Fault-tolerant Coding for Quantum Communication

56
Fault-tolerant Coding for Quantum Communication Matthias Christandl 1, * and Alexander Müller-Hermes 2, 3, 1 Department of Mathematical Sciences, University of Copenhagen, Universitetsparken 5, 2100 Copenhagen, Denmark 2 Institut Camille Jordan, Université Claude Bernard Lyon 1, 43 boulevard du 11 novembre 1918, 69622 Villeurbanne cedex, France 3 Department of Mathematics, University of Oslo, P.O. box 1053, Blindern, 0316 Oslo, Norway (Dated: February 23, 2022) Designing encoding and decoding circuits to reliably send messages over many uses of a noisy channel is a central problem in communication theory. When studying the optimal transmission rates achievable with asymptotically vanishing error it is usually assumed that these circuits can be implemented using noise-free gates. While this assumption is satisfied for classical machines in many scenarios, it is not expected to be satisfied in the near term future for quantum machines where decoherence leads to faults in the quantum gates. As a result, fundamental questions regarding the practical relevance of quantum channel coding remain open. By combining techniques from fault-tolerant quantum computation with techniques from quantum communication, we initiate the study of these questions. We introduce fault-tolerant versions of quantum capacities quantifying the optimal communication rates achievable with asymptotically vanishing total error when the encoding and decoding circuits are affected by gate errors with small probability. Our main results are threshold theorems for the classical and quantum capacity: For every quantum channel T and every > 0 there exists a threshold p(, T ) for the gate error probability below which rates larger than C - are fault-tolerantly achievable with vanishing overall communication error, where C denotes the usual capacity. Our results are not only relevant in communication over large distances, but also on-chip, where distant parts of a quantum computer might need to communicate under higher levels of noise than affecting the local gates. Contents I. Introduction 2 A. Communication setting with gate errors 3 B. Nature of the problem 3 C. Our approach and main results 3 II. Preliminaries 6 A. Notation 6 B. Quantum circuits and noise models 6 C. Analyzing noisy quantum circuits 7 D. Concatenated quantum error correcting codes and the threshold theorem 12 E. Capacities of classical-quantum and quantum channels 13 III. Fault-tolerant techniques for quantum communication 15 A. Interfaces for concatenated codes 15 B. Encoding and decoding of concatenated codes and effective communication channels 21 IV. Capacities under arbitrarily varying perturbations 27 A. Arbitrarily varying perturbations 27 B. Separable environment states 28 C. Unrestricted environment states 30 V. Fault-tolerant capacities 34 A. Fault-tolerant capacity of a classical-quantum channel 34 * Electronic address: [email protected] Electronic address: [email protected], [email protected] arXiv:2009.07161v2 [quant-ph] 22 Feb 2022

Transcript of Fault-tolerant Coding for Quantum Communication

Page 1: Fault-tolerant Coding for Quantum Communication

Fault-tolerant Coding for Quantum Communication

Matthias Christandl1, ∗ and Alexander Müller-Hermes2, 3, †

1Department of Mathematical Sciences, University of Copenhagen,Universitetsparken 5, 2100 Copenhagen, Denmark

2Institut Camille Jordan, Université Claude Bernard Lyon 1,43 boulevard du 11 novembre 1918, 69622 Villeurbanne cedex, France

3Department of Mathematics, University of Oslo,P.O. box 1053, Blindern, 0316 Oslo, Norway

(Dated: February 23, 2022)

Designing encoding and decoding circuits to reliably send messages over many uses of a noisychannel is a central problem in communication theory. When studying the optimal transmissionrates achievable with asymptotically vanishing error it is usually assumed that these circuits can beimplemented using noise-free gates. While this assumption is satisfied for classical machines in manyscenarios, it is not expected to be satisfied in the near term future for quantum machines wheredecoherence leads to faults in the quantum gates. As a result, fundamental questions regarding thepractical relevance of quantum channel coding remain open.By combining techniques from fault-tolerant quantum computation with techniques from quantum

communication, we initiate the study of these questions. We introduce fault-tolerant versions ofquantum capacities quantifying the optimal communication rates achievable with asymptoticallyvanishing total error when the encoding and decoding circuits are affected by gate errors withsmall probability. Our main results are threshold theorems for the classical and quantum capacity:For every quantum channel T and every ε > 0 there exists a threshold p(ε, T ) for the gate errorprobability below which rates larger than C−ε are fault-tolerantly achievable with vanishing overallcommunication error, where C denotes the usual capacity. Our results are not only relevant incommunication over large distances, but also on-chip, where distant parts of a quantum computermight need to communicate under higher levels of noise than affecting the local gates.

Contents

I. Introduction 2A. Communication setting with gate errors 3B. Nature of the problem 3C. Our approach and main results 3

II. Preliminaries 6A. Notation 6B. Quantum circuits and noise models 6C. Analyzing noisy quantum circuits 7D. Concatenated quantum error correcting codes and the threshold theorem 12E. Capacities of classical-quantum and quantum channels 13

III. Fault-tolerant techniques for quantum communication 15A. Interfaces for concatenated codes 15B. Encoding and decoding of concatenated codes and effective communication channels 21

IV. Capacities under arbitrarily varying perturbations 27A. Arbitrarily varying perturbations 27B. Separable environment states 28C. Unrestricted environment states 30

V. Fault-tolerant capacities 34A. Fault-tolerant capacity of a classical-quantum channel 34

∗Electronic address: [email protected]†Electronic address: [email protected], [email protected]

arX

iv:2

009.

0716

1v2

[qu

ant-

ph]

22

Feb

2022

Page 2: Fault-tolerant Coding for Quantum Communication

2

B. Fault-tolerant capacities of quantum channels 37C. Specific coding schemes from asymptotically good codes 41

VI. Conclusion and open problems 43

VII. Acknowledgements 44

A. Concatenated quantum error correcting codes 441. 7-qubit Steane code 442. Code concatenation 453. Explicit interface from level 0 to level 1 46

B. Chasing constants in weak typicality 47

C. Explicit error bound in the HSW-theorem 51

D. Explicit error bound in the LSD-theorem 54

References 54

I. INTRODUCTION

Shannon’s theory of communication [48] from the 1940s is the theoretical foundation for the communi-cation infrastructure we use today. While it is extremely successful, it also turned out to be incomplete asa theory of all communication physically possible, since it did not consider the transmission of quantumparticles (e.g., photons). This was first noted by Holevo in the 1970s [27, 28] and since led to a theoryof quantum communication, called quantum Shannon theory, which includes Shannon’s communicationtheory as a special case [42, 53].A basic problem in both classical and quantum Shannon theory is the transmission of messages via

noisy communication channels. To maximize the transmission rate while maintaining a high probabilityof correct transmission the sender may encode the message to make it more resilient against the noiseintroduced by the channel. The receiver then uses a decoder to guess the transmitted message. This isdepicted in Figure 1.

Encoder DecoderPhysicalChannel

MessageEstimatedMessage

FIG. 1: Basic communication setting.

In Shannon theory it is often assumed that messages can be encoded into multiple uses of the samechannel. Surprisingly, for many communication channels it is then possible to transmit messages at apositive rate and with a success probability approaching 1 in the asymptotic limit of infinitely manychannel uses. The communication channels for which this is possible and the optimal communicationrates are precisely characterized by Shannon’s famous capacity formula [48]. For proving this formula it isassumed that the encoder and the decoder can be executed perfectly, i.e., without introducing additionalerrors. This assumption is realistic in many applications of classical Shannon theory since the error ratesof individual logic gates in modern computers are effectively zero (at least in non-hostile environmentsand at the time-scales relevant for communication [41]).Different generalizations of Shannon’s capacity are studied in quantum Shannon theory, since messages

could be classical or quantum, or assisted by entanglement. All those scenarios have in common thatencoding and decoding operations are assumed to be implemented without errors as in Shannon’s originalwork [42, 53]. However, this is not a realistic assumption: At least in the near term future it is not expectedthat the error rates of individual quantum logic gates will effectively vanish [44]. The practical relevanceof capacities in quantum Shannon theory is therefore still unclear, and it is an open problem at whatrates information can be send over a noisy quantum channel in the presence of gate errors.

Page 3: Fault-tolerant Coding for Quantum Communication

3

A. Communication setting with gate errors

Throughout this article, we will consider communication settings where the encoder and decoder(cf. Figure 1) are quantum circuits, i.e., compositions of elementary operations such as unitary operationsfrom a universal gate set, measurements, preparations, and partial traces. Each elementary operationin a circuit is a location where a fault could happen with a certain probability. For simplicity we willconsider the noise model of i.i.d. Pauli noise, which we denote by Fπ(p), commonly used in the literatureon fault tolerance (see [3, 5]). Here, locations are chosen i.i.d. at random with some fixed probabilityp > 0 and at each chosen location a random Pauli error is applied (see Definition II.1 and DefinitionII.2 below). Now, the basic problem is to ensure high communication rates in the presence of gate errors(happening with some fixed probability p > 0 at each location in the coding circuits) while the probabilityof correct transmission of data converges towards 1 in the limit of asymptotically many channel uses.We define the fault-tolerant classical capacity CFπ(p)(T ) of a classical-quantum or quantum channel T ,and the fault-tolerant quantum capacity QFπ(p)(T ) of a quantum channel T as the suprema of achievablecommunication rates in the presence of gate errors modelled by the noise model Fπ(p).

B. Nature of the problem

The importance of analyzing noise in encoding and decoding operations has previously been noted inthe context of long distance quantum communication [12] and it is an important subject in the practicalimplementation of quantum communication [19, 20, 29, 40]. However, in these works the overall successprobability of the considered protocols does not approach 1 at a finite communication rate in the limitof infinitely many channel uses. Therefore, it does not resolve the aforementioned issues.To deal with non-vanishing gate errors in the context of quantum computation, the notion of fault-

tolerance has been developed using quantum error correcting codes. In particular, if the error rates ofindividual quantum logic gates are below a certain threshold, then it is possible to achieve a vanishingoverall error rate for a quantum computation with only a small overhead [3, 5]. While our approach willuse these techniques from fault-tolerance, we should emphasize that they will not directly lead to highcommunication rates over quantum channels in the presence of gate errors, and in most cases they cannoteven be directly applied to the problem. The latter point is exemplified by the qubit depolarizing channelρ 7→ (1− p)ρ+ p(σxρσx + σyρσy + σzρσz)/3 where σx, σy, σz denote the Pauli matrices and p ∈ [0, 1]. Ithas been shown in [22] that this channel has non-zero quantum capacity for p ∼ 0.19, but fault-toleranceunder this noise has only been shown [14, 43] with thresholds p ∼ 0.001. Therefore, it is not possible totreat the communication channel as circuit noise since it might cause faults with probabilities orders ofmagnitude above the threshold required for fault-tolerance.Even if we want to send information over a depolarizing channel with parameter p below the threshold it

is not clear that the codes used in fault-tolerance would achieve non-vanishing communication rates. Theconcatenated codes considered in [3, 5] encode a single logical qubit in a growing number of physical qubitsto achieve increasingly better protection against depolarizaing noise with parameter p below threshold, butthis scheme has a vanishing communication rate. While it might be possible to send quantum informationat non-vanishing communication rates in the presence of gate errors by directly using constant-overheadfault-tolerance schemes [21, 24], this would again only work for communication channels close to identity.Finally, it might be possible to achieve good communication rates in the presence of gate errors by

concatenating specific codes for a communication channel with quantum error correcting codes allowingfor fault-tolerance. However, there is a fundamental problem with this approach and it is unclear how toresolve it: When the inner code is the code for the communication channel (which might be a randomcode of large dimension) it is in general unclear how to implement quantum circuits fault-tolerantly onthe concatenated code. On the other hand, if the code for the communication channel is the outer code,then its codewords are encoded in the inner code and they cannot be directly transmitted through thequantum channel. Hence, it is not clear whether such schemes can lead to high communication ratesin the presence of gate errors, when the success probability of the overall communication scheme shouldapproach 1 in the limit of asymptotically many channel uses.

C. Our approach and main results

To construct encoders and decoders for sending information over quantum channels at high rates in thepresence of gate errors, we will use fault-tolerance, but combine it with techniques from quantum Shan-

Page 4: Fault-tolerant Coding for Quantum Communication

4

non theory. Specifically, we consider a quantum error correcting code enabling fault-tolerant quantumcomputation and use it to implement certain encoding and decoding circuits fault-tolerantly. Note thatthis is similar to the idea of concatenating a channel code (as outer code) with a circuit code (as innercode) outlined above. We will then use interface quantum circuits translating between logical qubits andphysical qubits (i.e., the qubits that the communication channel acts on). The overall setup is depictedin Figure 2. Note that the interfaces are quantum circuits themself and therefore affected by gate errors.

Encodingcircuit

Decodingcircuit

PhysicalChannelMessage

EstimatedMessageQuantum

computation Interface Interface Quantum computation

Error correcting code 1 Error correcting code 2

Effective communication channel

FIG. 2: Our approach: Encoder and decoder are implemented in error correcting codes and interfaces are usedto convert logical to physical states and vice-versa.

To find good lower bounds on the fault-tolerant capacities CFπ(p)(T ) and QFπ(p)(T ) of a quantumchannel T under the i.i.d. Pauli noise model Fπ(p), we take the following steps:

1. We show that there are interface circuits for the concatenated 7-qubit Steane code used in [5], thatonly fail with a probability linear in p (regardless of the concatenation level), when affected by thenoise model Fπ(p).

2. We define an effective communication channel by combining the channel T with the noisy interfaces(see Figure 2). This channel will be close to the original channel T with distance linear in p.

3. We consider encoders and decoders achieving rates close to the capacity of the effective communi-cation channel and by implementing them in the concatenated 7-qubit Steane code we constructthe overall coding scheme.

4. By continuity of the ideal capacities C(T ) and Q(T ) (i.e., the classical and quantum capacity of Twithout gate errors) we conclude that the achieved rates are also close to these quantities.

There are some pitfalls in the steps outlined above. Most importantly, a careful analysis of the interfacecircuits under noise will show that the effective communication channel might not be a tensor power of awell-defined quantum channel, i.e., it might not be i.i.d. in general even though both the physical channeland circuit noise model are i.i.d. The reason for this are correlated errors produced within the quantumcircuit at the encoder, e.g., through non-local CNOT gates. While such errors are correctable by thecircuit code, failures in the interface (which happen with probabilities proportional to the gate errorprobability p) might cause non-i.i.d. correlations to emerge in the effective channel. Instead, the effectivechannel takes the form of a fully-quantum arbitrarily varying channel [10] with some additional structure,which we refer to as a quantum channel under arbitrarily varying perturbation (AVP). A quantum channelT under AVP can be seen as a perturbation of the original channel T by a quantum channel interactingwith an environment system. When taking tensor powers of such a channel the environment systemscorresponding to the tensor factors can be in an entangled quantum state leading to correlations betweenthe different applications of the channel. We will analyze classical and quantum capacities of quantumchannels under AVP in two settings: The first setting allows only classical correlations in the environmentstate and the second setting allows for arbitrary (including entangled) environment states. In bothscenarios, we find coding schemes achieving rates close to the standard capacities by applying techniquesfrom the quantum Shannon theory of non-i.i.d. channel models (e.g., codes for arbitrarily varying channels,and postselection techniques). Finally, we show that the capacities under AVP are lower bounds to thefault-tolerant capacities which implies our main results, which are threshold theorems for fault-tolerantcapacities of classical-quantum and quantum channels:

Page 5: Fault-tolerant Coding for Quantum Communication

5

• For every ε > 0 and every dimension d ≥ 2 there exists a threshold p(ε, d) > 0 such that

CFπ(p)(T ) ≥ C(T )− ε

for all 0 ≤ p ≤ p(ε, d) and for all classical-quantum channels T : A →Md.

• For every ε > 0 and every quantum channel T :Md1 →Md2 there exists a p(ε, T ) > 0 such that

CFπ(p)(T ) ≥ C(T )− ε and QFπ(p)(T ) ≥ Q(T )− ε

for all 0 ≤ p ≤ p(ε, T ).

Our results show that communication at good rates and with vanishing communication error is possiblein non-trivial cases and in realistic scenarios where local gates are affected by noise. This is an importantvalidation of the practical relevance of quantum Shannon theory. Moreover, our results are not onlyrelevant for communication over large distances, but also for applications within a quantum computer.Here, communication between distant parts of a quantum computing chip may be affected by a higherlevel of noise than the local gates. Our results could then be used to optimize the design of quantumhardware with on-chip communication.Our article is organized as follows:

• In Section II we will go through preliminaries needed for the rest of the article.

– In Section IIA we introduce the most basic notation.– In Section II B we review the circuit model of quantum computation and we explain how tomodel the noise affecting quantum circuits.

– In Section IIC we slightly simplify the formalism from [5], and we explain how it can be usedto analyze quantum circuits affected by noise.

– In Section IID we review the threshold theorem from [5] for concatenated quantum codes.– In Section II E we define the most basic capacities considered in quantum Shannon theory:The classical capacity of a classical-quantum or quantum channel, and the quantum capacityof a quantum channel.

• In Section III we study quantum communication in the framework of fault-tolerant quantum circuits.

– In Section IIIA we define interfaces between physical qubits and logical qubits encoded instabilizer codes. For the concatenated 7-qubit Steane code we analyze how an interface isaffected by i.i.d. Pauli noise.

– In Section III B we use interfaces for the concatenated 7-qubit Steane code to study quantumprotocols for communication over a quantum channel. In particular, we identify the effectivecommunication channel (cf. Figure 2).

• In Section IV we study transmission of classical or quantum information via quantum channelsunder arbitrarily varying perturbation (AVP). This includes the effective communication problemfrom Section III B as a special case.

– In Section IVA we introduce quantum channels under AVP and we define the correspondingclassical and quantum capacity.

– In Section IVB we study quantum channels under AVP with fully separable environmentstates. This channel model reduces to the arbitrarily varying quantum channels studied in [4].

– In Section IVC we study quantum channels under AVP with unrestricted environment states(including highly entangled states). We apply a postselection technique to prove a lower boundon these quantities.

• In Section V we introduce and study fault-tolerant versions of the classical capacity and the quantumcapacity.

– In Section VA we study the fault-tolerant classical capacity CFπ(p)(T ) under i.i.d. Pauli noiseof a classical-quantum channel.

– In Section VB we study the fault-tolerant classical capacity CFπ(p)(T ) and the fault-tolerantquantum capacity QFπ(p)(T ) under i.i.d. Pauli noise of a quantum channel. We show thatthese capacities are lower bounded by capacities under AVP.

Page 6: Fault-tolerant Coding for Quantum Communication

6

– In Section VC we show how asymptotically good codes can be used to design fault-tolerantcoding schemes for quantum communication for quantum channels arising as convex combina-tions with the identity channel.

• In Section VI we conclude our article with some ideas for further research.

II. PRELIMINARIES

A. Notation

We will denote byMd :=M(Cd)the matrix algebra of complex d× d-matrices and by D(Cd) the set

of d-dimensional quantum states, i.e., the set of positive semidefinite matrices inMd with trace equal to1. As is common in the mathematical literature, we will define quantum channels as completely positiveand trace-preserving maps T : MdA → MdB . In analogy, we define channels with classical output aspositive and entry-sum-preserving maps into Cd, where we identify probability distributions on the set1, . . . , d with entrywise-positive vectors in Cd such that the sum of their entries equals 1 (which isequivalent to diagonal matrices in D(Cd)). By the same reasoning, we may define channels with classicalinput as linear maps on Cd, which are positive with respect to the entrywise-positive order on Cd andthe order on the output, and such that they yield normalized outputs on vectors with entries summingto 1. Sometimes, we will also use the more common definition as maps from a finite alphabet A, andextended linearly to the space of probability distributions on A denoted by P(A).

B. Quantum circuits and noise models

A quantum channel Γ :M⊗n2 →M⊗m2 is called a quantum circuit if it can be written as a compositionof elementary operations [42]. These elementary operations are:

• Elementary qubit gates: Pauli gates σx, σy and σz, Hadamard gate H, and the T -gate.

• Identity gate corresponding to a resting qubit.

• Controlled-not (CNOT) gate.

• Measurements and preparations in the computational basis.

• Qubit trace, i.e., discarding a qubit.

It is well known that the quantum circuits form a dense subset of the set of quantum channels T :M⊗n2 →M⊗m2 (see for instance [7, 11, 42]). Note that there may be many ways to construct the samequantum circuit (viewed as a quantum channel) from elementary operations. Moreover, after physicallyimplementing the quantum circuit, its performance under noise might depend on the specific construction.To simplify our discussion we will assume every quantum circuit Γ to be specified by a particular circuitdiagram GΓ. Formally, GΓ is an acyclic directed graph with vertices colored by elementary operations,and edges corresponding to qubits interacting at elementary gates. We define the set of locations of Γdenoted by Loc(Γ) as the set of vertices of GΓ, and we will denote by dout(l) ∈ 0, 1, 2 the outdegree ofa location l ∈ Loc(Γ). Note that dout(l) = 0 if the location l is a measurement or a trace, dout(l) = 1 ifthe location l is an elementary qubit gate, identity gate, or a preparation, and dout(l) = 2 if the locationl is a CNOT gate.Different models of noise affecting quantum circuits have been studied in the literature. Here, we are

restricting to the simplest case of these models where Pauli noise affects locations of the circuit locally. Tomodel such noise we will select subsets of locations in the circuit diagram where we will insert Pauli noisechannels either before or after the executed elementary operation. Formally, this introduces a secondcoloring of the vertices of the circuit diagram with colors representing different Pauli channels describingthe noise. It should be emphasized that whether noise channels are inserted before or after the elementaryoperations is a convention (except when elementary operations are preparations or measurements). Here,we choose the same convention as in [5], but other choices might be reasonable as well.

Definition II.1 (Pauli fault patterns and faulty circuits). Consider the single qubit Pauli channelsAdσx ,Adσy ,Adσz : M2 → M2 defined as Adσi (X) = σiXσi for i ∈ x, y, z and any X ∈ M2.

Page 7: Fault-tolerant Coding for Quantum Communication

7

A Pauli fault pattern affecting a quantum circuit Γ : M⊗n2 → M⊗m2 is a function F : Loc(Γ) →id, x, y, z ∪ (id, x, y, z × id, x, y, z) such that F (l) ∈ id, x, y, z if dout(l) ∈ 0, 1 and F (l) ∈id, x, y, z × id, x, y, z if dout(l) = 2. We denote by [Γ]F : M⊗n2 → M⊗m2 the quantum channelobtained by inserting the particular noise channels in the execution of the circuit diagram of the circuitΓ. Specifically, we do the following:

• We apply the Pauli channel AdσF (l) directly before the location l ∈ Loc(Γ) if dout(l) = 0.

• We apply the Pauli channel AdσF (l) directly after the location l ∈ Loc(Γ) if dout(l) = 1.

• We apply the Pauli channel AdσF (l)1⊗AdσF (l)2

directly after the location l ∈ Loc(Γ) if dout(l) = 2.

When given a Pauli fault pattern F affecting a composition Γ1 Γ2 of quantum circuits, we will oftenabuse notation and write [Γ1]F and [Γ2]F to denote the restriction of F (as a function) to the locationsof the circuits Γ1 and Γ2, respectively.

In our simplified treatment, a noise model F specifies a probability distribution over fault patterns tooccur in a given quantum circuit Γ. We will denote by [Γ]F the circuit affected by the noise model, i.e.,the quantum channel obtained after selecting a fault pattern at random according to the noise modeland inserting it into the circuit diagram of the circuit. Again it will be convenient to sometimes abusenotation, and use the notation [·]F for combinations of quantum circuits and quantum communicationchannels. For example, we might write [ΓD T⊗m ΓE ]F , where ΓE and ΓD are quantum circuits andT denotes a quantum channel that is not viewed as a quantum circuit. By this notation we mean theoverall quantum channel obtained from applying the noise model to the quantum circuits ΓE and ΓDthat have well-defined locations, but the quantum channel T is not affected by this.In the following, we will restrict to a very basic type of noise.

Definition II.2 (I.i.d. Pauli noise model). Consider a quantum circuit Γ : M⊗n2 → M⊗m2 .The i.i.d. Pauli noise model Fπ(p) selects a Pauli fault pattern F : Loc(Γ) → id, x, y, z ∪(id, x, y, z × id, x, y, z) with the probability

P (F ) = (1− p)lid(p/3)lx(p/3)ly (p/3)lz

where

li :=∣∣∣l ∈ Loc (Γ) : dout(l) ∈ 0, 1 and F (l) = i

∣∣∣+∣∣∣l ∈ Loc (Γ) : dout(l) = 2 and F (l)1 = i

∣∣∣+∣∣∣l ∈ Loc (Γ) : dout(l) = 2 and F (l)2 = i

∣∣∣,for any i ∈ id, x, y, z.

It is straightforward to define other examples of i.i.d. noise models or more general local noise modelssimilar to the previous definitions. The i.i.d. Pauli noise model corresponds to inserting the depolarizingchannel Dp :M2 →M2 given by Dp(X) = (1−p)X+p

∑i∈x,y,z σiXσi on the output systems of every

location in the circuit. Our main results also hold (with slightly modified proofs) for noise models wherearbitrary qubit quantum channels of the form X 7→ (1− p)X + pN(X) are inserted after each location,and where the qubit quantum channel N may be different for each location. Different techniques mightbe required for local noise models that are not i.i.d. , or for noise even more exotic. We will comment onthis further at the appropriate places.

C. Analyzing noisy quantum circuits

To protect a quantum circuit against noise it can be implemented in a quantum error correcting code.Here, we will describe how to analyze noisy quantum circuits following the ideas of [5]. We will firstintroduce idealized encoding and decoding operations that will select a specified basis in which faultsare interpreted. These operations are unitary and can be inserted into a quantum circuit affected by aspecific fault pattern. When the fault pattern is nice enough it will then be possible to transform thefaulty quantum circuit into the ideal quantum circuit by decoupling the data from the noise encoded in aquantum state corresponding to possible syndromes. In our discussion we will restrict to stabilizer codesencoding a single qubit where these constructions can be done in a starightforward way.

Page 8: Fault-tolerant Coding for Quantum Communication

8

Let C ⊂ (C2)⊗K denote the code space of a 2-dimensional stabilizer code, i.e., the common eigenspacefor eigenvalue +1 of a collection g1, . . . , gK−1 of commuting product-Pauli matrices generating a matrixgroup not containing −1. In such a setting we denote byWs ⊂ (C2)⊗K for s ∈ FK−1

2 the space of commoneigenvectors for the eigenvalues (−1)si with respect to each gi. By definition it is clear, that Ws ⊥ Ws′

for s 6= s′ and that dim (Ws) = dim (C) = 2 for any s ∈ FK−12 . Therefore, we have(

C2)⊗K =⊕

s∈FK−12

Ws

by a simple dimension counting argument, and C = W(1,1,...,1).We will denote by |0〉, |1〉 ⊂ C an orthonormal basis of the code space, i.e., the encoded computational

basis. For each s ∈ FK−12 we can select a product-Pauli operator1 Es :

(C2)⊗K → (

C2)⊗K such that

Ws = Es (C) .

Note that in this way W00···0 = C. To follow the usual convention we will call the operator Es the Paulierror associated with the syndrome s ∈ FK−1

2 . By the previous discussion the set⋃s∈FK−1

2

Es|0〉, Es|1〉 (1)

forms a basis of(C2)⊗K and it will be convenient to analyze noisy quantum circuits with respect to this

basis. This approach follows closely the analysis of [5, Section 5.2.2], but makes it slightly more precise.We start by defining a linear map D : (C2)⊗K → C2 ⊗ (C2)⊗K−1 acting as

D(Es|i〉

)= |i〉 ⊗ |s〉,

on the basis from (1) and extend it linearly. Here we used the abbreviation |s〉 := |s1〉⊗|s2〉⊗· · ·⊗|sK−1〉for s ∈ FK−1

2 . Clearly, D is a unitary change of basis when identifying C2 ⊗ (C2)⊗K−1 ' (C2)⊗K . Wecan therefore define the unitary quantum channel Dec∗ :M⊗K2 →M⊗K2 by

Dec∗(X) = DXD†, (2)

and its inverse Enc∗ :M⊗K2 →M⊗K2 by

Enc∗(X) = D†XD, (3)

for any X ∈ M⊗K2 . The quantum channel Dec∗ is also called the ideal decoder, and Enc∗ is called theideal encoder. Note that these quantum channels in the case of concatenated codes appear in [5, p.17],where they are called the “k−∗decoder” and the “k−∗encoder”. Finally we define a quantum channelEC∗ :M⊗K2 →M⊗K2 by

EC∗ = Enc∗ [id2 ⊗ (|0〉〈0|Tr)] Dec∗, (4)

where |0〉 ∈ (C2)⊗(K−1) corresponds to the zero syndrome, and Tr : M⊗(K−1)2 → C. The quantum

channel EC∗ corrects errors on the data and is called the ideal error correction. In the following, we willsometimes use a superscript ∗ in order to identify ideal operations or gates. Such operations or gatesshould be thought of as either mathematical objects appearing in the analysis of noisy quantum circuits(in the case of Enc∗, Dec∗ and EC∗), or as logical operations or logical gates acting on logical states. Inany case, these ideal objects are never subject to noise.To implement quantum circuits in a stabilizer code with code space C ⊂ (C2)K of dimension dim(C) = 2

we will assume that there are quantum circuits, called gadgets, implementing the elementary operationon the code space.

Definition II.3 (Implementation of a quantum circuit). Let C ⊂ (C2)⊗K satisfying dim(C) = 2 be thecode space of a stabilizer code, and let |0〉〈0| ∈ M⊗(K−1)

2 denote the pure state corresponding to the zerosyndrome. We assume that certain elementary quantum circuits called gadgets are given:

1 using that product-Pauli operators either commute or anticommute

Page 9: Fault-tolerant Coding for Quantum Communication

9

1. For each elementary single qubit operation G∗ :M2 →M2, we have a gadget G :M⊗K2 →M⊗K2such that

Dec∗ G Enc∗ (· ⊗ |0〉〈0|) = G∗ (·)⊗ |0〉〈0|.

2. For the CNOT gate G∗CNOT :M4 →M4, we have a gadget GCNOT :M⊗K2 ⊗M⊗K2 →M⊗K2 ⊗M⊗K2such that

(Dec∗)⊗2 GCNOT (Enc∗)⊗2 (· ⊗ (|0〉〈0|)⊗2) = G∗CNOT (·)⊗ (|0〉〈0|)⊗2.

3. For a measurement G∗ : M2 → Diag2 in the computational basis we have a gadget G : M⊗K2 →Diag2 such that

G Enc∗ (· ⊗ |0〉〈0|) = G∗(·).

4. For a preparation in the computational basis G∗ : C →M2 we have a gadget G : C →M⊗K2 suchthat

Dec∗ G = G∗ ⊗ |0〉〈0|.

5. For the trace G∗ :M2 → C we have a gadget G :M⊗K2 → C such that

G Enc∗ (· ⊗ |0〉〈0|) = G∗(·).

Besides the gadgets defined above, we consider a quantum circuit EC : M⊗K2 → M⊗K2 realizing thequantum channel EC∗ : M⊗K2 → M⊗K2 from (4). Then, we define the rectangle of an elementaryoperation G∗ with corresponding gadget G to be the quantum circuit given by

RG =

EC G if G∗ is a single qubit operation, or a preparation.EC⊗2 G if G∗ is a CNOT gate.G if G∗ is a measurement or a trace.

Given a quantum circuit Γ : M⊗n2 → M⊗m2 we define its implementation ΓC : M⊗nK2 → M⊗mK2 asthe quantum circuit obtained by replacing each qubit in the circuit Γ by a block of K qubits, and eachelementary operation in the circuit diagram of Γ by the corresponding rectangle.

Implementations of quantum circuits can be analyzed using the ideal encoder Enc∗ and the ideal decoderDec∗ introduced previously. To illustrate this, consider a quantum error correcting code C ⊂ (C2)Ksatisfying dim(C) = 2 and an elementary single qubit operation G∗ : M2 → M2 with correspondingrectangle RG. Using that Dec∗ and Enc∗ are inverse to each other, we can compute

Dec∗ RG Enc∗ (· ⊗ |0〉〈0|) = Dec∗ EC G Enc∗ (· ⊗ |0〉〈0|)= Dec∗ EC Enc∗ Dec∗ G Enc∗ (· ⊗ |0〉〈0|)= G∗(·)⊗ |0〉〈0|,

where we used the assumptions from Definition II.3 and the fact that as quantum channels and withoutnoise we have EC = EC∗ with the ideal error correction from (4). In a similar way, it can be checkedthat

(Dec∗)⊗m ΓC (Enc∗)⊗n(· ⊗ |0〉〈0|⊗n

)= Γ (·)⊗ |0〉〈0|⊗m

for any quantum circuit Γ :M⊗n2 →M⊗m2 , where ΓC denotes its implementation as in Definition II.3.So far, we have not considered any noise affecting the elementary gates of a quantum circuit. In

Definition II.3 we have only described how to implement a given quantum circuit within a certain quantumerror correcting code. It should be noted that the gadgets required for this construction are always easy toconstruct. However, it is more challenging to construct these gadgets in a way such that implementationsof quantum circuits become fault-tolerant, which is one of the main achievements of [5]. We will not gointo the details of these constructions, but only review the concepts needed for our main results.

Page 10: Fault-tolerant Coding for Quantum Communication

10

Intuitively, a quantum circuit affected by noise should be called “correct” if its behaviour matchesthat of the same circuit without noise. Moreover, the probability of a quantum circuit being “correct”under the noise model that is considered should be high after implementing the circuit in a quantumerror correcting code. It is the central idea in [5] to derive “correctness” of noisy quantum circuits fromconditions that are satisfied with high probability by the rectangles making up the circuit. However, thisapproach requires some care. When the input to a rectangle is unconstrained it might already includea high level of noise. Then, a single additional fault occurring with probability p under the noise modelFπ(p) in the rectangle could cause the accumulated faults to be uncorrectable and the overall circuitnot to agree with its desired output. As a consequence the quantum circuit containing the rectanglewould not be correct. To avoid this problem it makes sense to derive “correctness” from properties ofextended rectangle combining the rectangle and the error correction preceeding it. By designing the errorcorrection in a certain way [5] its output can be controlled even if there are faults, and a meaningfulcondition can be defined. In the following, we will make these notions precise.

Definition II.4 (Extended rectangles). Let C ⊂ (C2)⊗K satisfying dim(C) = 2 be the code space ofa stabilizer code where a gadget is defined for every elementary operation and for the error correctionoperation as in Definition II.3. The extended rectangle corresponding to an elementary operation G∗

with corresponding gadget G is the quantum circuit given by

EG =

EC G EC if G∗ is a single qubit operation.EC⊗2 G EC⊗2 if G∗ is a CNOT gate.G EC if G∗ is a measurement or a partial trace.EC G if G∗ is a preparation.

In [5, p.10] a combinatorial condition called “goodness” is introduced for extended rectangles affectedby fault patterns. This condition behaves as follows: First, using the ideal decoder and encoder it canbe shown that an implemented quantum circuit with classical input and output and affected by noisecan be transformed into the ideal quantum circuit without noise whenever all its extended rectangles are“good”. Secondly, the probability of an extended rectangle being “good” under the Pauli i.i.d. noise modelFπ(p) is very high. By using the union bound, it can then be concluded that an implemented quantumcircuit affected by the noise model Fπ(p) is “correct” with high probability. Unfortunately, the precisedefinition of “goodness” is quite cumbersome, and we have chosen to avoid it here. Instead, we definewhen extended rectangles are well-behaved under a fault-pattern, which is inspired by the transformationrules stated in [5, p.11] for “good” extended rectangles. In particular, a well-behaved quantum circuit,i.e., a quantum circuit in which all extended rectangles are well-behaved, can be transformed in the sameway as in [5] leading to an ideal quantum circuit when input and output are classical. Moreover, thenotion of “goodness” from [5, Section 3.1.] implies our notion of “well-behaved”.

Definition II.5 (Well-behaved extended rectangles and quantum circuits). Let C ⊂ (C2)⊗K satisfyingdim(C) = 2 be the code space of a stabilizer code where a gadget is defined for every elementary operationas in Definition II.3. Let G∗ be an elementary operation with corresponding extended rectangle EGaccording to Definition II.4, and let F be a Pauli fault pattern on the quantum circuit EG. We will callthe extended rectangle EG well-behaved under the fault pattern F if the corresponding condition holds:

1. The operation G∗ is a single qubit operation and we have

Dec∗ [EG]F = (G∗ ⊗ SFG) Dec∗ [EC]F ,

for some quantum channel SFG :M⊗(K−1)2 →M⊗(K−1)

2 on the syndrome space.

2. The operation G∗ is a CNOT gate and we have

(Dec∗)⊗2 [EG]F = (G∗ ⊗ SFG) (Dec∗)⊗2 [(EC)⊗2]

F,

for some quantum channel SFG :M⊗2(K−1)2 →M⊗2(K−1)

2 on the syndrome space.

3. The operation G∗ is a measurement or a trace and we have

[EG]F = (G∗ ⊗ Tr) Dec∗ [EC]F .

Page 11: Fault-tolerant Coding for Quantum Communication

11

4. The operation G∗ is a preparation and we have

Dec∗ [EG]F = G∗ ⊗ σFG,

for some quantum state σFG ∈M⊗(K−1)2 .

Similarily, we will call the implementation ΓC :M⊗nK2 →M⊗mK2 of a quantum circuit Γ :M⊗n2 →M⊗m2well-behaved under the Pauli fault pattern F affecting the quantum circuit ΓC if all extended rectanglescontained in [ΓC ]F are well-behaved.

To analyze a faulty but well-behaved implementation of a quantum circuit we can use the transformationrules from Definition II.5 repeatedly. First, we either introduce an ideal decoder after the final step of theimplemented circuit, or if the quantum circuit has classical output we use the transformation rules formeasurements or traces in its final step and thereby obtain an ideal decoder. Second, we move the idealdecoder towards the beginning of the quantum circuit using the transformation rules from Definition II.5repeatedly. In Figure 3 we depict a schematic of this process.

EC EC ECG G Dec1 2

*

extended rectangle 1 extended rectangle 2

= EC EC

S

G

G

Dec1

2*

*

= EC

S

G

Dec1

*

* G2*

2

S21

FIG. 3: Transforming faulty but well-behaved extended rectangles using Definition II.5.

Finally, if the quantum circuit has classical input can use the transformation rule for a preparation(depending on the classical data) from Definition II.5 to remove the ideal decoder in the initial step of thequantum circuit, or we keep the ideal decoder before the first error correction appearing in the quantumcircuit. The argument just given is implicitly contained in [5, Lemma 4] and the next lemma will makethe conclusion more precise:

Lemma II.6 (Transformation of well-behaved implementations). Let C ⊂ (C2)⊗K satisfying dim(C) = 2be the code space of a stabilizer code where a gadget is defined for every elementary operation as inDefinition II.3, and let Γ : M⊗n2 → M⊗m2 be a quantum circuit. If the quantum circuit ΓC EC⊗n iswell-behaved under a Pauli fault pattern F , then we have

(Dec∗)⊗m [ΓC EC⊗n

]F

=(Γ⊗ SFΓ

) (Dec∗)⊗n

[EC⊗n

]F,

for some quantum channel SFΓ : M⊗n(K−1)2 → M⊗m(K−1)

2 acting on the syndrome space. Moreover, ifΓ : C2n → C2m is a quantum circuit with classical input and output, and F is a Pauli fault pattern underwhich the quantum circuit ΓC is well-behaved, then we have [ΓC ]F = Γ.

Proof. We will only state the proof for the first part of the lemma, since the second part is an obviousmodification. By Definition II.3 the final part of the circuit ΓC are error corrections, measurements ortraces, depending on the final elementary operations in the circuit diagram of Γ. In the following, wedenote by Γ :Mn

2 →Mm′

2 the quantum circuit obtained by removing the final measurements and partialtraces from Γ leaving m′ ≥ m qubits in the output. Without loss of generality we can assume qubits1, . . . ,m in the output of Γ to correspond to the output of Γ. The extended rectangles corresponding to

Page 12: Fault-tolerant Coding for Quantum Communication

12

any measurement or partial trace are well-behaved by assumption, and replacing them as in DefinitionII.5 shows that

(Dec∗)⊗m [ΓC EC⊗n

]F

=

m⊗j=1

(id2 ⊗ idSj )⊗m′⊗

i=m+1(G∗i ⊗ TrSi)

(Dec∗)⊗m′[ΓC EC⊗n

]F,

where G∗i denote the ideal measurements or partial traces applied in the final step of Γ and where theTrSi acts on the syndrome system belonging to this code block. Using that every extended rectangleis well-behaved in the circuit

[ΓC EC⊗n

]F

affected by fault pattern F , we can apply Definition II.5successively to transform each extended rectangle into the corresponding ideal operation. By doing so,the ideal decoder Dec∗ moves towards the beginning of the circuit and in the final step we leave it directlyafter the initial error corrections. Collecting the quantum channels and partial traces acting on thesyndrome space, and the quantum states on the syndrome space emerging from well-behaved extendedrectangles corresponding to preparations into a single quantum channel SFΓ : Mn(K−1)

2 → Mm(K−1)2

finishes the proof.

Lemma II.6 is only valid under the assumption that every extended rectangle in a potentially largecircuit is well-behaved. Without any further assumptions this event might be very unlikely. In thenext section we will restrict to quantum error correcting codes for which the extended rectangles arewell-behaved with very high probability. Note that formally this is a property of both the code andthe implementation of elementary gadgets (cf. Definition II.3). In the following we will restrict to theconcatenations [31] of the 7-qubit Steane code for which elementary gadgets have been constructedin [5]. Using these gadgets it is possible to prove the threshold theorem of [5] showing that fault-tolerantimplementations of quantum circuits are possible.

D. Concatenated quantum error correcting codes and the threshold theorem

A major result on fault-tolerant implementations of quantum circuits using quantum error correctingcodes is the threshold theorem by Aharonov and Ben-Or [1–3]. Here, we will focus our discussion onconcatenated quantum error correcting codes [31] constructed from the 7-qubit Steane code and onthe version of the threshold theorem stated in [5, Theorem 1]. For convenience we have collected theconstruction and basic properties of this family of quantum error correcting codes in Appendix A, butin the following we will not need to go through these details. We will only state the threshold theoremfrom [5] using the terminology introduced in the previous section. This will be sufficient to prove theresults in the rest of our article.For any l ∈ N let Cl ⊂ (C2)⊗7l denote the lth level of the concatenated 7-qubit Steane code. Note that

each level defines a quantum error correcting code as introduced in the previous section. In particular,we use the following terminology throughout our article:

• We denote by Enc∗l , Dec∗l , and EC∗l the ideal operations introduced in (3), (2), and (4) respectivelyfor the lth level of the concatenated 7-qubit Steane code.

• We refer to the elementary gadgets, error corrections, rectangles, and extended rectangles (seeDefinition II.3 and Definition II.4) at the lth level of the concatenated 7-qubit Steane code asl-gadgets, l-error corrections, l-rectangles, and l-extended rectangles respectively. In formulas lindices will indicate the level.

• All gadgets are constructed as explained in [5, Section 7].In [5, p.10] the notion of “good” extended rectangles is introduced, which implies the transformation

rules of Definition II.5. As a consequence an extended rectangle is well-behaved whenever it is “good”.Therefore, the following lemma follows directly from [5, Lemma 2].

Lemma II.7 (Threshold lemma). For l ∈ N let Cl ⊂ (C2)⊗7l denote the lth level of the concatenated 7-qubit Steane code. For every l there exist gadgets implementing the elementary operations as in DefinitionII.3 such that the following holds: There is a threshold p0 ∈ (0, 1] such that for any 0 ≤ p < p0, anyl ∈ N, and any l-extended rectangle E(l)

G we have

P([E

(l)G

]F

is not well-behaved)≤ p0

(p

p0

)2l

,

Page 13: Fault-tolerant Coding for Quantum Communication

13

where P denotes the probability distribution over Pauli fault patterns F induced by the fault model Fπ(p).

Using the threshold theorem, long quantum computations can be protected from noise. The probabilityof any extended rectangle in a given quantum circuit to be not well-behaved can be upper bounded usingthe union bound. Combining this with Lemma II.6 leads to the following theorem, which can be seen asyet another version of the threshold theorem [5, Theorem 1]):

Theorem II.8 (Threshold theorem II.). For l ∈ N let Cl ⊂ (C2)⊗7l denote the lth level of the concatenated7-qubit Steane code with threshold p0 ∈ (0, 1] as in Lemma II.7. For any 0 ≤ p < p0, any l ∈ N thefollowing statements hold:

1. For any quantum circuit Γ :M⊗n2 →M⊗m2 we have

P(An extended rectangle in

[ΓCl EC⊗nl

]F

is not well-behaved)≤ C

(p

p0

)2l

|Loc (Γ) |,

where C ∈ R+ is a constant independent of l ∈ N and Γ, and P denotes the probability distributionover Pauli fault patterns F induced by the fault model Fπ(p).

2. For a quantum circuit Γ : C2n → C2m with classical input and classical output we have

‖Γ− [ΓCl ]Fπ(p) ‖1→1 ≤ 2C(p

p0

)2l

|Loc (Γ) |.

It should be emphasized that in the previous discussion we did not have to define any notion of “correct”.In [5, p.11] correctness of rectangles under fault patterns is defined via the same transformation rules usedin Definition II.5 restricted to the rectangles contained in the corresponding extended rectangles (i.e.,omitting the initial error correction). Since the proof of [5, Theorem 1] and all the proofs in our articleonly use this notion together with well-behaved extended rectangles (or “good” ones in the case of [5]) itis sufficient to only use the stronger notion. An exception is our discussion of interfaces in Section IIIAwhere we do define a notion of “correctness”, which is related but different from the notion used in [5].

E. Capacities of classical-quantum and quantum channels

Arguably, the simplest quantum communication scenario is the transmission of classical informationvia a classical-quantum channel (cq-channel) T : A → Md. Here, A = 1, . . . , |A| is a classical inputalphabet and for each i ∈ A the output T (i) ∈ D(Cd) is a quantum state. We define the classicalcommunication error of a classical channel K : Cd → Cd as

εcl(K) := 1− 1d

d∑i=1

(K (|i〉))i , (5)

Now, we can state the following definition:

Definition II.9 (Coding schemes for cq-channels). For n,m ∈ N and ε ∈ R+ an (n,m, ε)-coding schemefor the cq-channel T : A →Md consists of a map E : 1, . . . , 2n → Am and a channel D :M⊗md → C2n

with classical output such that the classical communication error

εcl(D T⊗m E

)≤ ε.

With the previous definition we define the capacity of a cq-channel as follows:

Definition II.10 (Capacity of a cq-channel). The classical capacity of a cq-channel T : A → Md isdefined as

C(T ) = supR ≥ 0 achievable,

Page 14: Fault-tolerant Coding for Quantum Communication

14

where R ≥ 0 is called achievable if for every m ∈ N there exists nm ∈ N such that there are (nm,m, εm)-coding schemes with limm→∞ εm = 0 and

R ≤ lim infm→∞

nmm.

Given an ensemble of quantum states pi, ρii, i.e., such that (pi)i is a probability distribution and ρiare quantum states, the Holevo quantity [28] is given by

χ (pi, ρii) = S

(∑i

piρi

)−∑i

piS(ρi). (6)

Here, S(ρ) = −Tr (ρ log2(ρ)) denotes the von-Neumann entropy. We now recall the following theorem:Theorem II.11 (Holevo,Schumacher,Westmoreland [26, 47]). For any cq-channel T : A →Md we have

C(T ) = supp∈P(A)

χ (pi, T (i)i) .

For quantum channels we will consider the classical capacity and the quantum capacity quantifying theoptimal transmission rates of classical information or quantum information respectively via the quan-tum channel. Again, we will start by defining the coding schemes considered for these communicationproblems. In addition to the classical communication error from (5) we need to define the quantumcommunication error of a quantum channel T :Md →Md by

εq(T ) := 1−min|ψ〉〈ψ|T (|ψ〉〈ψ|)|ψ〉,

where the minimum is over pure quantum states |ψ〉 ∈ Cd with 〈ψ|ψ〉 = 1.Definition II.12 (Coding schemes). Let T : MdA → MdB denote a quantum channel. For n,m ∈ Nand ε ∈ R+ an (n,m, ε) coding scheme for

• classical communication consists of a cq-channel E : C2n →M⊗mdA and a channel D :M⊗mdB → C2n

with classical output such that εcl (D T⊗m E) ≤ ε.

• quantum communication consists of a quantum channel E :M⊗n2 →M⊗mdA and a quantum channelD :M⊗mdB →M

⊗n2 such that εq (D T⊗m E) ≤ ε.

With this we can state the following definition.Definition II.13 (Classical and quantum capacity). Let T :MdA →MdB denote a quantum channel.We call R ≥ 0 an achievable rate for classical (quantum) communication if for every m ∈ N there existsan nm ∈ N and an (nm,m, εm) coding scheme for classical (quantum) communication with εm → 0 asm→∞ and

lim infm→∞

nmm≥ R.

The classical capacity of T is given by

C(T ) = supR ≥ 0 achievable rate for classical communication,

and the quantum capacity of T is given by

Q(T ) = supR ≥ 0 achievable rate for quantum communication.

The classical and quantum capacity can be related to entropic quantities. Given a quantum channelT :MdA →MdB we define

χ(T ) := suppi,ρii

χ (pi, T (ρi)i) , (7)

with the Holevo quantity from (6) and where the supremum is over all ensembles pi, ρii with quantumstates ρi ∈ D

(CdA

). The following theorem is a major achievement of quantum Shannon theory.

Page 15: Fault-tolerant Coding for Quantum Communication

15

Theorem II.14 (Holevo, Schumacher, Westmoreland [26, 47]). For any quantum channel T : MdA →MdB we have

C(T ) = limk→∞

1kχ(T⊗k).

For a bipartite quantum state ρAB ∈ D(CdA ⊗ CdB

)we define the coherent information as

Icoh (A;B)ρ := S(ρB)− S(ρAB),

where ρB ∈ D(CdB

)denotes the marginal ρB = TrA (ρAB). For a quantum channel T : MdA → MdB

we define

Icoh (T ) := maxρAA′

Icoh (A;B)(idA⊗T )(ρAA′ ), (8)

where the maximum is over quantum states ρAA′ ∈ D(CdA ⊗ CdA

). The following theorem is a major

achievement of quantum information theory building on earlier results by [8, 45, 46]:

Theorem II.15 (Lloyd, Shor, Devetak [18, 35, 50]). For any quantum channel T : MdA → MdB wehave

Q(T ) = limk→∞

1kIcoh

(T⊗k

).

III. FAULT-TOLERANT TECHNIQUES FOR QUANTUM COMMUNICATION

A. Interfaces for concatenated codes

To study capacities of quantum channels with encoding and decoding operations protected by quantumerror correcting codes we need interfaces between the code space and the physical systems where thequantum channel acts (see Figure 2). For simplicity we will only consider interfaces between physicalqubits and the code spaces composed from many qubits. It is straightforward to adapt our definitionsand results to quantum error correcting codes using qudits.

Definition III.1 (Interfaces). Let C ⊂ (C2)⊗K be a stabilizer code with dim (C) = 2, such that |0〉〈0| ∈M⊗(K−1)

2 denotes the pure state corresponding to the zero syndrome, and Enc∗ :M2⊗M⊗(K−1)2 →M⊗K2

denotes the ideal encoding operation and Dec∗ : M⊗K2 → M2 ⊗M⊗(K−1)2 denotes the ideal decoding

operation (see Section II C). An interface for C is given by an encoding quantum circuit Enc :M2 →M⊗K2and a decoding quantum circuit Dec :M⊗K2 →M2 such that the following conditions hold:

1. The quantum circuit EC from Definition II.3 is applied at the end of Enc.

2. We have

Dec∗ Enc = id2 ⊗ |0〉〈0|.

3. We have

Dec Enc∗(· ⊗ |0〉〈0|) = id2(·).

The quantum circuits Enc and Dec are related to the ideal quantum channels Enc∗ and Dec∗ asspecified in the previous definition. However, Enc and Dec should be seen as quantum circuits that areimplemented physically (and will be subject to noise), while Enc∗ and Dec∗ are ideal (mathematical)objects that only appear in the formal analysis of quantum circuits. Next, we define correctness forinterfaces affected by noise.

Page 16: Fault-tolerant Coding for Quantum Communication

16

Definition III.2 (Correctness of interfaces). Let C ⊂ (C2)⊗K be a stabilizer code with dim (C) = 2, andlet Enc :M2 →M⊗K2 and Dec :M⊗K2 →M2 denote an interface for C.

• We say that Enc is correct under a Pauli fault pattern F if

Dec∗ [Enc]F = id2 ⊗ σF , (9)

for a quantum state σF ∈M⊗(K−1)2 on the syndrome space depending on F .

• We say that the quantum circuit Dec EC is correct under a Pauli fault pattern F if

[Dec EC]F = (id2 ⊗ TrS) Dec∗ [EC]F , (10)

where TrS :M⊗(K−1)2 → C traces out the syndrome space.

Note that in the above definition, correctness is defined differently for the encoder and the decoder, sincewe want to use the transformation rules from Definition II.5 and Lemma II.6 to analyze these interfaces.For example, we need to consider the whole circuit Dec EC including an initial error correction in (10),since otherwise we could not decompose the interface into extended rectangles and use the notion ofwell-behavedness as in Definition II.5 and Lemma II.6. We do not have this problem for the encoder,since it ends in a final error correction due to Definition III.1.Interfaces should be robust against noise when they are used for quantum information processing.

Clearly, the probability of a fault occuring in an interface between a quantum error correcting code anda single qubit is at least as large as the probability of a fault in a single qubit operation, which couldhappen at the end of Dec or at the beginning of Enc. Fortunately, it is possible for concatenated quantumerror correcting codes to construct interfaces that only fail with a probability slightly larger than thislower bound. This result has previously been established in [38] (see also [36] for a similar discussionfor various topological codes). Since the proof given in [38] seems to neglect certain details, we will herepresent a more rigorous version of the argument using the formalism stated in the previous section. Itshould be emphasized that the main ideas of the argument are the same as in [38].

Theorem III.3 (Nice interface for concatenated 7-qubit code). For l ∈ N we denote by Cl ⊂ (C2)⊗7l

the lth level of the concatenated 7-qubit Steane code with threshold p0 ∈ (0, 1] (see Lemma II.7). Thereexists a constant c > 0, encoding quantum circuits Encl : M2 → M⊗7l

2 and decoding quantum circuitsDecl : M⊗7l

2 → M2 for each l ∈ N forming an interface for the stabilizer code Cl, such that for any0 ≤ p ≤ p0/2 we have

P ([Encl]F not correct) ≤ 2cp,

and

P ([Decl]F not correct) ≤ 2cp,

where P denotes the probability distribution over Pauli fault patterns induced by the fault model Fπ(p).

The proof of the previous theorem will construct the desired interfaces successively by defining partialencoding (and decoding) quantum circuits between different levels of the code. Before presenting theproof, we will state some lemmas:

Lemma III.4 (Successive interfaces). Let (Cl)l∈N be a family of stabilizer codes such that for everyl ∈ N we have Cl ⊂ (C2)⊗Kl , dim (Cl) = 2, and the state |0〉〈0|l ∈ M⊗(Kl−1)

2 denotes the pure statecorresponding to the zero syndrome of Cl. Assume that for every l ∈ N there are quantum circuitsEnc(l−1)→l : M⊗Kl−1

2 →M⊗Kl2 and Decl→(l−1) : M⊗Kl2 →M⊗Kl−12 , where K0 := 1, with the following

properties:

1. The pair Enc0→1 and Dec1→0 is an interface for C1.

2. For any l ≥ 2 we have

Dec∗l Enc(l−1)→l Enc∗l−1 (· ⊗ |0〉〈0|l−1) = id2(·)⊗ |0〉〈0|l.

Page 17: Fault-tolerant Coding for Quantum Communication

17

3. For any l ≥ 2 we have

Decl→(l−1) Enc∗l (· ⊗ |0〉〈0|l) = Enc∗l−1(· ⊗ |0〉〈0|l−1).

Then, for any l ∈ N the quantum circuits Encl :M2 →M⊗Kl2 and Decl :M⊗Kl2 →M2 given by

Encl := Enc(l−1)→l · · · Enc1→2 Enc0→1

and

Decl := Dec1→0 · · · Dec(l−1)→(l−2) Decl→(l−1)

are an interface for Cl according to Definition III.1.Proof. The proof proceeds by induction on l ∈ N. For l = 1 we have Enc1 = Enc0→1 and Dec1 = Dec1→0,which is an interface for C1 by assumption. By definition we have

Encl+1 = Encl→(l+1) Encl,

and

Decl+1 = Decl Dec(l+1)→l .

Assuming that for l ≥ 1 the pair of quantum circuits Encl and Decl is an interface for Cl we find that

Dec∗l+1 Encl+1 = Dec∗l+1 Encl→(l+1) Encl= Dec∗l+1 Encl→(l+1) Enc∗l Dec∗l Encl= Dec∗l+1 Encl→(l+1) Enc∗l (· ⊗ |0〉〈0|l)= id2 ⊗ |0〉〈0|l+1

where we used that Enc∗l and Dec∗l are inverse to each other in the second equality sign and the inductionhypothesis together with the Property 2 from Definition III.1 in the third equality sign. In the last step,we used the Assumption 2 on Encl→(l+1). Similarly, we find that

Decl+1 Enc∗l+1(· ⊗ |0〉〈0|l+1) = Decl Dec(l+1)→l Enc∗l+1(· ⊗ |0〉〈0|l+1)= Decl Enc∗l (· ⊗ |0〉〈0|l)= id2(·),

where we used Assumption 3 in the second equality sign and the inductive hypothesis in the last equalitysign. This shows that the pair Encl and Decl form an interface for Cl according to Definition III.1.

To construct an interface for the concatenated 7-qubit Steane code, we will first construct interfacecircuits Enc0→1 and Dec1→0 for the first level, i.e., for the 7-qubit Steane code. By implementing Enc0→1and Dec1→0 in higher levels of the concatenated code, we will then obtain general interface circuits. Theconstruction outlined below works in general whenever the quantum circuits Enc0→1 and Dec1→0 definean interface for the 7-qubit Steane code, and only the size of these circuits will appear in the main results.For convenience we have included an explicit construction of interfaces Enc0→1 and Dec1→0 in AppendixA 3 using a simple teleportation circuit.In the following, let Enc0→1 and Dec1→0 denote a fixed interface for the 7-qubit Steane code. For l ∈ N

we define the following quantum circuits as implementations as in Definition II.3:

Enc(l−1)→l := (Enc0→1)Cl−1and Decl→(l−1) := (Dec1→0)Cl−1

. (11)

Using (A6) from Appendix A 2 it is easy to verify the conditions of Lemma III.4 for the quantum circuitsfrom (11). Therefore, for any l ∈ N, the quantum circuits

Encl := Enc(l−1)→l · · · Enc1→2 Enc0→1 (12)

and

Decl := Dec1→0 · · · Dec(l−1)→(l−2) Decl→(l−1) (13)

form an interface for the l-th level of the concatenated 7-qubit Steane code. It remains to analyze howthis interface behaves under the i.i.d. Pauli noise model as introduced after Definition II.2. It will behelpful to extend the notion of well-behavedness from Definition II.5 to interfaces:

Page 18: Fault-tolerant Coding for Quantum Communication

18

Definition III.5 (Well-behaved interfaces). For each i ∈ 0, . . . , l − 1 we denote by Enci→(i+1) andDec(i+1)→i the quantum circuits from (11) and by ECi the error correction of the ith level of the concate-nated 7-qubit Steane code, where we set EC0 = id2.

1. We will call the quantum circuit Enci→(i+1) ECi well-behaved under the Pauli fault pattern F ifall i-extended rectangles in

[Enci→(i+1) ECi

]F

are well-behaved.

2. We will call the quantum circuit Dec(i+1)→i ECi+1 well-behaved under the Pauli fault pattern F

if all i-extended rectangles in[Dec(i+1)→i ECi+1

]F

are well-behaved.

Moreover, for l ∈ N we will call the quantum circuit Encl or Decl ECl well-behaved under the Paulifault pattern F if the partial circuits Enci→(i+1) ECi or Dec(i+1)→i ECi+1 respectively are well-behavedunder the restrictions of F for every i ∈ 0, . . . , l − 1.

The following lemma gives an upper bound on the probability that interfaces are well-behaved.

Lemma III.6 (Probability of an interface to be well-behaved). For l ∈ N consider Γl ∈Encl,Decl ECl from (12) and (13). Then, we have

P (Γl not well-behaved under F ) ≤ cp0

l−1∑i=0

(p

p0

)2i

where p0 ∈ (0, 1] denotes the threshold probability (see Lemma II.7), and where

c = max (|Loc (Enc0→1) |, |Loc (Dec1→0 EC1) |) ,

and P denotes the probability distribution over Pauli fault patterns induced by the fault model Fπ(p) andrestricted to Γl.

Proof. For l ∈ N we will state the proof in the case Γl = Encl. The remaining case follows in the sameway. Consider the fault model Fπ(p) restricted to Encl. Let Ni ∈ N denote the number of i-extendedrectangles in the quantum circuit Enci→(i+1) ECi for i ∈ 0, . . . , l − 1, where 0-extended rectanglesare the elementary operations and EC0 = id2. By (11) we have Ni = N0 = |Loc (Enc0→1) |, and by theunion bound and Lemma II.7 we have

P(All i-extended rectangles in

[Enci→(i+1) ECi

]F

are well-behaved)≤ p0N0

(p

p0

)2i

for any i ∈ 0, . . . , l − 1. By comparing with Definition III.5 we find that

P (Encl not well-behaved) ≤ cp0

l−1∑i=0

(p

p0

)2i

,

where c = |Loc (Enc0→1) |.

The following lemma analyzes how the successive interfaces defined above behave under the faultpatterns introduced in Defintion III.5.

Lemma III.7 (Successive interfaces under noise). For each l ∈ N let Cl ⊂ (C2)⊗7l denote the lth levelof the concatenated 7-qubit Steane code, and let Enc(l−1)→l :M⊗7l−1

2 →M⊗7l2 and Decl→(l−1) :M⊗7l

2 →M⊗7l−1

2 denote the quantum circuits from (11). Whenever the quantum circuit Enc(l−1)→l ECl−1 orDecl→(l−1) ECl is well-behaved under a Pauli fault pattern F we have

Dec∗l [Enc(l−1)→l ECl−1

]F

= (id2 ⊗ S(l−1)→lF ) Dec∗l−1 [ECl−1]F

or

Dec∗l−1 [Decl→(l−1) ECl

]F

= (id2 ⊗ Sl→(l−1)F ) Dec∗l [ECl]F ,

respectively, for quantum channels S(l−1)→lF : M⊗(Kl−1−1)

2 → M⊗(Kl−1)2 and S

l→(l−1)F : M⊗(Kl−1)

2 →M⊗(Kl−1−1)

2 between the syndrome spaces of Cl−1 and Cl.

Page 19: Fault-tolerant Coding for Quantum Communication

19

Proof. Recall that by construction of the concatenated 7-qubit Steane code (see (A5)) we have

Dec∗l =(

Dec∗1⊗id⊗7(7l−1−1)2

)(Dec∗l−1

)⊗7. (14)

Let F denote a Pauli fault pattern under which every (l−1)-extended rectangle in[Enc(l−1)→l ECl−1

]F

is well-behaved. Treating Enc(l−1)→l ECl−1 as the implementation of a quantum circuit at level (l− 1)of the concatenated 7 qubit Steane code outputting 7 registers, we have(

Dec∗l−1)⊗7

[Enc(l−1)→l ECl−1

]F

= (Enc0→1⊗SF ) Dec∗l−1 [ECl−1]F ,

by Lemma II.6 and for some quantum channel SF : M⊗(7l−1−1)2 → M⊗7(7l−1−1)

2 . Using (14) and thatEnc0→1 is an interface for C1 (see Definition III.1) we find that

Dec∗l [Enc(l−1)→l ECl−1

]F

= (Dec∗1 Enc0→1⊗SF ) Dec∗l−1 [ECl−1]F=(id2 ⊗ SF

)Dec∗l−1 [ECl−1]F ,

where SF = SF ⊗ |0〉〈0| for the pure state |0〉〈0| ∈ M⊗62 corresponding to the 0 syndrome of C1. This

verifies the first statement of the lemma.For the second statement, let F denote a Pauli fault pattern under which every (l − 1)-extended

rectangle in[Decl→(l−1) ECl

]F

is well-behaved. Since ECl = (EC1)Cl−1the quantum circuit ECl has

error corrections EC⊗7l−1 in its final step, and we can decompose ECl = EC⊗7

l−1 Γ for some quantum circuitΓ. Treating Decl→(l−1) EC⊗7

l−1 as the implementation of a quantum circuit at level l − 1 with 7 inputregisters, we find

Dec∗l−1 [Decl→(l−1) ECl

]F

= Dec∗l−1 [Decl→(l−1) EC⊗7

l−1 Γ]F

= (Dec1→0⊗SF ) (Dec∗l−1

)⊗7 [ECl]F ,

for some quantum channel SF :M⊗7(7l−1−1)2 →M⊗(7l−1−1)

2 . Finally, we can use (14) and that Dec1→0is an interface for C1 (see Definition III.1) to prove

Dec∗l−1 [Decl→(l−1) ECl

]F

= ((Dec1→0 Enc∗1 Dec∗1)⊗ SF ) (Dec∗l−1

)⊗7 [ECl]F= (id2 ⊗ SF ) Dec∗l [ECl]F .

Now, we are in the position to prove Theorem III.3.

Proof of Theorem III.3. For each l ∈ N consider the interfaces given by Encl and Decl from (12) and (13)defined via the quantum circuits Enci→(i+1) and Dec(i+1)→i from (11). We will now show that Encl iscorrect as in Definition III.2 under a fault pattern F whenever it is well-behaved under F as in DefinitionIII.5. Our proof will use induction on the level l ∈ N. Clearly, we have

Dec∗1 [Enc1]F = Dec∗1 [Enc0→1]F = id2 ⊗ |0〉〈0|1, (15)

for any fault pattern F under which Enc1 is well-behaved, since then every elementary gate in Enc0→1 isideal by Definition III.5. Next, consider a fault pattern F under which Encl+1 is well-behaved for somel ∈ N. Note that by Definition III.5 also Encl (arising as a part of Encl+1) is well-behaved under F whenrestricted to that quantum circuit. Therefore, we can first apply Lemma III.7 and then the inductionhypothesis to compute

Dec∗l+1 [Encl+1]F = Dec∗l+1 [Encl→(l+1) ECl ˜Encl

]F

(16)

= (id2 ⊗ Sl→(l+1)F ) Dec∗l [Encl]F (17)

= id2 ⊗ σF . (18)

Here, ˜Encl denotes the quantum circuit obtained from Encl by removing the final error correction ECl,and

Sl→(l+1)F :M⊗(Kl−1)

2 →M⊗(Kl+1−1)2

Page 20: Fault-tolerant Coding for Quantum Communication

20

denotes a quantum channel between the syndrome spaces of Cl and Cl+1, and

σF =l−1∏i=1

Si→(i+1)F (|0〉〈0|1)

is the final syndrome state. This shows that [Encl]F is correct as in Definition III.2 whenever Encl iswell-behaved under F according to Definition III.5. By Lemma III.6 we find that

P ([Encl]F not correct) ≤ P (Encl not well-behaved under F ) ≤ cp0

l−1∑i=0

(p

p0

)2i

,

where

c = max (|Loc (Enc0→1) |, |Loc (Dec1→0 EC1) |) .

In the case where 0 ≤ p ≤ p0/2 we upper bound the previous sum using a geometric series and obtain

P ([Encl]F not correct) ≤ 2cp.

To deal with Decl we will again employ induction on the level l ∈ N to show that the quantum circuitDecl ECl is correct under a Pauli fault pattern F as in Definition III.2 whenever it is well-behaved underF as in Definition III.5. Let F denote a Pauli fault pattern under which Dec1 EC1 is well-behaved andnote that by Definition III.5 every elementary gate is then ideal. Using that an error correction withoutany faults coincides with the ideal error correction from (4) as a quantum channel, we find

[Dec1 EC1]F = Dec1→0 EC∗1= Dec1→0 EC∗1 EC∗1= Dec1→0 Enc∗1 (id2 ⊗ |0〉〈0|1 Tr) Dec∗1 EC∗1= (id2 ⊗ Tr) Dec∗1 [EC1]F ,

where we used that the ideal error correction EC∗1 is a projection (see (4)) and that Dec1→0 is an interface(cf. Definition III.1). Now, consider a fault pattern F under which [Decl+1 ECl+1]F is well-behaved forsome l ∈ N. By Definition III.5, Decl ECl (arising as a part of the quantum circuit Decl+1 ECl+1) iswell-behaved under the restriction of F . Applying Lemma III.7 and the induction hypothesis we compute

[Decl+1 ECl+1]F =[Decl ECl Dec(l+1)→l ECl+1

]F

= (id2 ⊗ TrS) Dec∗l [Dec(l+1)→l ECl+1

]F

= (id2 ⊗ TrS) (id2 ⊗ S(l+1)→lF ) Dec∗l+1 [ECl+1]F

= (id2 ⊗ TrS) Dec∗l+1 [ECl+1]F ,

where Dec(l+1)→l in the first line denotes the quantum circuit obtained from Dec(l+1)→l by removingthe final error correction ECl. By induction it follows that [Decl ECl]F is correct as in Definition III.2whenever Decl ECl is well-behaved under F . Again, we can apply Lemma III.6 to conclude

P ([Decl]F not correct) ≤ P (Decl ECl not well-behaved under F ) ≤ cp0

l−1∑i=0

(p

p0

)2i

where

c = max (|Loc (Enc0→1) |, |Loc (Dec1→0 EC1) |) .

Finally, whenever 0 ≤ p ≤ p0/2 we can upper bound the previous sum using a geometric series as beforeand obtain

P ([Encl]F not correct) ≤ 2cp.

Page 21: Fault-tolerant Coding for Quantum Communication

21

B. Encoding and decoding of concatenated codes and effective communication channels

To study quantum capacities in a fault-tolerant setting we will need to consider quantum circuits wheresome of the lines are replaced by quantum channels describing stronger noise than the usual noise model.These lines might describe wires connecting physically separate locations where quantum computersare located. Naturally, the noise affecting a qubit during transmission through such a wire might bemuch larger than the noise affecting the gates locally in the quantum computer. Here we will developa framework to deal with this situation. We will start with a lemma combining the quantum interfacesintroduced in the previous section with general quantum circuits.

Lemma III.8 (Noisy interfaces and quantum circuits). Let Γ1 : C2N →M⊗m2 be a quantum circuit withclassical input, and Γ2 : M⊗m2 → C2M be a quantum circuit with classical output. For each l ∈ N letCl ⊂ (C2)⊗7l denote the lth level of the concatenated 7-qubit Steane code with threshold p0 ∈ (0, 1] (seeLemma II.7). Moreover, we denote by Encl :M2 →M⊗7l

2 and Decl :M⊗7l2 →M2 the interface circuits

from (12) and (13), and by c > 0 the constant from Theorem III.3. Then, the following two statementshold for any 0 ≤ p ≤ p0/2:

1. For any l ∈ N there exists a quantum channel Nl :M2 ⊗M⊗(7l−1)2 →M2 acting on a data qubit

and the syndrome space and only depending on l and the interface circuit Decl, and a quantumstate σS on the syndrome space such that

∥∥∥ [Dec⊗ml Γ1Cl]Fπ(p) −

(Ndec,lp

)⊗m (Γ1 ⊗ σS) ∥∥∥

1→1≤ 2C1

(p

p0

)2l

|Loc(Γ1)|+ 2C2

(p

p0

)2l−1

m,

where Ndec,lp :M2 ⊗M⊗(7l−1)

2 →M2 is the quantum channel given by

Ndec,lp = (1− 2cp)id2 ⊗ TrSl +2cpNl

acting on a data qubit and the syndrome space.

2. For any l ∈ N there exists a quantum channel N ′l :M2 →M2 only dependent on l and the interfacecircuit Encl such that

∥∥∥ [Γ2Cl Enc⊗ml

]Fπ(p) − Γ2

(Nenc,lp

)⊗m ∥∥∥1→1≤ 2C

(p

p0

)2l

|Loc(Γ2)|,

where Nenc,lp :M2 →M2 is the quantum channel given by

Nenc,lp = (1− 2cp)id2 + 2cpN ′l .

In the above, C1, C2, C > 0 denote constants not depending on m, l or the quantum circuits involved.

The proof of Lemma III.8 will be based on Theorem III.3 showing that the probability of an interfacenot being correct under the noise model Fπ(p) is upper bounded by an expression linear in p. However, amajor difficulty arises from the fact, that the two notions “well-behavedness” (of the quantum circuits Γ1

Cland Γ2

Cl) and “correctness” (of the interfaces Encl and Decl) refer to overlapping parts of the combinedcircuits Dec⊗ml Γ1

Cl and Γ2Cl Enc⊗ml . To be precise, the circuits in question overlap in joined error

corrections. To obtain the i.i.d. structure of Lemma III.8 we have to deal with this overlap, which willtake the largest part of the following proof. It should be noted that doing so is slightly more difficult inthe first part of the proof considering the interface Decl.

Proof.

1. The quantum circuit Γ1Cl ends in error corrections according to Definition II.3, and we can write

Dec⊗ml Γ1Cl = (Decl ECl)⊗m Γ1

Cl ,

for some quantum circuit Γ1Cl . Every fault pattern F affecting the circuit (Decl ECl)⊗m Γ1

Cl canbe decomposed into F = F1 ⊕ F2 ⊕ F3 with fault pattern F1 affecting Dec⊗ml , fault pattern F2

Page 22: Fault-tolerant Coding for Quantum Communication

22

Decl

Decl

Decl

ECl

ECl

ECl

Γ1Cl~

F

FF 11

F12

F1m

2

1

F22

F2m

3 F2 F1FIG. 4: Partition of fault patterns in the first part of the Proof of Lemma III.8.

affecting EC⊗ml and F3 affecting Γ1Cl . Because F1 and F2 are local faults acting on tensor products

of quantum circuits, we can decompose F1 = F 11 ⊕ · · · ⊕ Fm1 and F2 = F 1

2 ⊕ · · · ⊕ Fm2 , where thefault pattern F ki acts on the kth line of the m-fold tensor product (Decl ECl)⊗m. See Figure 4 fora schematic of how the fault patterns are labeled.Let B denote the set of all fault patterns F2 ⊕ F3 such that every l-extended rectangle in thequantum circuit [

(ECl)⊗m Γ1Cl

]F2⊕F3

is well-behaved. By the threshold lemma (Lemma II.7) and the union bound we have

ε = P (F2 ⊕ F3 /∈ B) ≤ C1

(p

p0

)2l

|Loc(Γ1)|, (19)

and by Lemma II.6 (for n = 0) we have

(Dec∗l )⊗m

[EC⊗ml Γ1

Cl]F2⊕F3

= Γ1 ⊗ σF2⊕F3 (20)

for any F2 ⊕ F3 ∈ B, and where σF2⊕F3 denotes some quantum state on the syndrome space. Forevery k ∈ 1, . . . ,m we define the function

β(F k1 ⊕ F k2

)=

0 if Decl ECl well-behaved under F k1 ⊕ F k21 otherwise

,

where “well-behaved” is used as in Definition III.5. Since any well-behaved interface is correct asin Definition III.2 (cf. proof of Theorem III.3) we can transform the circuit [Decl ECl]Fk1 ⊕Fk2 asdescribed in Definition III.2 whenever β

(F k1 ⊕ F k2

)= 0. If β

(F k1 ⊕ F k2

)= 1, then we can insert an

identity id = Enc∗l Dec∗l directly after the final error correction. This shows that

[Decl ECl]Fk1 ⊕Fk2 = IFk1β(Fk1 ⊕Fk2 ) Dec∗l [ECl]Fk2 , (21)

for quantum channels

IFk10 := id2 ⊗ TrS and I

Fk11 := [Decl]Fk1 Enc∗l .

Using (21) we can rewrite

[Dec⊗ml

]F1[EC⊗ml

]F2

=(

m⊗k=1

(IFk1β(Fk1 ⊕Fk2 ) Dec∗l

))[EC⊗ml

]F2,

Page 23: Fault-tolerant Coding for Quantum Communication

23

where we decomposed F1 = F 11 ⊕ · · · ⊕ Fm1 and F2 = F 1

2 ⊕ · · · ⊕ Fm2 as explained above. Now, wecompute[

(Decl ECl)⊗m Γ1Cl

]Fπ(p)

=∑

F1⊕F2⊕F3

P (F1 ⊕ F2 ⊕ F3)[Dec⊗ml

]F1[EC⊗ml

]F2[Γ1Cl]F3

=∑

F1⊕F2⊕F3

P (F1 ⊕ F2 ⊕ F3)(

m⊗k=1

(IFk1β(Fk1 ⊕Fk2 ) Dec∗l

))[EC⊗ml

]F2[Γ1Cl]F3

=∑

F1⊕F2⊕F3F2⊕F3∈B

P (F1 ⊕ F2 ⊕ F3)(

m⊗k=1

(IFk1β(Fk1 ⊕Fk2 ) Dec∗l

))[EC⊗ml

]F2[Γ1Cl]F3

+ εE

=∑

F1⊕F2⊕F3F2⊕F3∈B

P (F1 ⊕ F2 ⊕ F3)(

m⊗k=1

IFk1β(Fk1 ⊕Fk2 )

)(Γ1 ⊗ σF2⊕F3

)+ εE, (22)

where we used (19) and (20) in the last two equalities. Here, E denotes a quantum channel, andP denotes the probability distribution on fault patterns according to the i.i.d. Pauli fault modelFπ(p).Next, we introduce

β(F k1 ) =

0 if there exists F k2 such that Decl ECl is well-behaved under F k1 ⊕ F k21 otherwise

.

Note that β(F k1 ) = 1 implies β(F k1 ⊕F k2 ) = 1 for any fault pattern F k2 . Therefore, the fault patternsF k1 ⊕ F k2 such that β(F k1 ) 6= β(F k1 ⊕ F k2 ) have to satisfy β(F k1 ) = 0 and β(F k1 ⊕ F k2 ) = 1. This isonly possible if the well-behavedness condition from Definition III.5 fails only in the last step of thecircuit containing the error correction affected by F k2 , i.e., this part of the circuit has to containan (l − 1)-extended rectangle that is not well-behaved under the fault pattern F k1 ⊕ F k2 . By thisreasoning and the union bound we have

δ := P(∃k ∈ 1, . . . ,m : β(F k1 ) 6= β(F k1 ⊕ F k2 )

)≤

m∑k=1

P(β(F k1 ) 6= β(F k1 ⊕ F k2 )

)≤

m∑k=1

P

(An (l − 1)-extended rectangle in

[Deckl→(l−1) ECkl

]Fk1 ⊕Fk2

is not well-behaved)

≤ mC2

(p

p0

)2l−1

,

where we used the threshold lemma (Lemma II.7) in the two last inequalities and introduced thenumber δ. Note that C2 > 0 depends only on the number of (l − 1)-extended rectangles in thequantum circuit Decl→(l−1) ECl, but not on l.Using (22) and twice the triangle inequality we have

‖[(Decl ECl)⊗m Γ1

Cl

]Fπ(p)

−∑

F1⊕F2⊕F3F2⊕F3∈B

P (F1 ⊕ F2 ⊕ F3)m⊗k=1

(IFk1β(Fk1 )

)(Γ1 ⊗ σF2⊕F3

)‖1→1

≤ ‖∑

F1⊕F2⊕F3F2⊕F3∈B

P (F1 ⊕ F2 ⊕ F3)[m⊗k=1

(IFk1β(Fk1 ⊕Fk2 )

)−

m⊗k=1

(IFk1β(Fk1 )

)](Γ1 ⊗ σF2⊕F3

)‖1→1 + ε

≤∑

F1⊕F2⊕F3F2⊕F3∈B

P (F1 ⊕ F2 ⊕ F3)∥∥∥[ m⊗

k=1

(IFk1β(Fk1 ⊕Fk2 )

)−

m⊗k=1

(IFk1β(Fk1 )

)]∥∥∥1→1

+ ε.

Page 24: Fault-tolerant Coding for Quantum Communication

24

The 1 → 1-norm in the final expression of the previous computation is either 0 when β(F k1 ⊕ F k2 )and β(F k1 ) coincide for every k ∈ 1, . . . ,m, or it can be upper bounded by 2 since its argumentis the difference of two quantum channels. Therefore, we find

‖[(Decl ECl)⊗m Γ1

Cl

]Fπ(p)

−∑

F1⊕F2⊕F3F2⊕F3∈B

P (F1 ⊕ F2 ⊕ F3)m⊗k=1

(IFk1β(Fk1 )

)(Γ1 ⊗ σF2⊕F3

)‖1→1

≤ 2P(∃k ∈ 1, . . . ,m : β(F k1 ) 6= β(F k1 ⊕ F k2 ) and F2 ⊕ F3 ∈ B

)+ ε

≤ 2P(∃k ∈ 1, . . . ,m : β(F k1 ) 6= β(F k1 ⊕ F k2 )

)+ ε

= 2δ + ε.

where we used elementary probability theory and δ from above. Since faults F1, F2 and F3 affectingthe three parts of the circuit act independently, we can write

∑F1⊕F2⊕F3F2⊕F3∈B

P (F1 ⊕ F2 ⊕ F3)m⊗k=1

(IFk1β(Fk1 )

)(Γ1 ⊗ σF2⊕F3

)

=∑

F1⊕F2⊕F3F2⊕F3∈B

P1(F1)P2(F2)P3(F3)m⊗k=1

(IFk1β(Fk1 )

)(Γ1 ⊗ σF2⊕F3

)

= (1− ε)m⊗k=1

∑Fk1

P1(F k1 )IFk1

β(Fk1 )

(Γ1 ⊗ σS)

= (1− ε)(Ndec,lq1

)⊗m (Γ1 ⊗ σS),

where P1, P2 and P3 denote the probability distributions on fault patterns according to thei.i.d. Pauli fault model Fπ(p) restricted to the three partial circuits. Furthermore, σS denotesa quantum state on the syndrome space given by

σS = 11− ε

∑F2⊕F3∈B

P2(F2)P3(F3)σF2⊕F3 ,

and we introduced a quantum channel Ndec,lq1

:M2 ⊗M⊗(7l−1)2 →M2 as

Ndec,lq1

= (1− q1)id2 ⊗ TrS +q1Nl

for

q1 := P(β(F k1 ) = 1

)and the quantum channel Nl :M2 ⊗M⊗(7l−1)

2 →M2 defined as

Nl = 1q1

∑F1 s.th. β(F1)=1

P1(F1) [Decl]F1 Enc∗l .

Finally, by the aforementioned properties of β and β and the same reasoning as in the proof ofTheorem III.3 (and with the same constant c > 0) we have

q1 = P(β(F k1 ) = 1

)≤ P

(β(F k1 ⊕ F k2 ) = 1

)≤ 2cp,

where p is the local noise parameter of the i.i.d. Pauli noise model Fπ(p). Finally, we have

Ndec,lq1

= (1− 2cp)id2 ⊗ TrS +2cpNl =: Ndec,lp

for some quantum channel Nl :M2 ⊗M⊗(7l−1)2 →M2 finishing the proof.

Page 25: Fault-tolerant Coding for Quantum Communication

25

2. The quantum circuit Encl from (12) ends in an error correction ECl of the code Cl (cf. DefinitionIII.1) and we can write

Γ2Cl Enc⊗ml = Γ2

Cl EC⊗ml Enc⊗ml ,

where Encl denotes the quantum circuit obtained from Encl by removing the final error correction.In the following, we will write F = F1⊕F2⊕F3 to denote a fault pattern F affecting the quantumcircuit Γ2

Cl Enc⊗ml that is composed of fault patterns F1, F2 and F3 affecting the quantum circuitsΓ1Cl , EC⊗ml and Enc⊗ml respectively. Let A denote the set of fault patterns F = F1 ⊕ F2 such that

every l-extended rectangle in the quantum circuit[Γ2Cl EC⊗ml

]F1⊕F2

is well-behaved. By Lemma II.6 we have[Γ2Cl EC⊗ml

]F1⊕F2

=(Γ2 ⊗ TrS

) (Dec∗l )⊗m

[EC⊗ml

]F2.

for any F1 ⊕ F2 ∈ A, and the threshold lemma (Lemma II.7) shows that

ε := P (F1 ⊕ F2 /∈ A) ≤ C(p

p0

)2l

|Loc(Γ2)|,

where P denotes the probability distribution of fault patterns according to the i.i.d. Pauli faultmodel Fπ(p). Now, we can compute[

Γ2Cl Enc⊗ml

]Fπ(p)

=∑

F fault patternP (F )

[Γ2Cl EC⊗ml Enc⊗ml

]F

=∑

F=F1⊕F2⊕F3F1⊕F2∈A

P1(F1)P2(F2)P3(F3)[Γ2Cl EC

⊗ml Enc⊗ml

]F

+ εE

=(Γ2 ⊗ TrS

) (Dec∗l )⊗m

∑F2⊕F3

P2(F2)P3(F3)

∑F1

F1⊕F2∈A

P1(F1)

[Enc⊗ml]F2⊕F3

+ εE,

for some quantum channel E, and where P1, P2 and P3 denote the probability distributions on faultpatterns according to the i.i.d. Pauli fault model Fπ(p) restricted to the three partial circuits. Bynormalization we have

∑F2⊕F3

P2(F2)P3(F3)

∑F1

F1⊕F2∈A

P1(F1)

= 1− ε.

By definition of the i.i.d. Pauli fault model Fπ(p) we have[Enc⊗ml

]Fπ(p) =

∑F2⊕F3

P2(F2)P3(F3)[Enc⊗ml

]F2⊕F3

.

Using the triangle inequality and that(Γ2 ⊗ TrS

) (Dec∗l )⊗m is a quantum channel, we find that

‖[Γ2Cl Enc⊗ml

]Fπ(p) −

(Γ2 ⊗ TrS

) (Dec∗l )⊗m

[Enc⊗ml

]Fπ(p) ‖1→1

≤ ε+∑F2⊕F3

P2(F2)P3(F3)

1−∑F1

F1⊕F2∈A

P1(F1)

= 2ε.

Page 26: Fault-tolerant Coding for Quantum Communication

26

Finally, we note that by Definition III.2 we have∑F s.th. [Encl]F correct

P (F ) Dec∗l [Encl]F = (1− q1)id2 ⊗ σ,

where

q2 := P ([Encl]F not correct) ≤ 2cp,

using Theorem III.3 (and with the same constant c > 0). Moreover, the quantum state σ on thesyndrome space is given by

σ = 11− q2

∑F s.th. [Encl]F correct

P (F )σF ,

for quantum states σF depending on the specific fault patterns. Dividing the fault patterns F intotwo sets, one where [Encl]F is correct and one where it is not, leads to

(id2 ⊗ TrS) Dec∗l [Encl]Fπ(p) = (1− q2)id2 + q2N′l ,

where N ′l :M2 →M2 is given by

N ′l = 1q2

∑F s.th. [Encl]F not correct

P (F ) (id2 ⊗ TrS) Dec∗l [Encl]F .

Finally, we can rewrite

(1− q2)id2 + q2N′l = (1− 2cp)id2 + 2cpN ′l =: N enc,l

p

for some quantum channel N ′l :M2 →M2 finishing the proof.

The previous lemma shows that using a quantum error correcting code to protect a general codingscheme for information transmission via a physical channel leads to a modified effective channel betweenthe data subspaces. To make this precise, we state the following theorem which is a direct consequenceof Lemma III.8.

Theorem III.9 (Effective communication channel). Let T :M⊗j12 →M⊗j2

2 denote a quantum channel.Furthermore, let Γ1 : C2N → M⊗m2 be a quantum circuit with classical input, and Γ2 : M⊗m2 → C2M

a quantum circuit with classical output. For each l ∈ N let Cl ⊂ (C2)⊗7l denote the lth level of theconcatenated 7-qubit Steane code with threshold p0 ∈ (0, 1] (see Lemma II.7). Moreover, we denote byEncl :M2 →M⊗7l

2 and Decl :M⊗7l2 →M2 the interface circuits from (12) and (13) and by c > 0 the

constant from Theorem III.3. For any 0 ≤ p ≤ min(p0/2, (2(j1 + j2)c)−1) and any l ∈ N we have∥∥∥ [Γ2Cl

(Enc⊗j2

l T Dec⊗j1l

)⊗m Γ1Cl

]Fπ(p)

− Γ2 T⊗mp,l (Γ1 ⊗ σS)∥∥∥

1→1

≤ C(p

p0

)2l (|Loc(Γ1)|+ |Loc(Γ2)|+ j1m

),

where σS denotes some quantum state on the syndrome space of the last j1m lines depending on l ∈ Nthe quantum circuit Γ2 and the interface circuit Decl. The effective quantum channel Tp,l : M⊗j1

2 ⊗M⊗j1(7l−1)

2 →M⊗j22 is of the form

Tp,l = (1− 2(j1 + j2)cp)T ⊗ TrS +2(j1 + j2)cpNl,

where the partial trace is acting on the syndrome space, and the quantum channel Nl : M⊗j12 ⊗

M⊗j1(7l−1)2 → M⊗j2

2 depends on T , the level l and the interface circuits Encl and Decl. In the above,C > 0 denotes a constant not depending on m, j1, j2, l or any of the occurring quantum circuits orquantum channels.

Page 27: Fault-tolerant Coding for Quantum Communication

27

It should be noted that the right hand side of the inequality in Theorem III.9 has to be small for theencoded quantum circuits Γ1

Cl and Γ2Cl to be well-behaved with high probability. If these circuits implement

a coding scheme for transmitting information (classical or quantum) over the quantum channel T , thento function correctly under noise, they also have to be a coding scheme for the effective quantum channelTp,l taking as input the data qubits and a syndrome state that might be entangled over multiple copiesof the quantum channel. Note also that in general the number of locations |Loc(Γ1)| and |Loc(Γ2)| ofthe quantum circuits may scale exponentially in the number of channel uses m. Even in this case, theright hand side of the inequality in Theorem III.9 can be made arbitrarily small by choosing the level llarge enough.The entanglement in the syndrome state σS in Theorem III.9 and how it might affect the effective

quantum channel has to be studied more carefully in the future. It is certainly possible for σS to behighly entangled between multiple communication lines and still correspond to a correctable syndrome.However, we have not actually shown this to be the case in practice. Difficulties arise from the fact thatthe structure of σS depends on the quantum circuits in question, and that high levels of the concatenated7-qubit Steane code and the corresponding interface circuits are quite complicated. In the following, wehave therefore adopted the approach of finding coding schemes for the worst-case scenario, where σS ishighly entangled. These coding schemes will likely be applicable for noise models beyond i.i.d. Pauli noiseFπ(p) including correlations between multiple communication lines.Note that Theorem III.9 is formulated for quantum channels T : M⊗j1

2 → M⊗j22 between quantum

systems composed of qubits. This is only for notational convenience since we consider interfaces betweenlogical qubits and physical qubits. A general quantum channel can always be embedded into a quantumchannel between systems composed of qubits, and then Theorem III.9 applies. We will use this fact inthe next section to obtain more general results.

IV. CAPACITIES UNDER ARBITRARILY VARYING PERTURBATIONS

Before studying fault-tolerant capacities we want to discuss a different kind of capacity which werefer to capacities under arbitrarily varying perturbations. This will turn out to capture the effectivecommunication problems, i.e., for channel models of the same form as the effective communication channelin Theorem III.9, which emerge from applying interface circuits in the scenarios of classical and quantumcommunication over quantum channels.

A. Arbitrarily varying perturbations

Given a quantum channel T :MdA →MdB , we define a quantum channel Tp,N :MdA⊗MdE →MdB

by

Tp,N = (1− p)T ⊗ TrE + pN,

for any dE ∈ N and any quantum channel N :MdA⊗MdE →MdB . We will now consider communicationproblems for the quantum channels

M⊗mdA 3 X 7→ T⊗mp,N (X ⊗ σE) ∈M⊗mdB , (23)

where σE ∈ D((CdE )⊗m

)is some environment state. We call this channel model channels under arbi-

trarily varying perturbation (AVP) motivated by similar channel models [10]. In the following, we willdefine capacities under AVP and we will start by defining the corresponding coding schemes. It shouldbe emphasized that in our definitions, the environment states σE will have unconstrained dimension (butthey might belong to a certain subset of quantum states). Therefore, previous results such as [10], wherethis dimension is finite, are not directly applicable to these definitions.Definition IV.1 (Coding schemes for classical communication under AVP). Let T : MdA → MdB

denote a quantum channel and fix p ∈ [0, 1]. For each m, dE ∈ N consider a set Sm(dE) ⊆ D((CdE )⊗m

)of environment states. For n,m ∈ N and ε ∈ R+, an (n,m, ε) coding scheme for classical communicationover T under AVP of strength p consists of a classical-quantum channel E : C2n →M⊗mdA and a quantum-classical channel D :M⊗m2 → C2n such that

supdE ,σE ,Nm

εcl

(D T⊗mp,Nm (E ⊗ σE)

)≤ ε,

Page 28: Fault-tolerant Coding for Quantum Communication

28

where the supremum goes over the environment dimension dE ∈ N, states σE ∈ Sm(dE) and quantumchannels Nm :MdA ⊗MdE →MdB . By εcl we denote the classical communication error from (5).

We have a similar definition in the case of quantum communication:

Definition IV.2 (Coding scheme for quantum communication under AVP). Let T : MdA → MdB

denote a quantum channel and fix p ∈ [0, 1]. For each m, dE ∈ N consider a set Sm(dE) ⊆ D((CdE )⊗m

)of quantum states. For n,m ∈ N and ε ∈ R+, an (n,m, ε) coding scheme for quantum communicationover T under AVP of strength p consists of a quantum channel E : C2n →M⊗mdA and a quantum channelD :M⊗m2 → C2n such that

supdE ,σE ,Nm

‖id⊗n2 −D T⊗mp,Nm (E ⊗ σE) ‖ ≤ ε,

where the supremum goes over the dimension dE ∈ N, states σE ∈ Sm(dE) and quantum channelsNm :MdA ⊗MdE →MdB .

Having defined coding schemes for quantum channels under AVP, we can define the correspondingcapacities as usual:

Definition IV.3 (Classical and quantum capacity under AVP). Let T :MdA →MdB denote a quantumchannel and fix p ∈ [0, 1]. For each m, dE ∈ N consider a set Sm(dE) ⊆ D

((CdE )⊗m

)of quantum states.

We call R ≥ 0 an achievable rate for classical (or quantum) communication under AVP of strength p iffor every m ∈ N there exists an nm ∈ N and an (nm,m, εm) coding scheme under AVP of strength p forclassical (or quantum) communication over T with εm → 0 as m→∞ and

lim infm→∞

nmm≥ R.

The classical capacity under arbitrarily varying perturbations of T is given by

CAVPSm(dE)(p, T ) = supR ≥ 0 achievable rate for classical communication under AVP of strength p,

and the quantum capacity under arbitrarily varying perturbations of T is given by

QAVPSm(dE)(p, T ) = supR ≥ 0 achievable rate for quantum communication under AVP of strength p.

We will use a special notation for two special cases of these capacities:

• We write CAVPSEP (p, T ) and QAVP

SEP (p, T ), when the set Sm(dE) of environment states consists of allfully separable states in D

((CdE )⊗m

)for each m ∈ N.

• We write CAVPALL (p, T ) and QAVP

ALL(p, T ), when the set Sm(dE) of environment states is the entire setD((CdE )⊗m

)for each m ∈ N.

B. Separable environment states

In the case of separable environment states, the capacities under AVP coincide with certain capacitiesfor arbitrarily varying quantum channels (we refer to [4] for the precise definition of this channel model).Specifically, we have the following lemma:

Lemma IV.4. For any quantum channel T :MdA →MdB we have

CAVPSEP (p, T ) = C ((1− p)T + pSS) and QAVP

SEP (p, T ) = Q ((1− p)T + pSS) ,

where (1− p)T + pSS denotes the arbitrarily varying quantum channel consisting of quantum channels(1− p)T + pS for arbitrary quantum channels S :MdA →MdB .

Proof. To see that the respective left-hand sides are larger than the right-hand sides in the identies ofthe lemma, consider a fixed dimension dE ∈ N a quantum channel N : MdA ⊗MdE → MdB and thequantum channel

M⊗mdA 3 X 7→ T⊗mp,N (X ⊗ σE) ∈M⊗mdB , (24)

Page 29: Fault-tolerant Coding for Quantum Communication

29

for any fixed fully-separable quantum state σE ∈ D((CdE )⊗m

). By assumption, we have

σE =k∑i=1

qiτ(1)i ⊗ · · · ⊗ τ (m)

i ,

for a probability distribution qiki=1 and quantum states τ (1)i , . . . , τ

(m)i ∈ D

(CdE

). Inserting this state

into (24) shows that

T⊗mp,N (X ⊗ σS) =k∑i=1

qi

m⊗j=1

((1− p)T + pNi,j) ,

where Ni,j : MdA → MdB is the quantum channel given by Ni,j(X) = N(X ⊗ τ (j)

i

). By convexity of

the error measures used in Defintion IV.1 and Definition IV.2, we conclude that any communication ratethat is achievable for the arbitrarily varying channel (1−p)T +pSS parametrized by quantum channelsS :MdA →MdB is also achievable under AVP in the case of fully separable environment states.

To see the other directions, consider any m-tuple of quantum channels S1, . . . , Sm :MdA →MdB . Itis easy to see that

m⊗j=1

((1− p)T + pSj) (X) = T⊗mp,N (X ⊗ σE) ,

by choosing N :MdA ⊗MdE →MdB for dE = m as

N(Y ) =m∑i=1

Ni ((1dA ⊗ 〈i|)Y (1dA ⊗ |i〉)) ,

for any Y ∈ MdA ⊗ MdE , and σE =⊗m

j=1 |j〉〈j|. This shows that the arbitrarily varying quantumchannel (1 − p)T + pSS can be simulated by quantum channels of the form (24). We conclude thatcommunication rates under AVP, in the case where the environment states are fully separable, are alsoachievable for the arbitrarily varying quantum channel (1− p)T + pSS .

Now, we have the following lower bound:

Theorem IV.5 (Lower bounds on capacities under AVP with separable environment states). For anyquantum channel T :MdA →MdB with classical capacity C(T ) > 0 there exists a p0 ∈ [0, 1] such that

QAVPSEP (p, T ) ≥ Q(T )− 2p log(dB)− (1 + p)h2

(p

1 + p

),

for any p ≤ p0.

Proof. By Lemma IV.4 and the quantum version of the Ahlswede dichotomy [4, Theorem 5] we have

QAVPSEP (p, T ) = Q ((1− p)T + pSS) ≥ lim

k→∞

1k

maxφA′

infSIcoh

(φA′ , ((1− p)T + pS)⊗k

),

with optimizations over quantum states φA′ ∈ D((CdA)⊗k

)and quantum channels S : MdA → MdB ,

whenever CAVPSEP (p, T ) > 0. Using Theorem IV.7 and the fact that C(T ) > 0, there exists p0 ∈ [0, 1] such

that CAVPSEP (p, T ) ≥ CAVP

ALL (p, T ) > 0. Finally, by the continuity bound [49, Proposition 3A] (see also [33])we find that

QAVPSEP (p, T ) ≥ lim

k→∞

1k

maxφA′

infSIcoh

(φA′ , ((1− p)T + pS)⊗k

)≥ Q(T )− 2p log(dB)− (1 + p)h2

(p

1 + p

),

for any p ≤ p0. This finishes the proof.

We are certain that a similar bound as in the previous theorem should also hold for the classical capacityunder AVP with separable environment states. However, the analogue of the Ahlswede dichotomy forthe classical capacity does not seem to appear in the literature.

Page 30: Fault-tolerant Coding for Quantum Communication

30

C. Unrestricted environment states

To deal with arbitrary environment states, we will use the following postselection technique:Lemma IV.6 (Simple postselection technique). Consider m, dE ∈ N and a quantum state σE ∈D((CdE )⊗m

). For p ∈ [0, 1], and quantum channels T :MdA →MdB and N :MdA ⊗MdE →MdB we

define the quantum channel Tp,N :MdA ⊗MdE →MdB asTp,N = (1− p)T ⊗ TrS +pN

We denote by T⊗mp,N (· ⊗ σE) :M⊗mdA →M⊗mdB

the quantum channel given by

M⊗mdA 3 X 7→ T⊗mp,N (X ⊗ σE) ∈M⊗mdB ,

where each Tp,N acts partially on the state σE. Then, for any δ > 0 we have

T⊗mp,N (· ⊗ σE) ≤ dm(p+δ)B T⊗mp + exp(−mδ2

3p )S

for a quantum channel S :M⊗mdA →M⊗mdB

and where

Tp = (1− p)T + p1dBdB

Tr .

Here, we write S1 ≤ S2 for linear maps S1 and S2 when S2 − S1 is completely positive.Proof. Consider a quantum state σE ∈ D

((CdE )⊗m

), m ∈ N, and some δ > 0. Setting S0 = T ⊗TrE and

S1 = N we findT⊗mp,N (· ⊗ σE)

=m∑k=0

(1− p)m−kpk∑

i1+···+im=k

m⊗s=1

Sis (· ⊗ σE)

=bm(p+δ)c∑k=0

(1− p)m−kpk∑

i1+i2+···im=k

m⊗s=1

Sis (· ⊗ σE) + P(

1m

m∑i=1

Xi > p+ δ

)S,

where S :M⊗mdA →M⊗mdB

denotes a quantum channel collecting the second part of the sum. Furthermore,we introduced independent and identically distributed 0, 1-valued random variables Xi with P (X1 =1) = q. Note that each of the product channels in the remaining sum can be upper bounded as

m⊗s=1

Sis (· ⊗ σE) ≤ dbm(p+δ)cB

m⊗s=1

Sis ,

with S0 = T and S1 = 1dBdB

Tr, where we used that i1 + · · · + im ≤ bm(p + δ)c and that every quantumchannel N ′ :M⊗kdA →M

⊗kdB

satisfies

N ′ ≤ dkB(1dBdB

Tr)⊗k

.

By the Chernoff bound we have

P(

1m

m∑i=1

Xi > p+ δ

)≤ exp

(−mδ2

3p

),

and combining the previous equations, we find that

T⊗mp (· ⊗ σE) ≤ dbm(p+δ)cB

bm(p+δ)c∑k=0

(1− p)m−kpk∑

i1+i2+···im=k

m⊗s=1

Sis + exp(−mδ2

3p

)S

≤ dbm(p+δ)cB

m∑k=0

(1− p)m−kpk∑

i1+i2+···im=k

m⊗s=1

Sis + exp(−mδ2

3p

)S

= dbm(p+δ)cB T⊗mp + exp

(−mδ2

3p

)S,

Page 31: Fault-tolerant Coding for Quantum Communication

31

where we added a completely positive term in the second inequality. Since bm(p + δ)c ≤ m(p + δ) theproof is finished.

We will show the following theorem:

Theorem IV.7 (Lower bounds on capacities under AVP). Let T : MdA → MdB denote a quantumchannel, and fix p ∈ [0, 1). In the classical case, we have

CAVPALL (p, T ) ≥ 1

kχ(T⊗k

)− (2kp log2(dB))1/2| log

(p

dB

)| − 3p log2(dB)− (1 + p)h2

(p

1 + p

),

for any k ∈ N. In the quantum case, we have

QAVPALL(p, T )

≥ 1kIcoh

(T⊗k

)− η(pk)− 3(kp log2(dB))1/2| log

(min

(p

dA,p2

dB

))| − 4p log2(dB)− (1 + p)h2

(p

1 + p

),

for any k ∈ N such that

pk ≤ (2 +√

3)/4 ≈ 0.067.

In the above bounds, h2 denotes the binary entropy and η : [0, 1]→ R+ is given by

η(ε) = (2ε+ 4√ε(1− ε)) log2(dA) + h2(2ε) + h2(2

√ε(1− ε)).

Note that the lower bounds in the previous theorem cannot be regularized, i.e., for fixed p > 0 the limitk → ∞ leads to a vanishing lower bound. However, they are still strong enough to show the followingpoint-wise continuity statement.

Theorem IV.8 (Point-wise continuity of capacities under AVP). For every quantum channel T :MdA →MdB and every ε > 0, there exists a threshold p(ε, T ) > 0 such that

CAVPALL (p, T ) ≥ C(T )− ε,

and

QAVPALL(p, T ) ≥ Q(T )− ε,

for all 0 ≤ p ≤ p(ε, T ). In particular, we have

limp0

CAVPALL (p, T ) = C(T ) and lim

p0QAVP

ALL(p, T ) = Q(T ),

for all quantum channels T :MdA →MdB .

Proof. We will only state the proof for the classical capacity C(T ) since it is the same for the quantumcapacity Q(T ). Let T : MdA → MdB denote a fixed quantum channel. For every ε > 0 there exists akε ∈ N such that 1

kεχ(T⊗kε

)≥ C(T )− ε/2. Using Theorem IV.7, we then find p(ε, T ) ∈ [0, 1] such that

CAVPALL (p, T ) ≥ 1

kεχ(T⊗kε

)− ε

2 ≥ C(T )− ε,

for all 0 ≤ p ≤ p(ε, T ). This finishes the proof.

The main difficulty in the proof of Theorem IV.7 comes from the fact that the quantum channelunder AVP (see (23)) partially acts on the environment state σE , which, in general, is entangled overthe applications of the channel. Therefore, we are in a setting not covered by standard i.i.d. quantuminformation theory. We will apply the following strategy:

1. Find a coding scheme for classical or quantum communication for the quantum channel Tp fromLemma IV.6 at a fixed blocklength k ∈ N.

2. Use Lemma IV.6 to show that the coding scheme from 1. is a coding scheme under AVP of strengthp for the original quantum channel T .

Page 32: Fault-tolerant Coding for Quantum Communication

32

3. Apply a continuity inequality for the quantities χ and Icoh to relate the resulting capacity boundinvolving Tp to a similar bound involving the original channel T .

Note that step 1. in the previous strategy is straightforward using standard techniques from quantumShannon theory (i.e., random code constructions). To execute step 2. we need to know precise errorbounds in the coding theorems used for step 1., because these errors have to vanish quickly enough tocompensate the exponentially growing factor arising from Lemma IV.6. See Appendix C and AppendixD for a review of the explicit error bounds we are using in our proof. Step 3. is again straightforward.

Proof of Theorem IV.7. For p ∈ [0, 1), we will focus on the lower bound for the quantum capacityQAVP

ALL(p, T ) stated in the theorem. The lower bound on the effective classical capacity CAVPALL (p, T ) follows

along the same lines avoiding some additional technicalities as explained at the end of this proof.Consider the quantum channel Tp :MdA →MdB given by

Tp = (1− p)T + p1dBdB

Tr .

Fix k ∈ N such that the condition from the theorem is satisfied. Applying Corollary D.2 for the quantumchannel T⊗kp , some pure state |φAA′〉 ∈ CdkA ⊗ CdkA , some R > 0 and each m ∈ N shows the existence ofencoders Em :M⊗(Rkm−1)

2 →M⊗kmdAand decoders Dm :M⊗kmdB

→M⊗(Rkm−1)2 such that

F(Dm T⊗kmp Em

)≥ 1− εm

with

εm = 8√

3 exp(− mδ2

log (µmin (φA′))2

)+ 2 · 2−mk2 ( 1

k Icoh(φA′ ,T⊗kp )−R−3 δk ) (25)

and

µmin(φA′) = minλ > 0 : λ ∈ spec (φA′) ∪ spec(T⊗kp (φA′)

)∪ spec

((T⊗kp

)c (φA′)).

Here, we use the minimum fidelity (cf. [32])

F (S) = min〈ψ|S(|ψ〉〈ψ|)|ψ〉 : |ψ〉 ∈ Cd, 〈ψ|ψ〉 = 1,

to quantify the distance between a quantum channel S : Md → Md and the identity channel. Fix anysequence of quantum channels Nm : MdA ⊗MdE,m → MdB for any sequence of dimensions dE,m ∈ N.We will now use the coding scheme for the quantum channel Tp to send quantum information over thequantum channel T under AVP (cf., Definition IV.2) with the quantum channels Tp,Nm :MdA⊗MdE,m →MdB given by

Tp,Nm = (1− p)T ⊗ TrE +pNm.

For any syndrome state σE,m ∈ D((CdE,m)⊗m

)we may use the Fuchs-van-de-Graaf inequalities to

compute

‖id⊗(Rkm−1)2 −Dm T⊗kmp,Nm

(Em ⊗ σE,m)‖1→1 ≤ 2(

1− F(Dm T⊗kmp,Nm

(Em ⊗ σE,m)))1/2

≤ 2(

2km log2(dB)(p+δ) (1− F (Dm T⊗kmp Em))

+ exp(−mk δ

2

3p

))1/2

, (26)

where we used Lemma IV.6 in the final line, and where δ > 0 may be chosen as small as we need it tobe. With (25) we have

2km log2(dB)(p+δ) (1− F (Dm T⊗kmp Em))→ 0 (27)

as m→∞ when

2−m(

δ2 log2(e)log(µmin(φA′ ))2−k log2(dB)(p+δ)

)→ 0 (28)

Page 33: Fault-tolerant Coding for Quantum Communication

33

and

2−mk2 ( 1k Icoh(φA′ ,T⊗kp )−R−3 δk−2 log2(dB)(p+δ)) → 0 (29)

as m→∞. To guarantee (28) we choose

δ = (kp log2(dB))1/2| log (µmin (φA′)) |,

and δ > 0 sufficiently small. Now, (29) is satisfied for δ > 0 sufficiently small whenever

R <1kIcoh

(φA′ , T

⊗kp

)− 3k1/2 (p log2(dB))1/2| log (µmin (φA′)) | − 2p log2(dB). (30)

We have constructed a sequence of (mkR − 1,mk, εm) coding schemes under AVP of strength p as inDefinition IV.2 for any R satisfying (30) and some sequence of communication errors (εm)m∈N such thatεm → 0 as m→∞. Using [32, Lemma 7.1] it is easy to find for any m ∈ N a (mR,m, ε′m) coding schemesunder AVP of strength p such that ε′m → 0 as m→∞. This shows that any rate R satisfying (30) is anachievable rate under AVP of strength p as in Definition IV.2. We conclude that

QAVPALL(p, T ) ≥ 1

kIcoh

(φA′ , T

⊗kp

)− 3k1/2 (p log2(dB))1/2| log (µmin (φA′)) | − 2p log2(dB), (31)

for any pure state |φAA′〉 ∈ CdkA ⊗CdkA . To obtain an expression in terms of the channel T , we fix a purestate |ψAA′〉 ∈ CdkA ⊗ CdkA such that

1kIcoh

(ψA′ , T

⊗k) = max|φAA′ 〉

1kIcoh

(φA′ , T

⊗k) =: 1kIcoh

(T⊗k

).

For ε = pk, we consider a purification |ψεAA′〉 ∈ CdkA ⊗ CdkA of the state

ψεA′ = (1− 2ε)ψA′ + ε

(1dAdA

)⊗k+ ε|τA′〉〈τA′ |,

for some fixed pure state |τA′〉 ∈ CdkA , such that

‖ψAA′ − ψεAA′‖1 ≤ 4√ε(1− ε),

which is possible by Uhlmann’s theorem [52, Theorem 3.22] and the Fuchs-van-de-Graaf inequality [52,Theorem 3.33]. Using the definition of the coherent information and the Fannes-Audenaert inequality(see [52, Theorem 5.26]), we find that

|1kIcoh

(ψA′ , T

⊗kp

)− 1kIcoh

(ψεA′ , T

⊗kp

)| ≤ (2ε+ 4

√ε(1− ε)) log2(dA) + h2(2ε) + h2(2

√ε(1− ε))

=: η(ε), (32)

where we have used that 2√ε(1− ε) ≤ 1/2 by the assumption on k and p. Furthermore, we note that

µmin (ψεA′) ≥(

min(p

dA,p2

dB

))k, (33)

since

minλ > 0 : λ ∈ spec (ψεA′) ≥ε

dkA=(p

dA

)k,

minλ > 0 : λ ∈ spec(T⊗kp (ψεA′)

) ≥

(p

dB

)k,

minλ > 0 : λ ∈ spec((T⊗kp

)c (ψεA′)) ≥ ε

(p

dB

)k=(p2

dB

)k.

Page 34: Fault-tolerant Coding for Quantum Communication

34

For the first two estimates we used the special form of ψεA′ and Tp, and for the final estimate the factthat the non-zero spectrum of

(T⊗kp

)c (|τA′〉〈τA′ |) and(T⊗kp

)(|τA′〉〈τA′ |) coincides (see [30, Theorem 3]).

After inserting the state ψεA′ into (31) and using both (32) and (33) we find that

QAVPALL(p, T ) ≥ 1

kIcoh

(ψA′ , T

⊗kp

)− η(ε)− 3(kp log2(dB))1/2| log

(min

(p

dA,p2

dB

))| − 2p log2(dB). (34)

By the continuity bound [49, Proposition 3A] (see also [33]) we find that

1k

∣∣Icoh(ψA′ , T

⊗kp

)− Icoh

(T⊗k

) ∣∣ ≤ 2p log2(dB) + (1 + p)h2

(p

1 + p

).

Combining this bound with (34) leads to

QAVPALL(p, T )

≥ 1kIcoh

(T⊗k

)− η(ε)− 3(kp log2(dB))1/2| log

(min

(p

dA,p2

dB

))| − 4p log2(dB)− (1 + p)h2

(p

1 + p

)This finishes the proof.The lower bound on the classical capacity CAVP

ALL (p, T ) under AVP follows along the same lines as theproof above. To construct a coding scheme for the quantum channel Tp it is convenient to use the standardtechniques outlined in Appendix C. In particular, the error bound from Theorem C.2 replaces the errorbound from (25) in the previous proof. There are two simplifications compared to the quantum case: First,the classical communication error εcl already has the required monotonicity property under the partialorder ≤ appearing in Lemma IV.6 and we do not need to relate it to another distance measure. Second,both values λmin and µmin appearing in Theorem C.2 can be lower bounded by (p/dB)k for the specialchannel T⊗kp . Therefore, the analogue of (33) can be obtained much easier without introducing anotherstate like ψεA′ in the previous proof. Finally, we applied the continuity bound from [49, Proposition 3A](see also [33]) to make the final estimate.

V. FAULT-TOLERANT CAPACITIES

Definition II.10 of the classical capacity C(T ) of a cq-channel T assumes that the decoder can beapplied without faults. This assumption might not be realistic in practice, since the decoder necessarilyperforms a measurement of a potentially large quantum state. This reasoning applies also to the classicalcapacity C(T ) and the quantum capacity Q(T ) of a quantum channel T from Definition II.13 consideringcoding schemes even more involved. In this section, we will introduce fault-tolerant versions of theaforementioned capacities. Since our circuit model (including the noise model) is based on qubits, wewill state definitions of fault-tolerant coding schemes for cq-channels of the form T : A →M⊗j2 and forquantum channels of the form T :M⊗j1

2 →M⊗j22 , i.e., with input and output quantum systems composed

from qubits. However, these definitions also apply for general quantum channels by simply embeddingthem into a multi-qubit quantum channel. Our results are therefore stated for general channels.

A. Fault-tolerant capacity of a classical-quantum channel

To define the fault-tolerant capacity of a cq-channel, we will first define fault-tolerant coding schemestaking into account the faults occuring in quantum circuits executed by the receiver.

Definition V.1 (Fault-tolerant coding scheme for cq-channels). For p ∈ [0, 1] let Fπ(p) denote thei.i.d. Pauli noise model from Definition II.2. For n,m ∈ N and ε ∈ R+, an (n,m, ε)-fault tolerantcoding scheme for classical communication over the cq-channel T : A → M⊗j2 under the noise modelFπ(p) consists of a (classical) map E : 1, . . . , 2n → Am and a quantum circuit with classical output∆ :

(M⊗j2

)⊗m→ C2n such that the classical communication error (see (5)) satisfies

εcl

([∆]Fπ(p) T

⊗m E)≤ ε.

Page 35: Fault-tolerant Coding for Quantum Communication

35

Now, we can define the fault-tolerant capacity as follows:

Definition V.2 (Fault-tolerant capacity of a cq-channel). For a cq-channel T : A → Md, and thei.i.d. Pauli noise model Fπ(p) with p ∈ [0, 1] (see Definition II.2) we define the fault-tolerant classicalcapacity as

CFπ(p)(T ) = supR ≥ 0 fault-tolerantly achievable rate for T.

Here, R ≥ 0 is called an fault-tolerantly achievable rate if for every m ∈ N there exists an nm ∈ N andan (nm,m, εm)-fault-tolerant coding scheme for classical communication over the cq-channel T under thenoise model Fπ(p) such that εm → 0 as m→∞ and

R ≤ lim infm→∞

nmm.

Although Definition V.1 allows for arbitrary decoding circuits ∆, we will consider decoding circuits ofthe form ∆ = ΓDC Enc⊗jmC . Here, C is the concatenated 7-qubit Steane code at some concatentationlevel and EncC the interface circuit constructed in Section IIIA. The effective channel for this code andinterface (cf. Theorem III.9) will have a tensor product structure allowing the use of coding schemes forcompound channels [9, 17, 39]) to construct the decoding circuit ΓD. Using more advanced quantumerror correcting codes [21, 24] might lead to more complicated effective channels, but possibly to higherinformation transmission rates. We leave this for further investigation. We will show the following:

Theorem V.3 (Lower bound on the fault-tolerant capacity of a cq-channel). Let p0 denote the thresholdof the concatenated 7-qubit Steane code (see Lemma II.7) and c > 0 the constant from Theorem III.3.For any cq-channel T : A →Md and any p ≤ min(p0/2, (2cdlog2(d)e)−1) we have

CFπ(p)(T ) ≥ C(T )− 2cpdlog2(d)e2 − 2 (1 + 2cpdlog2(d)e)h2

(2cpdlog2(d)e

1 + 2cpdlog2(d)e

),

where h2 denotes the binary entropy.

The previous theorem directly implies the following threshold-type result.

Theorem V.4 (Threshold theorem for the fault-tolerant capacity of a cq-channel). For every ε > 0 andevery d ∈ N there exists a threshold p(ε, d) > 0 such that

CFπ(p)(T ) ≥ C(T )− ε,

for all cq-channels T : A →Md and all 0 ≤ p ≤ p(ε, d). In particular, we have

limp0

CFπ(p)(T ) = C(T ).

We will now prove the lower bound on the fault-tolerant capacity of a cq-channel stated above.

Proof of Theorem V.3. Without loss of generality we may assume that T : A →M⊗j2 with j = dlog2(d)e.Our fault-tolerant coding scheme will use the concatenated 7-qubit Steane code (see Appendix A) as aquantum circuit code. For each l ∈ N let Cl ⊂ (C2)⊗7l denote the lth level of the concatenated 7-qubitSteane code with threshold p0 ∈ (0, 1] (see Lemma II.7). Moreover, we denote by Encl : M2 → M⊗7l

2the interface circuit from (12), and recall Theorem III.3 introducing a constant c > 0.

We will start by constructing a fault-tolerant coding scheme as in Definition V.1. For any cq-channelN : A →M⊗j2 and p ∈

[0, (2cj)−1] we denote by Tp,N : A →M⊗j2 the cq-channel

Tp,N = (1− 2cpj)T + 2cpjN.

Next, we fix a probability distribution q ∈ P (A) and a rate

R < infNχ (qi, Tp,N (i)) , (35)

Page 36: Fault-tolerant Coding for Quantum Communication

36

with infimum running over cq-channels N : A → M⊗j2 . Applying [39, Theorem IV.18.] we obtain asequence nm ∈ N satisfying

R ≤ lim infm→∞

nmm,

and maps Em : 1, . . . , 2nm → Am and quantum-classical channels Dm :M⊗jm2 → C⊗2nm such that

supNεcl

(Dm T⊗mp,N Em

)→ 0 (36)

as m→∞ and with εcl as in (5). For each m ∈ N let ΓD,m :M⊗jm2 → C⊗2nm denote a quantum circuitsatisfying

‖ΓD,m −Dm‖1→1 ≤1m, (37)

and choose lm ∈ N such that

2c(p

p0

)2lm

|Loc(ΓD,m)| ≤ 1m. (38)

By the previous bound and the second case of Lemma III.8 (using Bernoulli’s inequality) we find that

‖[ΓD,mClm Enc⊗jmlm

]Fπ(p)

− ΓD,m (N lmp )⊗m‖1→1 ≤

1m,

where N lmp :M⊗j2 →M⊗j2 denotes a quantum channel of the form

N lmp = (1− 2cpj)id2 + 2cpjNlm ,

for a quantum channel Nlm :M⊗j2 →M⊗j2 depending on lm ∈ N. Using the particular form of the codingerror εcl from (5) and the estimate (37) leads to

εcl

([ΓD,mClm Enc⊗jmlm

]Fπ(p)

T⊗m Em)≤ εcl

(ΓD,m

(Tp,Nlm

)⊗m Em)+ 1m

≤ εcl(Dm

(Tp,Nlm

)⊗m Em)+ 2m

−→ 0 as m→∞, (39)

where we used (36) in the final line. We have shown that any rate R chosen as in (35) is fault-tolerantlyachievable in the sense of Definition V.2. We conclude that

CFπ(p)(T ) ≥ infNχ (qi, Tp,N (i)) , (40)

for any probability distribution q ∈ P (A). Finally, we use the continuity bound from [49, Proposition 5]to estimate

|χ (qi, Tp,N (i))− χ (qi, T (i)) | ≤ ε0j + 2(1 + ε0)h2

(ε0

1 + ε0

)≤ 2cpj2 + 2(1 + 2cpj)h2

(2cpj

1 + 2cpj

),

where we used that the function x 7→ (1 + x)h2

(x

1+x

)is monotonically increasing and that

ε0 := 12∑i

‖qiTp,N (i)− qiT (i)‖1 = cpj∑i

qi‖T (i)−N(i)‖1 ≤ 2cpj.

Combining this estimate with (40) we have

CFπ(p)(T ) ≥ χ (qi, T (i))− 2cpj2 − 2(1 + 2cpj)h2

(2cpj

1 + 2cpj

)for any probability distribution q ∈ P (A). Maximizing over q and using the Holevo-Schumacher-Westmoreland theorem (see Theorem II.11) finishes the proof.

Page 37: Fault-tolerant Coding for Quantum Communication

37

The previous proof used a coding scheme for a so-called compound channel [9, 17, 39] in (36). The sameproof would also work for a coding scheme that is constructed for the specific sequence of tensor powers(Tp,Nlm

)⊗m appearing in the first and second line of (39). However, due to the dependence of the localchannel on m through the concatenation level lm of the concatenated code, this is not the tensor powerof a fixed qubit channel. Standard techniques from i.i.d. quantum information theory might thereforenot apply in this setting, and similar constructions as for compound channels might be needed here ingeneral.

B. Fault-tolerant capacities of quantum channels

Next, we consider fault-tolerant capacities of quantum channels. We will focus on the classical capacityand the quantum capacity introduced in Section II E. We will begin by introducing fault-tolerant codingschemes for transmitting classical information over quantum channels.

Definition V.5 (Fault-tolerant coding schemes for classical communication). For p ∈ [0, 1] let Fπ(p)denote the i.i.d Pauli noise model from Definition II.2 and consider a quantum channel T : M⊗j1

2 →M⊗j2

2 . For n,m ∈ N and ε ∈ R+, an (n,m, ε) fault-tolerant coding scheme for classical communicationover T under the noise model Fπ(p) consists of a quantum circuit E : C2n →M⊗j1m

2 with classical inputand a quantum circuit ∆ :M⊗j2m

2 → C2n with classical output such that

εcl

([∆ T⊗m E

]Fπ(p)

)≤ ε,

where εcl denotes the classical communication error from (5).

To define fault-tolerant coding schemes for quantum communication, it will be important to choose asuitable way to measure the communication error. Note that the usual equivalences between differenterror measures commonly used to define the quantum capacity (see for instance the discussion in [32])might not hold in a setting where noise is affecting the coding operations. Our definition is motivated bypossible applications of fault-tolerant capacities in quantum computing, where a quantum computation isperformed on two separate quantum computers connected by a communication line. Intuitively, the fault-tolerant quantum capacity should measure the optimal rates at which quantum information can be sendwithin any quantum computation such that the overall quantum computation, i.e., the combination of thetwo parts of the executed circuit and the communication scheme, can be implemented fault-tolerantly. Fortechnical reasons, which will become clear in the discussion after the definition, we will need the followingconsequence of the Solovay-Kitaev theorem and the approximation of multi-qubit unitary operations byuniversal gate sets (see the discussion in [42]): For each m ∈ N and δ > 0, we denote by N(m, δ) ∈ N thesmallest number such that for any classical-quantum channel R : C2k → M⊗m2 there exists a quantumcircuit ΓR with |Loc (ΓR) | ≤ N(m, δ) and ‖ΓR−R‖1→1 ≤ δ, and such that the analogous approximationholds for any quantum-classical channel R′ : M⊗m2 → C2k as well. Note that N(m, δ) is monotonicallydecreasing in δ. Now, we can state our definition:

Definition V.6 (Fault-tolerant coding scheme for quantum communication). For p ∈ [0, 1] let Fπ(p)denote the i.i.d. Pauli noise model from Definition II.2 and consider a quantum channel T : M⊗j1

2 →M⊗j2

2 . For n,m ∈ N and ε > 0 an (n,m, ε)-fault-tolerant coding scheme for quantum communicationover T under the noise model Fπ(p) consists of a quantum error correcting code C (depending on n andm) of dimension dim (C) on KC qubits and quantum circuits

ΓE(C) :M⊗nKC2 →M⊗mj12 and ΓD(C) :M⊗mj2

2 →M⊗nKC2 ,

such that

‖Γ2 Γ1 −[Γ2C ΓD(C) T⊗m ΓE(C) Γ1

C]Fπ(p) ‖1→1 ≤ ε.

for any k1, k2 ∈ N and all quantum circuits Γ1 : C2k1 →M⊗n2 and Γ2 :M⊗n2 → C2k2 satisfying

max(|Loc

(Γ1) |, |Loc (Γ2) |) ≤ N(n, δ),

for some δ ≤ ε.

Page 38: Fault-tolerant Coding for Quantum Communication

38

We should emphasize that the previous definition can be stated in the same way with quantum errorcorrecting codes C of dimensions dim (C) > 2 (including the case where multiple qubits are encoded inthe same code). We have chosen to restrict to dimension dim (C) = 2 because we have only introducedimplementations of circuits in this case (see Definition II.3). While a more general definition is possibleit would be more involved and beyond the scope of this article. Note, however, that Definition V.5 offault-tolerant coding schemes for classical information is as general as possible.Having defined fault-tolerant coding schemes, we can define fault-tolerant capacities as usual:

Definition V.7 (Fault-tolerant classical and quantum capacity). For p ∈ [0, 1] let Fπ(p) denote thei.i.d. Pauli noise model from Definition II.2 and consider a quantum channel T : M⊗j1

2 → M⊗j22 . We

call R ≥ 0 an fault-tolerantly achievable rate for classical (or quantum) communication if for everym ∈ N there exists an nm ∈ N and an (nm,m, εm) fault-tolerant coding scheme for classical (or quantum)communication over T under the noise model Fπ(p) with εm → 0 as m→∞ and

lim infm→∞

nmm≥ R.

The fault-tolerant classical capacity of T is given by

CFπ(p)(T ) = supR ≥ 0 fault-tolerantly achievable rate for classical communication,

and the fault-tolerant quantum capacity of T is given by

QFπ(p)(T ) = supR ≥ 0 fault-tolerantly achievable rate for quantum communication.

It should also be emphasized that the noiseless capacity Q(T ) is recovered from the previous definitionwhen p = 0. Indeed, it is easy to see how fault-tolerant coding schemes under the noise model Fπ(0)can be constructed from coding scheme for quantum communication over the channel T (see DefinitionII.12) by using a trivial quantum error correcting code and quantum circuits approximating the codingoperations. This shows that Q(T ) ≤ QFπ(0)(T ). To see the remaining inequality, we choose for every purestate |ψ〉 ∈ (C2)⊗n a quantum circuit Γ1,ψ : C→M⊗n2 preparing |ψ〉〈ψ| up to an error ε in 1→ 1-normand a quantum circuit Γ2,ψ : M⊗n2 → C2 measuring the POVM |ψ〉〈ψ|,1⊗n2 − |ψ〉〈ψ| up to an errorε in 1 → 1-norm, and such that each of these circuits has less than N(n, ε) many locations (which ispossible by the Solovay-Kitaev theorem). Assume that there exists a (n,m, ε)-coding scheme consistingof an error correcting code C and quantum circuits ΓE(C) and ΓD(C) as in Definition V.6 such that

‖Γ2,ψ Γ1,ψ − Γ2,ψC ΓD(C) T⊗m ΓE(C) Γ1,ψ

C ‖1→1 ≤ ε.

Since there are no faults affecting the circuits, we may apply the transformation rules in Definition II.3to conclude that

‖Γ2,ψ Γ1,ψ − Γ2,ψ (id⊗n2 ⊗ TrS

)Dec∗ ΓD(C) T⊗m ΓE(C) Enc∗

(Γ1,ψ ⊗ |0〉〈0|

)‖1→1 ≤ ε,

for any pure state |ψ〉 ∈ (C2)⊗n, and where |0〉〈0| denotes the syndrome state corresponding to the zerosyndrome. Using the minimum fidelity

F (S) = min〈ψ|S(|ψ〉〈ψ|)|ψ〉 : |ψ〉 ∈ Cd1 , 〈ψ|ψ〉 = 1,

of a quantum channel S :Md1 →Md2 , we conclude that

F((id⊗n2 ⊗ TrS

)Dec∗ ΓD(C) T⊗m ΓE(C) Enc∗ (· ⊗ |0〉〈0|)

)≥ 1− 5ε.

By the equivalences of error criteria from [32], we conclude that the pair of quantum channels

D =(id⊗n2 ⊗ TrS

)Dec∗ ΓD(C),

and

E = ΓE(C) Enc∗ (· ⊗ |0〉〈0|) ,

defines an (n,m, 5ε)-coding scheme. After going through the definitions of the capacities, we find thatQ(T ) ≥ QFπ(0)(T ).

Page 39: Fault-tolerant Coding for Quantum Communication

39

The previous definitions of fault-tolerant capacities and their coding schemes allow for arbitrary quan-tum error correcting codes to be used by the sender and the receiver. As in Section VA, we will restrictto concatenated codes with the same concatenation level and the interface circuits constructed in SectionIIIA. As before, we should emphasize that more advanced quantum error correcting codes [21, 24] mightlead to more complicated effective channels, but possibly to higher information transmission rates. Weleave this for further investigation. By using concatenated codes, we can use the effective channel fromTheorem III.9 to show that any coding scheme for communication over T under AVP (see DefinitionIV.1) can be used to construct a fault-tolerant coding scheme. This leads to the following lower boundsin the same spirit as Theorem V.3.

Theorem V.8 (Lower bounds on fault-tolerant capacities). Let p0 denote the threshold of the concate-nated 7-qubit Steane code (see Lemma II.7) and c > 0 the constant from Theorem III.3. For any quantumchannel T :Md1 →Md2 and any

p ≤ min(p0/2, (2(j1 + j2)c)−1),

with j1 = dlog2(d1)e and j2 = dlog2(d2)e, we have

CFπ(p)(T ) ≥ CAVPALL (2(j1 + j2)cp, T )

and

QFπ(p)(T ) ≥ QAVPALL(2(j1 + j2)cp, T ).

An explicit lower bound can be obtained from Theorem IV.7.

Together with Theorem IV.8, the previous theorem implies the following threshold-type result.

Theorem V.9 (Threshold theorem for fault-tolerant capacities). For every quantum channel T :Md1 →Md2 and every ε > 0, there exists a threshold p(ε, T ) > 0 such that

CFπ(p)(T ) ≥ C(T )− ε,

and

QFπ(p)(T ) ≥ Q(T )− ε,

for all 0 ≤ p ≤ p(ε, T ). In particular, we have

limp0

CFπ(p)(T ) = C(T ) and limp0

QFπ(p)(T ) = Q(T ),

for all quantum channels T :Md1 →Md2 .

It should be emphasized that the threshold in Theorem V.9 does depend on the quantum channel Tand it is not uniform as in the case of cq-channels (cf. Theorem V.4). This issue is closely connected tothe fact that our bounds in Theorem IV.7 do not regularize (i.e., the limit k → ∞ leads to a vanishinglower bound for any p > 0). This issue might be resolved by either using a different fault-tolerant codingscheme, or by showing that the syndrome states σS occuring in Theorem III.9 are fully separable. In thelatter case, we could lower bound the fault-tolerant capacities by the capacities under AVP with separableenvironment states CSEP

ALL(2(j1 +j2)cp, T ) and CSEPALL(2(j1 +j2)cp, T ) (see Definition IV.3). Using Theorem

IV.5 (and a potential analogue of this theorem for the classical capacity) the lower bounds on the fault-tolerant capacities could then be improved and a threshold not depending on the particular channel (asin Theorem V.4) could be obtained.By restricting our statements to classes of quantum channels where the regularization in the capacity

formulas is not required, it is also possible to obtain uniform thresholds using our techniques. Forexample, there is a threshold p(ε, dA, dB) only depending on ε and the dimensions dA, dB such thatQFπ(p)(T ) ≥ Q(T ) − ε holds for every degrabable quantum channel T : MdA → MdB (see [16] for thedefinition and properties of this class of quantum channels) and for every 0 ≤ p ≤ p(ε, dA, dB).

Proof of Theorem V.8. We will focus on the lower bound for the fault-tolerant quantum capacityQFπ(p)(T ) stated in the theorem. The lower bound on the fault-tolerant classical capacity CFπ(p)(T )follows along the same lines.

Page 40: Fault-tolerant Coding for Quantum Communication

40

Without loss of generality we may assume that T :M⊗j12 →M⊗j2

2 with

j1 = dlog2(d1)e and j2 = dlog2(d1)e,

and we fix any rate R < QAVPALL(2(j1 + j2)cp, T ). Our fault-tolerant coding scheme (as in Definition

V.6) achieving this rate will use the concatenated 7-qubit Steane code (see Appendix A) as a quantumcircuit code. For each l ∈ N let Cl ⊂ (C2)⊗7l denote the lth level of the concatenated 7-qubit Steanecode with threshold p0 ∈ (0, 1] (see Lemma II.7). Moreover, we denote by Encl : M2 → M⊗7l

2 andDecl : M⊗7l

2 → M2 the interface circuits from (12) and (13), and recall Theorem III.3 introducingthe constant c > 0. Consider a sequence of (nm,m, εm)-coding schemes for quantum communicationover T under AVP of strength 2(j1 + j2)cp given by quantum channels Em : M⊗nm2 → M⊗j1m

2 andDm : M⊗j2m

2 → M⊗nm2 for each m ∈ N (as in Definition IV.2) such that R ≤ lim infm→∞ nm/m andlimm→∞ εm = 0.To construct a fault-tolerant coding scheme, we consider quantum circuits ΓE,m : M⊗nm2 → M⊗j1m

2and ΓD,m :M⊗j2m

2 →M⊗nm2 such that

‖ΓE,m − Em‖ ≤1m, (41)

and

‖ΓD,m −Dm‖ ≤1m, (42)

for every m ∈ N. Recall the numbers N(k, δ) ∈ N defined in the text before Definition V.6 specifyingthe number of locations needed in quantum circuits to approximate any classical-quantum and quantum-classical channels with output and input, respectively, consisting of k qubits up to an error δ > 0 in1→ 1-norm. After choosing lm ∈ N such that(

p

p0

)2lm (|Loc

(ΓE,m

)|+ |Loc

(ΓD,m

)|+ 2N(nm, εm) + j1m

)≤ 1m, (43)

for every m ∈ N, we define quantum circuits

ΓE,m (Clm) =(

Dec⊗j1lm

)⊗m ΓE,mClm ,

and

ΓD,m (Clm) = ΓD,mClm (

Enc⊗j2lm

)⊗m.

For any pair of quantum circuits Γ1,m : C2k1 →Mnm2 and Γ2,m :Mnm

2 → C2k2 satisfying

max(|Loc

(Γ1,m) |, |Loc (Γ2,m) |) ≤ N(nm, εm),

for any m ∈ N, we can compute

‖Γ2,m Γ1,m −[Γ2,mClm ΓD,m (Clm) T⊗m ΓE,m (Clm) Γ1,m

Clm

]Fπ(p)

‖1→1

≤ C

m+ ‖Γ2,m Γ1,m − Γ2,m ΓD,m T⊗mp,lm ΓE,m (Γ1,m ⊗ σS,m)‖1→1

≤ C + 2m

+ ‖Γ2,m Γ1,m − Γ2,m Dm T⊗mp,lm Em (Γ1,m ⊗ σS,m)‖1→1

≤ C + 2m

+ ‖id⊗Rkm2 −Dm T⊗mp,lm (Em ⊗ σS,m)‖ ≤C + 2m

+ εm. (44)

Here, we used Theorem III.9 together with (43) in the first inequality, (41) and (42) together with thetriangle inequality in the second inequality, and finally monotonicity of the 1→ 1-norm under quantumchannels in the third inequality. Here, σS,m is some quantum state on the syndrome space depending onΓ1,m and ΓE,m, and Tp,lm (· ⊗ σS,m) is the effective quantum channel introduced in Theorem III.9, whichis a special case of the channel T under AVP (see (23)) and it is therefore corrected up to an error εm

Page 41: Fault-tolerant Coding for Quantum Communication

41

by the coding scheme defined by Em and Dm. This implies the final inequality. We have shown thateach R < QAVP

ALL(2(j1 + j2)cp, T ) is a fault-tolerantly achievable rate for quantum communication over thequantum channel T under the noise model Fπ(p) and we conclude thatQFπ(p)(T ) ≥ QAVP

ALL(2(j1+j2)cp, T ).The lower bound on the fault-tolerant classical capacity CFπ(p)(T ) follows along the same lines as

the previous proof by using a coding scheme for classical communication over T under AVP of strength2(j1 + j2)cp. Again, we can approximate the coding maps Em and Dm by quantum circuits and use theconcatenated 7-qubit Steane code at a sufficiently high level to protect these circuits from noise.

C. Specific coding schemes from asymptotically good codes

In this section, we will show how to construct fault-tolerant coding schemes for certain quantumchannels from asymptotically good codes. For our purposes it will be sufficient to consider such codesbetween systems where the dimension is a power of two.Definition V.10 (Asymptotically good codes). Let d = 2j be a power of two. An asymptotically goodcode of rate R > 0 and goodness α ∈ (0, 1) is given by a sequence ((Em, Dm))m∈N of encoding operationsEm : M⊗nm2 → M⊗md and decoding operations Dm : M⊗md → M⊗nm2 such that there is a sequence(tm)m∈N ∈ NN satisfying the following:

1. lim infm→∞ nmm > R.

2. lim infm→∞ tmm > α.

3. For any quantum channel N :M⊗md →M⊗md acting non-trivially on only tm qudits, we have

Dm N Em = id⊗nm2 .

Asymptotically good codes were first constructed for d = 2 by Calderbank and Shor in [13]. For

α0 := minα ∈ [0, 1] : h2(2α) = 12 (45)

and any goodness α ∈ (0, α0) these codes achieve a rate

R(α) := 1− 2h2(2α),

where h2 denotes the binary entropy h2(p) = −(1− p) log2(1− p)− p log2(p). By now, many families ofasymptotically good codes are known [6, 15, 34, 37]. We will show the following theorem:Theorem V.11 (Fault-tolerant coding schemes from asymptotically good codes). Let d = 2j be apower of 2 and let p0 denote the threshold of the concatenated 7-qubit Steane code (see Lemma II.7).For a sequence (nm)m∈N consider quantum operations Em :M⊗nm2 →M⊗md and Dm :M⊗md →M⊗nm2defining an asymptotically good code of rate R > 0 and goodness α ∈ (0, 1). Moreover, consider p, q ∈ [0, 1]such that

p < min(p0/2, c−1) and x := 4jcp+ q < α, (46)

where c > 0 is the constant from Theorem III.3. For any δ > 0 with x+ δ < α there is an m0 ∈ N suchthat for any m ≥ m0 there exists an (ε′m, nm,m) fault-tolerant coding scheme for quantum communicationunder the noise model Fπ(p) via the quantum channel

Iq = (1− q)idd + qT,

with any fixed quantum channel T :Md →Md. Here, we have

ε′m = εm + 3 exp(−mδ2

3x

),

for any sequence εm > 0 with limm→∞ εm = 0. In particular, the rate R is fault-tolerantly achievable andwe have

QFπ(p)(Iq) ≥ R.

Page 42: Fault-tolerant Coding for Quantum Communication

42

Proof. For each l ∈ N let Cl ⊂ (C2)⊗7l denote the lth level of the concatenated 7-qubit Steane code withthreshold p0 ∈ (0, 1] (see Lemma II.7). Moreover, we denote by Encl :M2 →M⊗7l

2 and Decl :M⊗7l2 →

M2 the interface circuits from (12) and (13), and recall Theorem III.3 introducing a constant c > 0. For0 ≤ p < min(p0/2, c−1) and q ∈ [0, 1] we choose some δ > 0 such that x+ δ < α, where x < α was definedin (46). Finally, note again that d = 2j throughout the proof.Consider any sequence (εm)m∈N such that εm > 0 for all m ∈ N and limm→∞ εm = 0. To construct a

fault-tolerant coding scheme, we consider quantum circuits ΓEm :M⊗nm2 →M⊗md and ΓDm :M⊗md →M⊗nm2 such that

max(‖ΓEm − Em‖, ‖ΓDm −Dm‖

)≤ εm. (47)

For each m ∈ N we choose the code Clm with lm ∈ N large enough such that

C

(p

p0

)2lm (|Loc(ΓDm)|+ |Loc(ΓEm)|+ 2N(nm, εm) +m

)≤ exp

(−mδ2

3x

), (48)

where C > 0 is the constant from Theorem III.9. Next, we define quantum circuits

ΓE,m (Clm) =(

Dec⊗j1lm

)⊗m ΓEmClm ,

and

ΓD,m (Clm) = ΓDmClm (

Enc⊗j2lm

)⊗m,

for each m ∈ N. Now, consider a pair of quantum circuits Γ1 : C2k1 →M⊗nm2 and Γ2 :M⊗nm2 → C2k2

for some k1, k2 ∈ N such that

max(|Loc(Γ1)|, |Loc(Γ2)|

)≤ N(nm, εm).

By Theorem III.9 and (48) we have∥∥∥ [Γ2Clm ΓD,m (Clm) I⊗mq ΓE,m (Clm) Γ1

Clm

]Fπ(p)

− Γ2 ΓDm I⊗mx,lm (ΓEm Γ1 ⊗ σS)∥∥∥

1→1

≤ exp(−mδ2

3x

), (49)

for every m ∈ N, where the quantum channel Ix,lm :Md ⊗M⊗(7lm−1)2 →Md is of the form

Ix,lm = (1− x)idd ⊗ TrS +xNlm

for some quantum channel Nlm : M2 ⊗M⊗(7lm−1)2 → M2 acting on a data qubit and the syndrome

space, and x < α as in (46). For any m ∈ N and any syndrome state σS ∈(M⊗(7lm−1)

2

)⊗mwe denote

by I⊗mx,lm (· ⊗ σS) :Md →Md the quantum channel given by

Md 3 X 7→ I⊗mx,lm (X ⊗ σS) ∈Md,

where each Ix,lm acts partially on some part of the state σS . Now, we compute

I⊗mx,lm (· ⊗ σS) =∑

i1,...,im∈0,1

(1− x)m−∑m

j=1ijx

∑m

j=1ij

⊗j|ij=0

id2

⊗ N (i1...,im)lm

=b(x+δ)mc∑

s=0(1− x)m−sxs

∑i1+i2+···im=s

⊗j|ij=0

id2

⊗ N (i1...,im)lm

+ P(

1m

m∑i=1

Xi > x+ δ

)N, (50)

Page 43: Fault-tolerant Coding for Quantum Communication

43

where N (i1...,im)lm

:M⊗∑

jij

2 →M⊗∑

jij

2 denotes a quantum channel acting on the tensor factors corre-sponding to the j for which ij = 1 (constructed from tensor powers of Nlm acting partially on the stateσS), and N : M⊗m2 → M⊗m2 denotes a quantum channel collecting the second part of the sum. Fur-thermore, we introduced independent and identically distributed 0, 1-valued random variables Xi withP (X1 = 1) = x. Finally, we can use that ΓEm :M⊗nm2 →M⊗m2 and ΓDm :Mm

2 →M⊗nm2 approximate

the coding operations of an asymptotically good code with goodness α. Let m0 ∈ N be large enough suchthat b(x+ δ)mc ≤ αm ≤ tm for any m ≥ m0, where (tm)m∈N denotes the sequence as in Definition V.10for the asymptotically good code given by (Em, Dm). Using first the triangle inequality with (47) andthen property 3. from Definition V.10 together with (50) we find

‖idm2 − ΓDm I⊗mx,lm (ΓEm ⊗ σS)‖ ≤ ‖idm2 −Dm I⊗mx,lm (Em ⊗ σS)‖ + 2εm

≤ 2P(

1m

m∑i=1

Xi > x+ δ

)+ 2εm

≤ 2exp(−mδ2

3x

)+ 2εm, (51)

where the final estimate is the Chernoff bound. Combining (51) with (49) using the triangle inequality,we find

‖Γ2 Γ1 −[Γ2Clm ΓD,m (Clm) I⊗mq ΓE,m (Clm) Γ1

Clm

]Fπ(p)

‖1→1

≤ 3exp(−mδ2

3x

)+ 2εm.

Therefore, we find that the codes Clm and the pairs (ΓE,m (Clm) ,ΓD,m (Clm)) define a sequence of(nm,m, ε′m) fault-tolerant coding schemes as in Definition V.6.

Using the good codes constructed by Calderbank and Shor in [13] we obtain the following corollary:Corollary V.12 (Lower bound from good codes). Let p0 denote the threshold of the concatenated 7-qubitSteane code (see Lemma II.7), c > 0 the constant from Theorem III.3, and α0 the constant from (45).For p, q ∈ [0, 1] such that p ≤ min(p0/2, c−1) and 4cp+ q ≤ α0 we have

QFπ(p) ((1− q)id2 + qT ) ≥ 1− 2h2 (8cp+ 2q) ,for any quantum channel T :M2 →M2.

VI. CONCLUSION AND OPEN PROBLEMS

By combining techniques from fault-tolerant quantum computation and quantum Shannon theory, wehave initiated the study of fault-tolerant quantum Shannon theory. We introduced fault-tolerant capac-ities for classical and quantum communication via quantum channels, and for classical communicationvia classical-quantum channels. These capacities take into account that the encoding and decoding op-erations in the usual definitions of capacities are inherently affected by noise. We proved threshold-typetheorems for the fault-tolerant capacities showing that rates ε-close to the usual capacities can be obtainfor non-vanishing gate error probabilities below some threshold value depending on ε and the communi-cation channel. In the case of classical-quantum channels T : A → Md the threshold only depends onε and the output dimension d. We leave open the question whether such “uniform” threshold theoremsalso hold for the classical and quantum capacity of a quantum channel.Although we have focused on capacities and optimal achievable communication rates our results also

apply for specific codes. As an example we considered fault-tolerant quantum communication schemesbased on asymptotically good codes and via quantum channels of a specific form. Similar to the thresholdtheorem from [5], protecting a specific coding scheme against Pauli i.i.d. noise (of strength below thethreshold) requires only a polylogarithmic overhead in the size of the quantum circuit implementing it.It will then yield a fault-tolerant coding scheme if the ideal coding scheme corrects the errors introducedby the effective quantum channel induced by the interfaces (cf. Theorem III.9).In future research it would be interesting to extend our results to other communication scenarios, such as

private communication, quantum communication assisted by classical communication, or communicationscenarios between multiple parties. Finally, it would also be interesting to study the effects of differentcircuit noise models on the corresponding fault-tolerant capacities.

Page 44: Fault-tolerant Coding for Quantum Communication

44

VII. ACKNOWLEDGEMENTS

We thank Gorjan Alagic and Hector Bombin for their contributions during early stages of this project.Moreover, we thank Paula Belzig, Omar Fawzi and Milan Mosonyi for helpful comments on an earlier ver-sion of this article. MC acknowledges financial support from the European Research Council (ERC GrantAgreement No. 81876), VILLUM FONDEN via the QMATH Centre of Excellence (Grant No.10059) andthe QuantERA ERA-NET Cofund in Quantum Technologies implemented within the European Union’sHorizon 2020 Programme (QuantAlgo project) via the Innovation Fund Denmark. AMH acknowledgesfunding from the European Union’s Horizon 2020 research and innovation programme under the MarieSkłodowska-Curie Action TIPTOP (grant no. 843414).

Appendix A: Concatenated quantum error correcting codes

In this appendix we review basic facts about concatenated quantum codes [31].

1. 7-qubit Steane code

Let V : C2 → (C2)⊗7 denote the encoding isometry for the 7-qubit Steane code [51] given by

|0〉 = V |0〉 = 18(|0000000〉+ |0001111〉+ |0110011〉+ |1010101〉

+ |0111100〉+ |1011010〉+ |1100110〉+ |1101001〉)

and

|1〉 = V |1〉 = 18(|1111111〉+ |1110000〉+ |1001100〉+ |0101010〉

+ |1000011〉+ |0100101〉+ |0011001〉+ |0010110〉).

This code is a stabilizer code with the following stabilizer generators:

g1 = 1⊗ 1⊗ 1⊗X ⊗X ⊗X ⊗Xg2 = 1⊗X ⊗X ⊗ 1⊗ 1⊗X ⊗Xg3 = X ⊗ 1⊗X ⊗ 1⊗X ⊗ 1⊗Xg4 = 1⊗ 1⊗ 1⊗ Z ⊗ Z ⊗ Z ⊗ Zg5 = 1⊗ Z ⊗ Z ⊗ 1⊗ 1⊗ Z ⊗ Zg6 = Z ⊗ 1⊗ Z ⊗ 1⊗ Z ⊗ 1⊗ Z.

As a stabilizer code, the codewords |0〉 and |1〉 arise as the common eigenvectors for the eigenvalue +1 ofthe commuting set of Hermitian involutions g1, g2, . . . , g6. As explained in Section IIC, we can define asubspace Ws ⊂ (C2)⊗7 for each syndrome s ∈ F6

2 as the space of common eigenvectors for the eigenvalues(−1)si with respect to each gi. Then, we find that(

C2)⊗7 =⊕s∈F6

2

Ws, (A1)

and we can introduce the error basis ⋃s∈F6

2

Es|0〉, Es|1〉

of(C2)⊗7 with Pauli operators Es associated to the syndrome s ∈ F6

2 and such that

Ws = spanEs|0〉, Es|1〉. (A2)

Finally, we define the unitary map D : (C2)⊗7 → C2 ⊗ (C2)⊗6 by

D(Es|i〉

)= |i〉 ⊗ |s〉,

Page 45: Fault-tolerant Coding for Quantum Communication

45

extended linearily giving rise to the ideal decoder Dec∗ :M⊗72 →M2 ⊗M⊗6

2 via

Dec∗ = AdD,

and its inverse, the ideal encoder, Enc∗ :M2 ⊗M⊗62 →M⊗7

2 via

Enc∗ = AdD† .

For more details on quantum error correcting codes and the stabilizer formalism see [23].

2. Code concatenation

The 7-qubit Steane code can be used to construct a concatenated code [31] achieving higher protectionagainst noise. To define this concatenated code we recursively define the encoding isometry V (l) : C2 →C7l encoding a single (logical) qubit into 7l (physical) qubits via

V (1) = V and V (l) = V ⊗7l−1 V (l−1) for any l ≥ 2.

Note that in this way we have

V (l) = V ⊗7l−1 V ⊗7l−2

· · · V ⊗7 V,

and regrouping of this equation leads to the identity

V ⊗7l−1 V (l−1) = (V (l−1))⊗7 V. (A3)

Recall that the syndrome space of the 7-qubit Steane code consist of 6 qubits. By the previous discussion,we see that the syndrome space of the kth level of the concatenated 7-qubit Steane code consists of

6l−1∑i=0

7i = 7l − 1

qubits. This is not surprising since we encode a single qubit into 7l qubits using a code satisfying (A1)and (A2). Next, we recursively define the ideal operations

Enc∗l :M2 ⊗M⊗(7l−1)2 →M⊗7l

2 and Dec∗l :M⊗7l2 →M2 ⊗M⊗(7l−1)

2 ,

for every level l ∈ N by

Enc∗1 = Enc∗, and Enc∗l = [Enc∗1]⊗7l−1[Enc∗l−1⊗id

⊗(7l−16)2

]for any l ≥ 2, (A4)

and

Dec∗1 = Dec∗, and Dec∗l =[Dec∗1⊗id

⊗7(7l−1−1)2

][Dec∗l−1

]⊗7 for any l ≥ 2, (A5)

where we reordered the tensor factors such that ideal operations executed later in the circuit do not acton the syndrome output of operations executed earlier in the circuit. Again we can expand the previousrecursions and verify by regrouping that

[Enc∗1]⊗7l−1[Enc∗l−1⊗id⊗7l−16

2

]=[Enc∗l−1

]⊗7 [Enc∗1⊗id

⊗7(7l−1−1)2

], (A6)

and [Dec∗1⊗id

⊗7(7l−1−1)2

][Dec∗l−1

]⊗7 =[Dec∗l−1⊗id⊗7l−16

2

] [Dec∗1]⊗7l−1

, (A7)

where we again ordered the tensor factors appropriately.

Page 46: Fault-tolerant Coding for Quantum Communication

46

H

Ui

i

EC0

0

ψ

FIG. 5: The circuit Enc0→1 encoding the unknown state |ψ〉 into the first level of the 7-qubit Steane code. Notethat the logical CNOT gate is implemented in this code by applying elementary CNOT gates to each physicalqubit. Here, Ui denotes a certain Pauli gate (implemented in the 7-qubit Steane code) depending on the outcomei ∈ 0, 1, 2, 3 of the Bell measurement, and EC denotes the error correction of the code.

3. Explicit interface from level 0 to level 1

In this section, we will explicitly construct an interface for the first level of the 7-qubit Steane codeusing a simple teleportation circuit. We should emphasize that this construction is certainly well-knownand it works also for more general quantum error correcting codes. We state it here for convenience, sothat the constructions in Section IIIA can be made explicit.Let ω(0,1) ∈

(M2 ⊗M⊗7

2)+ denote a maximally entangled state between a physical qubit and the

7-qubit Steane code (i.e., the first level of the concatenated code). Specifically, we define ω(0,1) :=|Ω(0,1)〉〈Ω(0,1)| for

|Ω(0,1)〉 = 1√2(|0〉 ⊗ |01〉+ |1〉 ⊗ |11〉

), (A8)

where |i1〉1i=0 denotes the computational basis in the 7-qubit Steane code. Note that ω(0,1) can beprepared using elementary gates (Hadamard and CNOT gates) only, and we will denote this preparationcircuit by Prep

(ω(0,1)

). Now, we define

Enc0→1 (·) := Λ1(· ⊗ Prep

(ω(0,1)

)). (A9)

Here, Λ1 : M2 ⊗M2 ⊗M⊗72 → M⊗7

2 denotes the teleportation protocol, i.e., measuring the two firstregisters in the Bell basis

|φ1〉 = |Ω2〉,|φ2〉 = (12 ⊗ σz)|Ω2〉,|φ3〉 = (12 ⊗ σx)|Ω2〉,|φ4〉 = (12 ⊗ σxσz)|Ω2〉,

and then depending on the measurement outcome i ∈ 1, 2, 3, 4 performing the 1-rectangle correspondingto the unitary gate Ui ∈ U2 on the 7-qubit Steane code space, where

U1 = 12, U2 = σz, U3 = σx, U4 = σxσz.

Again, we note that Λ1 is a quantum circuit ending in an error correction. See Figure 5 for a circuitdiagram of the quantum circuit Enc0→1.

Page 47: Fault-tolerant Coding for Quantum Communication

47

H Ui

i

0

0

ψ

Prep(ω )(0,1) Λ 2

FIG. 6: The circuit Dec0→1 decoding the unknown state |ψ〉 from the first level of the 7-qubit Steane code to aphysical qubit. Here, Ui denotes a certain Pauli gate (applied to the physical qubit) depending on the outcomei ∈ 0, 1, 2, 3 of the Bell measurement.

Similarily, we denote by Λ2 :M2 ⊗M⊗72 ⊗M

⊗72 →M2 the teleportation protocol measuring the two

final registers in the Bell basis (implemented in the 7-qubit Steane code) and performing the unitary gatestated above depending on the measurement outcome on the remaining qubit system. Then, we define

Dec1→0 (·) := Λ2(Prep

(ω(0,1)

)⊗ ·). (A10)

Again, this is a quantum circuit, and the circuit diagram can be seen in Figure 6.

Appendix B: Chasing constants in weak typicality

To prove our main results, we need precise error bounds in the direct parts of the classical and quantumcapacity theorems (cf. Theorem II.14 and Theorem II.15). Such bounds can be obtained by chasing theconstants appearing in the proofs of these results, and by identifying the precise dependence of the finalerror on the number of channel uses. Most proofs of Theorem II.14 and Theorem II.15 in the literature(see e.g., [18, 25, 52, 53]) are based on different notions of typicality. In this appendix we summarizethe neccessary results and obtain explicit error bounds needed to chase the constants in the capacitytheorems. We should emphasize that the following results are well-known, and we merely extracted themfrom the literature (in particular from [53]) making some bounds explicit.For each n ∈ N let Xn = (X1, . . . , Xn) be an n-tuple of i.i.d. random variables Xi ∼ X with values in

a finite set A. We denote by supp(X) = x ∈ A : pX(x) 6= 0 and the Shannon entropy by

H(X) = −∑x∈A

pX(x) log(pX(x)).

The set of δ-typical sequences of length n ∈ N is given by

TXn

δ := xn ∈ An |∣∣∣− 1

nlog (pXn(xn))−H(X)

∣∣∣ ≤ δ,where pXn(xn) = Πn

i=1pX(xi). We first note the following simple property that follows immediately fromthe definition:

Page 48: Fault-tolerant Coding for Quantum Communication

48

Theorem B.1. For any δ > 0 we have

pXn(xn) ≤ 2−n(H(X)−δ),

for any xn ∈ TXnδ .

The following theorem shows that with high probability the random variable Xn takes values in thetypical set TXnδ .

Theorem B.2. For pmin = minx∈supp(X) pX (x) and any δ > 0 we have

P(Xn ∈ TX

n

δ

)≥ 1− exp

(−2nδ2

log(pmin)2

).

Proof. For each x ∈ A we define an i.i.d. sequence of indicator random variables as

Ix(Xi) =

1, if Xi = x

0, else.

Now consider the n-tuples of i.i.d. random variables Zn = (Z1, . . . , Zn) defined as

Zi = −∑

x∈supp(X)

Ix (Xi) log(pX(x)).

Note that E(Zi) = H(X) and Zi ∈ [0,− log (pmin)] almost surely. By Hoeffding’s inequality we have

P

(∣∣∣∣∣ 1nn∑i=1

Zi −H(X)∣∣∣∣∣ > δ

)≤ exp

(−2nδ2

log(pmin)2

).

Now note that xn ∈ An with pXn(xn) > 0 satisfies xn ∈ TXnδ if and only if

− 1n

log (pXn (xn)) = − 1n

n∑i=1

log (pX(xi))

= − 1n

n∑i=1

∑x∈supp(X)

Ix(xi) log (pX(x)) ∈ [H(X)− δ,H(X) + δ] .

Therefore, we have

P(Xn ∈ TX

n

δ

)= 1− P

(Xn /∈ TX

n

δ

)= 1− P

− 1n

n∑i=1

∑x∈supp(X)

Ix(Xi) log (pX(x)) /∈ [H(X)− δ,H(X) + δ]

= 1− P

(∣∣∣∣∣ 1nn∑i=1

Zi −H(X)∣∣∣∣∣ > δ

)

≥ 1− exp(−2nδ2

log(pmin)2

).

For each n ∈ N let (X,Y )n = ((X1, Y1), . . . , (Xn, Yn)) be an n-tuple of i.i.d. pairs of random variables(Xi, Yi) ∼ (X,Y ) with values in the finite product set A × B. We define supp (X,Y ) = (x, y) ∈A× B : pX,Y (x, y) 6= 0 and the joint Shannon entropy by

H(X,Y ) = −∑

(x,y)∈A×B

pX,Y (x, y) log(p(X,Y )(x, y)).

Page 49: Fault-tolerant Coding for Quantum Communication

49

The set of conditional δ-typical sequences of length n ∈ N conditioned onto a sequence xn ∈ An satisfyingpXn (xn) > 0 is given by

TY n|xnδ := yn ∈ Bn |

∣∣∣− 1n

log(pXn,Y n(xn, yn)

pXn (xn)

)−H(X,Y ) +H(X)

∣∣∣ ≤ δ.Here, pX denotes the marginal probability distribution pX(x) =

∑y∈B pX,Y (x, y).

Theorem B.3. For rmin = min(x,y)∈supp(X,Y )

(pX,Y (x,y)pX(x)

)and any δ > 0 we have

EXnPY n|Xn(Y n ∈ TY

n|Xnδ

)≥ 1− exp

(−2nδ2

log(rmin)2

).

Proof. For each pair (x, y) ∈ A× B define the sequence of indicator random variables

Ix,y (Xi, Yi) =

1, if (Xi, Yi) = (x, y)0, else.

Now define n-tuples of i.i.d. random variables Zn = (Z1, . . . , Zn) given by

Zi = −∑

(x,y)∈supp(X,Y )

I(x,y) (Xi, Yi) [log (pX,Y (x, y))− log(pX (x))] .

Note that E (Zi) = H(X,Y ) −H(X) and Zi ∈ [0,− log(rmin)] almost surely. By Hoeffding’s inequalitywe have

P

(∣∣∣ 1n

n∑i=1

Zi −H(X,Y ) +H(X)∣∣∣ > δ

)≤ exp

(−2nδ2

log(rmin)2

).

Note that for a given n-tuple xn ∈ An satisfying pXn(xn) > 0 and yn ∈ Bn satisfying pXn,Y n(xn, yn) > 0we have yn ∈ TY

n|xnδ if and only if

− 1n

log(pXn,Y n (xn, yn)

pXn (xn)

)= − 1

n

n∑i=1

[log(pX,Y (xi, yi))− log(pX(xi))]

= − 1n

n∑i=1

∑(x,y)∈supp(X,Y )

I(x,y)(xi, yi) [log(pX,Y (x, y))− log(pX(x))]

∈ [H(X,Y )−H(X)− δ,H(X,Y )−H(X) + δ] =: Iδ.

Using the law of total probability, we obtain

EXnPY n|Xn(Y n ∈ TY

n|Xnδ

)=

∑xn∈supp(Xn)

PXn(xn)PY n|Xn(Y n ∈ TY

n|Xnδ

)= P

(Y n ∈ TY

n|Xnδ

)= 1− P

(Y n /∈ TY

n|Xnδ

)= 1− P

− 1n

n∑i=1

∑(x,y)∈supp(X,Y )

I(x,y)(Xi, Yi) [log(pX,Y (x, y))− log(pX(x))] /∈ Iδ

= 1− P

(∣∣∣ 1n

n∑i=1

Zi −H(X,Y ) +H(X)∣∣∣ > δ

)≥ 1− exp

(−2nδ2

log(rmin)2

).

Page 50: Fault-tolerant Coding for Quantum Communication

50

Theorem B.4 (Size bound for conditional typical subset).For every n-tuple xn ∈ An and any δ > 0 we have∣∣∣TY |xnδ

∣∣∣ ≤ 2n(H(X,Y )−H(X)+δ)

Proof. Without loss of generality we can assume that xi ∈ supp(X) since otherwise TY |xn

δ = ∅. We knowthat for each yn ∈ TY |x

n

δ we have

pY n|xn (yn) =n∏i=1

pY |xi (yi) =n∏i=1

pX,Y (xi, yi)pX(xi)

≥ 2−n(H(X,Y )−H(X)+δ).

Therefore, we have

1 ≥∑

yn∈TY |xn

δ

pY n|xn (yn) ≥∣∣∣TY |xnδ

∣∣∣2−n(H(X,Y )−H(X)+δ).

This finishes the proof.

Now, we will use the previous theorems from this section to introduce quantum typicality. Let ρ =∑x∈A pX(x)|x〉〈x| denote a quantum state, where we introduced a random variable X with values in A

distributed according to the spectrum of ρ. Note that H(X) = S(ρ), i.e., the von-Neumann entropy ofthe quantum state ρ. For δ > 0 we define the δ-typical projector with respect to ρ⊗n as

Πnδ =

∑xn∈TXn

δ

|xn〉〈xn|.

The following theorem follows easily from Theorem B.1

Theorem B.5. For any δ > 0 and n ∈ N we have

Πnδ ρ⊗nΠn

δ ≤ 2−n(S(ρ)−δ)Πnδ .

From Theorem B.2 we easily get the following theorem:

Theorem B.6 (Quantum typicality). For λ∗min(ρ) := minλ ∈ spec(ρ) \ 0 and any δ > 0 we have

Tr(Πnδ ρ⊗n) ≥ 1− exp

(−2nδ2

log(λ∗min(ρ))2

).

Now let pX(x), ρxx∈A denote an ensemble of quantum states, and for each x ∈ A we have theeigendecomposition ρx =

∑y∈B pY |X(y|x)|yx〉〈yx| defining the random variable Y . For δ > 0 and an n-

tuple xn ∈ An we define the conditional δ-typical projector of the ensemble pX(x), ρxx∈A conditionedon xn by

ΠBn|xnδ =

∑ynxn∈TY

n|xnδ

|ynxn〉〈ynxn |.

Note that [ΠBn|xnδ , ρxn

]= 0,

and by Theorem B.3 we have the following result:

Theorem B.7 (Conditional quantum typicality). For µ∗min = minx∈supp(X) minλ ∈ spec (ρx)\0 andany δ > 0 we have ∑

xn∈AnpXn(xn) Tr

(ΠBn|xnδ ρxn

)≥ 1− exp

(−2nδ2

log(µ∗min)2

).

Page 51: Fault-tolerant Coding for Quantum Communication

51

Finally, by Theorem B.4 we have the following theorem:

Theorem B.8. For any δ > 0 we have

Tr(

ΠBn|xnδ

)≤ 2n(H(X,Y )−H(X)+δ).

Appendix C: Explicit error bound in the HSW-theorem

For the proof of Theorem V.8 we need explicit error bounds in the direct part of the proof of theHSW-theorem (cf. Theorem II.14). To make our article selfcontained we derive these bounds in thisappendix. We will start with the following lemma, which is a version of the well-known packing lemma(see for instance [53, Lemma 16.3.1]). Its proof combines the general strategy used in [53, Lemma 16.3.1]with some insights from [52, Section 8.1.2] leading to a slightly better error estimate.

Lemma C.1 (Packing lemma). Let pi, σiLi=1 be an ensemble of quantum states σi ∈M+d with ensemble

average σ =∑Li=1 piσi. Let Π : Cd → Cd and Πi : Cd → Cd for each i ∈ 1, . . . , L denote projectors

such that the following conditions hold for ε1, ε2 > 0 and A,B ∈ N:

1. Tr [Πσ] ≥ 1− ε1.

2.∑Li=1 pi Tr [Πiσi] ≥ 1− ε2.

3. [Πi, σi] = 0 for any i ∈ 1, . . . , L.

4. Tr [Πi] ≤ A for any i ∈ 1, . . . , L.

5. ΠσΠ ≤ 1BΠ.

Then, for any M ≤ L there exists an M -tuple I = (i1, . . . , iM ) ∈ 1, . . . , LM , and a POVM (Λs)Ms=1 ∈(M+

d )M such that

1M

M∑s=1

Tr (Λsσis) ≥ 1− 4ε1 − 2ε2 − 4M A

B.

Proof. Let M ≤ L be fixed in the following. For each i ∈ 1, . . . , L define the operators

Yi = ΠΠiΠ.

For any M -tuple I = (i1, . . . , iM ) ∈ 1, . . . , LM we define a POVM (Λs)Ms=1 by

Λs =(

M∑s′=1

Yis′

)− 12

Yis

(M∑s′=1

Yis′

)− 12

, for s ∈ 1, . . . ,M,

and the decoding error of symbol s ∈ 1, . . . ,M by

pe(s, I) = Tr ((1d − Λs)σis) .

Recall the Hayashi-Nagaoka inequality (see [52, Lemma 8.28])

1d − (P +Q)−12 P (P +Q)−

12 ≤ 2 (1d − P ) + 4Q,

for any pair of positive matrices P,Q ∈M+d satisfying P ≤ 1d. Applying this inequality for P = Yis and

Q =∑s′ 6=s Yis gives the estimate

pe(s, I) ≤ 2 [1− Tr (Yisσis)] + 4∑s′ 6=s

Tr(Yis′σis

).

Page 52: Fault-tolerant Coding for Quantum Communication

52

Following [52] we apply the operator equality

ABA = AB +BA−B + (1d −A)B (1d −A) ,

and using assumption 3. from above we obtain

pe(s, I)

= 2 [1− 2 Tr (ΠΠisσis) + Tr (Πisσis)− Tr ((1d −Π)Πis(1d −Π)σis)] + 4∑s′ 6=s

Tr(Yis′σis

)≤ 2 [1− Tr ((2Π− 1d)Πisσis)] + 4

∑s′ 6=s

Tr(Yis′σis

). (C1)

By an elementary computation and using that 2Π− 1d ≤ 1d we find that

1− Tr ((2Π− 1d)Πisσis) = 1− Tr ((2Π− 1d)σis) + Tr ((2Π− 1d)(1d −Πis)σis)≤ 1− Tr ((2Π− 1d)σis) + Tr ((1d −Πis)σis)= 3− 2 Tr (Πσis)− Tr (Πisσis) .

Combining this with (C1) leads to

pe(s, I) ≤ 6− 4 Tr (Πσis)− 2 Tr (Πisσis) + 4∑s′ 6=s

Tr(Yis′σis

).

For fixed I = i1, . . . , iM we define the average decoding error by

pe (I) = 1M

M∑s=1

pe(s, I).

Now define a 1, . . . , L-valued random variable Z distributed according to the probability distributionpiLi=1. Choosing the index set I at random according to M i.i.d. copies of Z leads to the followingupper bound on the expected value of pe:

EZ1,...,ZM [pe (Z1, . . . , ZM)]

≤ 1M

M∑s=1

EZ1,...,ZM

6− 4 Tr (ΠσZs)− 2 Tr (ΠZsσZs) + 4∑s′ 6=s

Tr(YZs′σZs

)= 6 + 1

M

M∑s=1

−4EZs Tr (ΠσZs)− 2EZs Tr (ΠZsσZs) + 4∑s′ 6=s

EZs,Zs′ Tr(YZs′σZs

)Using the assumptions 1. ,2. ,4. , and 5. from above, and that EZσZ = σ we find

EZ1,...,ZM [pe (Z1, . . . , ZM)]

≤ 6 + 1M

M∑s=1

−4 Tr (Πσ)− 2EZs Tr (ΠZsσZs) + 4∑s′ 6=s

EZs′ Tr(ΠZs′ΠσΠ

)≤ 6 + 1

M

M∑s=1

−4(1− ε1)− 2(1− ε2) + 4∑s′ 6=s

EZs′ Tr(

ΠZs′1B

Π)

≤ 4ε1 + 2ε2 + 4M A

B

The previous estimate shows the existence of an M -tuple I = (i1, . . . , iM ) such that

pe (I) ≤ 4ε1 + 2ε2 + 4M A

B.

Choosing this index set for the coding scheme gives the result of the lemma.

Page 53: Fault-tolerant Coding for Quantum Communication

53

To prove the direct part of the Holevo-Schumacher-Westmoreland theorem (following the strategypresented in [53]) we can use typical projectors in the packing lemma. Specifically, let pX(x), σxLx=1be an ensemble of quantum states with average state σ =

∑Lx=1 pX(x)σx and such that each σx has

eigenvalues pY |X(y|x) for y ∈ 1, . . . , L. For any δ > 0 let Πnδ be the δ-typical projector with respect

to σ⊗n and for any xn ∈ 1, . . . , Ln satisfying pXn(xn) > 0 let ΠBn|xnδ denote the conditional typical

projector with respect to the ensemble pXn(xn), ρxn conditioned onto xn. Applying Theorem B.5,Theorem B.6, Theorem B.7 and Theorem B.8 shows that

1. Tr [Πnδ σ⊗n] ≥ 1− exp

(−2nδ2

log(λmin)2

).

2.∑xn pXn(xn) Tr

[ΠBn|xnδ σxn

]≥ 1− exp

(−2nδ2

log(µmin)2

).

3.[ΠBn|xnδ , σxn

]= 0 for any xn ∈ 1, . . . , Ln.

4. Tr[ΠBn|xnδ

]≤ 2n(H(X,Y )−H(X)+δ) for any xn ∈ 1, . . . , Ln.

5. Πnδ σ⊗nΠn

δ ≤ 2−n(S(σ)−δ)Πnδ .

Here, we used

λmin := minλ ∈ spec(σ) \ 0,

and

µmin = minx∈supp(X)

minλ ∈ spec (σx) \ 0

Applying Lemma C1 shows that for every M ≤ Ln there exist I = (x(1)n, . . . , x(M)n) ∈ (1, . . . , Ln)M ,and a POVM (Λm)Ms=1 ∈ ((M⊗nL )+)M such that

1M

M∑s=1

Tr(Λsσx(s)n

)≥ 1− 4 exp

(−2nδ2

log(λmin)2

)− 2 exp

(−2nδ2

log(µmin)2

)− 4M2−n(χ(pX(x),σxLx=1)−2δ),

with the Holevo quantity of the ensemble pX(x), σxLx=1 given by

χ(pX(x), σxLx=1

)= S(σ) +H(X)−H(X,Y ).

Given a quantum channel T : Md1 → Md2 and an ensemble pX(x), ρxLi=1 we can apply the abovereasoning to the ensemble pX(x), T (ρx)Li=1 which leads to a coding scheme for the cq-channel x 7→ T (ρx)and the following theorem:

Theorem C.2 (Error bound classical capacity). Let T : Md1 → Md2 be a quantum channel,pX(x), ρxLi=1 an ensemble of quantum states on Cd1 , and σ =

∑x pX(x)T (ρx) the average state at the

channel output. For any M = 2nR with R ≤ log(L) and any δ > 0, there exists I = (x(1)n, . . . , x(M)n) ∈(1, . . . , Ln)M , and a POVM (Λs)Ms=1 ∈ ((M⊗nL )+)M such that

1M

M∑s=1

Tr(ΛsT⊗n

(ρx(s)n

))≥ 1− 4 exp

(−2nδ2

log(λmin)2

)− 2 exp

(−2nδ2

log(µmin)2

)− 4 · 2n(R−χ(pX(x),T (ρx))+2δ)

with

λmin = minλ ∈ spec(σ) \ 0,

and

µmin = minx∈supp(X)

minλ ∈ spec (T (ρx)) \ 0.

Page 54: Fault-tolerant Coding for Quantum Communication

54

Appendix D: Explicit error bound in the LSD-theorem

The following theorem follows from analyzing the coding scheme from [25].

Theorem D.1 (Decoupling with error bound). For any quantum channel T : Md1 → Md2 , any purestate |φAA′〉 ∈ Cd1 ⊗ Cd1 , any m ∈ N, any δ > 0, and any rate R > 0 there exists an encoder Em :M⊗Rm2 →M⊗md1

and a decoder Dm :M⊗md2→M⊗Rm2 such that

F(|Ω⊗Rm2 〉,

(id⊗Rm2 ⊗Dm T⊗m Em

) (ω⊗Rm2

))≥ 1− εm,

where ω2 = |Ω2〉〈Ω2| denotes the maximally entangled qubit state with |Ω2〉 = (|00〉+ |11〉)/√

2, and

εm = 4√

3 exp(− mδ2

log (µmin (φA′))2

)+ 2−m2 (Icoh(φA′ ,T )−R−3δ)

for

µmin(φA′) = minλ > 0 : λ ∈ spec (φA′) ∪ spec (T (φA′)) ∪ spec (T c(φA′)),

where T c denotes the complementary channel of T .

The error measure in Theorem D.1 is called the entanglement generation fidelity or channel fidelity(cf. [32]). This quantity can be related to the minimum fidelity

F (T ) = min〈ψ|T (|ψ〉〈ψ|)|ψ〉 : |ψ〉 ∈ Cd1 , 〈ψ|ψ〉 = 1.

Specifically, we can use [32, Proposition 4.5.] modifying the coding scheme slightly to find the followingcorollary.

Corollary D.2. For any quantum channel T : Md1 → Md2 , any pure state |φAA′〉 ∈ Cd1 ⊗ Cd1 , anyrate R > 0 and any m ∈ N such that Rm > 1, there exists an encoder Em :M⊗(Rm−1)

2 →M⊗md1and a

decoder Dm :M⊗md2→M⊗(Rm−1)

2 such that

F(Dm T⊗m Em

)≥ 1− εm

with

εm = 1− 2εm,

and εm as in Theorem D.1. Note the coding scheme given by Em and Dm still achieves the rate R.

[1] D. Aharonov and M. Ben-Or, “Fault Tolerant Quantum Computation with Constant Error,” https://arxiv.org/abs/quant-ph/9611025 , 2008.

[2] ——, “Fault-Tolerant Quantum Computation with Constant Error Rate,” https://arxiv.org/abs/quant-ph/9906129 , 1999.

[3] ——, “Fault-Tolerant Quantum Computation with Constant Error Rate,” SIAM Journal on Computing,2008.

[4] R. Ahlswede, I. Bjelaković, H. Boche, and J. Nötzel, “Quantum capacity under adversarial quantum noise:Arbitrarily varying quantum channels,” Communications in Mathematical Physics, vol. 317, no. 1, pp. 103–156, 2013.

[5] P. Aliferis, D. Gottesman, and J. Preskill, “Quantum accuracy threshold for concatenated distance-3 codes,”Quantum Information & Computation, vol. 6, no. 2, pp. 97–165, 2006.

[6] A. Ashikhmin, S. Litsyn, and M. A. Tsfasman, “Asymptotically good quantum codes,” Physical Review A,vol. 63, no. 3, p. 032311, 2001.

[7] A. Barenco, C. H. Bennett, R. Cleve, D. P. DiVincenzo, N. Margolus, P. Shor, T. Sleator, J. A. Smolin, andH. Weinfurter, “Elementary gates for quantum computation,” Physical review A, vol. 52, no. 5, p. 3457, 1995.

[8] H. Barnum, M. A. Nielsen, and B. Schumacher, “Information transmission through a noisy quantum channel,”Physical Review A, vol. 57, no. 6, p. 4153, 1998.

Page 55: Fault-tolerant Coding for Quantum Communication

55

[9] I. Bjelakovic and H. Boche, “Classical capacities of compound and averaged quantum channels,” IEEETransactions on Information theory, vol. 55, no. 7, pp. 3360–3374, 2009.

[10] H. Boche, C. Deppe, J. Nötzel, and A. Winter, “Fully quantum arbitrarily varying channels: Random codingcapacity and capacity dichotomy,” in 2018 IEEE International Symposium on Information Theory (ISIT).IEEE, 2018, pp. 2012–2016.

[11] P. O. Boykin, T. Mor, M. Pulver, V. Roychowdhury, and F. Vatan, “On universal and fault-tolerant quantumcomputing: a novel basis and a new constructive proof of universality for Shor’s basis,” in 40th AnnualSymposium on Foundations of Computer Science (Cat. No. 99CB37039). IEEE, 1999, pp. 486–494.

[12] H.-J. Briegel, W. Dür, J. I. Cirac, and P. Zoller, “Quantum repeaters: the role of imperfect local operationsin quantum communication,” Physical Review Letters, vol. 81, no. 26, p. 5932, 1998.

[13] A. R. Calderbank and P. W. Shor, “Good quantum error-correcting codes exist,” Physical Review A, vol. 54,no. 2, p. 1098, 1996.

[14] C. Chamberland, T. Jochym-O’Connor, and R. Laflamme, “Thresholds for universal concatenated quantumcodes,” Physical review letters, vol. 117, no. 1, p. 010501, 2016.

[15] H. Chen, S. Ling, and C. Xing, “Asymptotically good quantum codes exceeding the Ashikhmin-Litsyn-Tsfasman bound,” IEEE Transactions on Information Theory, vol. 47, no. 5, pp. 2055–2058, 2001.

[16] T. S. Cubitt, M. B. Ruskai, and G. Smith, “The structure of degradable quantum channels,” Journal ofMathematical Physics, vol. 49, no. 10, p. 102104, 2008.

[17] N. Datta and T. Dorlas, “Classical capacity of quantum channels with general markovian correlated noise,”Journal of Statistical Physics, vol. 134, no. 5-6, pp. 1173–1195, 2009.

[18] I. Devetak, “The private classical capacity and quantum capacity of a quantum channel,” IEEE Transactionson Information Theory, vol. 51, no. 1, pp. 44–55, 2005.

[19] W. Dür, H.-J. Briegel, J. I. Cirac, and P. Zoller, “Quantum repeaters based on entanglement purification,”Physical Review A, vol. 59, no. 1, p. 169, 1999.

[20] M. Epping, H. Kampermann, and D. Bruß, “On the error analysis of quantum repeaters with encoding,”Applied Physics B, vol. 122, no. 3, p. 54, 2016.

[21] O. Fawzi, A. Grospellier, and A. Leverrier, “Constant overhead quantum fault-tolerance with quantum ex-pander codes,” in 2018 IEEE 59th Annual Symposium on Foundations of Computer Science (FOCS). IEEE,2018, pp. 743–754.

[22] J. Fern and K. B. Whaley, “Lower bounds on the nonzero capacity of pauli channels,” Physical Review A,vol. 78, no. 6, p. 062335, 2008.

[23] D. Gottesman, “Stabilizer codes and quantum error correction,” https://arxiv.org/abs/quant-ph/9705052 ,1997.

[24] ——, “Fault-tolerant quantum computation with constant overhead,” Quantum Information & Computation,vol. 14, no. 15-16, pp. 1338–1372, 2014.

[25] P. Hayden, M. Horodecki, A. Winter, and J. Yard, “A decoupling approach to the quantum capacity,” OpenSystems & Information Dynamics, vol. 15, no. 01, pp. 7–19, 2008.

[26] A. S. Holevo, “The capacity of the quantum channel with general signal states,” IEEE Transactions onInformation Theory, vol. 44, no. 1, pp. 269–273, 1998.

[27] ——, “On the mathematical theory of quantum communication channels,” Problemy Peredachi Informatsii,vol. 8, no. 1, pp. 62–71, 1972.

[28] ——, “Bounds for the quantity of information transmitted by a quantum communication channel,” ProblemyPeredachi Informatsii, vol. 9, no. 3, pp. 3–11, 1973.

[29] L. Jiang, J. M. Taylor, K. Nemoto, W. J. Munro, R. Van Meter, and M. D. Lukin, “Quantum repeater withencoding,” Physical Review A, vol. 79, no. 3, p. 032325, 2009.

[30] C. King, K. Matsumoto, M. Nathanson, and M. B. Ruskai, “Properties of conjugate channels with applicationsto additivity and multiplicativity,” arXiv preprint quant-ph/0509126, 2005.

[31] E. Knill and R. Laflamme, “Concatenated quantum codes,” https://arxiv.org/abs/quant-ph/9608012 , 1996.[32] D. Kretschmann and R. F. Werner, “Tema con variazioni: quantum channel capacity,” New Journal of

Physics, vol. 6, no. 1, p. 26, 2004.[33] D. Leung and G. Smith, “Continuity of quantum channel capacities,” Communications in Mathematical

Physics, vol. 292, no. 1, pp. 201–215, 2009.[34] Z. Li, L. Xing, and X. Wang, “A family of asymptotically good quantum codes based on code concatenation,”

IEEE transactions on information theory, vol. 55, no. 8, pp. 3821–3824, 2009.[35] S. Lloyd, “Capacity of the noisy quantum channel,” Physical Review A, vol. 55, no. 3, p. 1613, 1997.[36] J. Łodyga, P. Mazurek, A. Grudka, and M. Horodecki, “Simple scheme for encoding and decoding a qubit in

unknown state for various topological codes,” Scientific reports, vol. 5, p. 8975, 2015.[37] R. Matsumoto, “Improvement of Ashikhmin-Litsyn-Tsfasman bound for quantum codes,” IEEE Transactions

on Information Theory, vol. 48, no. 7, pp. 2122–2124, 2002.[38] P. Mazurek, A. Grudka, M. Horodecki, P. Horodecki, J. Łodyga, Ł. Pankowski, and A. Przysiężna, “Long-

distance quantum communication over noisy networks without long-time quantum memory,” Physical ReviewA, vol. 90, no. 6, p. 062311, 2014.

[39] M. Mosonyi, “Coding theorems for compound problems via quantum rényi divergences,” IEEE Transactionson Information Theory, vol. 61, no. 6, pp. 2997–3012, 2015.

[40] S. Muralidharan, J. Kim, N. Lütkenhaus, M. D. Lukin, and L. Jiang, “Ultrafast and fault-tolerant quantum

Page 56: Fault-tolerant Coding for Quantum Communication

56

communication across long distances,” Physical review letters, vol. 112, no. 25, p. 250501, 2014.[41] M. Nicolaidis, Soft errors in modern electronic systems. Springer Science & Business Media, 2010, vol. 41.[42] M. Nielsen, I.L. Chuang, “Quantum Computation and Quantum Information,” Cambridge University Press,

2007.[43] A. Paetznick and B. W. Reichardt, “Fault-tolerant ancilla preparation and noise threshold lower boudds for

the 23-qubit golay code,” Quantum Information & Computation, vol. 12, no. 11-12, pp. 1034–1080, 2012.[44] J. Preskill, “Quantum Computing in the NISQ era and beyond,” Quantum, vol. 2, p. 79, 2018.[45] B. Schumacher, “Sending entanglement through noisy quantum channels,” Physical Review A, vol. 54, no. 4,

p. 2614, 1996.[46] B. Schumacher and M. A. Nielsen, “Quantum data processing and error correction,” Physical Review A,

vol. 54, no. 4, p. 2629, 1996.[47] B. Schumacher and M. D. Westmoreland, “Sending classical information via noisy quantum channels,” Phys-

ical Review A, vol. 56, no. 1, p. 131, 1997.[48] C. E. Shannon, “A mathematical theory of communication,” The Bell system technical journal, vol. 27, no. 3,

pp. 379–423, 1948.[49] M. E. Shirokov, “Tight uniform continuity bounds for the quantum conditional mutual information, for the

holevo quantity, and for capacities of quantum channels,” Journal of Mathematical Physics, vol. 58, no. 10,p. 102202, 2017.

[50] P. W. Shor, “The quantum channel capacity and coherent information,” in Lecture notes, MSRI Workshopon Quantum Computation, 2002.

[51] A. Steane, “Multiple-particle interference and quantum error correction,” Proceedings of the Royal Society ofLondon. Series A: Mathematical, Physical and Engineering Sciences, vol. 452, no. 1954, pp. 2551–2577, 1996.

[52] J. Watrous, The Theory of Quantum Information. Cambridge University Press, 2018.[53] M. M. Wilde, Quantum Information Theory. Cambridge University Press, 2017.