification: λx\lambda_{x} with λ~wn\tilde{\lambda}_{w^{n}}, σx\sigma_{x} with θwn\theta_{w^{n}}, 𝒳\mathcal{X} with 𝒯δ(n)(W)\mathcal{T}_{\delta}^{(n)}(W), 𝒳¯\bar{\mathcal{X}} with 𝔽pn\mathbb{F}_{p}^{n}, σ\sigma with θ~\ensurestackMath\stackon[1pt]=Δwnλ~wnθwn\widetilde{\theta}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\sum_{w^{n}}\tilde{\lambda}_{w^{n}}{\theta}_{w^{n}}, Π\Pi with Πθ\Pi_{\theta}, Πx\Pi_{x} with Πwnθ\Pi_{w^{n}}^{\theta}, and μx=1pn\mu_{x}=\frac{1}{p^{n}} for all x𝒳¯x\in\bar{\mathcal{X}}. Since the collection of random variables {Wn,(μ)(a,m)}\{W^{n,(\mu)}(a,m)\} are generated using Unionized Coset Codes, we have

(𝟙{Wn,(μ)(a,m)=wn}=1)=1pn,for allwn𝔽pn.\displaystyle\mathbb{P}\left(\mathbbm{1}_{\{W^{n,(\mu)}(a,m)=w^{n}\}}=1\right)=\frac{1}{p^{n}},\quad\text{for all}\quad w^{n}\in\mathbb{F}_{p}^{n}.

Note that λ~wn1/pn2n(S(W)σθlog(p)δw)\frac{\tilde{\lambda}_{w^{n}}}{1/p^{n}}\leq 2^{-n{(S(W)_{\sigma_{\theta}}-\log{p}-\delta_{w})}} for all wn𝔽pnw^{n}\in\mathbb{F}_{p}^{n}, where δw(δ)0\delta_{w}(\delta)\searrow 0 as δ0\delta\searrow 0, and σθ\sigma_{\theta} is defined in the statement of the lemma. With these, we check the hypotheses of Lemma 2. Firstly, using the pinching arguments described in [7, Property 15.2.7], we have Tr(Πθθwn)1ϵ\Tr{\Pi_{\theta}\theta_{w^{n}}}\geq 1-\epsilon for all ϵ(0,1),δ>0\epsilon\in(0,1),\delta>0 and sufficiently large nn, satisfying hypothesis (9a). Secondly, (9b) and (9e) are satisfied from the construction of Πwnθ\Pi_{w^{n}}^{\theta}. Next, we consider the hypothesis (9c). We have

Πθθ~1\displaystyle\left\|\Pi^{\theta}\sqrt{\tilde{\theta}}\right\|_{1}\!\! =Tr(Πθθ~Πθ)\displaystyle=\Tr{\sqrt{\Pi^{\theta}\tilde{\theta}\Pi^{\theta}}}
1(1ε)Tr(ΠθθnΠθ)2n2(S(R)σθ+δw),\displaystyle\leq\frac{1}{\sqrt{(1-\varepsilon)}}\Tr{\sqrt{\Pi^{\theta}\theta^{\otimes n}\Pi^{\theta}}}\leq 2^{\frac{n}{2}(S(R)_{\sigma_{\theta}}+\delta_{w}^{\prime})},

where the first inequality above follows from the fact that wnλ~wnθwn1(1ε)wnλwnθwn=θn(1ε)\sum_{w^{n}}\tilde{\lambda}_{w^{n}}\theta_{w^{n}}\leq\frac{1}{(1-\varepsilon)}\sum_{w^{n}}\lambda_{w^{n}}\theta_{w^{n}}=\frac{\theta^{\otimes n}}{(1-\varepsilon)} and using the operator monotonicty of the square-root function (Theorem 2.62.6 in [41]). The second inequality follows from the property of the typical projector for some δw\delta_{w}^{\prime} such that δw0\delta_{w}^{\prime}\searrow 0 as δ0\delta\searrow 0. This gives

D=2n(S(R)σθ+δw).{D}={2^{n(S(R)_{\sigma_{\theta}}+\delta_{w}^{\prime})}}.

Finally, the hypotheses (9d) is satisfied from the property of conditional typical projectors for d=2n(S(R|W)σθδw′′)d=2^{n(S(R|W)_{\sigma_{\theta}}-\delta_{w}^{\prime\prime})}, where δw′′0\delta_{w}^{\prime\prime}\searrow 0 as δ0\delta\searrow 0 (see [7, Property 15.2.6]). Next we check the pairwise independence of Wn,(μ)(a,m)W^{n,(\mu)}(a,m) and Wn,(μ)(a~,m~)W^{n,(\mu)}(\tilde{a},\tilde{m}). Since these are constructed using randomly and uniformly generated GG and h(μ)h^{(\mu)}, we have {Wn,(μ)(a,m)}a𝔽pk,m𝔽pl,μ[1:N]\{W^{n,(\mu)}(a,m)\}_{a\in\mathbb{F}^{k}_{p},m\in\mathbb{F}_{p}^{l},\mu\in[1:N^{\prime}]} to be pairwise independent for each (see [37] for details). Therefore, employing Lemma 2 we get

𝔼[L2]\displaystyle\mathbb{E}[L_{2}] 2n(S(R)σθ+δw)2n(S(W)σθlog(p)δw)Npk+l2n(S(R|W)σθδw′′)+8ϵ\displaystyle\leq\sqrt{\frac{2^{n(S(R)_{\sigma_{\theta}}+\delta_{w}^{\prime})}2^{-n{(S(W)_{\sigma_{\theta}}-\log{p}-\delta_{w})}}}{N^{\prime}p^{k+l}2^{n(S(R|W)_{\sigma_{\theta}}-\delta_{w}^{\prime\prime})}}}+8\sqrt{\epsilon}
exp2[n2(k+lnlog(p)+1nlog(N)I(R;W)σθ\displaystyle\leq\text{exp}_{2}\bigg{[}-\frac{n}{2}\bigg{(}\frac{k+l}{n}\log{p}+\frac{1}{n}\log{N^{\prime}}-I(R;W)_{\sigma_{\theta}}
log(p)+S(W)σθδwδwδw′′)]+8ϵ,\displaystyle\hskip 30.0pt-\log{p}+S(W)_{\sigma_{\theta}}-\delta_{w}-\delta_{w}^{\prime}-\delta_{w}^{\prime\prime}\bigg{)}\bigg{]}+8\sqrt{\epsilon}, (59)

where exp(x)2\ensurestackMath\stackon[1pt]=Δ2x.{}_{2}(x)\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}2^{x}. As for L3L_{3}, taking expectation and using 𝔼[𝟙{Wn,(μ)(a,m)=wn}]=1pn\mathbb{E}[\mathbbm{1}_{\{W^{n,(\mu)}(a,m)=w^{n}\}}]=\frac{1}{p^{n}} gives

𝔼[L3]η+ε(1+η)+ε(1+η)=η+2ε1+η.\displaystyle\mathbb{E}[L_{3}]\leq\frac{\eta+\varepsilon}{(1+\eta)}+\frac{\varepsilon}{(1+\eta)}=\frac{\eta+2\varepsilon}{1+\eta}. (60)

Combining the bounds from (58),(59)\eqref{eq:L1Term},\eqref{eq:L2Term} and (60)\eqref{eq:L3Term} gives the desired result.

A.4 Proof of Lemma 6

We begin by using the Hólder’s inequality [39, 41] for operator norm, i.e., (AB1AB1\|AB\|_{1}\leq\|A\|_{\infty}\|B\|_{1}), and defining Λ^wn=ρn1ρ~wnρn1\hat{\Lambda}_{w^{n}}=\sqrt{\rho^{\otimes n}}^{-1}\tilde{\rho}_{w^{n}}\sqrt{\rho^{\otimes n}}^{-1}. This gives us

wnγwn(μ)ρn(A¯wn(μ)Awn(μ))ρn1=\displaystyle\sum_{w^{n}}\!\gamma_{w^{n}}^{(\mu)}\left\|\sqrt{\rho^{\otimes n}}\!\left(\!\bar{A}_{w^{n}}^{(\mu)}-A_{w^{n}}^{(\mu)}\!\right)\!\sqrt{\rho^{\otimes n}}\right\|_{1}= wnαwnγwn(μ)ΠρρnΛ^wnρnΠρΠρρnΠμΛ^wnΠμρnΠρ1\displaystyle\sum_{w^{n}}\alpha_{w^{n}}\gamma_{w^{n}}^{(\mu)}\left\|\Pi_{\rho}\sqrt{\rho^{\otimes n}}\hat{\Lambda}_{w^{n}}\sqrt{\rho^{\otimes n}}\Pi_{\rho}-\Pi_{\rho}\sqrt{\rho^{\otimes n}}\Pi^{\mu}\hat{\Lambda}_{w^{n}}\Pi^{\mu}\sqrt{\rho^{\otimes n}}\Pi_{\rho}\right\|_{1}
\displaystyle\leq wnαwnγwn(μ)Πρρn2Λ^wnΠμΛ^wnΠμ1\displaystyle\sum_{w^{n}}\alpha_{w^{n}}\gamma_{w^{n}}^{(\mu)}\left\|\Pi_{\rho}\sqrt{\rho^{\otimes n}}\right\|_{\infty}^{2}\left\|\hat{\Lambda}_{w^{n}}-\Pi^{\mu}\hat{\Lambda}_{w^{n}}\Pi^{\mu}\right\|_{1}
\displaystyle\leq  2n(S(ρ)δρ)wnαwnγwn(μ)2Tr((ΠρAΠμ)Λ^wn)Tr(Λ^wn),\displaystyle\;2^{-n(S(\rho)-\delta_{\rho})}\sum_{w^{n}}\alpha_{w^{n}}\gamma_{w^{n}}^{(\mu)}2\sqrt{\Tr{(\Pi_{\rho_{A}}-\Pi^{\mu})\hat{\Lambda}_{w^{n}}}\Tr{\hat{\Lambda}_{w^{n}}}},

where the equality follows from the fact that Πρ\Pi_{\rho} and Πμ\Pi^{\mu} commute, the first inequality follows from the Hólder’s inequality, and the second inequality uses the following bounds

\displaystyle\bigg{\|} Λ^wnΠμΛ^wnΠμ1\displaystyle\hat{\Lambda}_{w^{n}}-\Pi^{\mu}\hat{\Lambda}_{w^{n}}\Pi^{\mu}\bigg{\|}_{1}
Λ^wnΠμΛ^wn1+ΠμΛ^wnΠμΛ^wnΠμ1\displaystyle\leq\left\|\hat{\Lambda}_{w^{n}}-\Pi^{\mu}\hat{\Lambda}_{w^{n}}\right\|_{1}+\left\|\Pi^{\mu}\hat{\Lambda}_{w^{n}}-\Pi^{\mu}\hat{\Lambda}_{w^{n}}\Pi^{\mu}\right\|_{1}
=Tr{|(ΠρΠμ)Λ^wnΛ^wn|}\displaystyle=\Tr\left\{\left|(\Pi_{\rho}-\Pi^{\mu})\sqrt{\hat{\Lambda}_{w^{n}}}\sqrt{\hat{\Lambda}_{w^{n}}}\right|\right\}
+Tr{|ΠμΛ^wnΛ^wn(ΠρΠμ)|}\displaystyle\hskip 72.26999pt+\Tr\left\{\left|\Pi^{\mu}\sqrt{\hat{\Lambda}_{w^{n}}}\sqrt{\hat{\Lambda}_{w^{n}}}(\Pi_{\rho}-\Pi^{\mu})\right|\right\}
Tr{(ΠρΠμ)2Λ^wn}Tr{Λ^wn}\displaystyle\leq\sqrt{\Tr\left\{(\Pi_{\rho}-\Pi^{\mu})^{2}{\hat{\Lambda}_{w^{n}}}\right\}\Tr\left\{{\hat{\Lambda}_{w^{n}}}\right\}}
+Tr{ΠμΛ^wn}Tr{Λ^wn(ΠρΠμ)2}\displaystyle\hskip 61.42993pt+\sqrt{\Tr\left\{\Pi^{\mu}{\hat{\Lambda}_{w^{n}}}\right\}\Tr\left\{{\hat{\Lambda}_{w^{n}}}(\Pi_{\rho}-\Pi^{\mu})^{2}\right\}}
2Tr((ΠρΠμ)Λ^wn)Tr(Λ^wn),\displaystyle\leq 2\sqrt{\Tr{(\Pi_{\rho}-\Pi^{\mu})\hat{\Lambda}_{w^{n}}}\Tr{\hat{\Lambda}_{w^{n}}}},

where the second inequality uses Cauchy-Schwarz inequality along with the polar decomposition (see the usage in [39, Lemma 9.4.2]) and the last inequality uses the arguments: (i) Πμ\Pi^{\mu} is a projector onto a subspace of Πρ\Pi_{\rho} and (ii) Tr(ΠμΛ^wn)Tr(Λ^wn)\Tr{\Pi^{\mu}\hat{\Lambda}_{w^{n}}}\leq\Tr{\hat{\Lambda}_{w^{n}}}. Further, using the fact that for wn𝒯δ(n)(W),w^{n}\in\mathcal{T}_{\delta}^{(n)}(W),

Tr{Λ^wn}\displaystyle\Tr\{\hat{\Lambda}_{w^{n}}\} =ΠρΛ^wnΠρ1\displaystyle=\|\Pi_{\rho}\hat{\Lambda}_{w^{n}}\Pi_{\rho}\|_{1}
Πρρn1ρ~wn11Πρρn1\displaystyle\leq\|\Pi_{\rho}\sqrt{\rho^{\otimes n}}^{-1}\|_{\infty}\underbrace{\|\tilde{\rho}_{w^{n}}\|_{1}}_{\leq 1}\|\Pi_{\rho}\sqrt{\rho^{\otimes n}}^{-1}\|_{\infty}
Πρρn122n(S(ρ)+δρ),\displaystyle\leq\|\Pi_{\rho}\sqrt{\rho^{\otimes n}}^{-1}\|_{\infty}^{2}\leq 2^{n(S(\rho)+\delta_{\rho})},

it follows that

wnγwn(μ)ρn(A¯wn(μ)Awn(μ))ρn1\displaystyle\sum_{w^{n}}\gamma_{w^{n}}^{(\mu)}\left\|\sqrt{\rho^{\otimes n}}\!\left(\bar{A}_{w^{n}}^{(\mu)}-A_{w^{n}}^{(\mu)}\right)\!\sqrt{\rho^{\otimes n}}\right\|_{1}
22n2(S(ρ)4δρ)wnαwnγwn(μ)Tr((ΠρΠμ)Λ^wn)\displaystyle\leq 2\cdot 2^{-\frac{n}{2}(S(\rho)-4\delta_{\rho})}\sum_{w^{n}}{\alpha_{w^{n}}\gamma_{w^{n}}^{(\mu)}}\sqrt{\Tr{(\Pi_{\rho}-\Pi^{\mu})\hat{\Lambda}_{w^{n}}}}
223nδρΔ(μ)wnαwnγwn(μ)Δ(μ)Tr((ΠρΠμ)ρ~wn)\displaystyle\leq 2\cdot{2^{3n\delta_{\rho}}}\Delta^{(\mu)}\sqrt{\sum_{w^{n}}\frac{\alpha_{w^{n}}\gamma_{w^{n}}^{(\mu)}}{\Delta^{(\mu)}}\Tr{(\Pi_{\rho}-\Pi^{\mu})\tilde{\rho}_{w^{n}}}}
=223nδρ(Δ(μ)𝔼[Δ(μ)]+𝔼[Δ(μ)])\displaystyle=2\cdot{2^{3n\delta_{\rho}}}\left(\Delta^{(\mu)}-\mathbb{E}[\Delta^{(\mu)}]+\mathbb{E}[\Delta^{(\mu)}]\right)
×Tr((ΠρΠμ)wnαwnγwn(μ)Δ(μ)ρ~wn)\displaystyle\hskip 65.04256pt\times\sqrt{\Tr{(\Pi_{\rho}-\Pi^{\mu})\sum_{w^{n}}\frac{\alpha_{w^{n}}\gamma_{w^{n}}^{(\mu)}}{\Delta^{(\mu)}}\tilde{\rho}_{w^{n}}}}
223nδρ𝔼[Δ(μ)]Tr((ΠρΠμ)wnαwnγwn(μ)Δ(μ)ρ~wn)\displaystyle\leq 2\cdot{2^{3n\delta_{\rho}}}\mathbb{E}[\Delta^{(\mu)}]\sqrt{\Tr{(\Pi_{\rho}-\Pi^{\mu})\sum_{w^{n}}\frac{\alpha_{w^{n}}\gamma_{w^{n}}^{(\mu)}}{\Delta^{(\mu)}}\tilde{\rho}_{w^{n}}}}
+223nδρ|Δ(μ)𝔼[Δ(μ)]|H0\displaystyle\hskip 122.85876pt+2\cdot 2^{3n\delta_{\rho}}\underbrace{\left|\Delta^{(\mu)}-\mathbb{E}[\Delta^{(\mu)}]\right|}_{H_{0}}
223nδρ(H0+(1ε)(1+η)H1+H2+H3),\displaystyle\leq 2\cdot{2^{3n\delta_{\rho}}}\left(H_{0}+\frac{\sqrt{(1-\varepsilon)}}{(1+\eta)}\sqrt{H_{1}+H_{2}+H_{3}}\right),

where the second inequality above follows by defining Δ(μ)=wn𝒯δ(n)(W)αwnγwn(μ)\Delta^{(\mu)}=\sum_{w^{n}\in\mathcal{T}_{\delta}^{(n)}(W)}{\alpha_{w^{n}}\gamma_{w^{n}}^{(\mu)}} and using the concavity of the square-root function, the third inequality follows by using the fact that

wnαwnγwn(μ)Δ(μ)\displaystyle\sum_{w^{n}}\frac{\alpha_{w^{n}}\gamma_{w^{n}}^{(\mu)}}{\Delta^{(\mu)}} Tr((ΠρΠμ)ρ~wn)\displaystyle\Tr{(\Pi_{\rho}-\Pi^{\mu})\tilde{\rho}_{w^{n}}}
wnαwnγwn(μ)Δ(μ)Tr(ρ~wn)1,\displaystyle\leq\sum_{w^{n}}\frac{\alpha_{w^{n}}\gamma_{w^{n}}^{(\mu)}}{\Delta^{(\mu)}}\Tr{\tilde{\rho}_{w^{n}}}\leq 1, (61)

and defining H0H_{0} as above. and the last one follows by first using 𝔼[Δ(μ)]=(1ε)(1+η)\mathbb{E}[\Delta^{(\mu)}]=\frac{(1-\varepsilon)}{(1+\eta)} and then defining H1,H2H_{1},H_{2} and H3H_{3} as in the statement of the lemma and using the inequality Tr(Λ(ωσ))Λ(ωσ)1Λωσ1\Tr{\Lambda(\omega-\sigma)}\leq\|\Lambda(\omega-\sigma)\|_{1}\leq\|\Lambda\|_{\infty}\|\omega-\sigma\|_{1}. This completes the proof.

Appendix B Proof of Propositions

B.1 Proof of Proposition 1

Applying the triangle inequality on S~1\widetilde{S}_{1} gives S~1S~11+S~12\widetilde{S}_{1}\leq\widetilde{S}_{11}+\widetilde{S}_{12}, where

S~11\displaystyle\widetilde{S}_{11} \ensurestackMath\stackon[1pt]=Δ1Nμwnλwnρ^wnwnαwnγwn(μ)ρ^wn1,\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}{\frac{1}{N}\sum_{\mu}\left\|\sum_{w^{n}}\lambda_{w^{n}}\hat{\rho}_{w^{n}}-\sum_{w^{n}}\alpha_{w^{n}}\gamma_{w^{n}}^{(\mu)}\hat{\rho}_{w^{n}}\right\|_{1}},
S~12\displaystyle\widetilde{S}_{12} \ensurestackMath\stackon[1pt]=Δ1Nμwnαwnγwn(μ)ρ^wnρ~wn1.\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}{\frac{1}{N}\sum_{\mu}\sum_{w^{n}}\alpha_{w^{n}}\gamma_{w^{n}}^{(\mu)}\left\|\hat{\rho}_{w^{n}}-\tilde{\rho}_{w^{n}}\right\|_{1}}.

For the first term S~11\widetilde{S}_{11}, we use Lemma 5, and identify θwn\theta_{w^{n}} with ρ^wn\hat{\rho}_{w^{n}} and N=1N^{\prime}=1. Using this lemma, we obtain the following: For any ϵ>0\epsilon>0, and any η,δ(0,1)\eta,\delta\in(0,1) sufficiently small and any nn sufficiently large, 𝔼[S~11]ϵ\mathbb{E}[\widetilde{S}_{11}]\leq\epsilon, if the k+lnlogp>I(W;R)σS(W)σ+log(p)\frac{k+l}{n}\log p>I(W;R)_{\sigma}-S(W)_{\sigma}+\log{p}, where σ\sigma is defined in the statement of the theorem. As for the second term S~12\widetilde{S}_{12}, we use the gentle measurement lemma and bound its expected value as

𝔼\displaystyle\mathbb{E} [1Nμwnαwnγwn(μ)ρ^wnρ~wn1]\displaystyle\left[\frac{1}{N}\sum_{\mu}\sum_{w^{n}}\alpha_{w^{n}}\gamma_{w^{n}}^{(\mu)}\left\|\hat{\rho}_{w^{n}}-\tilde{\rho}_{w^{n}}\right\|_{1}\right]
wn𝒯δ(n)(W)λwn(1+η)ρ^wnρ~wn1+wn𝒯δ(n)(W)λwn(1+η)ϵS~12,\displaystyle\leq\sum_{w^{n}\in\mathcal{T}_{\delta}^{(n)}(W)}\frac{\lambda_{w^{n}}}{(1+\eta)}\left\|\hat{\rho}_{w^{n}}-\tilde{\rho}_{w^{n}}\right\|_{1}+\sum_{w^{n}\notin\mathcal{T}_{\delta}^{(n)}(W)}\frac{\lambda_{w^{n}}}{(1+\eta)}\leq\epsilon_{\scriptscriptstyle\widetilde{S}_{12}},

where the inequality is based on the repeated usage of the average gentle measurement lemma by setting ϵS~12=(1ε)(1+η)(2ε+2ε′′)\epsilon_{\scriptscriptstyle\widetilde{S}_{12}}=\frac{(1-\varepsilon)}{(1+\eta)}(2\sqrt{\varepsilon^{\prime}}+2\sqrt{\varepsilon^{\prime\prime}}) with ϵS~120\epsilon_{\scriptscriptstyle\widetilde{S}_{12}}\searrow 0 as nn\rightarrow\infty and ε=εp+2εp\varepsilon^{\prime}=\varepsilon^{\prime}_{p}+2\sqrt{\varepsilon^{\prime}_{p}} and ε′′=2εp+2εp\varepsilon^{\prime\prime}=2\varepsilon^{\prime}_{p}+2\sqrt{\varepsilon^{\prime}_{p}} for εp\ensurestackMath\stackon[1pt]=Δ1min{Tr(Πρρ^wn),Tr(Πwnρ^wn),1ε}\varepsilon^{\prime}_{p}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}1-\min\left\{\Tr{\Pi_{\rho}\hat{\rho}_{w^{n}}},\Tr{\Pi_{w^{n}}\hat{\rho}_{w^{n}}},1-\varepsilon\right\} (see (35) in [3] for more details).

B.2 Proof of Proposition 2

To provide a bound for S~2\widetilde{S}_{2}, we individually bound the terms corresponding to H0H_{0} and H~\widetilde{H} in an expected sense. Let us first consider H~\widetilde{H}. To provide a bound for H~\tilde{H} we use Lemma 2 with the following identification: λx\lambda_{x} with λwn(1ε)\frac{\lambda_{w^{n}}}{(1-\varepsilon)}, σx\sigma_{x} with ρ^wn\hat{\rho}_{w^{n}}, 𝒳\mathcal{X} with 𝒯δ(n)(W)\mathcal{T}_{\delta}^{(n)}(W), 𝒳¯\mathcal{\bar{X}} with 𝔽pn\mathbb{F}_{p}^{n}, Π\Pi with Πρ\Pi_{\rho}, Πx\Pi_{x} with Πwn\Pi_{w^{n}}, and μx\mu_{x} with 1pn\frac{1}{p^{n}}.

Firstly, we have λwn1/pn2n(S(W)σlog(p)δw)\frac{\lambda_{w^{n}}}{1/p^{n}}\leq 2^{-n{(S(W)_{\sigma}-\log{p}-\delta_{w})}} for all wn𝔽pnw^{n}\in\mathbb{F}_{p}^{n}, where δw(δ)0\delta_{w}(\delta)\searrow 0 as δ0\delta\searrow 0, which gives

κ=2n(S(W)σlog(p)δw).\kappa=2^{-n{(S(W)_{\sigma}-\log{p}-\delta_{w})}}.

With these, we check the hypotheses of Lemma 2. As for the first hypothesis (9a), using the pinching arguments described in [7, Property 15.2.7], we have Tr(Πρρ^wn)1ϵ\Tr{\Pi_{\rho}\hat{\rho}_{w^{n}}}\geq 1-\epsilon for all ϵ(0,1),δ>0\epsilon\in(0,1),\delta>0 and sufficiently large nn. Then the hypotheses (9b) and (9e) are satisfied from the construction of Πwn\Pi_{w^{n}}. Next, consider the hypothesis (9c). We have

Πρ(wn𝒯δ(n)(W)λwn(1ε)ρ^wn)1\displaystyle\Bigg{\|}\Pi_{\rho}\sqrt{\bigg{(}\sum_{w^{n}\in\mathcal{T}_{\delta}^{(n)}(W)}\frac{\lambda_{w^{n}}}{(1-\varepsilon)}\hat{\rho}_{w^{n}}\bigg{)}}\Bigg{\|}_{1}
11εTr(ΠρρnΠρ)2n2(S(R)σ+δρ),\displaystyle\leq\frac{1}{\sqrt{1-\varepsilon}}\Tr{\sqrt{\Pi_{\rho}\rho^{\otimes n}\Pi_{\rho}}}\leq 2^{\frac{n}{2}(S(R)_{\sigma}+\delta_{\rho}^{\prime})},

where the first inequality above follows from using wn𝒯δ(n)(W)λwn(1ε)ρ^wn1(1ε)ρn\sum_{w^{n}\in\mathcal{T}_{\delta}^{(n)}(W)}\frac{\lambda_{w^{n}}}{(1-\varepsilon)}\hat{\rho}_{w^{n}}\leq\frac{1}{(1-\varepsilon)}\rho^{\otimes n} and the operator monotonicty of the square-root function. The second inequality follows from the property of the typical projector for some δρ\delta_{\rho}^{\prime} such that δw0\delta_{w}^{\prime}\searrow 0 as δ0\delta\searrow 0. This gives

D=2n(S(R)σ+δρ),{D}={2^{n(S(R)_{\sigma}+\delta_{\rho}^{\prime})}},

where σ\sigma is as defined in the statement of the theorem. Finally, the hypotheses (9d) is satisfied from the property of conditional typical projectors for d=2n(S(R|W)σδw′′)d=2^{n(S(R|W)_{\sigma}-\delta_{w}^{\prime\prime})}, where δw′′0\delta_{w}^{\prime\prime}\searrow 0 as δ0.\delta\searrow 0. Next we check the pairwise independence of Wn,(μ)(a,m)W^{n,(\mu)}(a,m) and Wn,(μ)(a~,m~)W^{n,(\mu)}(\tilde{a},\tilde{m}). Since these are constructed using randomly and uniformly generated GG and h(μ)h^{(\mu)}, we have {Wn,(μ)(a,m)}a𝔽pk,m𝔽pl,μ[1,N]\{W^{n,(\mu)}(a,m)\}_{a\in\mathbb{F}^{k}_{p},m\in\mathbb{F}_{p}^{l},\mu\in[1,N]} to be pairwise independent (see [37] for details). Therefore, employing inequality (11) of Lemma 2, we get

𝔼[H~]2n(S(R)σ+δρ)2n(S(W)σlog(p)δw)N2nS2n(S(R|W)σδw′′)\displaystyle\mathbb{E}[\tilde{H}]\leq\sqrt{\frac{2^{n(S(R)_{\sigma}+\delta_{\rho}^{\prime})}2^{-n{(S(W)_{\sigma}-\log{p}-\delta_{w})}}}{N2^{nS}2^{n(S(R|W)_{\sigma}-\delta_{w}^{\prime\prime})}}}
2n2(k+lnlog(p)+1nlog(N)I(R;W)σlog(p)+S(W)σδwδρδw′′).\displaystyle\;\leq 2^{-\frac{n}{2}\left(\frac{k+l}{n}\log{p}+\frac{1}{n}\log{N}-I(R;W)_{\sigma}-\log{p}+S(W)_{\sigma}-\delta_{w}-\delta_{\rho}^{\prime}-\delta_{w}^{\prime\prime}\right)}.

Next, consider H0H_{0} and perform the following simplification

𝔼[H0]\displaystyle\mathbb{E}[H_{0}] =(1ε)(1+η)𝔼|wn𝒯δ(n)(W)λwn(1ε)\displaystyle=\frac{(1-\varepsilon)}{(1+\eta)}\mathbb{E}\bigg{|}\sum_{w^{n}\in\mathcal{T}_{\delta}^{(n)}(W)}\!\!\!\!\frac{\lambda_{w^{n}}}{(1-\varepsilon)}
pnpk+lwn𝒯δ(n)(W)a,iλwn(1ε)𝟙{Wn,(μ)(a,i)=wn}|\displaystyle\hskip 10.0pt-\frac{p^{n}}{p^{k+l}}\!\!\!\!\sum_{w^{n}\in\mathcal{T}_{\delta}^{(n)}(W)}\sum_{a,i}\frac{\lambda_{w^{n}}}{(1-\varepsilon)}\mathbbm{1}_{\{W^{n,(\mu)}(a,i)=w^{n}\}}\bigg{|}
=(1ε)(1+η)𝔼wn𝒯δ(n)(W)λwn(1ε)ω0npnpk+l\displaystyle=\frac{(1-\varepsilon)}{(1+\eta)}\mathbb{E}\bigg{\|}\sum_{w^{n}\in\mathcal{T}_{\delta}^{(n)}(W)}\!\!\!\!\frac{\lambda_{w^{n}}}{(1-\varepsilon)}\omega_{0}^{\otimes n}-\frac{p^{n}}{p^{k+l}}
×wn𝒯δ(n)(W)a,iλwn(1ε)𝟙{Wn,(μ)(a,i)=wn}ω0n1,\displaystyle\hskip 5.0pt\times\sum_{w^{n}\in\mathcal{T}_{\delta}^{(n)}(W)}\sum_{a,i}\frac{\lambda_{w^{n}}}{(1-\varepsilon)}\mathbbm{1}_{\{W^{n,(\mu)}(a,i)=w^{n}\}}\omega_{0}^{\otimes n}\bigg{\|}_{1}, (62)

where ω0𝒟()\omega_{0}\in\mathcal{D}(\mathcal{H}) is any state independent of WW. We again apply Lemma 2 to the above term with the following identification: λx\lambda_{x} with λwn(1ε)\frac{\lambda_{w^{n}}}{(1-\varepsilon)}, σx\sigma_{x} with ω0n\omega_{0}^{\otimes n}, 𝒳\mathcal{X} with 𝒯δ(n)(W)\mathcal{T}_{\delta}^{(n)}(W), 𝒳¯\mathcal{\bar{X}} with 𝔽pn\mathbb{F}_{p}^{n}, Π\Pi and Πx\Pi_{x} with Identity operator II, and μx\mu_{x} with 1pn\frac{1}{p^{n}}. With this identification, κ\kappa remains as above, κ=2n(S(W)σlog(p)δw)\kappa=2^{-n{(S(W)_{\sigma}-\log{p}-\delta_{w})}} and D=d=1D=d=1. Hence, using in inequality (11) of Lemma 2, we obtain

𝔼[H0]\displaystyle\mathbb{E}[H_{0}] 2n2(k+lnlog(p)log(p)+S(W)σδw).\displaystyle\leq 2^{-\frac{n}{2}\left(\frac{k+l}{n}\log{p}-\log{p}+S(W)_{\sigma}-\delta_{w}\right)}.

This completes the proof.

B.3 Proof of Proposition 3

We begin using the definition of Awn(μ)A_{w^{n}}^{(\mu)} and applying triangle inequality to S2S_{2} to obtain

S2\displaystyle S_{2} 1(1+η)1Nμa,i>0wn,znλwnpnpk+l𝟙{aG+h(μ)(i)=wn}\displaystyle\leq\frac{1}{(1+\eta)}\frac{1}{N}\sum_{\mu}\sum_{a,i>0}\sum_{w^{n},z^{n}}\cfrac{\lambda_{w^{n}}p^{n}}{p^{k+l}}\mathbbm{1}_{\{aG+h^{(\mu)}(i)=w^{n}\}}
×ρnΠμρn1ρ~wnρn1Πμρn1\displaystyle\hskip 15.0pt\times{\left\|\sqrt{\rho^{\otimes n}}\Pi^{\mu}\sqrt{\rho^{\otimes n}}^{-1}\tilde{\rho}_{w^{n}}\sqrt{\rho^{\otimes n}}^{-1}\Pi^{\mu}\sqrt{\rho^{\otimes n}}\right\|_{1}}
×|PZ|Wn(zn|wn)PZ|Wn(zn|F(μ)(i))|\displaystyle\hskip 50.58878pt\times\left|P^{n}_{Z|W}(z^{n}|w^{n})-P^{n}_{Z|W}\left(z^{n}|F^{(\mu)}(i)\right)\right|
22nδρ(1+η)1Nμa,i>0wn,znλwnpnpk+l𝟙{aG+h(μ)(i)=wn}\displaystyle\leq\frac{2^{2n\delta_{\rho}}}{(1+\eta)}\frac{1}{N}\sum_{\mu}\sum_{a,i>0}\sum_{w^{n},z^{n}}\cfrac{\lambda_{w^{n}}p^{n}}{p^{k+l}}\mathbbm{1}_{\{aG+h^{(\mu)}(i)=w^{n}\}}
×|PZ|Wn(zn|wn)PZ|Wn(zn|F(μ)(i))|\displaystyle\hskip 20.0pt\times\left|P^{n}_{Z|W}(z^{n}|w^{n})-P^{n}_{Z|W}\left(z^{n}|F^{(\mu)}(i)\right)\right|
22nδρ(1+η)1Nμa,i>0wn2λwnpnpk+l\displaystyle\leq\frac{2^{2n\delta_{\rho}}}{(1+\eta)}\frac{1}{N}\sum_{\mu}\sum_{a,i>0}\sum_{w^{n}}2\cfrac{\lambda_{w^{n}}p^{n}}{p^{k+l}}
×𝟙{aG+h(μ)(i)=wn}𝟙(μ)(wn,i),\displaystyle\hskip 65.04256pt\times\mathbbm{1}_{\{aG+h^{(\mu)}(i)=w^{n}\}}\mathbbm{1}^{(\mu)}(w^{n},i), (63)

where the second inequality above uses the following arguments

ρnΠμρn1ρ~wnρn1Πμρn1\displaystyle\left\|\sqrt{\rho^{\otimes n}}\Pi^{\mu}\sqrt{\rho^{\otimes n}}^{-1}\tilde{\rho}_{w^{n}}\sqrt{\rho^{\otimes n}}^{-1}\Pi^{\mu}\sqrt{\rho^{\otimes n}}\right\|_{1}
=ρnΠρΠμρn1Πρρ~wnΠρρn1ΠμΠρρn1\displaystyle=\left\|\sqrt{\rho^{\otimes n}}\Pi_{\rho}\Pi^{\mu}\sqrt{\rho^{\otimes n}}^{-1}\Pi_{\rho}\tilde{\rho}_{w^{n}}\Pi_{\rho}\sqrt{\rho^{\otimes n}}^{-1}\Pi^{\mu}\Pi_{\rho}\sqrt{\rho^{\otimes n}}\right\|_{1}
ρnΠρΠμρn1Πρρ~wnΠρρn1Πμ1\displaystyle\leq\left\|\sqrt{\rho^{\otimes n}}\Pi_{\rho}\right\|_{\infty}\left\|\Pi^{\mu}\sqrt{\rho^{\otimes n}}^{-1}\Pi_{\rho}\tilde{\rho}_{w^{n}}\Pi_{\rho}\sqrt{\rho^{\otimes n}}^{-1}\Pi^{\mu}\right\|_{1}
×ρnΠρ\displaystyle\hskip 20.0pt\times\left\|\sqrt{\rho^{\otimes n}}\Pi_{\rho}\right\|_{\infty}
2n(S(ρ)δρ)Πμ2ρn1Πρρ~wnΠρρn11\displaystyle\leq 2^{-n(S(\rho)-\delta_{\rho})}\left\|\Pi^{\mu}\right\|^{2}_{\infty}\left\|\sqrt{\rho^{\otimes n}}^{-1}\Pi_{\rho}\tilde{\rho}_{w^{n}}\Pi_{\rho}\sqrt{\rho^{\otimes n}}^{-1}\right\|_{1}
22nδρρ~wn122nδρ,\displaystyle\leq 2^{2n\delta_{\rho}}\left\|\tilde{\rho}_{w^{n}}\right\|_{1}\leq 2^{2n\delta_{\rho}}, (64)

where the above inequalities follow from the Hólder’s inequality. Finally, the last inequality in (63) follows by defining 𝟙(μ)(wn,i)\mathbbm{1}^{(\mu)}(w^{n},i) as

𝟙(μ)(wn,i)\ensurestackMath\stackon[1pt]=Δ𝟙{\displaystyle\mathbbm{1}^{(\mu)}(w^{n},i)\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\mathbbm{1}\bigg{\{}\exists (w~n,a~n):w~n=a~nG+h(μ)(i),\displaystyle(\tilde{w}^{n},\tilde{a}^{n}):\tilde{w}^{n}=\tilde{a}^{n}G+h^{(\mu)}(i),
w~n𝒯δ(n)(W),w~nwn}.\displaystyle\tilde{w}^{n}\in\mathcal{T}_{{\delta}}^{(n)}(W),\tilde{w}^{n}\neq w^{n}\bigg{\}}.

Observe that

𝔼[𝟙(μ)(wn,i)𝟙{aG+h(μ)(i)=wn}]a~𝔽pkw~𝒯δ(n)(W)w~wn1pnpn,\displaystyle\mathbb{E}[\mathbbm{1}^{(\mu)}(w^{n},i)\mathbbm{1}_{\{aG+h^{(\mu)}(i)=w^{n}\}}]\leq\sum_{\tilde{a}\in\mathbb{F}_{p}^{k}}\sum_{\begin{subarray}{c}\tilde{w}\in\mathcal{T}_{{\delta}}^{(n)}(W)\\ \tilde{w}\neq w^{n}\end{subarray}}\frac{1}{p^{n}p^{n}},

which follows from the pairwise independence of the codewords. Using this, we obtain

𝔼[S2]\displaystyle\mathbb{E}[S_{2}] 2 22nδρ(1+η)2nRpk+lpnw~n𝒯δ(n)(W)wn𝒯δ(n)(W)λwn\displaystyle\leq\frac{2\;2^{2n\delta_{\rho}}}{(1+\eta)}\frac{2^{-nR}p^{k+l}}{p^{n}}\sum_{\tilde{w}^{n}\in\mathcal{T}_{{\delta}}^{(n)}(W)}\sum_{{w}^{n}\in\mathcal{T}_{{\delta}}^{(n)}(W)}\lambda_{w^{n}}
2 2n(k+lnlog(p)Rlog(p)+S(W)σ+δS2),\displaystyle\leq 2\;{2^{n(\frac{k+l}{n}\log{p}-R-\log{p}+S(W)_{\sigma}+\delta_{S_{2}})}},

where δS20\delta_{S_{2}}\searrow 0 as δ0\delta\searrow 0, and σ\sigma is as defined in the statement of the theorem. This completes the proof.

B.4 Proof of Proposition 4

Recalling S2S_{2}, we have S2S21+S22S_{2}\leq S_{21}+S_{22}, where

S21\displaystyle S_{21} \ensurestackMath\stackon[1pt]=Δ2N1N2μ¯1,μ¯2un,vnαunβvnγun(μ¯1)ζvn(μ¯2)Ωun,vn\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\frac{2}{N_{1}N_{2}}\sum_{\bar{\mu}_{1},\bar{\mu}_{2}}\sum_{u^{n},v^{n}}\alpha_{u^{n}}\beta_{v^{n}}\gamma_{u^{n}}^{(\bar{\mu}_{1})}\zeta_{v^{n}}^{(\bar{\mu}_{2})}\Omega_{u^{n},v^{n}}
×𝟙{(un,vn)𝒯δ(n)(U,V)},\displaystyle\hskip 115.63243pt\times\mathbbm{1}_{\{(u^{n},v^{n})\not\in\mathcal{T}_{\delta}^{(n)}(U,V)\}},
S22\displaystyle S_{22} \ensurestackMath\stackon[1pt]=Δ2N¯1N¯2μ¯1,μ¯2un,vnαunβvnγun(μ¯1)ζvn(μ¯2)Ωun,vn\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\frac{2}{\bar{N}_{1}\bar{N}_{2}}\sum_{\bar{\mu}_{1},\bar{\mu}_{2}}\sum_{u^{n},v^{n}}\alpha_{u^{n}}\beta_{v^{n}}\gamma_{u^{n}}^{(\bar{\mu}_{1})}\zeta_{v^{n}}^{(\bar{\mu}_{2})}\Omega_{u^{n},v^{n}}
×𝟙(μ¯1,μ¯2)(un+vn,i,j),\displaystyle\hskip 115.63243pt\times\mathbbm{1}^{(\bar{\mu}_{1},\bar{\mu}_{2})}(u^{n}+v^{n},i,j),

where Ωun,vn\Omega_{u^{n},v^{n}} and 𝟙(μ¯1,μ¯2)(wn,i,j)\mathbbm{1}^{(\bar{\mu}_{1},\bar{\mu}_{2})}(w^{n},i,j) are defined as

Ωun,vn\ensurestackMath\stackon[1pt]=ΔTr{[(ΠAμ¯1ΠBμ¯2)ρAnρBn1(ρ~unA\displaystyle\Omega_{u^{n},v^{n}}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\Tr\{\left[\left(\Pi_{A}^{\bar{\mu}_{1}}\otimes\Pi_{B}^{\bar{\mu}_{2}}\right)\sqrt{\rho^{\otimes n}_{A}\otimes\rho^{\otimes n}_{B}}^{-1}(\tilde{\rho}_{u^{n}}^{A}\otimes\right.
ρ~vnB)ρAnρBn1(ΠAμ¯1ΠBμ¯2)]ρABn},\displaystyle\hskip 65.04256pt\left.\tilde{\rho}_{v^{n}}^{B})\sqrt{\rho^{\otimes n}_{A}\otimes\rho^{\otimes n}_{B}}^{-1}\left(\Pi_{A}^{\bar{\mu}_{1}}\otimes\Pi_{B}^{\bar{\mu}_{2}}\right)\right]\rho^{\otimes n}_{AB}\Big{\}},
𝟙(μ¯1,μ¯2)(wn,i,j)\displaystyle\mathbbm{1}^{(\bar{\mu}_{1},\bar{\mu}_{2})}(w^{n},i,j)
\ensurestackMath\stackon[1pt]=Δ𝟙{(w~n,a~n):w~n=a~nG+h1(μ¯1)(i)+h2(μ¯2)(j),\displaystyle\hskip 10.0pt\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\mathbbm{1}\bigg{\{}\exists(\tilde{w}^{n},\tilde{a}^{n}):\tilde{w}^{n}=\tilde{a}^{n}G+h_{1}^{(\bar{\mu}_{1})}(i)+h_{2}^{(\bar{\mu}_{2})}(j),
w~n𝒯δ^(n)(U+V),w~nwn}.\displaystyle\hskip 108.405pt\tilde{w}^{n}\in\mathcal{T}_{\hat{\delta}}^{(n)}(U+V),\tilde{w}^{n}\neq w^{n}\bigg{\}}.

We begin by bounding the term corresponding to S21S_{21}. Consider the following argument.

S21\displaystyle S_{21} |2N¯1N¯2μ¯1,μ¯2un,vnαunβvnγun(μ¯1)ζvn(μ¯2)Ωun,vn𝟙{(un,vn)𝒯δ(n)(U,V)}(un,vn)𝒯δ(n)(UV)un𝒯δ(n)(U),vn𝒯δ(n)(V)2λun,vnAB|+(un,vn)𝒯δ(n)(U,V)2λun,vnAB\displaystyle\leq\Bigg{|}\frac{2}{\bar{N}_{1}\bar{N}_{2}}\sum_{\bar{\mu}_{1},\bar{\mu}_{2}}\sum_{u^{n},v^{n}}\alpha_{u^{n}}\beta_{v^{n}}\gamma_{u^{n}}^{(\bar{\mu}_{1})}\zeta_{v^{n}}^{(\bar{\mu}_{2})}\Omega_{u^{n},v^{n}}\mathbbm{1}_{\{(u^{n},v^{n})\not\in\mathcal{T}_{\delta}^{(n)}(U,V)\}}-\sum_{\begin{subarray}{c}(u^{n},v^{n})\not\in\mathcal{T}_{\delta}^{(n)}(UV)\\ u^{n}\in\mathcal{T}_{\delta}^{(n)}(U),v^{n}\in\mathcal{T}_{\delta}^{(n)}(V)\end{subarray}}2\lambda_{u^{n},v^{n}}^{AB}\Bigg{|}+\sum_{(u^{n},v^{n})\not\in\mathcal{T}_{\delta}^{(n)}(U,V)}2\lambda_{u^{n},v^{n}}^{AB}
(a)2un𝒰nvn𝒱n|λun,vnAB1N¯1N¯2μ¯1,μ¯2αunβvnγun(μ¯1)ζvn(μ¯2)Ωun,vn|+(un,vn)𝒯δ(n)(UV)2λu,vAB\displaystyle\overset{(a)}{\leq}2\sum_{u^{n}\in\mathcal{U}^{n}}\sum_{v^{n}\in\mathcal{V}^{n}}\left|\lambda^{AB}_{u^{n},v^{n}}-\frac{1}{\bar{N}_{1}\bar{N}_{2}}\sum_{\bar{\mu}_{1},\bar{\mu}_{2}}\alpha_{u^{n}}\beta_{v^{n}}\gamma^{(\bar{\mu}_{1})}_{u^{n}}\zeta^{(\bar{\mu}_{2})}_{v^{n}}\Omega_{u^{n},v^{n}}\right|+\sum_{(u^{n},v^{n})\not\in\mathcal{T}_{\delta}^{(n)}{(UV)}}2\lambda^{AB}_{u,v}
(b)2S~1+2(un,vn)𝒯δ(n)(UV)λun,vnAB,\displaystyle\overset{(b)}{\leq}2\tilde{S}_{1}+2\sum_{(u^{n},v^{n})\not\in\mathcal{T}_{\delta}^{(n)}{(UV)}}\lambda^{AB}_{u^{n},v^{n}},

where

S~1\displaystyle\tilde{S}_{1} \ensurestackMath\stackon[1pt]=Δ(idM¯AnM¯Bn)(ΨRABρ)n\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\bigg{\|}(\text{id}\otimes\bar{M}_{A}^{\otimes n}\otimes\bar{M}_{B}^{\otimes n})(\Psi^{\rho}_{RAB})^{\otimes n}-
1N¯1N¯2μ¯1,μ¯2(id[M1(μ¯1)][M2(μ¯2)])(ΨRABρ)n1,\displaystyle\hskip 15.0pt\frac{1}{\bar{N}_{1}\bar{N}_{2}}\sum_{\bar{\mu}_{1},\bar{\mu}_{2}}(\text{id}\otimes[M_{1}^{(\bar{\mu}_{1})}]\otimes[M_{2}^{(\bar{\mu}_{2})}])(\Psi^{\rho}_{RAB})^{\otimes n}\bigg{\|}_{1},

(a) follows by applying the triangle inequality, and (b) follows from the Lemma 9 given below. Note that in S~1\tilde{S}_{1}, the average over the entire common information sequence (μ¯1,μ¯2)(\bar{\mu}_{1},\bar{\mu}_{2}) is inside the norm.

Lemma 9.

We have

un𝒰n\displaystyle\sum_{u^{n}\in\mathcal{U}^{n}} vn𝒱n|λun,vnAB\displaystyle\sum_{v^{n}\in\mathcal{V}^{n}}\bigg{|}\lambda^{AB}_{u^{n},v^{n}}-
1N¯1N¯2μ¯1,μ¯2αunβvnγun(μ¯1)ζvn(μ¯2)Ωun,vn|S1.\displaystyle\frac{1}{\bar{N}_{1}\bar{N}_{2}}\sum_{\bar{\mu}_{1},\bar{\mu}_{2}}\alpha_{u^{n}}\beta_{v^{n}}\gamma_{u^{n}}^{(\bar{\mu}_{1})}\zeta_{v^{n}}^{(\bar{\mu}_{2})}\Omega_{u^{n},v^{n}}\bigg{|}\leq S_{1}. (65)
Proof.

The proof follows from Lemma 2 in [3]. ∎

Next we use Theorem 2 twice with (a) ρ=ρA\rho=\rho_{A}, M=M¯AM=\bar{M}_{A}, 𝒲=𝒰\mathcal{W}=\mathcal{U}, 𝒵=𝒰\mathcal{Z}=\mathcal{U} and PZ|W(z|w)=𝟙{z=w}P_{Z|W}(z|w)=\mathbbm{1}{\{z=w\}}, and (b) ρ=ρB\rho=\rho_{B}, M=M¯BM=\bar{M}_{B}, 𝒲=𝒱\mathcal{W}=\mathcal{V}, 𝒵=𝒱\mathcal{Z}=\mathcal{V} and PZ|W(z|w)=𝟙{z=w}P_{Z|W}(z|w)=\mathbbm{1}{\{z=w\}}, and [24, Lemma 5] to yield the following: for any ϵ(0,1)\epsilon\in(0,1), and any η,δ(0,1)\eta,\delta\in(0,1) sufficiently small, and any nn sufficiently large 𝔼[S1]2ϵ\mathbb{E}[S_{1}]\leq 2\epsilon if k+l1nlogp>I(U;RB)σ1S(U)σ3+logp\frac{k+l_{1}}{n}\log p>I(U;RB)_{\sigma_{1}}-S(U)_{\sigma_{3}}+\log p, k+l2nlogp>I(V;RA)σ2S(V)σ3+logp\frac{k+l_{2}}{n}\log p>I(V;RA)_{\sigma_{2}}-S(V)_{\sigma_{3}}+\log p, k+l1nlogp+1nlogN¯1>logp\frac{k+l_{1}}{n}\log p+\frac{1}{n}\log\bar{N}_{1}>\log p, k+l2nlogp+1nlogN¯2>logp\frac{k+l_{2}}{n}\log p+\frac{1}{n}\log\bar{N}_{2}>\log p, where σ1,σ2\sigma_{1},\sigma_{2} and σ3\sigma_{3} are defined as in the statement of the theorem. Consequently, we have 𝔼[S21]4ϵ\mathbb{E}[S_{21}]\leq 4\epsilon for all sufficiently large nn.

In regards to S22S_{22}, note that

𝔼[𝟙(μ¯1,μ¯2)(un+vn,i,j)𝟙{a1G+h1(μ¯1)(i)=un}\displaystyle\mathbb{E}\big{[}\mathbbm{1}^{(\bar{\mu}_{1},\bar{\mu}_{2})}(u^{n}+v^{n},i,j)\mathbbm{1}_{\{a_{1}G+h_{1}^{(\bar{\mu}_{1})}(i)=u^{n}\}}
𝟙{a2G+h2(μ¯2)(j)=vn}]a~𝔽pka~aw~𝒯δ^(n)(U+V)w~un+vn1pnpnpn.\displaystyle\hskip 30.0pt\mathbbm{1}_{\{a_{2}G+h_{2}^{(\bar{\mu}_{2})}(j)=v^{n}\}}\bigg{]}\leq\sum_{\begin{subarray}{c}\tilde{a}\in\mathbb{F}_{p}^{k}\\ \tilde{a}\neq a\end{subarray}}\;\;\sum_{\begin{subarray}{c}\tilde{w}\in\mathcal{T}_{\hat{\delta}}^{(n)}(U+V)\\ \tilde{w}\neq u^{n}+v^{n}\end{subarray}}\frac{1}{p^{n}p^{n}p^{n}}.

Using this, we obtain

𝔼[S22]\displaystyle\mathbb{E}[S_{22}] 2(1+η)2pk+l12nR1pnw~n𝒯δ^(n)(U+V)\displaystyle\leq\cfrac{2}{(1+\eta)^{2}}\frac{p^{k+l_{1}}2^{nR_{1}}}{p^{n}}\sum_{\tilde{w}^{n}\in\mathcal{T}_{\hat{\delta}}^{(n)}(U+V)}
un𝒯δ(n)(U)vn𝒯δ(n)(V)λunAλvnBΩun,vn\displaystyle\hskip 30.0pt\sum_{u^{n}\in\mathcal{T}_{\delta}^{(n)}(U)}\sum_{v^{n}\in\mathcal{T}_{\delta}^{(n)}(V)}\lambda_{u^{n}}^{A}\lambda_{v^{n}}^{B}\Omega_{u^{n},v^{n}}
2 2n(k+l1nlog(p)R1log(p)+S(U+V)σ3+δρAB+δ^W)(1+η)2,\displaystyle\leq\frac{2\;{2^{n(\frac{k+l_{1}}{n}\log{p}-R_{1}-\log{p}+S(U+V)_{\sigma_{3}}+\delta_{\rho_{AB}}+\hat{\delta}_{W})}}}{(1+\eta)^{2}},

where δ^W0\hat{\delta}_{W}\searrow 0 as δ0\delta\searrow 0 and the above inequality follows from the following lemma (Lemma 10). Hence, 𝔼[S21]ϵ\mathbb{E}[S_{21}]\leq\epsilon if the conditions in the proposition are satisfied.

Lemma 10.

For λunA,λvnB\lambda^{A}_{u^{n}},\lambda^{B}_{v^{n}} and Ωun,vn\Omega_{u^{n},v^{n}} as defined above, we have

un𝒯δ(n)(U)vn𝒯δ(n)(V)Ωun,vnλunAλvnB2nδρAB,\sum_{u^{n}\in\mathcal{T}_{\delta}^{(n)}(U)}\sum_{v^{n}\in\mathcal{T}_{\delta}^{(n)}(V)}\Omega_{u^{n},v^{n}}\lambda^{A}_{u^{n}}\lambda^{B}_{v^{n}}\leq 2^{n\delta_{\rho_{AB}}},

for some δρAB0\delta_{\rho_{AB}}\searrow 0 as δ0.\delta\searrow 0.

Proof.

Firstly, note that

un,vnΩun,vnλunAλvnB\displaystyle\sum_{\begin{subarray}{c}u^{n},v^{n}\end{subarray}}\Omega_{u^{n},v^{n}}\lambda^{A}_{u^{n}}\lambda^{B}_{v^{n}}
=Tr{[(ΠAμ¯1ΠBμ¯2)(ρAn1(unλunAρ~unA)ρAn1\displaystyle\hskip 1.0pt=\text{Tr}\bigg{\{}\bigg{[}\!\!\left(\Pi_{A}^{\bar{\mu}_{1}}\otimes\Pi_{B}^{\bar{\mu}_{2}}\right)\!\bigg{(}\!\sqrt{\rho_{A}^{\otimes n}}^{-1}\!\!\Big{(}\sum_{u^{n}}\lambda^{A}_{u^{n}}\tilde{\rho}_{u^{n}}^{A}\Big{)}\sqrt{\rho_{A}^{\otimes n}}^{-1}
ρBn1(vnλvnBρ~vnB)ρBn1)(ΠAμ¯1ΠBμ¯2)]ρABn}.\displaystyle\;\;\;\otimes\sqrt{\rho_{B}^{\otimes n}}^{-1}\!\!\Big{(}\sum_{v^{n}}\lambda^{B}_{v^{n}}\tilde{\rho}_{v^{n}}^{B}\Big{)}\sqrt{\rho_{B}^{\otimes n}}^{-1}\bigg{)}\!\!\left(\Pi_{A}^{\bar{\mu}_{1}}\otimes\Pi_{B}^{\bar{\mu}_{2}}\right)\!\!\bigg{]}\rho^{\otimes n}_{AB}\bigg{\}}. (66)

We know, unλunAρ~unA2n(S(ρA)δρA)ΠρA\sum_{u^{n}}\lambda^{A}_{u^{n}}\tilde{\rho}_{u^{n}}^{A}\leq 2^{-n(S(\rho_{A})-\delta_{\rho_{A}})}\Pi_{\rho_{A}}, where δρA0\delta_{\rho_{A}}\searrow 0 as δ0\delta\searrow 0. This implies,

ΠAμ¯1\displaystyle\Pi_{A}^{\bar{\mu}_{1}} ρAn1(unλunAρ~unA)ρAn1ΠAμ¯1\displaystyle\sqrt{\rho_{A}^{\otimes n}}^{-1}\left(\sum_{u^{n}}\lambda^{A}_{u^{n}}\tilde{\rho}_{u^{n}}^{A}\right)\sqrt{\rho_{A}^{\otimes n}}^{-1}\Pi_{A}^{\bar{\mu}_{1}}
2n(S(ρA)δρA)ΠAμ¯1ρAn1ΠρAρAn1ΠAμ¯1\displaystyle\leq 2^{-n(S(\rho_{A})-\delta_{\rho_{A}})}\Pi_{A}^{\bar{\mu}_{1}}\sqrt{\rho_{A}^{\otimes n}}^{-1}\Pi_{\rho_{A}}\sqrt{\rho_{A}^{\otimes n}}^{-1}\Pi_{A}^{\bar{\mu}_{1}}
22nδρAΠAμ¯1ΠρAΠAμ¯122nδρAΠAμ¯1,\displaystyle\leq 2^{2n\delta_{\rho_{A}}}\Pi_{A}^{\bar{\mu}_{1}}\Pi_{\rho_{A}}\Pi_{A}^{\bar{\mu}_{1}}\leq 2^{2n\delta_{\rho_{A}}}\Pi_{A}^{\bar{\mu}_{1}}, (67)

where the second inequality appeals to the fact that ρAn1ΠρAρAn12n(S(ρA)+δρA)ΠρA\sqrt{\rho_{A}^{\otimes n}}^{-1}\Pi_{\rho_{A}}\sqrt{\rho_{A}^{\otimes n}}^{-1}\leq 2^{n(S(\rho_{A})+\delta_{\rho_{A}})}\Pi_{\rho_{A}}. Similarly, using the same arguments above for the operators acting on B\mathcal{H}_{B}, we have

ΠBμ¯2ρBn1(vnλvnBρ~vnB)ρBn1ΠBμ¯222nδρBΠBμ¯2,\Pi_{B}^{\bar{\mu}_{2}}\sqrt{\rho_{B}^{\otimes n}}^{-1}\left(\sum_{v^{n}}\lambda^{B}_{v^{n}}\tilde{\rho}_{v^{n}}^{B}\right)\sqrt{\rho_{B}^{\otimes n}}^{-1}\Pi_{B}^{\bar{\mu}_{2}}\leq 2^{2n\delta_{\rho_{B}}}\Pi_{B}^{\bar{\mu}_{2}}, (68)

where δρB0\delta_{\rho_{B}}\searrow 0 as δ0\delta\searrow 0. Using (i) the simplifications in (67) and (68), and (ii) the fact that for A1B10A_{1}\geq B_{1}\geq 0 and A2B20A_{2}\geq B_{2}\geq 0, (A1A2)(B1B2)(A_{1}\otimes A_{2})\geq(B_{1}\otimes B_{2}) in (66), gives

un,vn\displaystyle\sum_{u^{n},v^{n}} Ωun,vnλunAλvnB\displaystyle\Omega_{u^{n},v^{n}}\lambda^{A}_{u^{n}}\lambda^{B}_{v^{n}}
22n(δρA+δρB)Tr((ΠAμ¯1ΠBμ¯2)ρABn)\displaystyle\leq 2^{2n(\delta_{\rho_{A}}+\delta_{\rho_{B}})}\Tr{\left(\Pi_{A}^{\bar{\mu}_{1}}\otimes\Pi_{B}^{\bar{\mu}_{2}}\right)\rho^{\otimes n}_{AB}}
22n(δρA+δρB)Tr(ρABn)=22n(δρA+δρB).\displaystyle\leq 2^{2n(\delta_{\rho_{A}}+\delta_{\rho_{B}})}\Tr{\rho^{\otimes n}_{{AB}}}=2^{2n(\delta_{\rho_{A}}+\delta_{\rho_{B}})}.

Substituting δρAB=2(δρA+δρB)\delta_{\rho_{AB}}=2(\delta_{\rho_{A}}+\delta_{\rho_{B}}) gives the result.

B.5 Proof of Proposition 5

We bound S~\widetilde{S} as S~S~2+S~3+S~4\widetilde{S}\leq\widetilde{S}_{2}+\widetilde{S}_{3}+\widetilde{S}_{4}, where

S~2\ensurestackMath\stackon[1pt]=Δ\displaystyle\widetilde{S}_{2}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}} 1N1N2μ1,μ2i>0ρABn(ΓiA,(μ1)Γ0B,(μ2))\displaystyle\bigg{\|}\frac{1}{N_{1}N_{2}}\sum_{\mu_{1},\mu_{2}}\sum_{i>0}\sqrt{\rho_{AB}^{\otimes n}}\left(\Gamma^{A,(\mu_{1})}_{i}\otimes\Gamma^{B,(\mu_{2})}_{0}\right)
×ρABnPZ|U+Vn(zn|w0n)1,\displaystyle\hskip 108.405pt\times\sqrt{\rho_{AB}^{\otimes n}}P^{n}_{Z|U+V}(z^{n}|w^{n}_{0})\bigg{\|}_{1},
S~3\ensurestackMath\stackon[1pt]=Δ\displaystyle\widetilde{S}_{3}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}} 1N1N2μ1,μ2j>0ρABn(Γ0A,(μ1)ΓjB,(μ2))\displaystyle\bigg{\|}\frac{1}{N_{1}N_{2}}\sum_{\mu_{1},\mu_{2}}\sum_{j>0}\sqrt{\rho_{AB}^{\otimes n}}\left(\Gamma^{A,(\mu_{1})}_{0}\otimes\Gamma^{B,(\mu_{2})}_{j}\right)
×ρABnPZ|U+Vn(zn|w0n)1,\displaystyle\hskip 108.405pt\times\sqrt{\rho_{AB}^{\otimes n}}P^{n}_{Z|U+V}(z^{n}|w^{n}_{0})\bigg{\|}_{1},
S~4\ensurestackMath\stackon[1pt]=Δ\displaystyle\widetilde{S}_{4}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}} 1N1N2μ1,μ2ρABn(Γ0A,(μ1)Γ0B,(μ2))\displaystyle\bigg{\|}\frac{1}{N_{1}N_{2}}\sum_{\mu_{1},\mu_{2}}\sqrt{\rho_{AB}^{\otimes n}}\left(\Gamma^{A,(\mu_{1})}_{0}\otimes\Gamma^{B,(\mu_{2})}_{0}\right)
×ρABnPZ|U+Vn(zn|w0n)1.\displaystyle\hskip 108.405pt\times\sqrt{\rho_{AB}^{\otimes n}}P^{n}_{Z|U+V}(z^{n}|w^{n}_{0})\bigg{\|}_{1}.

Analysis of S~2\widetilde{S}_{2}: We have

S~2\displaystyle\widetilde{S}_{2} 1N1N2μ1,μ2i>0znPZ|U+Vn(zn|w0n)ρABn(ΓiA,(μ1)Γ0B,(μ2))ρABn1\displaystyle\leq\frac{1}{N_{1}N_{2}}\sum_{\mu_{1},\mu_{2}}\sum_{i>0}\sum_{z^{n}}P^{n}_{Z|U+V}(z^{n}|w^{n}_{0})\left\|\sqrt{\rho_{AB}^{\otimes n}}\left(\Gamma^{A,(\mu_{1})}_{i}\otimes\Gamma^{B,(\mu_{2})}_{0}\right)\sqrt{\rho_{AB}^{\otimes n}}\right\|_{1}
1N1N2μ1,μ2ρBnΓ0B,(μ2)ρBn1\displaystyle\leq\frac{1}{N_{1}N_{2}}\sum_{\mu_{1},\mu_{2}}\left\|\sqrt{\rho_{B}^{\otimes n}}\Gamma^{B,(\mu_{2})}_{0}\sqrt{\rho_{B}^{\otimes n}}\right\|_{1}
1N2μ2vnλvnBρ^vnBvnρBnζvn(μ2)B¯vn(μ2)ρBn1+1N2μ2vnρBnζvn(μ2)(B¯vn(μ2)Bvn(μ2))ρBn1\displaystyle\leq\frac{1}{N_{2}}\sum_{\mu_{2}}\left\|\sum_{v^{n}}\lambda^{B}_{v^{n}}\hat{\rho}^{B}_{v^{n}}-\sum_{v^{n}}\sqrt{\rho_{B}^{\otimes n}}\zeta_{v^{n}}^{(\mu_{2})}\bar{B}^{(\mu_{2})}_{v^{n}}\sqrt{\rho_{B}^{\otimes n}}\right\|_{1}+\frac{1}{N_{2}}\sum_{\mu_{2}}\left\|\sum_{v^{n}}\sqrt{\rho_{B}^{\otimes n}}\zeta_{v^{n}}^{(\mu_{2})}\left(\bar{B}^{(\mu_{2})}_{v^{n}}-B^{(\mu_{2})}_{v^{n}}\right)\sqrt{\rho_{B}^{\otimes n}}\right\|_{1}
1N2μ2vnλvnBρ^vnB1(1+η)pnpk+l2vna2,jλvnBρ^vnB𝟙{Vn,(μ2)(a2,j)=vn}1S~21+1N2μ2vnβvnζvn(μ2)ρ^vnBρ~vnB1S~22\displaystyle\leq\underbrace{\frac{1}{N_{2}}\sum_{\mu_{2}}\left\|\sum_{v^{n}}\lambda^{B}_{v^{n}}\hat{\rho}^{B}_{v^{n}}-\cfrac{1}{(1+\eta)}\cfrac{p^{n}}{p^{k+l_{2}}}\sum_{v^{n}}\sum_{a_{2},j}\lambda^{B}_{v^{n}}\hat{\rho}^{B}_{v^{n}}\mathbbm{1}_{\{V^{n,(\mu_{2})}(a_{2},j)=v^{n}\}}\right\|_{1}}_{\widetilde{S}_{21}}+\underbrace{\frac{1}{N_{2}}\sum_{\mu_{2}}\sum_{v^{n}}\beta_{v^{n}}\zeta_{v^{n}}^{(\mu_{2})}\left\|\hat{\rho}^{B}_{v^{n}}-\tilde{\rho}_{v^{n}}^{B}\right\|_{1}}_{\widetilde{S}_{22}}
+1N2μ2vnρBnζvn(μ2)(B¯vn(μ2)Bvn(μ2))ρBn1S~23,\displaystyle\hskip 100.0pt+\underbrace{\frac{1}{N_{2}}\sum_{\mu_{2}}\sum_{v^{n}}\left\|\sqrt{\rho_{B}^{\otimes n}}\zeta_{v^{n}}^{(\mu_{2})}\left(\bar{B}^{(\mu_{2})}_{v^{n}}-B^{(\mu_{2})}_{v^{n}}\right)\sqrt{\rho_{B}^{\otimes n}}\right\|_{1}}_{\widetilde{S}_{23}}, (69)

where the first inequality uses triangle inequality. The next inequality follows by using Lemma 1 where we use the fact that i>0ΓiA,(μ1)I.\sum_{i>0}\Gamma^{A,(\mu_{1})}_{i}\leq I. Finally, the last two inequalities follows again from triangle inequality.

Regarding the first term in (69), using Lemma 5 we claim that for all ϵ>0\epsilon>0, and η,δ(0,1)\eta,\delta\in(0,1) sufficiently small, and any nn sufficiently large, 𝔼[S~21]<ϵ\mathbb{E}[\tilde{S}_{21}]<\epsilon, if k+l2nlog(p)I(V;RA)σ2S(V)σ3+log(p)\frac{k+l_{2}}{n}\log{p}\geq I(V;RA)_{\sigma_{2}}-S(V)_{\sigma_{3}}+\log{p}, where σ2,σ3\sigma_{2},\sigma_{3} are as defined in the statement of the theorem. As for the second term, we use the gentle measurement lemma (as in (76)) and bound its expected value as

𝔼[S~22]\displaystyle\mathbb{E}[\tilde{S}_{22}]\! =𝔼[1N2μ2vnβvnζvn(μ2)ρ^vnBρ~vnB1]\displaystyle=\mathbb{E}\left[\frac{1}{N_{2}}\sum_{\mu_{2}}\sum_{v^{n}}\beta_{v^{n}}\zeta_{v^{n}}^{(\mu_{2})}\left\|\hat{\rho}^{B}_{v^{n}}-\tilde{\rho}_{v^{n}}^{B}\right\|_{1}\right]
=vn𝒯δ(n)(V)λvnB(1+η)ρ^vnBρ~vnB1+vn𝒯δ(n)(V)λvnB(1+η)ρ^vnB1\displaystyle=\sum_{v^{n}\in\mathcal{T}_{\delta}^{(n)}(V)}\frac{\lambda^{B}_{v^{n}}}{(1+\eta)}\!\!\left\|\hat{\rho}^{B}_{v^{n}}-\tilde{\rho}_{v^{n}}^{B}\right\|_{1}+\sum_{v^{n}\notin\mathcal{T}_{\delta}^{(n)}(V)}\frac{\lambda^{B}_{v^{n}}}{(1+\eta)}\!\!\left\|\hat{\rho}^{B}_{v^{n}}\right\|_{1}
ϵS~21,\displaystyle\leq\epsilon_{\scriptscriptstyle\widetilde{S}_{21}},

where the inequality is based on the repeated usage of the Average Gentle Measurement Lemma and ϵS~210\epsilon_{\scriptscriptstyle\widetilde{S}_{21}}\searrow 0 as δ0\delta\searrow 0 (see (35) in [3] for more details). Finally, consider the last term. To simplify this term, we appeal to Lemma 6 in Section V.2. This gives us

S~232 23nδN2μ2=1N2(H0B+(1εB)(1+η)H1B+H2B+H3B),\displaystyle\tilde{S}_{23}\leq\frac{2\;{2^{3n\delta}}}{N_{2}}\sum_{\mu_{2}=1}^{N_{2}}\!\!\left(\!\!H^{B}_{0}\!+\!\frac{\sqrt{(1-\varepsilon_{B})}}{(1+\eta)}\sqrt{H_{1}^{B}+H_{2}^{B}+H_{3}^{B}}\right), (70)

where

H0B\displaystyle H_{0}^{B} \ensurestackMath\stackon[1pt]=Δ|ΔB(μ2)𝔼[ΔB(μ2)]|,\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\left|\Delta^{(\mu_{2})}_{B}-\mathbb{E}[\Delta^{(\mu_{2})}_{B}]\right|,
H1B\displaystyle H_{1}^{B} \ensurestackMath\stackon[1pt]=ΔTr((ΠρBΠBμ2)vnλvnBρ~vnB),\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\Tr{(\Pi_{\rho_{B}}-\Pi_{B}^{\mu_{2}})\sum_{v^{n}}\lambda^{B}_{v^{n}}\tilde{\rho}_{v^{n}}^{B}},
H2B\displaystyle H_{2}^{B} \ensurestackMath\stackon[1pt]=ΔvnλvnBρ~vnB(1εB)vnβvnζvn(μ2)𝔼[ΔB(μ)]ρ~vnB1,\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\left\|\sum_{v^{n}}\lambda^{B}_{v^{n}}\tilde{\rho}_{v^{n}}^{B}-(1-\varepsilon_{B})\sum_{v^{n}}\frac{\beta_{v^{n}}\zeta_{v^{n}}^{(\mu_{2})}}{\mathbb{E}[\Delta^{(\mu)}_{B}]}\tilde{\rho}_{v^{n}}^{B}\right\|_{1},
H3B\displaystyle H_{3}^{B} \ensurestackMath\stackon[1pt]=Δ(1εB)vnβvnζvn(μ2)ΔB(μ2)ρ~vnBvnβvnζvn(μ2)𝔼[ΔB(μ2)]ρ~vnB1,\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}(1-\varepsilon_{B})\left\|\sum_{v^{n}}\frac{\beta_{v^{n}}\zeta_{v^{n}}^{(\mu_{2})}}{\Delta^{(\mu_{2})}_{B}}\tilde{\rho}_{v^{n}}^{B}-\sum_{v^{n}}\frac{\beta_{v^{n}}\zeta_{v^{n}}^{(\mu_{2})}}{\mathbb{E}[\Delta^{(\mu_{2})}_{B}]}\tilde{\rho}_{v^{n}}^{B}\right\|_{1}, (71)

and ΔB(μ)\ensurestackMath\stackon[1pt]=Δvn𝒯δ(n)(V)βvnζvn(μ2)\Delta^{(\mu)}_{B}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\sum_{v^{n}\in\mathcal{T}_{\delta}^{(n)}(V)}\beta_{v^{n}}\zeta_{v^{n}}^{(\mu_{2})} and εB=vn𝒯δ(n)(V)λvnB\varepsilon_{B}=\sum_{v^{n}\notin\mathcal{T}_{\delta}^{(n)}(V)}\lambda^{B}_{v^{n}}.

Further, using the simplification performed in (V.2.4), (V.2.4), and (V.2.4), and the concavity of the square-root function, we obtain,

𝔼[S~23]\displaystyle\mathbb{E}[\tilde{S}_{23}] 2N223nδρBμ2=1N2(𝔼[H0B]+(1εB)(1+η)(22nδρBη+1)𝔼[H~B]+(1εB)(1+η)𝔼[H0B]),\displaystyle\leq\frac{2}{N_{2}}2^{3n\delta_{\rho_{B}}}\sum_{\mu_{2}=1}^{N_{2}}\left(\mathbb{E}[H_{0}^{B}]+{\frac{(1-\varepsilon_{B})}{(1+\eta)}}\sqrt{\left(\frac{2^{2n\delta_{\rho_{B}}}}{\eta}+1\right)\mathbb{E}[\widetilde{H}^{B}]}+\sqrt{\frac{(1-\varepsilon_{B})}{(1+\eta)}}\sqrt{\mathbb{E}[H_{0}^{B}]}\right),
where H~B\displaystyle\text{where }\;\widetilde{H}^{B} \ensurestackMath\stackon[1pt]=Δ1(1εB)vnλvnBρ~vnBpnpk+l2vna2,j>0λvnBρ~vnB(1εB)𝟙{Vn,(μ2)(a2,j)=vn}1.\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\left\|\frac{1}{(1-\varepsilon_{B})}\sum_{v^{n}}\lambda^{B}_{v^{n}}\tilde{\rho}_{v^{n}}^{B}-\frac{p^{n}}{p^{k+l_{2}}}\sum_{v^{n}}\sum_{a_{2},j>0}\frac{\lambda^{B}_{v^{n}}\tilde{\rho}_{v^{n}}^{B}}{(1-\varepsilon_{B})}\mathbbm{1}_{\{V^{n,(\mu_{2})}(a_{2},j)=v^{n}\}}\right\|_{1}. (72)

Using Proposition 2, for any ϵ(0,1)\epsilon\in(0,1), any η,δ(0,1)\eta,\delta\in(0,1) sufficiently small, and any nn sufficiently large, we have 𝔼[S~23]ϵ\mathbb{E}\left[\tilde{S}_{23}\right]\leq\epsilon if k+l2nlogp>I(V;RA)σ2+log(p)S(V)σ3\frac{k+l_{2}}{n}\log p>I(V;RA)_{\sigma_{2}}+\log{p}-S(V)_{\sigma_{3}}, where σ2\sigma_{2}, σ3\sigma_{3} are the auxiliary state defined in the statement of the theorem.

Analysis of S~3\widetilde{S}_{3}: Due to the symmetry in S~2\widetilde{S}_{2} and S~3\widetilde{S}_{3}, the analysis of S~3\widetilde{S}_{3} follows very similar arguments as that of S~2\widetilde{S}_{2} and hence we obtain the following, for any ϵ(0,1)\epsilon\in(0,1), any η,δ(0,1)\eta,\delta\in(0,1) sufficiently small, and any nn sufficiently large, we have 𝔼[S~3]ϵ\mathbb{E}\left[\tilde{S}_{3}\right]\leq\epsilon if S1>I(U;RB)σ1+log(p)S(U)σ3S_{1}>I(U;RB)_{\sigma_{1}}+\log{p}-S(U)_{\sigma_{3}}, where σ1\sigma_{1}, σ3\sigma_{3} are the auxiliary state defined in the statement of the theorem.

Analysis of S~4\widetilde{S}_{4}: We have

S~4\displaystyle\widetilde{S}_{4}\! 1N1N2μ1,μ2znPZ|U+Vn(zn|w0n)\displaystyle\leq\frac{1}{N_{1}N_{2}}\sum_{\mu_{1},\mu_{2}}\sum_{z^{n}}P^{n}_{Z|U+V}(z^{n}|w^{n}_{0})
ρABn(Γ0A,(μ1)Γ0B,(μ2))ρABn1\displaystyle\hskip 72.26999pt\left\|\sqrt{\rho_{AB}^{\otimes n}}\left(\Gamma^{A,(\mu_{1})}_{0}\otimes\Gamma^{B,(\mu_{2})}_{0}\right)\sqrt{\rho_{AB}^{\otimes n}}\right\|_{1}
1N1N2μ1,μ2ρABn(Γ0A,(μ1)I)ρABn1\displaystyle\leq\frac{1}{N_{1}N_{2}}\sum_{\mu_{1},\mu_{2}}\left\|\sqrt{\rho_{AB}^{\otimes n}}\left(\Gamma^{A,(\mu_{1})}_{0}\otimes I\right)\sqrt{\rho_{AB}^{\otimes n}}\right\|_{1}
+1N1N2μ1,μ2vnρABn(Γ0A,(μ1)Bvn(μ2))ρABn1,\displaystyle\hskip 5.0pt+\!\frac{1}{N_{1}N_{2}}\sum_{\mu_{1},\mu_{2}}\sum_{v^{n}}\left\|\sqrt{\rho_{AB}^{\otimes n}}\left(\Gamma^{A,(\mu_{1})}_{0}\otimes B_{v^{n}}^{(\mu_{2})}\right)\sqrt{\rho_{AB}^{\otimes n}}\right\|_{1}\!\!, (73)

where the inequalities above are obtained by a straight forward substitution and use of triangle inequality. Further, since 0Γ0A,(μ1)I0\leq\Gamma^{A,(\mu_{1})}_{0}\leq I and 0Γ0B,(μ2)I0\leq\Gamma^{B,(\mu_{2})}_{0}\leq I, this simplifies the first term in (73) as

1N1N2μ1,μ2\displaystyle\frac{1}{N_{1}N_{2}}\sum_{\mu_{1},\mu_{2}} ρABn(Γ0A,(μ1)I)ρABn1\displaystyle\left\|\sqrt{\rho_{AB}^{\otimes n}}\left(\Gamma^{A,(\mu_{1})}_{0}\otimes I\right)\sqrt{\rho_{AB}^{\otimes n}}\right\|_{1}
=1N1μ1ρAn(Γ0A,(μ1))ρAn1.\displaystyle=\frac{1}{N_{1}}\sum_{\mu_{1}}\left\|\sqrt{\rho_{A}^{\otimes n}}\left(\Gamma^{A,(\mu_{1})}_{0}\right)\sqrt{\rho_{A}^{\otimes n}}\right\|_{1}.

Similarly, the second term in (73) simplifies using Lemma 1 as

1N1N2μ1,μ2vn\displaystyle\frac{1}{N_{1}N_{2}}\sum_{\mu_{1},\mu_{2}}\sum_{v^{n}} ρABn(Γ0A,(μ1)Bvn(μ2))ρABn1\displaystyle\left\|\sqrt{\rho_{AB}^{\otimes n}}\left(\Gamma^{A,(\mu_{1})}_{0}\otimes B_{v^{n}}^{(\mu_{2})}\right)\sqrt{\rho_{AB}^{\otimes n}}\right\|_{1}
1N1μ1ρAn(Γ0A,(μ1))ρAn1.\displaystyle\leq\frac{1}{N_{1}}\sum_{\mu_{1}}\left\|\sqrt{\rho_{A}^{\otimes n}}\left(\Gamma^{A,(\mu_{1})}_{0}\right)\sqrt{\rho_{A}^{\otimes n}}\right\|_{1}.

Using these simplifications, we have

S~4\displaystyle\widetilde{S}_{4} 2N1μ1ρAn(Γ0A,(μ1))ρAn1.\displaystyle\leq\frac{2}{N_{1}}\sum_{\mu_{1}}\left\|\sqrt{\rho_{A}^{\otimes n}}\left(\Gamma^{A,(\mu_{1})}_{0}\right)\sqrt{\rho_{A}^{\otimes n}}\right\|_{1}.

The above expression is similar to the one obtained in the simplification of S~2\widetilde{S}_{2} and hence we can bound S~4\widetilde{S}_{4} using similar constraints as S~2\widetilde{S}_{2}, for sufficiently large nn.

B.6 Proof of Proposition 6

We start by applying triangle inequality to obtain J1J11+J12J_{1}\leq J_{11}+J_{12}, where

J11\displaystyle J_{11} \ensurestackMath\stackon[1pt]=Δzn,vnunρABn(Λ¯unAΛ¯vnB1N1μ1=1N1αunγun(μ1)λunAΛ¯unAΛ¯vnB)ρABnPZ|Wn(zn|un+vn)1,\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\sum_{z^{n},v^{n}}\left\|\sum_{u^{n}}\sqrt{\rho_{AB}^{\otimes n}}\left(\bar{\Lambda}^{A}_{u^{n}}\otimes\bar{\Lambda}^{B}_{v^{n}}-\frac{1}{N_{1}}\sum_{\mu_{1}=1}^{N_{1}}\frac{\alpha_{u^{n}}\gamma_{u^{n}}^{(\mu_{1})}}{\lambda^{A}_{u^{n}}}\bar{\Lambda}^{A}_{u^{n}}\otimes\bar{\Lambda}^{B}_{v^{n}}\right)\sqrt{\rho_{AB}^{\otimes n}}P^{n}_{Z|W}(z^{n}|u^{n}+v^{n})\right\|_{1},
J12\displaystyle J_{12} \ensurestackMath\stackon[1pt]=Δzn,vn1N1μ1=1N1unρABn(αunγun(μ1)λunAΛ¯unAΛ¯vnBγun(μ1)A¯un(μ1)Λ¯vnB)ρABnPZ|Wn(zn|un+vn)1,\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\sum_{z^{n},v^{n}}\left\|\frac{1}{N_{1}}\sum_{\mu_{1}=1}^{N_{1}}\sum_{u^{n}}\sqrt{\rho_{AB}^{\otimes n}}\left(\frac{\alpha_{u^{n}}\gamma_{u^{n}}^{(\mu_{1})}}{\lambda^{A}_{u^{n}}}\bar{\Lambda}^{A}_{u^{n}}\otimes\bar{\Lambda}^{B}_{v^{n}}-\gamma_{u^{n}}^{(\mu_{1})}\bar{A}_{u^{n}}^{(\mu_{1})}\otimes\bar{\Lambda}^{B}_{v^{n}}\right)\sqrt{\rho_{AB}^{\otimes n}}P^{n}_{Z|W}(z^{n}|u^{n}+v^{n})\right\|_{1},

Now with the intention of employing Lemma 5, we express J11J_{11} as

J11\displaystyle J_{11} =un,vn,znλun,vnABρ^un,vnABϕun,vn,zn\displaystyle=\left\|\sum_{u^{n},v^{n},z^{n}}\lambda^{AB}_{u^{n},v^{n}}\hat{\rho}^{AB}_{u^{n},v^{n}}\otimes\phi_{u^{n},v^{n},z^{n}}\right.
1(1+η)pnpk+l1N1μ1un,vn,zna1,i>0λunA\displaystyle\hskip 25.0pt\left.-\frac{1}{(1+\eta)}\frac{p^{n}}{p^{k+l_{1}}N_{1}}\sum_{\mu_{1}}\sum_{u^{n},v^{n},z^{n}}\sum_{a_{1},i>0}\lambda_{u^{n}}^{A}\right.
×𝟙{Un,(μ1)(a1,i)=un}λun,vnABλunAρ^un,vnABϕun,vn,zn1,\displaystyle\hskip 20.0pt\times\left.\mathbbm{1}_{\{U^{n,(\mu_{1})}(a_{1},i)=u^{n}\}}\frac{\lambda^{AB}_{u^{n},v^{n}}}{\lambda^{A}_{u^{n}}}\hat{\rho}^{AB}_{u^{n},v^{n}}\otimes\phi_{u^{n},v^{n},z^{n}}\right\|_{1}\!\!\!,

where the equality above is obtained by defining ϕun,vn,zn=PZ|Wn(zn|un+vn)|vnvn||znzn|\phi_{u^{n},v^{n},z^{n}}=P^{n}_{Z|W}(z^{n}|u^{n}+v^{n})\outerproduct{v^{n}}{v^{n}}\otimes\outerproduct{z^{n}}{z^{n}} and using the definitions of αun,γun(μ1)\alpha_{u^{n}},\gamma_{u^{n}}^{(\mu_{1})} and ρ^un,vnAB\hat{\rho}^{AB}_{u^{n},v^{n}}, followed by using the triangle inequality for the block diagonal operators. Note that the triangle inequality in this case becomes an equality.

Let us define 𝒯un\mathcal{T}_{u^{n}} as

𝒯un\ensurestackMath\stackon[1pt]=Δvn,znλun,vnABλunAρ^un,vnABϕun,vn,zn.\displaystyle\mathcal{T}_{u^{n}}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\sum_{v^{n},z^{n}}\frac{\lambda^{AB}_{u^{n},v^{n}}}{\lambda^{A}_{u^{n}}}\hat{\rho}^{AB}_{u^{n},v^{n}}\otimes\phi_{u^{n},v^{n},z^{n}}.

Note that in the above definition of 𝒯un\mathcal{T}_{u^{n}} we have 𝒯un0\mathcal{T}_{u^{n}}\geq 0 and Tr(𝒯un)=1\Tr{\mathcal{T}_{u^{n}}}=1 for all un𝔽pnu^{n}\in\mathbb{F}_{p}^{n}. Further, it contains all the elements in product form, and thus can be written as 𝒯un=i=1n𝒯ui.\mathcal{T}_{u^{n}}=\bigotimes_{i=1}^{n}\mathcal{T}_{u_{i}}. This simplifies J11J_{11} as

J11\displaystyle J_{11} =unλunA𝒯un1(1+η)pnpk+l11N1μ1\displaystyle=\bigg{\|}\sum_{u^{n}}\lambda^{A}_{u^{n}}\mathcal{T}_{u^{n}}-\frac{1}{(1+\eta)}\frac{p^{n}}{p^{k+l_{1}}}\frac{1}{N_{1}}\sum_{\mu_{1}}
una1,i>0λunA𝒯un𝟙{Un,(μ1)(a1,i)=un}1.\displaystyle\hskip 35.0pt\sum_{u^{n}}\sum_{a_{1},i>0}\lambda^{A}_{u^{n}}\mathcal{T}_{u^{n}}\mathbbm{1}_{\{U^{n,(\mu_{1})}(a_{1},i)=u^{n}\}}\bigg{\|}_{1}.

Using Lemma 5, we claim the following: for any ϵ(0,1)\epsilon\in(0,1), any η,δ(0,1)\eta,\delta\in(0,1) sufficiently small, and any nn sufficiently large, we have 𝔼[J11]ϵ\mathbb{E}[J_{11}]\leq\epsilon, if k+l1nlog(p)+1nlog(N1)>I(U;RZV)σ3S(U)σ3+log(p)\frac{k+l_{1}}{n}\log{p}+\frac{1}{n}\log{N_{1}}>I(U;RZV)_{\sigma_{3}}-S(U)_{\sigma_{3}}+\log{p}, where σ3\sigma_{3} is the auxiliary state defined in the statement of the theorem.

Now we consider the term corresponding to J12J_{12} and prove that its expectation with respect to the Alice’s codebook is small. Recalling J12J_{12}, we get

J12\displaystyle J_{12} 1N1μ1=1N1un,vnznPZ|Wn(zn|un+vn)ρABn(αunγun(μ1)λunAΛ¯unAΛ¯vnBγun(μ1)A¯un(μ1)Λ¯vnB)ρABn1,\displaystyle\leq\frac{1}{N_{1}}\sum_{\mu_{1}=1}^{N_{1}}\sum_{u^{n},v^{n}}\sum_{z^{n}}P^{n}_{Z|W}(z^{n}|u^{n}+v^{n})\left\|\sqrt{\rho_{AB}^{\otimes n}}\left(\frac{\alpha_{u^{n}}\gamma_{u^{n}}^{(\mu_{1})}}{\lambda^{A}_{u^{n}}}\bar{\Lambda}^{A}_{u^{n}}\otimes\bar{\Lambda}^{B}_{v^{n}}-\gamma_{u^{n}}^{(\mu_{1})}\bar{A}_{u^{n}}^{(\mu_{1})}\otimes\bar{\Lambda}^{B}_{v^{n}}\right)\sqrt{\rho_{AB}^{\otimes n}}\right\|_{1},
=1N1μ1=1N1un,vnαunγun(μ1)ρABn((1λunAΛ¯unAρAn1ρ~unAρAn1)Λ¯vnB)ρABn1,\displaystyle=\frac{1}{N_{1}}\sum_{\mu_{1}=1}^{N_{1}}\sum_{u^{n},v^{n}}\alpha_{u^{n}}\gamma_{u^{n}}^{(\mu_{1})}\left\|\sqrt{\rho_{AB}^{\otimes n}}\left(\left(\frac{1}{\lambda^{A}_{u^{n}}}\bar{\Lambda}^{A}_{u^{n}}-\sqrt{\rho_{A}^{\otimes n}}^{-1}\tilde{\rho}_{u^{n}}^{A}\sqrt{\rho_{A}^{\otimes n}}^{-1}\right)\otimes\bar{\Lambda}^{B}_{v^{n}}\right)\sqrt{\rho_{AB}^{\otimes n}}\right\|_{1},

where the inequality is obtained by using triangle and the next equality follows from the fact that znPZ|Wn(zn|un+vn)=1\sum_{z^{n}}P^{n}_{Z|W}(z^{n}|u^{n}+v^{n})=1 for all un𝒰nu^{n}\in\mathcal{U}^{n} and vn𝒱nv^{n}\in\mathcal{V}^{n} and using the definition of Aun(μ1)A_{u^{n}}^{(\mu_{1})}. By applying expectation of J12J_{12} over the Alice’s codebook, we get

𝔼[J12]\displaystyle\mathbb{E}{\left[J_{12}\right]} 1(1+η)unλunAvnρABn((1λunAΛ¯unA\displaystyle\leq\frac{1}{(1+\eta)}\sum_{\begin{subarray}{c}u^{n}\end{subarray}}\lambda^{A}_{u^{n}}\sum_{v^{n}}\left\|\sqrt{\rho_{AB}^{\otimes n}}\left(\left(\frac{1}{\lambda^{A}_{u^{n}}}\bar{\Lambda}^{A}_{u^{n}}-\right.\right.\right.
ρAn1ρ~unAρAn1)Λ¯vnB)ρABn1,\displaystyle\hskip 20.0pt\left.\left.\left.\sqrt{\rho_{A}^{\otimes n}}^{-1}\tilde{\rho}_{u^{n}}^{A}\sqrt{\rho_{A}^{\otimes n}}^{-1}\right)\otimes\bar{\Lambda}^{B}_{v^{n}}\right)\sqrt{\rho_{AB}^{\otimes n}}\right\|_{1},

where we have used the fact that 𝔼[αunγun(μ1)]=λunA(1+η)\mathbb{E}{[\alpha_{u^{n}}\gamma_{u^{n}}^{(\mu_{1})}]}=\frac{\lambda^{A}_{u^{n}}}{(1+\eta)}. To simplify the above equation, we employ Lemma 1 which completely discards the effect of Bob’s measurement. Since vnΛ¯vnB=I\sum_{v^{n}}\bar{\Lambda}^{B}_{v^{n}}=I, from Lemma 1 we have for every unu^{n},

vnρABn((1λunAΛ¯unA\displaystyle\sum_{v^{n}}\left\|\sqrt{\rho_{AB}^{\otimes n}}\left(\left(\frac{1}{\lambda^{A}_{u^{n}}}\bar{\Lambda}^{A}_{u^{n}}-\right.\right.\right.
ρAn1ρ~unAρAn1)Λ¯vnB)ρABn1\displaystyle\hskip 20.0pt\left.\left.\left.\sqrt{\rho_{A}^{\otimes n}}^{-1}\tilde{\rho}_{u^{n}}^{A}\sqrt{\rho_{A}^{\otimes n}}^{-1}\right)\otimes\bar{\Lambda}^{B}_{v^{n}}\right)\sqrt{\rho_{AB}^{\otimes n}}\right\|_{1}
=ρAn(1λunAΛ¯unAρAn1ρ~unAρAn1)ρAn1.\displaystyle\hskip 4.0pt=\left\|\sqrt{\rho_{A}^{\otimes n}}\left(\frac{1}{\lambda^{A}_{u^{n}}}\bar{\Lambda}^{A}_{u^{n}}-\sqrt{\rho_{A}^{\otimes n}}^{-1}\tilde{\rho}_{u^{n}}^{A}\sqrt{\rho_{A}^{\otimes n}}^{-1}\right)\sqrt{\rho_{A}^{\otimes n}}\right\|_{1}.

This simplifies 𝔼[J12]\mathbb{E}{\left[J_{12}\right]} as

𝔼[J12]\displaystyle\mathbb{E}{\left[J_{12}\right]} 1(1+η)unλunAρAn(1λunAΛ¯unA\displaystyle\leq\frac{1}{(1+\eta)}\sum_{\begin{subarray}{c}u^{n}\end{subarray}}\lambda^{A}_{u^{n}}\left\|\sqrt{\rho_{A}^{\otimes n}}\left(\frac{1}{\lambda^{A}_{u^{n}}}\bar{\Lambda}^{A}_{u^{n}}-\right.\right.
ρAn1ρ~unAρAn1)ρAn1\displaystyle\hskip 72.26999pt\left.\left.\sqrt{\rho_{A}^{\otimes n}}^{-1}\tilde{\rho}_{u^{n}}^{A}\sqrt{\rho_{A}^{\otimes n}}^{-1}\right)\sqrt{\rho_{A}^{\otimes n}}\right\|_{1}
1(1+η)un𝒯δ(n)(U)λunAρ^unA1+\displaystyle\leq\frac{1}{(1+\eta)}\!\!\!\sum_{\begin{subarray}{c}u^{n}\notin\mathcal{T}_{\delta}^{(n)}(U)\end{subarray}}\!\!\!\!\lambda^{A}_{u^{n}}\left\|\hat{\rho}^{A}_{u^{n}}\right\|_{1}+
1(1+η)un𝒯δ(n)(U)λunA(ρ^unAρ~unA)1\displaystyle\hskip 36.135pt\frac{1}{(1+\eta)}\!\sum_{\begin{subarray}{c}u^{n}\in\mathcal{T}_{\delta}^{(n)}(U)\end{subarray}}\!\!\lambda^{A}_{u^{n}}\left\|\left(\hat{\rho}^{A}_{u^{n}}-\tilde{\rho}_{u^{n}}^{A}\right)\right\|_{1}
εA+ϵJ12\displaystyle\leq\varepsilon_{A}+\epsilon_{\scriptscriptstyle J_{12}}^{\prime} (74)

where the last inequality is obtained by repeated usage of the Average Gentle Measurement Lemma and ϵJ120\epsilon_{J_{12}}^{\prime}\searrow 0 as δ0\delta\searrow 0 (see (35) in [3] for details). This completes the proof.

B.7 Proof of Proposition 7

Noting the similarity between J2J_{2} and the term S~2\tilde{S}_{2} defined in the proof of Theorem 2 (see Section V.2), we begin by further simplifying J2J_{2} using Lemma 6. This gives us

J2223nδρAN1μ1=1N1(H0A+(1εA)(1+η)H1A+H2A+H3A),\displaystyle J_{2}\leq\frac{2{2^{3n\delta_{\rho_{A}}}}}{N_{1}}\!\!\sum_{\mu_{1}=1}^{N_{1}}\!\!\!\left(\!H_{0}^{A}+\frac{\sqrt{(1-\varepsilon_{A})}}{(1+\eta)}\sqrt{H_{1}^{A}+H_{2}^{A}+H_{3}^{A}}\right)\!\!, (75)

where

H0A\displaystyle H_{0}^{A} \ensurestackMath\stackon[1pt]=Δ|ΔA(μ1)𝔼[ΔA(μ1)]|,\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\left|\Delta^{(\mu_{1})}_{A}-\mathbb{E}[\Delta^{(\mu_{1})}_{A}]\right|,
H1A\displaystyle H_{1}^{A} \ensurestackMath\stackon[1pt]=ΔTr((ΠρAΠAμ1)wnλunAρ~unA),\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\Tr{(\Pi_{\rho_{A}}-\Pi_{A}^{\mu_{1}})\sum_{w^{n}}\lambda^{A}_{u^{n}}\tilde{\rho}_{u^{n}}^{A}},
H2A\displaystyle H_{2}^{A} \ensurestackMath\stackon[1pt]=ΔunλunAρ~unA(1εA)unαunγun(μ1)𝔼[ΔA(μ1)]ρ~unA1,\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\|\sum_{u^{n}}\lambda^{A}_{u^{n}}\tilde{\rho}_{u^{n}}^{A}-(1-\varepsilon_{A})\sum_{u^{n}}\frac{\alpha_{u^{n}}\gamma_{u^{n}}^{(\mu_{1})}}{\mathbb{E}[\Delta^{(\mu_{1})}_{A}]}\tilde{\rho}_{u^{n}}^{A}\|_{1},
H3A\displaystyle H_{3}^{A} \ensurestackMath\stackon[1pt]=Δ(1εA)unαunγun(μ1)ΔA(μ1)ρ~unAunαunγun(μ1)𝔼[ΔA(μ1)]ρ~unA1,\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}(1-\varepsilon_{A})\|\sum_{u^{n}}\frac{\alpha_{u^{n}}\gamma_{u^{n}}^{(\mu_{1})}}{\Delta^{(\mu_{1})}_{A}}\tilde{\rho}_{u^{n}}^{A}-\sum_{u^{n}}\frac{\alpha_{u^{n}}\gamma_{u^{n}}^{(\mu_{1})}}{\mathbb{E}[\Delta^{(\mu_{1})}_{A}]}\tilde{\rho}_{u^{n}}^{A}\|_{1},

and ΔA(μ1)\ensurestackMath\stackon[1pt]=Δun𝒯δ(n)(U)αunγun(μ1),\Delta^{(\mu_{1})}_{A}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\sum_{u^{n}\in\mathcal{T}_{\delta}^{(n)}(U)}\alpha_{u^{n}}\gamma_{u^{n}}^{(\mu_{1})}, εA\ensurestackMath\stackon[1pt]=Δun𝒯δ(n)(U)λunA\varepsilon_{A}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\sum_{u^{n}\notin\mathcal{T}_{\delta}^{(n)}(U)}\lambda^{A}_{u^{n}}, and δρA(δ)0 as δ0\delta_{\rho_{A}}(\delta)\searrow 0\text{ as }\delta\searrow 0. Further, using the simplification performed in (V.2.4), (V.2.4), and (V.2.4), and the concavity of the square-root function, we obtain,

𝔼[J2]\displaystyle\mathbb{E}[J_{2}] 2N123nδρAμ1=1N1(𝔼[H0A]+(1εA)(1+η)\displaystyle\leq\frac{2}{N_{1}}2^{3n\delta_{\rho_{A}}}\sum_{\mu_{1}=1}^{N_{1}}\Bigg{(}\mathbb{E}[H_{0}^{A}]+\frac{(1-\varepsilon_{A})}{(1+\eta)}
×(22nδρAη+1)𝔼[H~A]+(1εA)(1+η)𝔼[H0A]),\displaystyle\times\sqrt{\left(\frac{2^{2n\delta_{\rho_{A}}}}{\eta}+1\right)\mathbb{E}[\widetilde{H}^{A}]}+\sqrt{\frac{(1-\varepsilon_{A})}{(1+\eta)}}\sqrt{\mathbb{E}[H_{0}^{A}]}\Bigg{)},

where

H~A\displaystyle\widetilde{H}^{A} \ensurestackMath\stackon[1pt]=ΔunλunA(1εA)ρ~unA\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\bigg{\|}\sum_{u^{n}}\frac{\lambda^{A}_{u^{n}}}{(1-\varepsilon_{A})}\tilde{\rho}_{u^{n}}^{A}-
pn2nS1una1,i>0λunA(1εA)ρ~unA𝟙{Un,(μ1)(a1,i)=un}1.\displaystyle\hskip 10.0pt\frac{p^{n}}{2^{nS_{1}}}\sum_{u^{n}}\sum_{a_{1},i>0}\frac{\lambda^{A}_{u^{n}}}{(1-\varepsilon_{A})}\tilde{\rho}_{u^{n}}^{A}\mathbbm{1}_{\{U^{n,(\mu_{1})}(a_{1},i)=u^{n}\}}\bigg{\|}_{1}.

The proof from here follows from Proposition 2.

B.8 Proof of Proposition 8

We start by adding and subtracting the following terms within Q2Q_{2}

(i)\displaystyle(i) un,vnρABn(Λ¯unAΛ¯vnB)ρABnPZ|Wn(zn|un+vn),\displaystyle\sum_{u^{n},v^{n}}\sqrt{\rho_{AB}^{\otimes n}}\left(\bar{\Lambda}^{A}_{u^{n}}\otimes\bar{\Lambda}^{B}_{v^{n}}\right)\sqrt{\rho_{AB}^{\otimes n}}P^{n}_{Z|W}(z^{n}|u^{n}+v^{n}),
(ii)\displaystyle(ii) un,vn1N2μ2=1N2ρABn(Λ¯unAβvnζvn(μ2)λvnBΛ¯vnB)\displaystyle\sum_{u^{n},v^{n}}\frac{1}{N_{2}}\sum_{\mu_{2}=1}^{N_{2}}\sqrt{\rho_{AB}^{\otimes n}}\left(\bar{\Lambda}^{A}_{u^{n}}\otimes\cfrac{\beta_{v^{n}}\zeta^{(\mu_{2})}_{v^{n}}}{\lambda^{B}_{v^{n}}}\bar{\Lambda}^{B}_{v^{n}}\right)
×ρABnPZ|Wn(zn|un+vn),\displaystyle\hskip 97.56493pt\times\sqrt{\rho_{AB}^{\otimes n}}P^{n}_{Z|W}(z^{n}|u^{n}+v^{n}),
(iii)\displaystyle(iii) un,vn1N1N2μ1,μ2ρABn(γun(μ1)Aun(μ1)βvnζvn(μ2)λvnBΛ¯vnB)\displaystyle\sum_{u^{n},v^{n}}\frac{1}{N_{1}N_{2}}\sum_{\mu_{1},\mu_{2}}\sqrt{\rho_{AB}^{\otimes n}}\left(\!\!\gamma_{u^{n}}^{(\mu_{1})}A_{u^{n}}^{(\mu_{1})}\otimes\cfrac{\beta_{v^{n}}\zeta^{(\mu_{2})}_{v^{n}}}{\lambda^{B}_{v^{n}}}\bar{\Lambda}^{B}_{v^{n}}\!\!\right)
×ρABnPZ|Wn(zn|un+vn),\displaystyle\hskip 97.56493pt\times\sqrt{\rho_{AB}^{\otimes n}}P^{n}_{Z|W}(z^{n}|u^{n}+v^{n}),
(iv)\displaystyle(iv) un,vn1N1N2μ1,μ2ρABn(γun(μ1)Aun(μ1)ζvn(μ2)B¯vn(μ2))\displaystyle\sum_{u^{n},v^{n}}\frac{1}{N_{1}N_{2}}\sum_{\mu_{1},\mu_{2}}\sqrt{\rho_{AB}^{\otimes n}}\left(\gamma_{u^{n}}^{(\mu_{1})}A_{u^{n}}^{(\mu_{1})}\otimes{\zeta^{(\mu_{2})}_{v^{n}}}\bar{B}_{v^{n}}^{(\mu_{2})}\right)
×ρABnPZ|Wn(zn|un+vn).\displaystyle\hskip 97.56493pt\times\sqrt{\rho_{AB}^{\otimes n}}P^{n}_{Z|W}(z^{n}|u^{n}+v^{n}).

This gives us Q2Q21+Q22+Q23+Q24+Q25Q_{2}\leq Q_{21}+Q_{22}+Q_{23}+Q_{24}+Q_{25}, where

Q21\displaystyle Q_{21} \ensurestackMath\stackon[1pt]=Δznun,vnρABn((1N1μ1=1N1γun(μ1)Aun(μ1))Λ¯vnBΛ¯unAΛ¯vnB)ρABnPZ|Wn(zn|un+vn)1,\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\sum_{z^{n}}\left\|\sum_{u^{n},v^{n}}\sqrt{\rho_{AB}^{\otimes n}}\left(\left(\frac{1}{N_{1}}\sum_{\mu_{1}=1}^{N_{1}}\gamma_{u^{n}}^{(\mu_{1})}A_{u^{n}}^{(\mu_{1})}\right)\otimes\bar{\Lambda}^{B}_{v^{n}}-\bar{\Lambda}^{A}_{u^{n}}\otimes\bar{\Lambda}^{B}_{v^{n}}\right)\sqrt{\rho_{AB}^{\otimes n}}P^{n}_{Z|W}(z^{n}|u^{n}+v^{n})\right\|_{1},
Q22\displaystyle Q_{22} \ensurestackMath\stackon[1pt]=Δznun,vnρABn(Λ¯unAΛ¯vnBΛ¯unA(1N2μ2=1N2βvnζvn(μ2)λvnBΛ¯vnB))ρABnPZ|Wn(zn|un+vn)1,\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\sum_{z^{n}}\left\|\sum_{u^{n},v^{n}}\sqrt{\rho_{AB}^{\otimes n}}\left(\bar{\Lambda}^{A}_{u^{n}}\otimes\bar{\Lambda}^{B}_{v^{n}}-\bar{\Lambda}^{A}_{u^{n}}\otimes\left(\frac{1}{N_{2}}\sum_{\mu_{2}=1}^{N_{2}}\cfrac{\beta_{v^{n}}\zeta^{(\mu_{2})}_{v^{n}}}{\lambda^{B}_{v^{n}}}\bar{\Lambda}^{B}_{v^{n}}\right)\right)\sqrt{\rho_{AB}^{\otimes n}}P^{n}_{Z|W}(z^{n}|u^{n}+v^{n})\right\|_{1},
Q23\displaystyle Q_{23} \ensurestackMath\stackon[1pt]=Δznun,vnρABn((Λ¯unA1N1μ1γun(μ1)Aun(μ1))(1N2μ2βvnζvn(μ2)λvnBΛ¯vnB))ρABnPZ|Wn(zn|un+vn)1,\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\sum_{z^{n}}\left\|\sum_{u^{n},v^{n}}\!\sqrt{\rho_{AB}^{\otimes n}}\left(\left(\bar{\Lambda}^{A}_{u^{n}}-\frac{1}{N_{1}}\sum_{\mu_{1}}\gamma_{u^{n}}^{(\mu_{1})}A_{u^{n}}^{(\mu_{1})}\right)\otimes\left(\frac{1}{N_{2}}\!\sum_{\mu_{2}}\cfrac{\beta_{v^{n}}\zeta^{(\mu_{2})}_{v^{n}}}{\lambda^{B}_{v^{n}}}\bar{\Lambda}^{B}_{v^{n}}\right)\right)\sqrt{\rho_{AB}^{\otimes n}}P^{n}_{Z|W}(z^{n}|u^{n}+v^{n})\right\|_{1},
Q24\displaystyle Q_{24} \ensurestackMath\stackon[1pt]=Δznun,vn1N1N2μ1,μ2ρABn(γun(μ1)Aun(μ1)(βvnζvn(μ2)λvnBΛ¯vnBζvn(μ2)B¯vn(μ2)))ρABnPZ|Wn(zn|un+vn)1,\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\sum_{z^{n}}\left\|\sum_{u^{n},v^{n}}\frac{1}{N_{1}N_{2}}\sum_{\mu_{1},\mu_{2}}\sqrt{\rho_{AB}^{\otimes n}}\left(\gamma_{u^{n}}^{(\mu_{1})}A_{u^{n}}^{(\mu_{1})}\otimes\left(\cfrac{\beta_{v^{n}}\zeta^{(\mu_{2})}_{v^{n}}}{\lambda^{B}_{v^{n}}}\bar{\Lambda}^{B}_{v^{n}}-\zeta_{v^{n}}^{(\mu_{2})}\bar{B}_{v^{n}}^{(\mu_{2})}\right)\right)\sqrt{\rho_{AB}^{\otimes n}}P^{n}_{Z|W}(z^{n}|u^{n}+v^{n})\right\|_{1},
Q25\displaystyle Q_{25} \ensurestackMath\stackon[1pt]=Δznun,vn1N1N2μ1,μ2ρABn(γun(μ1)Aun(μ1)(ζvn(μ2)B¯vn(μ2)ζvn(μ2)Bvn(μ2)))ρABnPZ|Wn(zn|un+vn)1.\displaystyle\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\sum_{z^{n}}\left\|\sum_{u^{n},v^{n}}\frac{1}{N_{1}N_{2}}\sum_{\mu_{1},\mu_{2}}\sqrt{\rho_{AB}^{\otimes n}}\left(\gamma_{u^{n}}^{(\mu_{1})}A_{u^{n}}^{(\mu_{1})}\otimes\left(\zeta_{v^{n}}^{(\mu_{2})}\bar{B}_{v^{n}}^{(\mu_{2})}-\zeta_{v^{n}}^{(\mu_{2})}B_{v^{n}}^{(\mu_{2})}\right)\right)\sqrt{\rho_{AB}^{\otimes n}}P^{n}_{Z|W}(z^{n}|u^{n}+v^{n})\right\|_{1}.

We start by analyzing Q21Q_{21}. Note that Q21Q_{21} is exactly same as Q1Q_{1} and hence using the same rate constraints as Q1Q_{1}, this term can be bounded. Next, consider Q22Q_{22}. Substitution of ζvn(μ2)\zeta^{(\mu_{2})}_{v^{n}} gives

Q22\displaystyle Q_{22} =un,vn,znλun,vnABρ^un,vnABψun,vn,zn\displaystyle=\bigg{\|}\sum_{u^{n},v^{n},z^{n}}\lambda^{AB}_{u^{n},v^{n}}\hat{\rho}^{AB}_{u^{n},v^{n}}\otimes\psi_{u^{n},v^{n},z^{n}}
1N2μ2un,vn,znβvna2,j>0𝟙{Vn,(μ2)(a2,j)=vn}\displaystyle\hskip 20.0pt-\frac{1}{N_{2}}\sum_{\mu_{2}}\!\sum_{u^{n},v^{n},z^{n}}\!\!\!\!\!\beta_{v^{n}}\!\!\!\sum_{a_{2},j>0}\!\!\!\mathbbm{1}_{\{V^{n,(\mu_{2})}(a_{2},j)=v^{n}\}}
×λun,vnABλvnBρ^un,vnABψun,vn,zn1,\displaystyle\hskip 83.11005pt\times\frac{\lambda^{AB}_{u^{n},v^{n}}}{\lambda^{B}_{v^{n}}}\hat{\rho}^{AB}_{u^{n},v^{n}}\otimes\psi_{u^{n},v^{n},z^{n}}\bigg{\|}_{1},

where ψun,vn,zn\psi_{u^{n},v^{n},z^{n}} is defined as ψun,vn,zn=PZ|Wn(zn|un+vn)|znzn|,\psi_{u^{n},v^{n},z^{n}}=P^{n}_{Z|W}(z^{n}|u^{n}+v^{n})\outerproduct{z^{n}}{z^{n}}, and the equality uses the triangle inequality for block operators. Now we use Lemma 5 to bound Q22Q_{22}. Let

𝒯vn\ensurestackMath\stackon[1pt]=Δun,znλun,vnABλvnBρ^un,vnABψun,vn,zn.\displaystyle\mathcal{T}_{v^{n}}\mathrel{\ensurestackMath{\stackon[1pt]{=}{\scriptstyle\Delta}}}\sum_{u^{n},z^{n}}\cfrac{\lambda^{AB}_{u^{n},v^{n}}}{\lambda^{B}_{v^{n}}}\hat{\rho}^{AB}_{u^{n},v^{n}}\otimes\psi_{u^{n},v^{n},z^{n}}.

Note that 𝒯vn\mathcal{T}_{v^{n}} can be written in tensor product form as 𝒯vn=i=1n𝒯vi\mathcal{T}_{v^{n}}=\bigotimes_{i=1}^{n}\mathcal{T}_{v_{i}}. This simplifies Q22Q_{22} as

Q22\displaystyle Q_{22} =vnλvnB𝒯vn1(1+η)pn2nS2N2μ2vn\displaystyle=\bigg{\|}\sum_{v^{n}}\lambda^{B}_{v^{n}}\mathcal{T}_{v^{n}}-\cfrac{1}{(1+\eta)}\cfrac{p^{n}}{2^{nS_{2}}N_{2}}\sum_{\mu_{2}}\sum_{v^{n}}
a2,j>0λvnB𝒯vn𝟙{Vn,(μ2)(a2,j)=vn}1.\displaystyle\hskip 72.26999pt\sum_{a_{2},j>0}\lambda_{v^{n}}^{B}\mathcal{T}_{v^{n}}\mathbbm{1}_{\{V^{n,(\mu_{2})}(a_{2},j)=v^{n}\}}\bigg{\|}_{1}.

Application of Lemma 5 gives the following: for any ϵ(0,1)\epsilon\in(0,1), any η,δ(0,1)\eta,\delta\in(0,1) sufficiently small, and any nn sufficiently large, we have 𝔼[Q22]ϵ\mathbb{E}[Q_{22}]\leq\epsilon if

k+l2nlog(p)+1nlog(N2)>I(V;RZ)σ3S(V)σ3+log(p).\displaystyle\frac{k+l_{2}}{n}\log{p}+\frac{1}{n}\log{N_{2}}>I(V;RZ)_{\sigma_{3}}-S(V)_{\sigma_{3}}+\log{p}.

Now, we move on to consider Q23Q_{23}. Taking expectation with respect G,h1(μ1),h2(μ2)G,h_{1}^{(\mu_{1})},h_{2}^{(\mu_{2})} gives

𝔼[Q23]\displaystyle\mathbb{E}\left[Q_{23}\right] 𝔼[zn,vn1N2μ2=1N2βvnζvn(μ2)λvnBunρABn(Λ¯unAΛ¯vnB)ρABnPZ|Wn(zn|un+vn)\displaystyle\leq\mathbb{E}\left[\sum_{z^{n},v^{n}}\frac{1}{N_{2}}\sum_{\mu_{2}=1}^{N_{2}}\cfrac{\beta_{v^{n}}\zeta^{(\mu_{2})}_{v^{n}}}{\lambda^{B}_{v^{n}}}\left\|\sum_{u^{n}}\sqrt{\rho_{AB}^{\otimes n}}\left(\bar{\Lambda}^{A}_{u^{n}}\otimes\bar{\Lambda}^{B}_{v^{n}}\right)\sqrt{\rho_{AB}^{\otimes n}}P^{n}_{Z|W}(z^{n}|u^{n}+v^{n})\right.\right.
unρABn(1N1μ1γun(μ1)Aun(μ1)Λ¯vnB)ρABnPZ|Wn(zn|un+vn)1]\displaystyle\hskip 130.0pt\left.\left.-\sum_{u^{n}}\sqrt{\rho_{AB}^{\otimes n}}\left(\frac{1}{N_{1}}\sum_{\mu_{1}}\gamma_{u^{n}}^{(\mu_{1})}A_{u^{n}}^{(\mu_{1})}\otimes\bar{\Lambda}^{B}_{v^{n}}\right)\sqrt{\rho_{AB}^{\otimes n}}P^{n}_{Z|W}(z^{n}|u^{n}+v^{n})\right\|_{1}\right]
=𝔼G,h1[zn,vn1N2μ2=1N2𝔼h2|G[βvnζvn(μ2)|G]λvnBunρABn(Λ¯unAΛ¯vnB)ρABnPZ|Wn(zn|un+vn)\displaystyle=\mathbb{E}_{G,h_{1}}\left[\sum_{z^{n},v^{n}}\frac{1}{N_{2}}\sum_{\mu_{2}=1}^{N_{2}}\cfrac{\mathbb{E}_{h_{2}|G}\left[\beta_{v^{n}}\zeta^{(\mu_{2})}_{v^{n}}|G\right]}{\lambda^{B}_{v^{n}}}\left\|\sum_{u^{n}}\sqrt{\rho_{AB}^{\otimes n}}\left(\bar{\Lambda}^{A}_{u^{n}}\otimes\bar{\Lambda}^{B}_{v^{n}}\right)\sqrt{\rho_{AB}^{\otimes n}}P^{n}_{Z|W}(z^{n}|u^{n}+v^{n})\right.\right.
unρABn(1N1μ1γun(μ1)Aun(μ1)Λ¯vnB)ρABnPZ|Wn(zn|un+vn)1]\displaystyle\hskip 130.0pt\left.\left.-\sum_{u^{n}}\sqrt{\rho_{AB}^{\otimes n}}\left(\frac{1}{N_{1}}\sum_{\mu_{1}}\gamma_{u^{n}}^{(\mu_{1})}A_{u^{n}}^{(\mu_{1})}\otimes\bar{\Lambda}^{B}_{v^{n}}\right)\sqrt{\rho_{AB}^{\otimes n}}P^{n}_{Z|W}(z^{n}|u^{n}+v^{n})\right\|_{1}\right]
=𝔼G,h1[zn,vn1(1+η)unρABn(Λ¯unAΛ¯vnB)ρABnPZ|Wn(zn|un+vn)\displaystyle=\mathbb{E}_{G,h_{1}}\left[\sum_{z^{n},v^{n}}\cfrac{1}{(1+\eta)}\left\|\sum_{u^{n}}\sqrt{\rho_{AB}^{\otimes n}}\left(\bar{\Lambda}^{A}_{u^{n}}\otimes\bar{\Lambda}^{B}_{v^{n}}\right)\sqrt{\rho_{AB}^{\otimes n}}P^{n}_{Z|W}(z^{n}|u^{n}+v^{n})\right.\right.
unρABn(1N1μ1γun(μ1)Aun(μ1)Λ¯vnB)ρABnPZ|Wn(zn|un+vn)1]\displaystyle\hskip 130.0pt\left.\left.-\sum_{u^{n}}\sqrt{\rho_{AB}^{\otimes n}}\left(\frac{1}{N_{1}}\sum_{\mu_{1}}\gamma_{u^{n}}^{(\mu_{1})}A_{u^{n}}^{(\mu_{1})}\otimes\bar{\Lambda}^{B}_{v^{n}}\right)\sqrt{\rho_{AB}^{\otimes n}}P^{n}_{Z|W}(z^{n}|u^{n}+v^{n})\right\|_{1}\right]
=𝔼[J(1+η)],\displaystyle=\mathbb{E}\left[\cfrac{J}{(1+\eta)}\right],

where the inequality above is obtained by using the triangle inequality, and the first equality follows from h1(μ1)h_{1}^{(\mu_{1})} and h2(μ2)h_{2}^{(\mu_{2})} being generated independently. The last equality follows from the definition of JJ as in (VI.4). Hence, we use the result obtained in bounding 𝔼[J].\mathbb{E}[J]. Next, we consider Q24Q_{24}.

Q24\displaystyle Q_{24} un,vnznPZ|Wn(zn|un+vn)1N1N2μ1,μ2ρABn(γun(μ1)Aun(μ1)βvnζvn(μ2)λvnBΛ¯vnB)ρABn\displaystyle\leq\sum_{u^{n},v^{n}}\sum_{z^{n}}P^{n}_{Z|W}(z^{n}|u^{n}+v^{n})\left\|\frac{1}{N_{1}N_{2}}\sum_{\mu_{1},\mu_{2}}\sqrt{\rho_{AB}^{\otimes n}}\left(\gamma_{u^{n}}^{(\mu_{1})}A_{u^{n}}^{(\mu_{1})}\otimes\cfrac{\beta_{v^{n}}\zeta^{(\mu_{2})}_{v^{n}}}{\lambda^{B}_{v^{n}}}\bar{\Lambda}^{B}_{v^{n}}\right)\sqrt{\rho_{AB}^{\otimes n}}\right.
1N1N2μ1,μ2ρABn(γun(μ1)Aun(μ1)βvnζvn(μ2)(ρB1ρ~vnBρB1))ρABn1\displaystyle\hskip 70.0pt-\left.\frac{1}{N_{1}N_{2}}\sum_{\mu_{1},\mu_{2}}\sqrt{\rho_{AB}^{\otimes n}}\left(\gamma_{u^{n}}^{(\mu_{1})}A_{u^{n}}^{(\mu_{1})}\otimes\beta_{v^{n}}\zeta^{(\mu_{2})}_{v^{n}}\left(\sqrt{\rho_{B}}^{-1}\tilde{\rho}_{v^{n}}^{B}\sqrt{\rho_{B}}^{-1}\right)\right)\sqrt{\rho_{AB}^{\otimes n}}\right\|_{1}
1N2μ2un,vnβvnζvn(μ2)ρABn(1N1μ1γun(μ1)Aun(μ1)1λvnBΛ¯vnB)ρABn\displaystyle\leq\frac{1}{N_{2}}\sum_{\mu_{2}}\sum_{u^{n},v^{n}}\beta_{v^{n}}\zeta^{(\mu_{2})}_{v^{n}}\left\|\sqrt{\rho_{AB}^{\otimes n}}\left(\frac{1}{N_{1}}\sum_{\mu_{1}}\gamma_{u^{n}}^{(\mu_{1})}A_{u^{n}}^{(\mu_{1})}\otimes\cfrac{1}{\lambda^{B}_{v^{n}}}\bar{\Lambda}^{B}_{v^{n}}\right)\sqrt{\rho_{AB}^{\otimes n}}\right.
ρABn(1N1μ1γun(μ1)Aun(μ1)(ρB1ρ~vnBρB1))ρABn1,\displaystyle\hskip 70.0pt-\left.\sqrt{\rho_{AB}^{\otimes n}}\left(\frac{1}{N_{1}}\sum_{\mu_{1}}\gamma_{u^{n}}^{(\mu_{1})}A_{u^{n}}^{(\mu_{1})}\otimes\left(\sqrt{\rho_{B}}^{-1}\tilde{\rho}_{v^{n}}^{B}\sqrt{\rho_{B}}^{-1}\right)\right)\sqrt{\rho_{AB}^{\otimes n}}\right\|_{1},

where the inequalities follow from the definition of B¯vn(μ2)\bar{B}_{v^{n}}^{(\mu_{2})} and using multiple triangle inequalities. Taking expectation of Q24Q_{24} with respect to h2(μ2)h_{2}^{(\mu_{2})}, we get

𝔼[Q24]\displaystyle\mathbb{E}\left[Q_{24}\right] 𝔼G,h1[un,vnλvnB(1+η)ρABn(1N1μ1γun(μ1)Aun(μ1)(1λvnBΛ¯vnBρB1ρ~vnBρB1))ρABn]\displaystyle\leq\mathbb{E}_{G,h_{1}}\left[\sum_{\begin{subarray}{c}u^{n},v^{n}\end{subarray}}\cfrac{\lambda^{B}_{v^{n}}}{(1+\eta)}\Bigg{\|}\sqrt{\rho_{AB}^{\otimes n}}\left(\frac{1}{N_{1}}\sum_{\mu_{1}}\gamma_{u^{n}}^{(\mu_{1})}A_{u^{n}}^{(\mu_{1})}\otimes\left(\cfrac{1}{\lambda^{B}_{v^{n}}}\bar{\Lambda}^{B}_{v^{n}}-\sqrt{\rho_{B}}^{-1}\tilde{\rho}_{v^{n}}^{B}\sqrt{\rho_{B}}^{-1}\right)\right)\sqrt{\rho_{AB}^{\otimes n}}\right]
𝔼G,h1[vnλvnB(1+η)ρBn(1λvnBΛ¯vnBρB1ρ~vnBρB1)ρBn1]\displaystyle\leq\mathbb{E}_{G,h_{1}}\left[\sum_{v^{n}}\cfrac{\lambda^{B}_{v^{n}}}{(1+\eta)}\left\|\sqrt{\rho_{B}^{\otimes n}}\left(\cfrac{1}{\lambda^{B}_{v^{n}}}\bar{\Lambda}^{B}_{v^{n}}-\sqrt{\rho_{B}}^{-1}\tilde{\rho}_{v^{n}}^{B}\sqrt{\rho_{B}}^{-1}\right)\sqrt{\rho_{B}^{\otimes n}}\right\|_{1}\right]
=vn𝒯δ(n)(V)λvnB(1+η)ρ^vnB1+vn𝒯δ(n)(V)λvnB(1+η)ρ^vnBρ~vnB1εB+ϵQ24,\displaystyle=\sum_{v^{n}\notin\mathcal{T}_{\delta}^{(n)}(V)}\cfrac{\lambda^{B}_{v^{n}}}{(1+\eta)}\left\|\hat{\rho}^{B}_{v^{n}}\right\|_{1}+\sum_{v^{n}\in\mathcal{T}_{\delta}^{(n)}(V)}\cfrac{\lambda^{B}_{v^{n}}}{(1+\eta)}\left\|\hat{\rho}^{B}_{v^{n}}-\tilde{\rho}_{v^{n}}^{B}\right\|_{1}\leq\varepsilon_{B}+\epsilon_{Q_{24}}^{\prime}, (76)

where the second inequality above follows by using Lemma 1 and the fact that 1N1μ1unγun(μ1)Aun(μ1)I,\frac{1}{N_{1}}\sum_{\mu_{1}}\sum_{u^{n}}\gamma_{u^{n}}^{(\mu_{1})}A_{u^{n}}^{(\mu_{1})}\leq I, and the last inequality follows by applying the Average Gentle Measurement Lemma repeated and ϵQ240\epsilon_{Q_{24}}^{\prime}\searrow 0 as δ0\delta\searrow 0 (see (35) in [3] for more details). This completes the proof for the term Q24Q_{24}. Finally, we move onto considering Q25Q_{25}. Simplifying Q25Q_{25} gives

Q25\displaystyle Q_{25} 1N1N2μ1,μ2znun,vnPZ|Wn(zn|un+vn)ρABn\displaystyle\leq\frac{1}{N_{1}N_{2}}\!\!\sum_{\mu_{1},\mu_{2}}\!\sum_{z^{n}}\!\sum_{u^{n},v^{n}}\!\!\!P^{n}_{Z|W}(z^{n}|u^{n}+v^{n})\bigg{\|}\sqrt{\rho_{AB}^{\otimes n}}
(γun(μ1)Aun(μ1)(ζvn(μ2)B¯vn(μ2)ζvn(μ2)Bvn(μ2)))ρABn1\displaystyle\hskip 15.0pt\left(\gamma_{u^{n}}^{(\mu_{1})}A_{u^{n}}^{(\mu_{1})}\otimes\left(\zeta_{v^{n}}^{(\mu_{2})}\bar{B}_{v^{n}}^{(\mu_{2})}-\zeta_{v^{n}}^{(\mu_{2})}B_{v^{n}}^{(\mu_{2})}\right)\right)\sqrt{\rho_{AB}^{\otimes n}}\bigg{\|}_{1}
1N2μ2vnρBn(ζvn(μ2)B¯vn(μ2)ζvn(μ2)Bvn(μ2))ρBn1\displaystyle\leq\frac{1}{N_{2}}\sum_{\mu_{2}}\sum_{v^{n}}\left\|\sqrt{\rho_{B}^{\otimes n}}\!\left(\!\zeta_{v^{n}}^{(\mu_{2})}\bar{B}_{v^{n}}^{(\mu_{2})}\!-\!\zeta_{v^{n}}^{(\mu_{2})}B_{v^{n}}^{(\mu_{2})}\!\right)\!\sqrt{\rho_{B}^{\otimes n}}\right\|_{1}
=S~23,\displaystyle=\tilde{S}_{23},

where the first inequality uses traingle inequality and the second inequality uses Lemma 1 to remove the affect of approximating Alice’s POVM on Bob’s approximation, and S~23\tilde{S}_{23} is defined in (69) in the proof of Proposition 5. Therefore, we have the following: for any ϵ(0,1)\epsilon\in(0,1), any η,δ(0,1)\eta,\delta\in(0,1) sufficiently small, and any nn sufficiently large, we have 𝔼[Q25]ϵ,\mathbb{E}[Q_{25}]\leq\epsilon, if S2I(V;RA)σ2S(V)σ3+log(p)S_{2}\geq I(V;RA)_{\sigma_{2}}-S(V)_{\sigma_{3}}+\log{p}. This completes the proof for Q25Q_{25} and hence for all the terms corresponding to Q2Q_{2}.

References

  • Devetak et al. [2008] I. Devetak, A. W. Harrow, and A. J. Winter, A resource framework for quantum shannon theory, IEEE Transactions on Information Theory 54, 4587 (2008).
  • Winter [2004] A. Winter, ”Extrinsic” and ”intrinsic” data in quantum measurements: asymptotic convex decomposition of positive operator valued measures, Communication in Mathematical Physics 244, 157 (2004).
  • Wilde et al. [2012] M. M. Wilde, P. Hayden, F. Buscemi, and M.-H. Hsieh, The information-theoretic costs of simulating quantum measurements, Journal of Physics A: Mathematical and Theoretical 45, 453001 (2012).
  • Devetak and Winter [2003] I. Devetak and A. Winter, Classical data compression with quantum side information, Physical Review A 68, 042301 (2003).
  • Shannon [1948] C. E. Shannon, A Mathematical Theory of Communication, Bell System Technical Journal 27, 379–423 (July 1948).
  • Ahlswede and Winter [2002] R. Ahlswede and A. Winter, Strong converse for identification via quantum channels, IEEE Transactions on Information Theory 48, 569 (2002).
  • Wilde [2011] M. M. Wilde, From classical to quantum shannon theory, arXiv preprint arXiv:1106.1445  (2011).
  • Groenewold [1971] H. J. Groenewold, A problem of information gain by quantal measurements, International Journal of Theoretical Physics 4, 327 (1971).
  • Lindblad [1972] G. Lindblad, An entropy inequality for quantum measurements, Communications in Mathematical Physics 28, 245 (1972).
  • Ozawa [1986] M. Ozawa, On information gain by quantum measurements of continuous observables, Journal of mathematical physics 27, 759 (1986).
  • Buscemi et al. [2008] F. Buscemi, M. Hayashi, and M. Horodecki, Global information balance in quantum measurements, Physical review letters 100, 210504 (2008).
  • Luo [2010] S. Luo, Information conservation and entropy change in quantum measurements, Physical Review A 82, 052103 (2010).
  • Shirokov [2011] M. E. Shirokov, Entropy reduction of quantum measurements, Journal of mathematical physics 52, 052202 (2011).
  • Berta et al. [2014] M. Berta, J. M. Renes, and M. M. Wilde, Identifying the information gain of a quantum measurement, IEEE Transactions on Information Theory 60, 7987 (2014).
  • Horodecki et al. [2005a] M. Horodecki, J. Oppenheim, and A. Winter, Partial quantum information, Nature 436, 673 (2005a).
  • Horodecki et al. [2007] M. Horodecki, J. Oppenheim, and A. Winter, Quantum state merging and negative information, Communications in Mathematical Physics 269, 107 (2007).
  • Christandl et al. [2009] M. Christandl, R. König, and R. Renner, Postselection technique for quantum channels with applications to quantum cryptography, Physical review letters 102, 020504 (2009).
  • Anshu et al. [2019] A. Anshu, R. Jain, and N. A. Warsi, Convex-split and hypothesis testing approach to one-shot quantum measurement compression and randomness extraction, IEEE Transactions on Information Theory 65, 5905 (2019).
  • Anshu et al. [2017] A. Anshu, V. K. Devabathini, and R. Jain, Quantum communication using coherent rejection sampling, Physical review letters 119, 120506 (2017).
  • Anshu et al. [2014] A. Anshu, V. K. Devabathini, and R. Jain, Quantum message compression with applications, arXiv preprint arXiv:1410.3031  (2014).
  • Renes and Renner [2012] J. M. Renes and R. Renner, One-shot classical data compression with quantum side information and the distillation of common randomness or secret keys, IEEE Transactions on Information Theory 58, 1985 (2012).
  • Tomamichel [2015] M. Tomamichel, Quantum information processing with finite resources: mathematical foundations, Vol. 5 (Springer, 2015).
  • Khatri and Wilde [2020] S. Khatri and M. M. Wilde, Principles of quantum communication theory: A modern approach, arXiv preprint arXiv:2011.04672  (2020).
  • Atif et al. [2019] T. A. Atif, M. Heidari, and S. S. Pradhan, Faithful simulation of distributed quantum measurements with applications in distributed rate-distortion theory, arXiv e-prints , arXiv (2019).
  • Bennett et al. [2002] C. H. Bennett, P. W. Shor, J. A. Smolin, and A. V. Thapliyal, Entanglement-assisted capacity of a quantum channel and the reverse shannon theorem, IEEE Transactions on Information Theory 48, 2637 (2002).
  • Bennett et al. [2009] C. H. Bennett, I. Devetak, A. W. Harrow, P. W. Shor, and A. Winter, Quantum reverse shannon theorem, arXiv preprint arXiv:0912.5537  (2009).
  • Berta et al. [2011] M. Berta, M. Christandl, and R. Renner, The quantum reverse shannon theorem based on one-shot information theory, Communications in Mathematical Physics 306, 579 (2011).
  • Horodecki et al. [2003] M. Horodecki, K. Horodecki, P. Horodecki, R. Horodecki, J. Oppenheim, A. Sen, U. Sen, et al., Local information as a resource in distributed quantum systems, Physical review letters 90, 100402 (2003).
  • Horodecki et al. [2005b] M. Horodecki, P. Horodecki, R. Horodecki, J. Oppenheim, A. Sen, U. Sen, B. Synak-Radtke, et al., Local versus nonlocal information in quantum-information theory: formalism and phenomena, Physical Review A 71, 062307 (2005b).
  • Devetak [2005] I. Devetak, Distillation of local purity from quantum states, Physical Review A 71, 062303 (2005).
  • Krovi and Devetak [2007] H. Krovi and I. Devetak, Local purity distillation with bounded classical communication, Physical Review A 76, 012321 (2007).
  • Korner and Marton [1979] J. Korner and K. Marton, How to encode the modulo-two sum of binary sources (corresp.), IEEE Transactions on Information Theory 25, 219 (1979).
  • Krithivasan and Pradhan [2011] D. Krithivasan and S. S. Pradhan, Distributed source coding using abelian group codes: A new achievable rate-distortion region, IEEE Transactions on Information Theory 57, 1495 (2011).
  • Nazer and Gastpar [2007] B. Nazer and M. Gastpar, Computation over multiple-access channels, IEEE Trans. on Info. Th. 53, 3498 (2007).
  • Philosof and Zamir [2009] T. Philosof and R. Zamir, On the loss of single-letter characterization: The dirty multiple access channel, IEEE Trans. on Info. Th. 55, 2442 (2009).
  • Jafarian and Vishwanath [2012] A. Jafarian and S. Vishwanath, Achievable rates for kk-user Gaussian interference channels, IEEE Transactions on information theory 58, 4367 (2012).
  • Pradhan et al. [2021] S. S. Pradhan, A. Padakandla, and F. Shirani, An Algebraic and Probabilistic Framework for Network Information Theory, Vol. 18 (Foundations and Trends in Communications and Information Theory, 2021) pp. 173–376.
  • Gallager [1968] R. G. Gallager, Information Theory and Reliable Communication (John Wiley & Sons, New York, 1968).
  • Wilde [2013] M. M. Wilde, Quantum information theory (Cambridge University Press, 2013).
  • Ziegler [2012] G. M. Ziegler, Lectures on polytopes, Vol. 152 (Springer Science & Business Media, 2012).
  • Carlen [2010] E. Carlen, Trace inequalities and quantum entropy: an introductory course, Entropy and the quantum 529, 73 (2010).