This site uses cookies. By continuing to use this site you agree to our use of cookies. To find out more, see our Privacy and Cookies policy.
Paper The following article is Open access

The quantum internet: an efficient stabilizer states distribution scheme

Published 28 December 2023 © 2023 The Author(s). Published by IOP Publishing Ltd
, , Citation Seid Koudia 2024 Phys. Scr. 99 015115 DOI 10.1088/1402-4896/ad1565

1402-4896/99/1/015115

Abstract

Quantum networks are a fundamental component of quantum technologies, playing a pivotal role in advancing distributed quantum computing and laying the groundwork for the future quantum internet. They offer a scalable modular architecture for quantum chips and support infrastructure for measurement-based quantum computing. Furthermore, quantum networks serve as the backbone of the quantum internet, ensuring high levels of security. Notably, the advantages of quantum networks in communication are contingent upon entanglement distribution, which faces challenges such as high latency in protocols relying on Bell pair distribution and bipartite entanglement swapping. Additionally, algorithms designed for multipartite entanglement routing encounter intractability issues, rendering them unsolvable within polynomial time. In this paper, we explore a novel approach to distribute graph states in quantum networks, leveraging local quantum coding (LQC) isometries and multipartite states transfer. We also present single-use bounds for stabilizer states distribution. Analogous to network coding, these bounds are attainable when appropriate isometries and stabilizer codes are selected for relay nodes, resulting in reduced latency in entanglement distribution. We further demonstrate the protocol's advantages across various network performance metrics.

Export citation and abstract BibTeX RIS

Original content from this work may be used under the terms of the Creative Commons Attribution 4.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

Quantum networks [1, 2], represent the foundation of numerous emerging quantum technologies and serve as the core of the envisioned quantum internet [35]. These networks harness distinctively quantum phenomena, enabling them to surpass classical approaches and accomplish unprecedented tasks [69]. Notably, quantum networks hold the promise of enhancing information security through various quantum key distribution protocols, alongside facilitating tasks such as quantum secret sharing, quantum leader election, distributed quantum computing, and quantum sensing [10].

The fundamental underpinning of most quantum network applications lies in quantum multipartite entanglement. Specifically, many multi-party protocols target multipartite states to unlock the potential of quantum networks, as mentioned in [11]. A critical class of entangled states with far-reaching implications is known as graph states. These states, also recognized in the realm of quantum error correcting codes as stabilizer states [12], not only enable interconnection between quantum nodes for creating modular distributed quantum computing architectures, but also, facilitate secure quantum computing, whether blind quantum computing in the cloud or measurement-based quantum computing in a distributed fashion.

However, the primary challenge facing quantum networks is the distribution of these valuable multipartite entangled states [8, 13]. Various protocols have addressed the entanglement distribution problem from different perspectives. Many of these protocols involve the initial distribution of Bell pair states, followed by local operations at repeater nodes. This process ultimately establishes connections between clients and a central node for teleporting the desired graph state, as elaborated in [14] and [15]. In today's quantum networks, repeater nodes typically perform a single function: forwarding or swapping bipartite entanglement. However, there is no inherent reason to limit nodes to this functionality, particularly in application-level overlay and multihop networks. Allowing nodes to perform a broader range of functions makes practical sense. The current configuration of quantum repeaters raises concerns about the scalability of these proposed protocols

Furthermore, existing entanglement routing protocols primarily focus on point-to-point entanglement establishment, as discussed in [1619]. However, for concrete applications of quantum networks, distributing multipartite entangled states, rather than just end-to-end Bell pair distribution, is essential. Many efforts have been dedicated to this direction but face significant challenges. Unlike end-to-end entanglement routing, which can be solved using algorithms akin to Dijkstra's algorithm, capable of finding an optimal route in polynomial time concerning the number of network nodes, multipartite entanglement distribution protocols encounter NP-hard problems like the Steiner tree [20]. While polynomial-time approximations exist, exact solutions are elusive.

On the other hand, optimizing entanglement distribution should consider the full capabilities of communication networks. Specifically, constrained quantum networks become relevant when the use of individual point-to-point channels forming the network is restricted or when fixed time slots are allocated for client-required entanglement distribution. These practical assumptions are motivated by various factors, including the cost constraints of quantum hardware such as quantum memories [21]. Furthermore, considering relatively shorter time slots for entanglement distribution can offer several advantages, including increased data rates, reduced latency, and enhanced protocol robustness against adversarial attacks on security schemes that rely on multipartite entanglement.

1.1. Outline and contribution

In the present paper, we consider quantum intermediate relay nodes where the contents of outgoing states are causal functions of the contents of received states, referring to this mapping as Local Quantum Coding. First, we show that Local Quantum Coding –LQC–in intermediate nodes, can be used to distribute the particular class of graph states in a distributed way among remote client nodes. This feature will be proved to be important for the following reasons:

  • It is useful to beat the scalability issues of the previous protocols.
  • Providing a distributed architecture for multipartite entanglement distirbution

Second, we show that the scheme of using LQC, is nothing but a particular network coding strategy [2224]. This allows to achieve the single-use capacity of the appropriate quantum network topologies, which normal entanglement routing fails to fulfill. To this aim:

  • Different bounds for stabilizer states distribution capabilities of the underlying network would be obtained in terms of min-cuts capacities corresponding to different bipartitions of the remote client nodes requiring the graph state.

By considering the quantum network as a tensor network with a particular instance of tensors being specific coding isometries, the advantages of the advantages of our scheme are given in terms of the following figure of merits and applications:

  • We show that by allowing the intermediate nodes in the quantum network to use LQC, an exponential latency reduction in the entanglement distribution time is achieved.
  • We show that the use of LQC reduces drastically the memory qubits overhead of the entanglement distribution cost.
  • We discuss how by allowing intermediate use of quantum error correcting codes, distributed storage can be achieved in quantum networks.

The paper is structured as follows. In section. 2 some preliminaries are be given. In particular, graph states and their entanglement structure will be highlighted. Additionally, the concept of tensor networks along with the notion of min-cut will be presented. In section. 3, our model of the distribution quantum network will be elaborated. An equivalence between our model and a network of stabilizer states contracted by Bell pairs is given in section. 4. In section. 5 our results comprising the different bounds on the consumed resources to single-use distribute a target graph state will be established. Section 6 illustrates the obtained results and bounds for some specific network topologies. In particular, the protocol will be benchmarked to Bell pair based protocols with the time complexity of the graph states distribution and the number of memory qubits taken as figure of merits. We also study the application of choosing the LQC schemes to be valid stabilizer quantum error correcting schemes in order to achieve distributed quantum storage within quantum networks. We finish the paper with conclusions and future work in section 7.

2. Preliminaries

In this section, we provide the necessary preliminaries and tools allowing for the understanding of the derived results.

2.1. Graph states

Graph states admit a simple description in terms of mathematical graphs [12]. A graph is a pair (G, V) of a finite set V = {1....,N} and a set E ⊂ [V]2. The elements of V are called the vertices of the graph and the elements of E are called its edges. Additionally, we denote by Na the set of neighboring vertices of the vertex a.

The graph state ∣G〉 that corresponds to the graph G is the pure state given as

Equation (1)

where

Equation (2)

Similarly, Uab is the controlled Z gate between the vertices a and b.

It is known that any graph state is equivalent to a stabilizer state under a certain class of local operations, the so-called local Clifford operations [12], which map the set of stabilizer states onto itself. Accordingly, graph states can be defined uniquely as the common eigenvector with eigenvalues equal to one in ${{\mathbb{C}}}^{V}$ to the set of independent commuting operators

Equation (3)

for all aV.

2.2. Entanglement structure of graph states

Generally, the many interesting entanglement structure properties of any multipartite state are captured by the reduced state after tracing out parts of the system. Specifically, for a given pure multipartite state ∣Ψ〉. For a specific bipartition of the state, namely {A, B}, the reduced state on A

Equation (4)

where ρ = ∣Ψ〉〈Ψ∣AB . The entanglement rank of the state ∣Ψ〉 in the bipartition {A, B}, quantifies the amount of Bell pairs needed to create such bipartition, and is given as [12]:

Equation (5)

2.3. The min-cut

In this work, we represent the quantum network by an undirected network. For simplicity, an undirected network is described by a tuple (G, V, E, d) where G is the underlying graph, V is the set of its nodes and E is the set of the point to point links in this network. To each link (u, v) ∈ E is assigned a value d(u,v) representing the dimension of the quantum system that can be carried through this link. For two subsets of nodes SV and TV in the network is defined the notion of the cut $(\tilde{S},\tilde{T})$. This is a partition of the set of nodes $V=\tilde{S}\cup \tilde{T}$ such that $S\in \tilde{S}$ and $T\in \tilde{T}$. The value of a cut is given by the product:

Equation (6)

which is the product of the dimensions of independent links between the two subsets $\tilde{S}$ and $\tilde{T}$. In particular, the minimum cut between two sets of nodes S and T designated by MC(S,T) is the cut which has the minimum value [22, 25], namely:

Equation (7)

To better illustrate the notion of the min-cut, we consider the illustrating network in figure 1. It is a network with identical link dimensions for all the edges du,v in the graph. It is clear that the minimum cut separating the two sets of nodes S and T given in light blue and dark blue respectively, is given by the partition $(\tilde{S},\tilde{T})$ which are the respectively given by the set of nodes surrounded by the red and green regions. This min cut has value equals to four. It is much more convenient in the remaining of the paper, to work with the min-cut with respect to log2 of the dimensions, hence the min-cut would be given by

Equation (8)

Working with this definition of the min-cut would allow us to work later with the entanglement ranks instead of the dimensions of the states distributed. The reason we are considering undirected networks is that entanglement networks are symmetric on different bipartitions, in the sense that, a sender and a receiver might be interchanged. Similarly, the notions of the min-cut and the entanglement entropies of states, which will be needed later, are also independent of interchanging bipartitions.

Figure 1.

Figure 1. An illustration of the notion of the min-cut splitting two sets of nodes in an undirected network whose capacities of the links are identical and equal to two. Left: Two sets of nodes in the network designated by S in light blue, and T in dark blue. Right: the cut $(\tilde{S},\tilde{T})$ such as $S\in \tilde{S}$ and $T\in \tilde{T}$ with minimum value equal to four.

Standard image High-resolution image

3. The model

In our model of quantum networks we are embedding all relay nodes with the possibility of establishing an LQC strategy to generate locally graph states, instead of just doing swapping operations and joint bipartite measurements. Formally, each node aV is able to generate a graph state ∣Ga 〉:

Equation (9)

following a coherent control strategy according to the isometry:

Equation (10)

where d is the dimension of the control degree of freedom. Indeed, after measurement of the control degree of freedom in the Fourier basis, the state in equation (9), as well as local unitary equivalent to it, are post-selected.

Each party of the obtained graph state ∣Ga 〉, locally generated at node a, would be distributed to the neighboring relay nodes {Na }. Each part serves as a control degree of freedom, in its turn, of an appropriate LQC process ${\sum }_{i}{U}_{b}^{i}\otimes | i\rangle \langle i| $, in the corresponding neighboring node b. This is illustrated in figure 2. A more specific scheme relying on indefinite causal ordering to generate entanglement has been established and analyzed in details in [7]. We should note that the corrections need to be performed on the state after different contractions are broadcasted from each intermediate node to its neighboring nodes after revealing the result of the contractionspecifically, one of the four possible Bell pairs. As a result, the necessary series of Pauli corrections are delivered to the appropriate clients to correct the global state locally. It is crucial to emphasize that the global state is a local Pauli equivalent to the state before correction. This is of utmost importance, especially in applications such as measurement-based quantum computing or quantum multipartite cryptography primitives, where the clients must accurately know the global state in order to carry out subsequent operations correctly. To illustrate how the scheme works we consider the non-trivial network in figure 3. In this network, four end nodes, representing the clients, request a four qubit linear cluster state to be distributed between them. To achieve this aim, the intermediate nodes, r0, r1 and r2 implements different LQC isometries. For instance, r0 would distribute a Bell pair to the neighboring nodes, r1 and r2. For instance, we the Bell state distributed be the one given by:

Equation (11)

Sequentially, r1 and r2 implement LQC1 and LQC2 given respectively by:

Equation (12)

By considering local qubits in each node, the global state after distribution the Bell pair from r0 and before applying the LQC schemes in r1 and r2 is:

Equation (13)

where the subscript c refer to the control degrees of freedom and r to the local qubits. After performing the LQC isometries, and the distribution of the respective shares to the client nodes in blue, the global state obtained is given by:

Equation (14)

which is indeed a linear graph state between the four client nodes. We note that the signs in the final state could be fixed upon the reception of the classical information describing which Bell state has been distributed by r0, along with the classical information from r1 and r2 telling which outcomes of the measurements has been obtained. Clearly, other equivalent linear cluster states can be obtained by the clients by exchanging classical communications followed by local clifford gates. In the next section, we show that by considering a network whose nodes {a}aV are endowed with coding isometries ${\{{S}_{a}\}}_{a\in V}$, indeed, a graph state can be distributed appropriately among the client nodes, by studying the equivalence to a dual model of the network.

Figure 2.

Figure 2. Scheme for distributing multipartite entanglement states in a quantum network. The network is composed by twelve clients. each served by a relay node. To the right is a magnification of a relay node, which exploits an LQC $S={\sum }_{i=1}^{d}{U}^{(i)}\otimes | i\rangle \langle i| $ to distribute multipartite entanglement to remote nodes.

Standard image High-resolution image
Figure 3.

Figure 3. An illustrating network for the distribution of a four-linear cluster state between the remote client nodes (in blue) where LQC isometries are implemented in the the intermediate nodes (in red).

Standard image High-resolution image

3.1. Other models for multipartite entanglement distribution

The distribution of entanglement in the central-node scheme hinges on establishing the most efficient star-shaped network configuration. In this setup, clients desiring a multipartite entangled state are positioned at the endpoints of this star, while the central node connects to each client through a chain of EPR pairs. Entanglement swapping is employed in nodes in the EPR chains to establish long-haul EPR pairs between the central node and the terminals. Thereafter, through local adjustments of these long-haul EPR pair connections, the central node can effectively transmit its multipartite entangled state to the remote clients [14, 15]. This central node approach is schematized in figure 4. This approach proves to be efficient as long as the connections between the central node and the terminals operate independently. To put it differently, when there are bottleneck nodes, which are nodes with a connectivity degree greater than two, the protocol demands additional resources in these particular nodes. This leads to sub-optimal performance, especially in scenarios featuring highly connected nodes, resulting in inefficiency concerning the available communication capacity and time.

Figure 4.

Figure 4. A scheme showing the central-node approach for entanglement distribution. The blue node is the central node distributing the multipartite entangling state. The clients requesting a target entangled state are in the terminals in yellow. The green nodes are repeater nodes performing entanglement swapping to achieve a long-haul EPR pair between the central node and the terminals. To achieve full long-haul connection between the central node and the clients, two entanglement distribution sessions are needed, which are highlighted in blue and red.

Standard image High-resolution image

An alternative method for disseminating multipartite entanglement, particularly in the case of graph states, involves identifying the minimum-cost Steiner tree that connects all clients located at its terminals using point-to-point links [20]. This approach works perfectly for efficiently distributing GHZ states using local Clifford gates as a tool of transforming distributively the state. This is illustrated in figure 5. However, when dealing with general target graph states, an enhanced protocol is necessary. This entails the discovery and utilization of various Steiner trees that span different subsets of the terminals. These trees are employed to distribute a decorated edge graph state, which can then be transformed into any graph state having the same number of clients. For instance, when considering the distribution of a four qubit linear cluster state between the client nodes of the network in figure 3, this protocol does not use the quantum network infrastructure ones, but many usages are required.

Figure 5.

Figure 5. A scheme showing the Steiner tree approach for entanglement distribution. The green nodes perform local CZ between all their shares of the EPR pairs creating a fully connected graph state equivalent to a GHZ state. After that, each one of them performs a local complementation map followed by a measurement of all the qubits. At the end, a global GHZ state is established between the remote clients in the terminals.

Standard image High-resolution image

4. Distributed graph states generation

In this section, we show formally, how the endowment of quantum networks with chosen LQC in its relay nodes, allows for the distributed generation of graph states.

Lemma 1.  Consider a quantum network in which the relay nodes are equipped with controlled-isometries. These isometries, denoted as ${\{{S}_{x}\}}_{x\in v}$ , are designed to have the capability of generating local graph states. This generation process relies on a control degree of freedom received from the immediately preceding nodes $x\in {N}_{y}$ and represented by

Equation (15)

with r being the number of –incoming–control qubits. This model of the quantum network is equivalent to a network whose vertices $\{x\}$ are endowed with a graph state $\{| {G}_{x}\rangle \}$ . Each state of these is contracted with the neighboring state according to the edges $\{e\}$ which represent EPR pairs, as

Equation (16)

where $| {\bar{G}}_{y}\rangle $ is the augmented graph state given by

Equation (17)

for all ${a}_{y}\in {V}_{y}$.

Proof. In what follows we give a sketch of the proof. In general any two graph states $| {G}_{x}\rangle $ and $| {G}_{y}\rangle $ have the following Schmidt decomposition on single qubit bipartitions [12]:

Equation (18)

with $\langle {G}_{i}/{a}_{i}| {\sigma }_{Z}^{{N}_{{a}_{i}}}| {G}_{i}/{a}_{i}\rangle =0$ if the underlying graph Gi is connected. Similarly, an augmentation graph state $| {\bar{G}}_{y}\rangle $ of the state $| {G}_{y}\rangle $ on node ay has the the decomposition

Equation (19)

On the one hand, by contracting the two states with a Bell pair, the resulting state would be given by

Equation (20)

On the other hand, we let one qubit from $| {G}_{x}\rangle $ to be sent to node y to control a LQC strategy Sy given as

Equation (21)

where

Equation (22)

generating the graph state $| {G}_{y}\rangle $ after measuring the distributed qubit from x in the coherent basis. Indeed, the overall state obtained before measuring the control is given by

Equation (23)

The procedure extends easily to r incoming qubits, in the same spirit, leading to the equivalence of the two approaches. □

To get a feel of this lemma lets suppose that the relay nodes x, y choose appropriate LQC strategies enabling the generation of GHZ states ∣GHZx and ∣GHZy respectively. By sending a single qubit from ∣GHZx in order to control the LQC in y, the overall state is given up to a local unitary by

Equation (24)

it is easy to notice that this state is equivalent to the state 〈EPRxy GHZx GHZy .

Theorem 1.  The state distributed between the clients in the quantum network where the relay nodes use appropriate LQC strategies of single qubit unitaries is given by

Equation (25)

This overall state is a graph state.

Proof. The state in (25) results directly from a direct application of lemma 1 to many contractions between the relay nodes leading to the overall state $| {\rm{\Psi }}\rangle $ between the clients. Therefore, we only give the proof for the statement about the fact that the over state is a graph state. First we should note that any graph state is Local Clifford equivalent to a stabilizer state. Hence, in what comes we will limit ourselves to stabilizer states. We should point out that any stabilizer state is characterized by the set of generators of its stabilizer group. These are elements of the local Pauli group that mutually commute, forming an abelian group. Let $| {\rm{\Phi }}{\rangle }_{{AB}}$ be a stabilizer state whose stabilizer group is S. Similarity, Let $| \chi {\rangle }_{A}$ be a stabilizer state with corresponding stabilizer group T. The contraction of the two states $| {\rm{\Psi }}{\rangle }_{B}=\langle \chi | {\rm{\Phi }}{\rangle }_{B}$ can be written in density operator notation as

Equation (26)

By using properties of the trace operator and of pure density matrices, we can write the density operator $| {\rm{\Psi }}\rangle \langle {\rm{\Psi }}{| }_{B}$ in a trace form as:

Equation (27)

To simplify the trace formula, we can use the fact that the density matrices of the stabilizer states $| \chi {\rangle }_{A}$ and $| {\rm{\Phi }}{\rangle }_{{AB}}$ might be written respectively as:

Equation (28)

Therefore, we have

Equation (29)

The trace in the last equality is non-vanishing if and only if

Equation (30)

This is due to the fact that stabilizer group elements are a product of traceless Pauli matrices, which make them traceless in their turn. This means that the sum is carried only over the elements sAB of the form ${t}_{A}\otimes {u}_{B}$. It is worth noting that both tA and uB are both Local Pauli operators with support on A and B respectively. By this, we conclude that:

Equation (31)

The factor ${2}^{| A| }$ is due to the fact that $\mathrm{Tr}({{\mathbb{I}}}^{A})={2}^{| A| }$, where ${{\mathbb{I}}}^{A}$ is the identity on the subspace A. Furthermore, because all the state $| {G}_{x}\rangle $ or $| e\rangle $ are graph states, they should have full rank stabilizer groups, therefore the full cardinality of the stabilizers S and T should be respectively given by:

Equation (32)

By plugging this into equation (31) we obtain:

Equation (33)

Now we can notice that in order for $| {\rm{\Psi }}\rangle \langle {\rm{\Psi }}{| }_{B}$ to be a graph state, the elements $\{{u}_{B}\}$ should form a group and satisfy

Equation (34)

The first condition is to have a full rank stabilizer group, and the second guarantees that the group is indeed abelian. The first condition requires that $| K| $ should satisfy:

Equation (35)

In order to complete the proof, we should show that $\{{u}_{B}\}$ form an abelian group of local Pauli operators, making the state $| {\rm{\Psi }}{\rangle }_{B}$ a stabilizer state.

As a matter of fact, we note that since $\{{t}_{A}\otimes {u}_{B}\}$ are elements of an abelian group, they should pairwise commute. In order to prove that the operators $\{{t}_{A}\otimes {u}_{B}\}$ form a group we note that for any two elements ${t}_{A}^{1}\otimes {u}_{B}^{1}$ and ${t}_{A}^{2}\otimes {u}_{B}^{2}$, their product is of the form ${t}_{A}^{1}{t}_{A}^{2}\otimes {u}_{B}^{1}{u}_{B}^{2}$, hence of the form ${t}_{A}\otimes {u}_{B}$. This proves the closure of the set. It is easy to note that any operator in this set is its own inverse, hence the set of operators $\{{t}_{A}\otimes {u}_{B}\}$ is an abelian group. In turn, since $\{{t}_{A}\}$ form an abelian group, similarly, $\{{t}_{A}\otimes {u}_{B}\}$ form abelian subgroup of the stabilizer S, $\{{u}_{B}\}$ should form an abelian group as well. Consequently, the abelian group $\{{u}_{B}\}$ should describe a stabilizer state. By considering $| {\rm{\Phi }}{\rangle }_{{AB}}$ in (25) to be the stabilizer state given by ${\otimes }_{e}^{| E| }| e\rangle $ and $| \chi {\rangle }_{A}$ to be the stabilizer state given by ${\otimes }_{x}^{| V| }| x\rangle $, and by applying the previous discussion, we conclude that the state $| {\rm{\Psi }}\rangle $ shared between the clients is indeed a graph state.□

We highlight that lemma 1 and theorem 1 shows an equivalence between two different types of flows of entanglement in the network. The LQC in its controlled isometry approach has a temporal order of entanglement flow within the network. Whereas LQC in local graph states contracted by EPR pairs, does not have this temporal order in the entanglement flow. This gives a flexibility in the design of LQC in future quantum networks. Importantly, we emphasise that since LQC relies on the generation of graph states locally, a graphical visualization of the entanglement flow from one node to another, can be drawn in terms of local complementation and different kind of local measurements.

We also mention that the proof of Theorem. 1, is device-independent, as it is solely relying on the group theoretical structure of stabilizer states. Thus, the proof holds independently of the technological platform implementing the LQC generating entanglement.

5. Single-use network capacities

In this section, we establish entanglement distribution bounds in quantum networks using network coding methods. we quantify the different resources needed for the establishment of target graph states with known entanglement ranks.

5.1. Entanglement distribution bounds

Theorem 2.

theorem  A graph state with a list of bipartite entanglement ranks $\{{r}_{{AB}}\}$ , corresponding to different possible bipartitions {A, B} of the end nodes, can be distributed in a quantum network by a single use of the latter if and only if:

Equation (36)

for every bipartition $\{A,B\}$ of the target graph state.

Proof. We prove first the sufficient part of the theorem. A bipartition of the state of the client nodes into $\{A,B\}$ would induce a bipartition of the network into a cut $(\tilde{S},\tilde{T})$ with the client nodes $A,B$ satisfying $A\subset \tilde{S}$ and $B\subset \tilde{T}$ respectively. We can regard A and B as source and sink nodes S and T respectively—in light of the discussion in section. 2 –and we can collect them in a single node as an effective network without affecting the communication capacities between the bipartitions. It is known that the entanglement that can be established between the source and sink nodes in a tensor network should satisfy [25]

Equation (37)

For bi-partite pure entangled states the entanglement that can be extracted ${E}_{D}(A,B)$ (entanglement distillation) and the entanglement needed to create the state EF (the entanglement of formation) are equal and they satisfy

Equation (38)

Accordingly, for graph states and by using equation (5), we have that

Equation (39)

We prove the necessary part of the theorem by contradiction. Suppose that a target graph state with entanglement entropy rAB satisfying:

Equation (40)

in a given bipartition $\{A,B\}$ can be distributed in the underlying quantum network. Then we can necessarily find rAB end-to-end independent links between the two sets A and B in the quantum network. In this case the min-cut between the two sets satisfies:

Equation (41)

which is in contradiction with the initial assumtion of ${r}_{{AB}}\gt {{MC}}_{\{A,B\}}$

This theorem underscores the presence of two distinct graph structures. One of these structures pertains to the physical network, ensuring connectivity among nodes, while the other represents the theoretical structure of the desired graph state. The pivotal aspect is the condition imposed by the theorem, emphasizing that these two graphs must meet specific criteria. To be precise, Theorem. 2 stipulates that the minimum cut capacities in various client bipartitions within the physical network must exceed the entanglement entropies in the corresponding bipartitions within the graph state, giving the possibility of the latter to be successfully distributed. We highlight that methods using entanglement entropies of entangled state have been used in a slightly different context in [26]. Although it is obvious, the fact that by distributing EPR pairs between a central node and clients on a star topology, any graph state can be distributed among the clients, can be derived from the previous theorem in network information theoretical settings.

Corollary 1.

corollary  The star network topology can distribute any graph state between the leafs of the network from the central node by a single use of the network.

Proof. On the one hand, we note that the entanglement rank rAB for any bipartition $\{A,B\}$ of a graph state is bounded from above as:

Equation (42)

On the other hand, the min-cut of any bipartition of the clients in the star network topology satisfies:

Equation (43)

Therefore, we have that

Equation (44)

and hence, by theorem 2 the star topology of the network allows for the distribution of any graph state. □

This corollary establishes the optimality of the central node protocol framework in a star network topology where there are no bottlenecks in the relay nodes. In fact, under these conditions, our protocol achieves identical performance to that of the central node protocol.

Similarly, we can use the results of Theorem. 2 to state the following corollary for GHZ states distribution in arbitrary quantum networks.

Corollary 2.

corollary  Any topology of a connected quantum network distributes, in a single use, a GHZ state between the clients

Proof. The proof is straightforward from Theorem 2 by noting that the entanglement ranks rAB on any bipartition $\{A,B\}$ of the GHZ state satisfy

Equation (45)

On the other hand, the min-cut of any bipartition of the clients in any connected network satisfies:

Equation (46)

Therefore, we have that

Equation (47)

and hence, by Theorem 2, any connected topology of the quantum network can distribute a GHZ state between the clients by single use of the network. □

This corollary demonstrates that the results achieved in [20] regarding the Steiner tree protocol are indeed optimal. It is evident that the protocol's optimality is a direct consequence of the minimum Steiner tree, which connects the clients at its endpoints, effectively optimizing both communication overhead and communication time. Certainly, when it comes to distributing GHZ states exclusively, it's evident that our protocol, utilizing LQC, is indistinguishable from this particular protocol, and no advantage can be discerned at this stage. However, as we expand our consideration to encompass other categories of graph states, the distinctions between the two protocols become apparent. In pursuit of this objective, our protocol leverages local isometries to generate graph states with different entanglement entropies corresponding the target min-cut bounds imposed by the communication network structure. Conversely, the preceding protocol relies on a sequential process of GHZ state distribution to establish a highly connected edge graph state among the clients, granting them the capability to extract the desired graph state.

5.2. Achieveability of the ultimate bounds for entanglement distribution

In order to better harness the capabilities of a given quantum network for entanglement distribution, the design of appropriate LQC strategies in relay nodes is needed. This is equivalent to finding the optimal coding strategies in relay nodes in network coding in order to achieve the capacity of a given network. Notably, our network architecture relying on LQC for entanglement distribution, might be regarded as a quantum network coding counterpart of the classical network coding paradigm.

Specifically, suppose we endow all the relay nodes in the quantum network with a Linear local quantum coding strategy (LLQC) given by:

Equation (48)

This strategy is equivalent to a quantum repetition code, generating GHZ states in each node. It is easy to notice that this strategy is identical to the isometry in equation (21) up to a local unitary given by IHn−1. Indeed this strategy can only generate entanglement ranks no higher than rAB = 1, making it not a suitable candidate for achieving the network capacity bound in general. To illustrate this, we consider different sub-network structures appearing in the network of figure. 2. For the chain subnetwork (blue nodes), the isometry in equation (48) performed in the blue dashed node

Equation (49)

leads to the same performance of an entanglement swapping. Formally, the effect of the LQC 49 in the dashed node is equivalent, by Lemma 1, to

Equation (50)

which is exactly entanglement swapping performed at node 0. Differently, the performance of the same isometry, in the dashed nodes of sub-networks (orange nodes) and (end nodes), is not equivalent to entanglement swapping, and leads to better entanglement distribution rates as will be discussed in the next section. This suggests that our multiparty protocol provides an advantage over the Bell pair based approach when the entanglement distribution network has high connectivity.

6. Advantages

In this section we illustrate the advantages of the use of intermediate quantum coding in quantum networks according to different figures of merit.

6.1. Exponential latency and memory overhead reduction

To illustrate these advantages we consider regular networks with different degrees of connectivity. For instance, we will consider a tree network where each node has the same number of children nodes. We will consider also that the end relay nodes are connected to the same number of clients. In such a network the number of relay nodes is given by NO(np−1), where p is the depth of the network –the smallest distance from the central node and the clients–. Accordingly, the complexities of the single use protocol with the isometry 48 in each relay node, and the Bell pair based, are given respectively by

Equation (51)

Clearly, we have an exponential time complexity reduction in the depth of the network for the distribution of GHZ like states. The results are illustrated in the plots of figure 6. We can easily see that the more the deeper the quantum network is the more clear the advantage of using a multipartite distribution strategy is.

Figure 6.

Figure 6. The time complexity of the distribution of a target stabilizer state in different depths of the networks. Left: The time complexity of a network of connectivity n = 3. Right: The time complexity of a network of connectivity n = 4.

Standard image High-resolution image

The similar investigation can be conducted for the advantage over the reduction of memory qubits in the network for GHZ. The figure of merit that we use in this case is the maximal number of memory qubits/per node used in the network to achieve the task of distributing a target graph state between the clients. For the LQC approach, the maximum number of memory qubits needed in each relay node is equal to the number of controlled qubits na c required to generate locally the entangled state with an appropriate LQC strategy. For the case of the previous regular network, this is maximized by the number of connectivity n + 1 of the node according to lemma 1. Thus, the maximum number of memory qubits nm satisfies

Equation (52)

Differently, in the Bell pair-based approach, the highest possible number of memory qubits nm in each node is of the order of the number of clients in the network

Equation (53)

The results are illustrated in figure. 7. We can notice how the entanglement distribution based on multipartite protocols is beneficial for networks with growing depth and connectivity.

Figure 7.

Figure 7. The number of memory qubits required for the distribution of a target graph states within networks of different depths. Left: The number of memory qubits required in a network of depth three. Right: The number of memory qubits required in a network of depth four.

Standard image High-resolution image

6.2. Noise resilience

As a noise model, we assign to every entangled state distributed in the quantum network a probability of failure p. Namely, the noise model for the entanglement distribution is given by

Equation (54)

where e is the error flag which will abort the distribution protocol. Accordingly, the error propagation in the network for the distribution of n-EPR pairs would be given by

Equation (55)

Clearly, the probability that the protocol does not abort is when k = 0 which is given by

Equation (56)

As we can easily notice, the less qubit channels are used to establish the target state between the clients the less noisy is the protocol.

For the case of the tree network of depth p = 2, our protocol outperforms the protocol relying on central node and EPR pairs distribution. The performance of the two protocols is illustrated in figure. 8. It is clear that for different target graph states we have

Equation (57)

showing that our protocol is more robust to noise than other protocols.

Figure 8.

Figure 8. A plot comparing the probability of success of the distribution of different graph states in the framework of LQC with the central-based approach in a a tree network of depth 2 and connectivity degree equals to three, where the terminal nodes are the clients.

Standard image High-resolution image

6.3. Distributed quantum storage in quantum networks

A distributed storage system is an infrastructure that can split data across multiple physical servers, and often across more than one data center. It typically takes the form of a cluster of storage units. The storage system includes features that serves as a mechanism for data synchronization and coordination between cluster storage nodes that are geographically dispersed. Moreover, it has the intelligence to detect and respond to failures and cyber attacks. The objective is to achieve very low latency by storing data physically near the location it will be used.

Fundamentally, the transition from classical distributed storage to quantum distributed storage of quantum data, cannot be smooth due to the no-cloning theorem. Although, our distributed stabilizer states generation discussed in section 4, using intermediate quantum encoding, can be harnessed to enhance quantum storage of quantum data significantly.

Consider the network in figure. 9, where the bulk nodes (orange) denoted by {bi } want to store their single qubit data $\{{q}_{{b}_{i}}\}$ into the neighboring cluster nodes (red) denoted as {ci j}. The best that each node can do to store its qubit, independently of the other bulk nodes, is to encode it in a three qubit code given by a GHZ-like state

Equation (58)

It can easily be noticed that by encoding in such codes, single qubit errors that might occur during the storage cannot be corrected by no means, making the encoding vulnerable to noise and adversarial attacks in the transmission. Differently, if the bulk nodes (orange) collaborates and look for local codes that harness the full capacity of the network as discussed in section 5, they might achieve better performance. to do so, let the bulk nodes choose to encode their single qubits in a five qubit code each. The five qubit code, denoted as [[5, 1, 3]], can be described by a cyclic five qubits graph state, which have maximal entanglement ranks among five qubit graph states. Equivalently, it might be described by the stabilizer generators given by [27, 28]:

Equation (59)

This can encode at most a single logical qubit, and can correct single qubit errors and up to two erasure errors.

Figure 9.

Figure 9. A network where the bulk nodes (orange) store a single qubit quantum data and wants to store it in its neighboring cluster nodes (red).

Standard image High-resolution image
Corollary 3.

corollary  By contracting the three $[[5,1,3]]$ codes with Bell pairs corresponding to edges according to equation ( 25 ), a nine qubit code $[[9,3,3]]$ is established between remote cluster nodes. This nine qubit code has as generators

Equation (60)

It can be simply checked that these nine qubit Pauli operators are commuting. Moreover, it is shown in Appendix 7 that they are independent of each other. Therefore, they provide a stabilizer code storing three logical qubits -encoded locally in the three bulk nodes-, in nine physical qubits-shared between the remote cluster nodes-. Furthermore, it is shown in details in the appendix that the established nine qubit code can correct for single qubit errors and up to two erasure errors. As a result, the remotely established code between remote cluster nodes, and storing the three qubits from the bulk nodes is a [[9, 3, 3]] code. Not surprisingly, the obtained code satisfies, indeed, the quantum singleton bound given by [27, 28]:

Equation (61)

We can generalize the previous example of quantum storage to regular networks with a fixed number of neighbors. As a matter of fact we have the following proposition:

Proposition 1.

proposition  Let a quantum storage network where the boundary B, containing the quantum registers, and a bulk containing m nodes each of which encodes k logical qubits in a stabilizer code $[[l,k,d]]$ of l physical qubits and distance d. Then, by contracting the different codes, a stabilizer code $[[| B| ,{mk},D]]$ , encoding the mk qubits, can be established between the quantum registers $| B| $ , where its distance D satisfies:

Equation (62)

Proof. Suppose we have a network with bulk nodes having k qubits encoded in a stabilizer quantum error correcting code $[[l,k,d]]$, where l is the number of neighboring nodes. The code-space of such code is given by

Equation (63)

where $\{{u}_{{b}_{i}}\}$ are the elements of the corresponding stabilizer group of node i. Suppose also that we have m nodes in the bulk of the network and we have the storing registers sitting at the boundary B of the network. According to theorem 1 the state on the boundary would be given by the projector

Equation (64)

The last two equalities follow from the derivation of Theorem. 1. with km is the number of logical qubits stored into the $| B| $ quantum registers. Additionally, according to the quantum singleton bound, the best distance D, i.e., minimum Pauli error weight that cannot be corrected, that this stabilizer code can have should satisfy:

Equation (65)

therefore:

Equation (66)

Differently, the single codes sitting in the bulk satisfy

Equation (67)

by adding the two inequalities (65) and (67) we get:

Equation (68)

and hence, D should satisfy:

Equation (69)

Proposition 1 can be regarded as a straightforward result of known entropic inequalities. To better understand this, let each of the m nodes containing K EPR pairs where half of these pairs are encoded into a l-qubits quantum code with distance d and the other half is kept as a purifying system Ri . We can split the encoded l qubits part into three partitions Ai Fi Ci where $\dim ({A}_{i})=\dim ({F}_{i})={2}^{d-1}$ and $\dim ({C}_{i})={2}^{l-2(d-1)}$. The global system is thus given by ${RAFC}={R}_{i}^{{\otimes }_{i=1}^{m}}{A}_{i}^{{\otimes }_{i=1}^{m}}{F}_{i}^{{\otimes }_{i=1}^{m}}{C}_{i}^{{\otimes }_{i=1}^{m}}$. In the one hand, it is known that as long as the quantum code has distance d, any partition of d − 1 qubits of the code-space is independent of the encoded state [27, 28]. Formally, we should have

Equation (70)

Accordingly, we can have

Equation (71)

Applying the subbaditivity property of the entropy to the last two inequalities we get

Equation (72)

Therefore

Equation (73)

On the other hand, we note that the contraction of the codes maps the entanglement in the bulk into the boundary B as RAFCRB with

Equation (74)

Provided that ∣B∣ > km and that the state on the boundary is indeed a quantum code according to Theorem. 1, and in the same spirit as before we have:

Equation (75)

where $\dim ({C}_{B})=| B| -2(D-1)$.

We can easily verify that the example of the Network in figure 9, with ∣B∣ = 9, l = 5, m = 3, d = 3 and k = 1 satisfy proposition 1 in the sense that we have the singleton bound D = 3 ≤ 4, which is the smallest distance code for ∣B∣ = 9 physical quantum registers and encoding k = 3 logical qubits.

Clearly, the collective efforts of the bulk nodes -by harnessing the full capacity of the network- to establish a common code for distributed storage is advantageous than encoding separately in a centralized way. This gives more resilience to errors that might occur during the storage time or during the transmission process, in the meanwhile, it provides more resilience to losses, or to dishonest participation, when the stored information is to be retrieved. Therefore, this shows the importance of choosing appropriate quantum network encoding for providing secure quantum cryptographical schemes for distributively storing quantum data. Indeed, this would not be possible if the network relies solely on bipartite entanglement distribution and bipartite entanglement swapping.

7. Conclusions and future work

In conclusion, we have introduced a novel approach for entanglement distribution within quantum networks, centered on the concept of local quantum coding (LQC) strategies implemented within relay nodes. This approach represents a multiparty strategy that exhibits notable advantages, particularly in terms of single-use capacities in quantum networks, closely aligning with classical network coding paradigms. Our performance analysis, focused on specific ordered network topologies, has revealed compelling benefits. For selected networks, we have demonstrated a substantial exponential reduction in time complexity concerning the depth parameter compared to the conventional approach relying on EPR pair distribution with a central node. Additionally, we have identified a polynomial advantage in terms of the connectivity parameter, reducing the maximum number of required memory qubits throughout the network. These performance metrics underscore the practical advantages of LQC in communication networks with high depth and connectivity.

Furthermore, our approach exhibits resilience to noise in certain models, where the success probability of distributing target graph states using LQC significantly surpasses that of the central-based approach. We have also illustrated how our proposed protocol, when employing quantum error-correcting codes as part of LQC, can be valuable for distributed quantum storage applications. The core objective of the LQC approach is to enhance the capabilities of repeater/relay nodes within quantum networks to generate arbitrary graph states while preserving their connectivity. This innovation carries multiple advantages, as elucidated throughout this paper. Most notably, it enables quantum communication resource optimization, encompassing channel usage, memory qubits, and EPR pairs, while also streamlining entanglement distribution times. Significantly, akin to classical network coding, the LQC approach confers fault tolerance to the communication network up to a specified threshold for qubit loss.

However, it is important to note that while we have explored the advantages of LQC assuming a given distribution network, real-world scenarios require careful entanglement routing sub-network determination algorithms, which will be the focus of future work. Moreover, it is clear that having a reliable and deterministic source of graph states is crucial for the scalability and efficiency of LQC-based quantum networks. Quantum dots, with their unique properties, offer a promising solution in this endeavor. Quantum dots can be precisely controlled and manipulated to generate entangled cluster states of different sizes. The deterministic nature of this process ensures a consistent and on-demand source of entanglement, which is essential for the operation of LQC-enabled relay nodes in quantum networks. By harnessing quantum dots as a means to produce cluster states, the way can be paved for the practical realization of LQC-based quantum communication systems, facilitating the efficient distribution of entanglement across network nodes and advancing the capabilities of quantum information processing technologies.

Data availability statement

No new data were created or analysed in this study.

: Appendix: proof of Corollary. 3

Equation (A1)

The proof of Corollary. 3 requires explicit computation of the contractions of the three five qubit codes [[5, 1, 3]] with EPR pairs. Without any loss of generality, we let the EPR pairs used for contraction to be given by

Equation (A2)

This is indeed a graph state with stabilizer group given explicitly by the elements:

Equation (A3)

Since 3 EPR pairs are needed for contraction withing the bulk nodes, the overall state for contraction is ∣e⨂3 with stabilizer group with elements given by the tensor product of the individual elements of the stabilizer group in (A3). The cardinality of this group is indeed given by 43 = 64 elements. We don't count all the possibilities of combinations of the elements, but we give an example that illustrates the procedure. If we allow the nodes to be contracted by the stabilizer elements:

Equation (A4)

on each of the edges of the bulk, then the global stabilizer element of the contraction is given by

Equation (A5)

where the under-braces refer to operators that should act on different bulk nodes. Moreover, the elements of stabilizer group of the code [[5, 1, 3]] are given by:

Equation (A6)

The under-braces refer to the parts of the codes to be contracted with the neighboring bulk nodes. There are two neighboring bulk nodes to each bulk node, therefore, two physical qubits have to be contracted. If we contract (A5) with the stabilizer group of the three individual codes (A6) the following nine qubits residual element remains:

Equation (A7)

In fact, we have 64 nine qubit residual elements after full contraction. They indeed form a group as we have shown in the proof of Theorem 1. In order to prove that this residual group encodes three qubits, we have to show that it contains six independent generators. Clearely, proving that the six elements

Equation (A8)

belonging to the residual group, are independent, would be sufficient. To do so, we switch to the symplectic representation of the stabilizer elements [27, 28]. Accordingly, the above elements might be represented by the matrix H given by (A1). It is clear that the last two rows of the matrix H cannot be written as any possible linear combination over GF(2) of the other four rows. In the meanwhile the last two rows are independent from each other. From the last three entries of the Z part of the matrix H we can notice that the first four rows are linearly independent. Therefore, the rows of the matrix H are linearly independent of each other. Hence, the corresponding stabilizer elements in (A8) are indeed generators of the residual stabilizer group.

Now we should prove that the nine qubit code generated by the stabilizer elements in (A8) has distance 3. To do this, we should show that any 9-qubit Pauli operator of weight two should anticommute with at least one of the generators (A8). Let gl be a generator of the residual nine qubit stabilizer group, and let OB be a two qubit Pauli acting on the nine qubits. Form the proof of Theorem. 1 the following is true

Equation (A9)

for some sBA , an element of the stabilizer group of the three five qubit codes and tA an element of the stabilizer group corresponding to the contracting edges. Accordingly, we have

Equation (A10)

Since {sBA } is a stabilizer group correcting for at most two Pauli errors, then there is at least one element sBA that satisfies

Equation (A11)

Hence, there is at least one generator gl of the residual stabilizer group satisfying

Equation (A12)

Therefore, the nine qubit code can correct for any single qubit error. The same argument applies to show that the distance of the code is indeed D = 3.

Please wait… references are loading.
10.1088/1402-4896/ad1565