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

Quantum Error-Control Codes

Martianus Frederic Ezerman

Chapter 0 Quantum Error-Control Codes

1 Introduction

Information is physical. It is sensible to use quantum mechanics as a basis of computation and information processing [19]. Here at the intersection of information theory, computing, and physics, mathematicians and computer scientists must think in terms of the quantum physical realizations of messages. The often philosophical debates among physicists over the nature and interpretations of quantum mechanics shift to harnessing its power for information processing and testing the theory for completeness.

One cannot directly access information stored and processed in massively entangled quantum systems without destroying the content. Turning large-scale quantum computing into practical reality is massively challenging. To start with, it requires techniques for error control that are much more complex than those implemented effectively in classical systems. As a quantum system grows in size and circuit depth, error control becomes ever more important.

Quantum error-control is a set of methods to protect quantum information from unwanted environmental interactions, known as decoherence. Classically, one encodes information-carrying vectors into a larger space to allow for sufficient redundancy for error detection and correction. In the quantum setup, information is stored in a subspace embedded in a larger Hilbert space, which is a finite dimensional, normed, vector space over the field of complex numbers \mathbb{C}. Codewords are quantum states and errors are operators.

The good news is that noise, if it can be kept below a certain level, is not an obstacle to resilient quantum computation. This crucial insight is arrived at based on seminal results that form the so-called treshold theorems. Theoretical references include the exposition of Knill et al.in [34], the work of Preskill in [42], the results shown by Steane in [49], and the paper of Ahoronov and Ben-Or [1]. A comprehensive review on related experiments is available in [9].

The possibility of correcting errors in quantum systems was shown, e.g., in the early works of Shor [46], Steane [47] and Laflamme et al.[35]. While the quantum codes that these pioneers proposed may nowadays seem to be rather trivial in performance, their construction highlighted both the main obstacles and their respective workarounds. Measurement collapses the information contained in the state into something useless. One should measure the error, not the data. Since repetition is ruled out due to the no-cloning theorem [53], we use redundancy from spreading the states to avoid repetition. There are multiple types of errors, as we will soon see. The key is to start by correcting the phase errors and, then, use the Hadamard transform to exchange the bit flips and the phase errors. Quantum errors are continuous. Controlling them seemed to be too daunting a task. It turned out that handling a set of discrete error operators, represented by tensor product of Pauli matrices, allows for the control of every \mathbb{C}-linear combination of these operators.

Advances continue to be made as effort intensifies to scale quantum computing up. Research in quantum error-correcting codes (QECs) has attracted the sustained attention of established researchers and students alike. Several excellent online lecture notes, surveys, and books are available. Developments up to 2011 have been well-documented in [36]. It is impossible to describe the technical details of every important research direction in QECs. We focus on quantum stabilizer codes and their variants. The decidedly biased take here is for the audience with more applied mathematics background, including coding theorist, information theorist, researchers in discrete mathematics and finite geometries. No knowledge of quantum mechanics is required beyond the very basic. This chapter is meant to serve as an entry point for those who want to understand and get involved in building upon this foundational aspect of quantum information processing and computation, which have been tipped to be indispensable in future technologies.

A quantum stabilizer code is designed so that errors with high probability of occuring transform information-carrying states to an error space which is orthogonal to the original space. The beauty lies in how natural the determination of the location and type of each error in the system is. Correction becomes a simple application of the type of error at the very location.

2 Preliminaries

Consider the field extensions 𝔽p\mathbb{F}_{p} to 𝔽q=pr\mathbb{F}_{q=p^{r}} to 𝔽qm=prm\mathbb{F}_{q^{m}=p^{rm}}, for positive integers rr and mm. For α𝔽qm\alpha\in\mathbb{F}_{q^{m}}, the trace mapping from 𝔽qm\mathbb{F}_{q^{m}} to 𝔽q\mathbb{F}_{q} is given by

Tr𝔽qm/𝔽q(α)=α+αq++αqm1𝔽q.\Tr_{\mathbb{F}_{q^{m}}/\mathbb{F}_{q}}(\alpha)=\alpha+\alpha^{q}+\ldots+\alpha^{q^{m-1}}\in\mathbb{F}_{q}.

The trace of α\alpha is the sum of its conjugates. If the extension 𝔽q\mathbb{F}_{q} of 𝔽p\mathbb{F}_{p} is contextually clear, the notation Tr\Tr is sufficient. Properties of the trace map can be found in standard textbooks, e.g., [37, Chapter 2]. The key idea is that Tr𝔽qm/𝔽q\Tr_{\mathbb{F}_{q^{m}}/\mathbb{F}_{q}} serves as a description for all linear transformations from 𝔽qm\mathbb{F}_{q^{m}} to 𝔽q\mathbb{F}_{q}.

Let GG be a finite abelian group, written multiplicatively, with the identity 1G1_{G}. Let UU be the multiplicative group of complex numbers of modulus 11, i.e., the unit circle of radius 11 on the complex plane \mathbb{C}. A character χ:GU\chi:G\mapsto U is a homomorphism. For any gGg\in G, the images of χ\chi are |G|\absolutevalue{G}-th roots of unity since (χ(g))|G|=χ(g|G|)=1\left(\chi(g)\right)^{\absolutevalue{G}}=\chi\left(g^{\absolutevalue{G}}\right)=1. Let c¯\overline{c} denote the complex conjugate of cc. Then χ(g1)=(χ(g))1=χ(g)¯\chi(g^{-1})=\left(\chi(g)\right)^{-1}=\overline{\chi(g)}. The only trivial character is χ0:g1\chi_{0}:g\mapsto 1 for all gGg\in G. One can associate χ\chi and χ¯\overline{\chi} by using χ¯(g)=χ(g)¯\overline{\chi}(g)=\overline{\chi(g)}. The set of all characters of GG forms, under composition \circ, an abelian group G^\widehat{G}.

For g,hGg,h\in G and χ,ΨG^\chi,\Psi\in\widehat{G} we have two orthogonality relations

gGχ(g)Ψ(g)¯={0, if χΨ|G|, if χ=ΨandχG^χ(g)χ(h1)={0, if gh|G|, if g=h.\sum_{g\in G}\chi(g)\,\overline{\Psi(g)}=\begin{cases}0\mbox{, if }\chi\neq\Psi\\ \absolutevalue{G}\mbox{, if }\chi=\Psi\end{cases}\mbox{and}\quad\sum_{\chi\in\widehat{G}}\chi(g)\,\chi(h^{-1})=\begin{cases}0\mbox{, if }g\neq h\\ \absolutevalue{G}\mbox{, if }g=h\end{cases}. (1)

The additive character χ1:=ce2πpTr(c)\chi_{1}:=c\mapsto e^{\frac{2\pi}{p}\Tr(c)}, for all c𝔽qc\in\mathbb{F}_{q}, is called the canonical character. For a chosen b𝔽qb\in\mathbb{F}_{q} and for all c𝔽qc\in\mathbb{F}_{q},

χb:=𝔽qU sending cχ1(bc)=e2πpTr(bc)\chi_{b}:=\mathbb{F}_{q}\rightarrow U\mbox{ sending }c\mapsto\chi_{1}(b\cdot c)=e^{\frac{2\pi}{p}\Tr(b\cdot c)}

is a character of (𝔽q,+)(\mathbb{F}_{q},+). Every character of (𝔽q,+)(\mathbb{F}_{q},+) can, in fact, be expressed in this manner. The extension to (𝔽qn,+)(\mathbb{F}_{q}^{n},+) is straightforward.

Theorem 2.1.

Let ζ:=e2πp\zeta:=e^{\frac{2\pi}{p}} and Tr\Tr be the trace map with q=pmq=p^{m}. Let 𝐚=(a1,a2,,an){\mathbf{a}}=(a_{1},a_{2},\ldots,a_{n}) and 𝐛=(b1,b2,,bn){\mathbf{b}}=(b_{1},b_{2},\ldots,b_{n}) be vectors in 𝔽qn\mathbb{F}_{q}^{n}. For each 𝐚{\mathbf{a}},

λ𝐚:𝔽qn{1,ζ,ζ2,,ζp1}, sending 𝐛ζTr(𝐚𝐛)=ζTr(a1b1++anbn),\lambda_{{\mathbf{a}}}:\mathbb{F}_{q}^{n}\mapsto\{1,\zeta,\zeta^{2},\ldots,\zeta^{p-1}\}\mbox{, sending }{\mathbf{b}}\mapsto\zeta^{\Tr({\mathbf{a}}\cdot{\mathbf{b}})}=\zeta^{\Tr(a_{1}b_{1}+\ldots+a_{n}b_{n})},

for all 𝐛𝔽q{\mathbf{b}}\in\mathbb{F}_{q}, is a character of (𝔽qn,+)(\mathbb{F}_{q}^{n},+). Hence, 𝔽qn^={λ𝐚:𝐚𝔽qn}\widehat{\mathbb{F}_{q}^{n}}=\{\lambda_{{\mathbf{a}}}:{\mathbf{a}}\in\mathbb{F}_{q}^{n}\}.

A qubit, a term coined by Schumacher in [45], is the canonical quantum system consisting of two distinct levels. The states of a qubit live in 2\mathbb{C}^{2} and are defined by their continuous amplitudes. A qudit refers to a system of q3q\geq 3 distinct levels, with a qutrit commonly used when q=3q=3. Physicists prefer the “bra|\ket{\cdot} and “ket|\bra{\cdot} notation to describe quantum systems. A |φ\ket{\varphi} is a (column) vector while ψ|\bra{\psi} is the vector dual of |ψ\ket{\psi}.

Definition 1 (Quantum systems).

A qubit is a nonzero vector in 2\mathbb{C}^{2}, usually with basis {|0,|1}\{\ket{0},\ket{1}\}. It is written in vector form as |φ:=α|0+β|1\ket{\varphi}:=\alpha\ket{0}+\beta\ket{1}, or in matrix form as [αβ]\displaystyle{\begin{bmatrix}\alpha\\ \beta\end{bmatrix}}, with α2+β2=1\norm{\alpha}^{2}+\norm{\beta}^{2}=1 .

An nn-qubit system or vector is a nonzero element in (2)n2n\left(\mathbb{C}^{2}\right)^{\otimes n}\cong\mathbb{C}^{2^{n}}. Let 𝐚=(a1,,an)𝔽2n{\mathbf{a}}=(a_{1},\ldots,a_{n})\in\mathbb{F}_{2}^{n}. The standard \mathbb{C}-basis is

{|a1a2an:=|a1|a2|an:𝐚𝔽2n}.\{\ket{a_{1}a_{2}\ldots a_{n}}:=\ket{a_{1}}\otimes\ket{a_{2}}\otimes\ldots\otimes\ket{a_{n}}:{\mathbf{a}}\in\mathbb{F}_{2}^{n}\}.

An arbitrary nonzero vector in 2n\mathbb{C}^{2^{n}} is written

|ψ=𝐚𝔽2nc𝐚|𝐚, with c𝐚 and 12n𝐚𝔽2nc𝐚2=1.\ket{\psi}=\sum_{{\mathbf{a}}\in\mathbb{F}_{2}^{n}}c_{{\mathbf{a}}}\ket{{\mathbf{a}}}\mbox{, with }c_{{\mathbf{a}}}\in\mathbb{C}\mbox{ and }\frac{1}{2^{n}}\sum_{{\mathbf{a}}\in\mathbb{F}_{2}^{n}}\norm{c_{{\mathbf{a}}}}^{2}=1.

The normalization is optional since |ψ\ket{\psi} and α|ψ\alpha\ket{\psi} are considered the same state for nonzero α\alpha\in\mathbb{C}.

The inner product of |ψ:=𝐚𝔽2nc𝐚|𝐚\ket{\psi}:=\sum_{{\mathbf{a}}\in\mathbb{F}_{2}^{n}}c_{{\mathbf{a}}}\ket{{\mathbf{a}}} and |φ:=𝐚𝔽2nb𝐚|𝐚\ket{\varphi}:=\sum_{{\mathbf{a}}\in\mathbb{F}_{2}^{n}}b_{{\mathbf{a}}}\ket{{\mathbf{a}}} is

ψ|φ=𝐚𝔽2nc𝐚¯b𝐚.\innerproduct{\psi}{\varphi}=\sum_{{\mathbf{a}}\in\mathbb{F}_{2}^{n}}\overline{c_{{\mathbf{a}}}}\,b_{{\mathbf{a}}}.

Their (Kronecker) tensor product is written as |φ|ψ\ket{\varphi}\otimes\ket{\psi} and is often abbreviated to |φψ\ket{\varphi\psi}. The states |ψ\ket{\psi} and |φ\ket{\varphi} are orthogonal or distinguishable if ψ|φ=0\innerproduct{\psi}{\varphi}=0. Let AA be a 2n×2n2^{n}\times 2^{n} complex unitary matrix with conjugate transpose AA^{\dagger}. The (Hermitian) inner product of |φ\ket{\varphi} and A|ψA\ket{\psi} is equal to that of A|φA^{\dagger}\ket{\varphi} and |ψ\ket{\psi}. Henceforth, i:=1i:=\sqrt{-1}.

Definition 2 (Qubit error operators).

A qubit error operator is a unitary \mathbb{C}-linear operator acting on 2n\mathbb{C}^{2^{n}} qubit by qubit. It can be expressed by a unitary matrix with respect to the basis {|0,|1}\{\ket{0},\ket{1}\}. The three nontrivial errors acting on a qubit are known as the Pauli matrices:

σx=[0110],σz=[1001],σy=iσxσz=[0ii0].\sigma_{x}=\begin{bmatrix}0&1\\ 1&0\end{bmatrix},\quad\sigma_{z}=\begin{bmatrix}1&0\\ 0&-1\end{bmatrix},\quad\sigma_{y}=i\,\sigma_{x}\,\sigma_{z}=\begin{bmatrix}0&-i\\ i&0\end{bmatrix}. (2)

The actions of the error operators on a qubit |v=α|0+β|12\ket{v}=\alpha\ket{0}+\beta\ket{1}\in\mathbb{C}^{2} can be considered based on their types. The trivial operator I2I_{2} leaves the qubit unchanged. The bit-flip error σx\sigma_{x} flips the probabilities

σx|φ=β|0+α|1 or σx[αβ]=[βα].\sigma_{x}\,\ket{\varphi}=\beta\ket{0}+\alpha\ket{1}\mbox{ or }\sigma_{x}\begin{bmatrix}\alpha\\ \beta\end{bmatrix}=\begin{bmatrix}\beta\\ \alpha\end{bmatrix}.

The phase-flip error σz\sigma_{z} modifies the angular measures

σz|φ=α|0β|1 or σz[αβ]=[αβ].\sigma_{z}\,\ket{\varphi}=\alpha\ket{0}-\beta\ket{1}\mbox{ or }\sigma_{z}\begin{bmatrix}\alpha\\ \beta\end{bmatrix}=\begin{bmatrix}\alpha\\ -\beta\end{bmatrix}.

The combination error σy\sigma_{y} contains both bit-flip and phase-flip, implying

σy|φ=iβ|0+iα|1 or σy[αβ]=[iβiα].\sigma_{y}\,\ket{\varphi}=-i\beta\ket{0}+i\alpha\ket{1}\mbox{ or }\sigma_{y}\begin{bmatrix}\alpha\\ \beta\end{bmatrix}=\begin{bmatrix}-i\beta\\ i\alpha\end{bmatrix}.

It is immediate to confirm that σx2=σy2=σz2=I2\sigma_{x}^{2}=\sigma_{y}^{2}=\sigma_{z}^{2}=I_{2} and σxσz=σzσx\sigma_{x}\sigma_{z}=-\sigma_{z}\sigma_{x}. The Pauli matrices generate a group of order 1616. Each of its elements can be uniquely represented as iλwi^{\lambda}w, with λ{0,1,2,3}\lambda\in\{0,1,2,3\} and w{I2,σx,σz,σy}w\in\{I_{2},\sigma_{x},\sigma_{z},\sigma_{y}\}.

3 The Stabilizer Formalism

The most common route from classical coding theory to QEC is via the stabilizer formalism, from which numerous specific constructions emerge. Classical codes can not be used as quantum codes but can model the error operators in some quantum channels. The capabilities of a QEC can then be inferred from the properties of the corresponding classical codes. The main tools come from character theory and symplectic geometry over finite fields. Our focus is on the qubit setup since it is the most deployment-feasible and because the general qudit case naturally follows from it.

Let 𝐚=(a1,a2,,an)𝔽2n{\mathbf{a}}=(a_{1},a_{2},\ldots,a_{n})\in\mathbb{F}_{2}^{n}, λ{0,1,2,3}\lambda\in\{0,1,2,3\}, and wj{I2,σx,σz,σy}w_{j}\in\{I_{2},\sigma_{x},\sigma_{z},\sigma_{y}\}. A quantum error operator on 2n\mathbb{C}^{2^{n}} is of the form E:=iλw1w2wn{\rm E}:=i^{\lambda}w_{1}\otimes w_{2}\otimes\ldots\otimes w_{n}. It is a \mathbb{C}-linear unitary operator acting on a 2n\mathbb{C}^{2^{n}}-basis {|𝐚=|a1|a2|an}\{\ket{{\mathbf{a}}}=\ket{a_{1}}\otimes\ket{a_{2}}\otimes\ldots\otimes\ket{a_{n}}\} by E|𝐚:=iλ(w1|a1w2|a2wn|an){\rm E}\ket{{\mathbf{a}}}:=i^{\lambda}\left(w_{1}\ket{a_{1}}\otimes w_{2}\ket{a_{2}}\otimes\ldots\otimes w_{n}\ket{a_{n}}\right). The set of error operators

n:={iλw1w2wn}{\mathcal{E}}_{n}:=\{i^{\lambda}w_{1}\otimes w_{2}\otimes\ldots\otimes w_{n}\}

is a non-abelian group of cardinality 4n+14^{n+1}. Given E:=iλw1w2wn{\rm E}:=i^{\lambda}w_{1}\otimes w_{2}\otimes\ldots\otimes w_{n} and E:=iλw1w2wn{\rm E}^{\prime}:=i^{\lambda^{\prime}}w_{1}^{\prime}\otimes w_{2}^{\prime}\otimes\ldots\otimes w_{n}^{\prime} in n{\mathcal{E}}_{n}, we have

EE\displaystyle{\rm E}\,{\rm E}^{\prime} =iλ+λ(w1w1)(w2w2)(wnwn)\displaystyle=i^{\lambda+\lambda^{\prime}}(w_{1}w_{1}^{\prime})\otimes(w_{2}w_{2}^{\prime})\otimes\ldots\otimes(w_{n}w_{n}^{\prime})
=iλ+λ+λ′′w1′′w2′′wn′′, where wjwj=iλj′′wj′′ and λ′′=j=1nλj′′.\displaystyle=i^{\lambda+\lambda^{\prime}+\lambda^{\prime\prime}}w_{1}^{\prime\prime}\otimes w_{2}^{\prime\prime}\otimes\ldots\otimes w_{n}^{\prime\prime}\mbox{, where }w_{j}w_{j}^{\prime}=i^{\lambda_{j}^{\prime\prime}}w_{j}^{\prime\prime}\mbox{ and }\lambda^{\prime\prime}=\sum_{j=1}^{n}\lambda_{j}^{\prime\prime}.

Expanding EE{\rm E}^{\prime}\,{\rm E} makes it clear that EE=±1EE{\rm E}\,{\rm E}^{\prime}=\pm 1~{}{\rm E}^{\prime}\,{\rm E}.

Example 1.

Given n=2n=2, E=I2σx{\rm E}=I_{2}\otimes\sigma_{x} and E=σzσy{\rm E}^{\prime}=\sigma_{z}\otimes\sigma_{y}, we have EE=σzσxσy=σziσz=iσzσz{\rm E}\,{\rm E}^{\prime}=\sigma_{z}\otimes\sigma_{x}\sigma_{y}=\sigma_{z}\otimes i\sigma_{z}=i\sigma_{z}\otimes\sigma_{z} and EE=σzσyσx=σzi3σz=i3σzσz{\rm E}^{\prime}\,{\rm E}=\sigma_{z}\otimes\sigma_{y}\sigma_{x}=\sigma_{z}\otimes i^{3}\sigma_{z}=i^{3}\sigma_{z}\otimes\sigma_{z}.  ∎

The center of n{\mathcal{E}}_{n} is 𝒞(n):={iλI2I2I2}\mathcal{C}({\mathcal{E}}_{n}):=\{i^{\lambda}I_{2}\otimes I_{2}\otimes\ldots\otimes I_{2}\}. Let n¯\overline{{\mathcal{E}}_{n}} denote the quotient group n/𝒞(n){\mathcal{E}}_{n}/\mathcal{C}({\mathcal{E}}_{n}) of cardinality |n¯|=4n\absolutevalue{\overline{{\mathcal{E}}_{n}}}=4^{n}. This group is an abelian 22-elementary group (𝔽22n,+)\cong(\mathbb{F}_{2}^{2n},+), since E¯2=I2I2=I2n\overline{{\rm E}}^{2}=I_{2}\otimes\cdots\otimes I_{2}=I_{2^{n}} for any E¯n¯\overline{{\rm E}}\in\overline{{\mathcal{E}}_{n}}.

We switch notation to define the product of error operators in terms of an inner product of their vector representatives. We write E=iλw1w2wn{\rm E}=i^{\lambda}w_{1}\otimes w_{2}\otimes\ldots\otimes w_{n} as E=iλ+ϵX(𝐚)Z(𝐛){\rm E}=i^{\lambda+\epsilon}X({\mathbf{a}})Z({\mathbf{b}}), where 𝐚=(a1,,an),𝐛=(b1,,bn)𝔽2n{\mathbf{a}}=(a_{1},\ldots,a_{n}),{\mathbf{b}}=(b_{1},\ldots,b_{n})\in\mathbb{F}_{2}^{n} and ϵ:=|{1in}:wi=σy|\epsilon:=\absolutevalue{\{1\leq i\leq n\}:w_{i}=\sigma_{y}}, by replacing (ai,bi)(a_{i},b_{i}) with (0,0)(0,0) if wi=I2w_{i}=I_{2}, by (1,0)(1,0) if wi=σxw_{i}=\sigma_{x}, by (0,1)(0,1) if wi=σzw_{i}=\sigma_{z}, and by (1,1)(1,1) if wi=σyw_{i}=\sigma_{y}.

The respective actions of X(𝐚)X({\mathbf{a}}) and Z(𝐛)Z({\mathbf{b}}) on any vector |𝐯2n\ket{{\mathbf{v}}}\in\mathbb{C}^{2^{n}}, for 𝐯𝔽2n{\mathbf{v}}\in\mathbb{F}_{2}^{n}, are X(𝐚)|𝐯=|𝐚+𝐯X({\mathbf{a}})\ket{{\mathbf{v}}}=\ket{{\mathbf{a}}+{\mathbf{v}}} and Z(𝐛)|𝐯=(1)𝐛𝐯|𝐯Z({\mathbf{b}})\ket{{\mathbf{v}}}=(-1)^{{\mathbf{b}}\cdot{\mathbf{v}}}\ket{{\mathbf{v}}}. The matrix for X(𝐚)X({\mathbf{a}}) is a symmetric {0,1}\{0,1\} matrix. It represents a permutation consisting of 2n12^{n-1} transpositions. The matrix for Z(𝐛)Z({\mathbf{b}}) is diagonal with diagonal entries ±1\pm 1. Hence, writing the operators in n{\mathcal{E}}_{n} as E:=iλX(𝐚)Z(𝐛){\rm E}:=i^{\lambda}X({\mathbf{a}})Z({\mathbf{b}}) and E:=iλX(𝐚)Z(𝐛){\rm E}^{\prime}:=i^{\lambda^{\prime}}X({\mathbf{a}}^{\prime})Z({\mathbf{b}}^{\prime}), one gets EE=(1)𝐚𝐛+𝐚𝐛EE{\rm E}\,{\rm E}^{\prime}=(-1)^{{\mathbf{a}}\cdot{\mathbf{b}}^{\prime}+{\mathbf{a}}^{\prime}\cdot{\mathbf{b}}}~{}{\rm E}^{\prime}\,{\rm E}. The symplectic inner product of (𝐚|𝐛)({\mathbf{a}}|{\mathbf{b}}) and (𝐚|𝐛)({\mathbf{a}}^{\prime}|{\mathbf{b}}^{\prime}) in 𝔽22n\mathbb{F}_{2}^{2n} is

(𝐚|𝐛),(𝐚|𝐛)s=𝐚𝐛+𝐚𝐛\langle({\mathbf{a}}|{\mathbf{b}}),({\mathbf{a}}^{\prime}|{\mathbf{b}}^{\prime})\rangle_{s}={\mathbf{a}}\cdot{\mathbf{b}}^{\prime}+{\mathbf{a}}^{\prime}\cdot{\mathbf{b}} (3)

or, in matrix form,

(𝐚|𝐛),(𝐚|𝐛)s=[𝐚𝐛][0InIn0][𝐚𝐛].\langle({\mathbf{a}}|{\mathbf{b}}),({\mathbf{a}}^{\prime}|{\mathbf{b}}^{\prime})\rangle_{s}=\begin{bmatrix}{\mathbf{a}}&{\mathbf{b}}\end{bmatrix}\,\begin{bmatrix}0&I_{n}\\ I_{n}&0\end{bmatrix}\,\begin{bmatrix}{\mathbf{a}}^{\prime}\\ {\mathbf{b}}^{\prime}\end{bmatrix}.

The symplectic dual of C𝔽22nC\subseteq\mathbb{F}_{2}^{2n} is Cs={𝐮𝔽22n:𝐮,𝐜s=0𝐜C}C^{\perp_{s}}=\{{\mathbf{u}}\in\mathbb{F}_{2}^{2n}:\langle{\mathbf{u}},{\mathbf{c}}\rangle_{s}=0\,\forall\,{\mathbf{c}}\in C\}. Thus, a subgroup GG of n{\mathcal{E}}_{n} is abelian if and only if G¯\overline{G} is a symplectic self-orthogonal subspace of n¯𝔽22n\overline{{\mathcal{E}}_{n}}\cong\mathbb{F}_{2}^{2n}.

Example 2.

Continuing from Example 1, we write E=X((0,1))Z((0,0)){\rm E}=X((0,1))Z((0,0)) and E=iX((0,1))Z((1,1)){\rm E}^{\prime}=iX((0,1))Z((1,1)). We choose the ordering (0,0),(0,1),(1,0),(1,1)(0,0),(0,1),(1,0),(1,1) of 𝔽22\mathbb{F}_{2}^{2} and the corresponding ordering for the basis of 4\mathbb{C}^{4}. The matrix for X((0,1))X((0,1)) agrees with I2σxI_{2}\otimes\sigma_{x}, the matrix for Z((0,0))Z((0,0)) is I4I_{4}, and the matrix for Z((1,1))Z((1,1)) is diagonal with diagonal entries 1,1,1,11,-1,-1,1. Multiplying matrices confirms that σzσy\sigma_{z}\otimes\sigma_{y} is indeed iX((0,1))Z((1,1))iX((0,1))Z((1,1)).  ∎

The quantum weight of an error operator E=iλX(𝐚)Z(𝐛)n{\rm E}=i^{\lambda}X({\mathbf{a}})Z({\mathbf{b}})\in{\mathcal{E}}_{n} is

wQ(E):=wQ(E¯)\displaystyle\operatorname{w_{Q}}({\rm E}):=\operatorname{w_{Q}}(\overline{{\rm E}}) =wQ(𝐚|𝐛)=|{1in:ai=1 or bi=1}|\displaystyle=\operatorname{w_{Q}}({\mathbf{a}}|{\mathbf{b}})=\absolutevalue{\{1\leq i\leq n:a_{i}=1\mbox{ or }b_{i}=1\}}
=|{1in:wiI2}|.\displaystyle=\absolutevalue{\{1\leq i\leq n:w_{i}\neq I_{2}\}}.

By definition, wQ(EE)wQ(E)+wQ(E)\operatorname{w_{Q}}({\rm E}\,{\rm E}^{\prime})\leq\operatorname{w_{Q}}({\rm E})+\operatorname{w_{Q}}({\rm E}^{\prime}), for any E,En{\rm E},{\rm E}^{\prime}\in{\mathcal{E}}_{n}. We can define the set of all error operators of weight at most δ\delta in n{\mathcal{E}}_{n} and determine its cardinality. Let

n(δ):={En:wQ(E)δ} and n¯(δ)={E¯n¯:wQ(E¯)δ}.{\mathcal{E}}_{n}(\delta):=\{{\rm E}\in{\mathcal{E}}_{n}:\operatorname{w_{Q}}({\rm E})\leq\delta\}\mbox{ and }\overline{{\mathcal{E}}_{n}}(\delta)=\{\overline{{\rm E}}\in\overline{{\mathcal{E}}_{n}}:\operatorname{w_{Q}}(\overline{{\rm E}})\leq\delta\}.

Then |n(δ)|=4j=0δ3j(nj)\absolutevalue{{\mathcal{E}}_{n}(\delta)}=4\sum_{j=0}^{\delta}3^{j}\binom{n}{j} and |n¯(δ)|=j=0δ3j(nj)\absolutevalue{\overline{{\mathcal{E}}_{n}}(\delta)}=\sum_{j=0}^{\delta}3^{j}\binom{n}{j}.

In the classical setup, both errors and codewords are vectors over the same field. In the quantum setup, errors are linear combinations of the tensor products of Pauli matrices. A qubit code Q2nQ\subseteq\mathbb{C}^{2^{n}} has three parameters: its length nn, dimension KK over \mathbb{C}, and minimum distance d=d(Q)d=d(Q). We use

((n,K,d)) or n,k,d with k=log2K((n,K,d))\mbox{ or }\llbracket n,k,d\rrbracket\mbox{ with }k=\log_{2}K

to signify that QQ describes the encoding of kk logical qubits as nn physical qubits, with dd being the smallest number of simultaneous errors that can transform a valid codeword into another.

Definition 3 (Knill-Laflamme condition [33]).

A quantum code QQ can correct up to \ell quantum errors if the followings hold. If |φ,|ψQ\ket{\varphi},\ket{\psi}\in Q are distinguishable, i.e., φ|ψ=0\innerproduct{\varphi}{\psi}=0, then φ|E1E2|ψ=0\bra{\varphi}{\rm E}_{1}{\rm E}_{2}\ket{\psi}=0, i.e., E1|φ{\rm E}_{1}\ket{\varphi} and E2|ψ{\rm E}_{2}\ket{\psi} must remain distinguishable, for all E1,E2n(){\rm E}_{1},{\rm E}_{2}\in{\mathcal{E}}_{n}(\ell). The minimum distance of QQ is d:=d(Q)d:=d(Q) if φ|E|ψ=0\bra{\varphi}{\rm E}\ket{\psi}=0 for all En(d1){\rm E}\in{\mathcal{E}}_{n}(d-1) and for all distinguishable |φ,|ψQ\ket{\varphi},\ket{\psi}\in Q.

Given an ((n,K,d))((n,K,d))-qubit code QQ and an En{\rm E}\in{\mathcal{E}}_{n}, EQ{\rm E}\,Q is a subspace of 2n\mathbb{C}^{2^{n}}. The fact that QQ corrects errors of weight up to =d12\ell=\lfloor\frac{d-1}{2}\rfloor does not imply that the subspaces {E¯Q:E¯n¯()}\{\overline{{\rm E}}Q:\overline{{\rm E}}\in\overline{{\mathcal{E}}_{n}}(\ell)\} are orthogonal to each other. It is possible that a codeword |𝐯\ket{{\mathbf{v}}} is fixed by some EI2n{\rm E}\neq I_{2^{n}}, say, when |𝐯\ket{{\mathbf{v}}} is an eigenvector of E{\rm E} satisfying E|𝐯=α|𝐯{\rm E}\ket{{\mathbf{v}}}=\alpha\ket{{\mathbf{v}}} for some nonzero α\alpha\in\mathbb{C}. If the subspaces {E¯Q:E¯n¯()}\{\overline{{\rm E}}\,Q:\overline{{\rm E}}\in\overline{{\mathcal{E}}_{n}}(\ell)\} are orthogonal to each other, then QQ is said to be pure. Otherwise, the code is degenerate or impure.

To formally define a qubit stabilizer code, we choose an abelian group GG, which is a subgroup of n{\mathcal{E}}_{n}, and associate GG with a classical code C𝔽22nC\subset\mathbb{F}_{2}^{2n}, which is self-orthogonal under the symplectic inner product. The action of GG partitions 2n\mathbb{C}^{2^{n}} into a direct sum of χ\chi-eigenspaces Q(χ)Q(\chi) with χG^\chi\in\widehat{G}. The properties of Q:=Q(χ)Q:=Q(\chi) follow from the properties of CC and CsC^{\perp_{s}}. The stabilizer formalism, first introduced by Gottesman in his thesis [23] and described in the language of group algebra by Calderbank et al.in [8], remains the most widely-studied approach to control quantum errors. Ketkar et al.generalized the formalism to qudit codes derived from classical codes over 𝔽q2\mathbb{F}_{q^{2}} in [31].

Let GG be a finite abelian group acting on a finite dimensional \mathbb{C}-vector space VV. Each gGg\in G is a Hermitian operator of VV and, for any g,gGg,g^{\prime}\in G and for all |𝐯V\ket{{\mathbf{v}}}\in V, (gg)|𝐯=g(g(|𝐯)) and gg1(|𝐯)=|𝐯(gg^{\prime})\ket{{\mathbf{v}}}=g(g^{\prime}(\ket{{\mathbf{v}}}))\mbox{ and }gg^{-1}(\ket{{\mathbf{v}}})=\ket{{\mathbf{v}}}. Let G^\widehat{G} be the character group of GG. For any χG^\chi\in\widehat{G}, the map Lχ:=1|G|gGχ¯(g)gL_{\chi}:=\frac{1}{\absolutevalue{G}}\sum_{g\in G}\overline{\chi}(g)\,g is a linear operator over VV. The set {Lχ:χG^}\{L_{\chi}:\chi\in\widehat{G}\} is the system of orthogonal primitive idempotent operators.

Proposition 3.1.

LχL_{\chi} is idempotent, i.e., Lχ2=LχL_{\chi}^{2}=L_{\chi} and LχLχ=0L_{\chi}L_{\chi^{\prime}}=0 if χχ\chi\neq\chi^{\prime}. The operators in the system sum to the identity χG^Lχ=1\sum_{\chi\in\widehat{G}}L_{\chi}=1. For all gGg\in G, we have gLχ=χ(g)Lχg\,L_{\chi}=\chi(g)\,L_{\chi}.

Proof.

In GG, let gh=agh=a, i.e., h=ag1h=ag^{-1}. Using χ¯=χ1\overline{\chi}=\chi^{-1} and the orthogonality of characters, we write

LχLχ=1|G|2gGχ¯(g)ghGχ¯(h)h=1|G|2a,gGχ¯(g)χ¯(ag1)a=1|G|2aGχ¯(a)agG(χ¯χ)(g).L_{\chi}L_{\chi^{\prime}}=\frac{1}{\absolutevalue{G}^{2}}\sum_{g\in G}\overline{\chi}(g)\,g~{}\sum_{h\in G}\overline{\chi^{\prime}}(h)\,h=\frac{1}{\absolutevalue{G}^{2}}\sum_{a,g\in G}\overline{\chi}(g)~{}\overline{\chi^{\prime}}(ag^{-1})~{}a\\ =\frac{1}{\absolutevalue{G}^{2}}\sum_{a\in G}\overline{\chi^{\prime}}(a)\,a~{}\sum_{g\in G}(\overline{\chi}\,\chi^{\prime})(g). (4)

The third equality comes from collecting terms that contain only aa and only gg. By the first orthogonality relation in Equation (1), one arrives at

LχLχ=1|G|aGχ¯(a)a={0, if χχ,Lχ, by definition.L_{\chi}L_{\chi^{\prime}}=\frac{1}{\absolutevalue{G}}\sum_{a\in G}\overline{\chi^{\prime}}(a)a=\begin{cases}0\mbox{, if }\chi\neq\chi^{\prime},\\ L_{\chi}\mbox{, by definition.}\end{cases}

We verify the second assertion by using the second orthogonality relation in Equation (1). Since χ¯(1)=1\overline{\chi}(1)=1, we obtain

χG^Lχ=1|G|χG^gGχ¯(g)g=1|G|gGgχG^χ¯(g)χ¯(1)=1.\sum_{\chi\in\widehat{G}}L_{\chi}=\frac{1}{\absolutevalue{G}}\sum_{\chi\in\widehat{G}}\sum_{g\in G}\overline{\chi}(g)\,g=\frac{1}{\absolutevalue{G}}\sum_{g\in G}g\sum_{\chi\in\widehat{G}}\overline{\chi}(g)\,\overline{\chi}(1)=1.

Using gh=ag\,h=a, the definition of LχL_{\chi}, and the equality χ¯(g1)=χ(g)\overline{\chi}(g^{-1})=\chi(g), one gets

gLχ\displaystyle g\,L_{\chi} =1|G|hGχ¯(h)gh=1|G|aGχ¯(ag1)a\displaystyle=\frac{1}{\absolutevalue{G}}\sum_{h\in G}\overline{\chi}(h)\,g\,h=\frac{1}{\absolutevalue{G}}\sum_{a\in G}\overline{\chi}(a\cdot g^{-1})\,a
=1|G|χ¯(g1)aGχ¯(a)a=χ(g)Lχ.\displaystyle=\frac{1}{\absolutevalue{G}}\overline{\chi}(g^{-1})\sum_{a\in G}\overline{\chi}(a)\,a=\chi(g)\,L_{\chi}.

Proposition 3.2.

For each χG^\chi\in\widehat{G}, let V(χ):=LχV={Lχ(|𝐯):|𝐯V}V(\chi):=L_{\chi}V=\{L_{\chi}(\ket{{\mathbf{v}}}):\ket{{\mathbf{v}}}\in V\}. For |𝐯V(χ)\ket{{\mathbf{v}}}\in V(\chi) and gGg\in G, we have g|𝐯=χ(g)|𝐯g\ket{{\mathbf{v}}}=\chi(g)\,\ket{{\mathbf{v}}}. Thus, V(χ)V(\chi) is a common eigenspace of all operators in GG. A direct decomposition V=χG^V(χ)V=\bigoplus_{\chi\in\widehat{G}}V(\chi) ensures that each |𝐯V\ket{{\mathbf{v}}}\in V has a unique expression

|𝐯=χG^|𝐯χ, where |𝐯χV(χ).\ket{{\mathbf{v}}}=\sum_{\chi\in\widehat{G}}\ket{{\mathbf{v}}}_{\chi}\mbox{, where }\ket{{\mathbf{v}}}_{\chi}\in V(\chi).
Proof.

For |𝐯V(χ)\ket{{\mathbf{v}}}\in V(\chi) and gGg\in G, there exists |𝐰V\ket{{\mathbf{w}}}\in V such that

g|𝐯=gLχ(|𝐰)=χ(g)Lχ(|𝐰)=χ(g)|𝐯,g\ket{{\mathbf{v}}}=g\,L_{\chi}(\ket{{\mathbf{w}}})=\chi(g)\,L_{\chi}(\ket{{\mathbf{w}}})=\chi(g)\ket{{\mathbf{v}}},

confirming the first assertion.

For each |𝐯V\ket{{\mathbf{v}}}\in V, we write |𝐯=(χG^Lχ)|𝐯=χG^|𝐯χ\ket{{\mathbf{v}}}=\left(\sum_{\chi\in\widehat{G}}L_{\chi}\right)\ket{{\mathbf{v}}}=\sum_{\chi\in\widehat{G}}\ket{{\mathbf{v}}}_{\chi}, where |𝐯χ:=Lχ(|𝐯)V(χ)\ket{{\mathbf{v}}}_{\chi}:=L_{\chi}(\ket{{\mathbf{v}}})\in V(\chi). On the other hand, if |𝐯=χG^|𝐮χ\ket{{\mathbf{v}}}=\sum_{\chi\in\widehat{G}}\ket{{\mathbf{u}}_{\chi}} for |𝐮χV(χ)\ket{{\mathbf{u}}_{\chi}}\in V(\chi), then |𝐮χ=Lχ|𝐰χ\ket{{\mathbf{u}}_{\chi}}=L_{\chi}\ket{{\mathbf{w}}_{\chi}} for some |𝐰χV\ket{{\mathbf{w}}_{\chi}}\in V. Since {Lχ:χG^}\{L_{\chi}:\chi\in\widehat{G}\} has the orthogonality property, for every χG^\chi\in\widehat{G}, we have

|𝐯χ\displaystyle\ket{{\mathbf{v}}}_{\chi} =Lχ|𝐯=Lχ(χG^|𝐮χ)=Lχ(χG^Lχ|𝐰χ)\displaystyle=L{\chi}~{}\ket{{\mathbf{v}}}=L_{\chi}\left(\sum_{\chi^{\prime}\in\widehat{G}}\ket{{\mathbf{u}}_{\chi^{\prime}}}\right)=L_{\chi}\left(\sum_{\chi^{\prime}\in\widehat{G}}L_{\chi^{\prime}}\ket{{\mathbf{w}}_{\chi^{\prime}}}\right)
=χG^(LχLχ|𝐰χ)=Lχ|𝐰χ=|𝐮χ.\displaystyle=\sum_{\chi^{\prime}\in\widehat{G}}\left(L_{\chi}L_{\chi^{\prime}}\ket{{\mathbf{w}}_{\chi^{\prime}}}\right)=L_{\chi}\ket{{\mathbf{w}}_{\chi}}=\ket{{\mathbf{u}}_{\chi}}.

Thus, V=χG^V(χ)V=\bigoplus_{\chi\in\widehat{G}}V(\chi). ∎

It is also a well-known fact that V(χ)V(\chi) and V(χ)V(\chi^{\prime}) are Hermitian orthogonal for all χχG^\chi\neq\chi^{\prime}\in\widehat{G} when all gGg\in G are unitary linear operators on VV.

All the tools to connect qubit stabilizer codes to classical codes are now in place. We choose G:=g1,g2,,gkG:=\left\langle g_{1},g_{2},\ldots,g_{k}\right\rangle to be an abelian subgroup of n{\mathcal{E}}_{n} with gj:=iλjX(𝐚j)Z(𝐛j)g_{j}:=i^{\lambda_{j}}~{}X({\mathbf{a}}_{j})~{}Z({\mathbf{b}}_{j}) for 1jk1\leq j\leq k, where 𝐚j,𝐛j𝔽2n{\mathbf{a}}_{j},{\mathbf{b}}_{j}\in\mathbb{F}_{2}^{n} and λj𝐚j𝐛j(mod2)\lambda_{j}\equiv{\mathbf{a}}_{j}\cdot{\mathbf{b}}_{j}\pmod{2}. Since σx\sigma_{x} and σz\sigma_{z} are Hermitian unitary matrices, X(𝐚j)X({\mathbf{a}}_{j}) and Z(𝐛j)Z({\mathbf{b}}_{j}) are also Hermitian matrices. The basis element gjg_{j} is Hermitian since

gj:=g¯j\displaystyle g_{j}^{\dagger}:=\overline{g}_{j}^{\top} =(i)λjZ(𝐛j)X(𝐚j)=(i)λjZ(𝐛j)X(𝐚j)\displaystyle=(-i)^{\lambda_{j}}~{}Z({\mathbf{b}}_{j})^{\top}~{}X({\mathbf{a}}_{j})^{\top}=(-i)^{\lambda_{j}}~{}Z({\mathbf{b}}_{j})~{}X({\mathbf{a}}_{j})
=iλj(1)𝐚j𝐛j(1)𝐚j𝐛jX(𝐚j)Z(𝐛j)=iλjX(𝐚j)Z(𝐛j).\displaystyle=i^{\lambda_{j}}~{}(-1)^{{\mathbf{a}}_{j}\cdot{\mathbf{b}}_{j}}~{}(-1)^{{\mathbf{a}}_{j}\cdot{\mathbf{b}}_{j}}~{}X({\mathbf{a}}_{j})~{}Z({\mathbf{b}}_{j})=i^{\lambda_{j}}~{}X({\mathbf{a}}_{j})~{}Z({\mathbf{b}}_{j}). (5)
Theorem 3.3.

Let CC be an nkn-k-dimensional self-orthogonal subspace of 𝔽22n\mathbb{F}_{2}^{2n} under the symplectic inner product. Let d:=wQ(CsC)=min{wQ(𝐯):𝐯CsC}d:=w_{Q}\left(C^{\perp_{s}}\setminus C\right)=\min\{w_{Q}({\mathbf{v}}):{\mathbf{v}}\in C^{\perp_{s}}\setminus C\}. Then there is an n,k,d\llbracket n,k,d\rrbracket-qubit stabilizer code QQ.

Proof.

We lift C:=G¯n¯C:=\overline{G}\in\overline{{\mathcal{E}}_{n}} to an abelian subgroup GG of n{\mathcal{E}}_{n}, with G𝔽2nkG\cong\mathbb{F}_{2}^{n-k}. Then 2n=χG^Q(χ)\mathbb{C}^{2^{n}}=\bigoplus_{\chi\in\widehat{G}}Q(\chi), where Q(χ)=Lχ2nQ(\chi)=L_{\chi}\mathbb{C}^{2^{n}}, is the subspace

{|𝐯2n:g|𝐯=χ(g)|𝐯 for all gG}.\{\ket{{\mathbf{v}}}\in\mathbb{C}^{2^{n}}:g\ket{{\mathbf{v}}}=\chi(g)\,\ket{{\mathbf{v}}}\mbox{ for all }g\in G\}.

Showing that each Q(χ)Q(\chi) is an n,k,d\llbracket n,k,d\rrbracket-qubit code means proving

dimQ(χ)=2k and d(Q(χ))wQ(CsC).\dim_{\mathbb{C}}Q(\chi)=2^{k}\mbox{ and }d(Q(\chi))\geq w_{Q}\left(C^{\perp_{s}}\setminus C\right).

Consider the action of n{\mathcal{E}}_{n} on {Q(χ):χG^}\{Q(\chi):\chi\in\widehat{G}\}. For any |𝐯Q(χ)\ket{{\mathbf{v}}}\in Q(\chi) and gGg\in G, we have g|𝐯=χ(g)|𝐯g\ket{{\mathbf{v}}}=\chi(g)\,\ket{{\mathbf{v}}}. Thus, for any En{\rm E}\in{\mathcal{E}}_{n} and any gGg\in G,

g(E|𝐯)=(1)g¯,E¯sE(g|𝐯)=(1)g¯,E¯sχ(g)E|𝐯.g({\rm E}\ket{{\mathbf{v}}})=(-1)^{\langle\overline{g},\overline{{\rm E}}\rangle_{s}}\,{\rm E}(g\ket{{\mathbf{v}}})=(-1)^{\langle\overline{g},\overline{{\rm E}}\rangle_{s}}\,\chi(g)\,{\rm E}\ket{{\mathbf{v}}}.

Since χE¯:G{±1}\chi_{\overline{{\rm E}}}:G\mapsto\{\pm 1\} and χE¯(g)=(1)g¯,E¯s\chi_{\overline{{\rm E}}}(g)=(-1)^{\langle\overline{g},\overline{{\rm E}}\rangle_{s}} is a character of GG, we have

g(E|𝐯)=χE¯(g)χ(g)E|𝐯=χ(g)E|𝐯, for all gG.g({\rm E}\ket{{\mathbf{v}}})=\chi_{\overline{{\rm E}}}(g)\,\chi(g)\,{\rm E}\ket{{\mathbf{v}}}=\chi^{\prime}(g)\,{\rm E}\ket{{\mathbf{v}}}\mbox{, for all }g\in G.

This implies E|𝐯Q(χ){\rm E}\ket{{\mathbf{v}}}\in Q(\chi^{\prime}) and E:Q(χ)Q(χ){\rm E}:Q(\chi)\mapsto Q(\chi^{\prime}), where χ:=χE¯χ\chi^{\prime}:=\chi_{\overline{{\rm E}}}\,\chi. Since n{\mathcal{E}}_{n} is a group, E{\rm E} is a bijection, making dimQ(χ)=dimQ(χ)\dim_{\mathbb{C}}Q(\chi)=\dim_{\mathbb{C}}Q(\chi^{\prime}). As E{\rm E} runs through n{\mathcal{E}}_{n}, χE¯\chi_{\overline{{\rm E}}} takes all characters of GG, ensuring that dimQ(χ)\dim_{\mathbb{C}}Q(\chi) is the same for all χG^\chi\in\widehat{G}. Thus, dimQ(χ)=2n(nk)=2k\dim_{\mathbb{C}}Q(\chi)=2^{n-(n-k)}=2^{k} for any χG^\chi\in\widehat{G}.

We now show that, if Ed1{\rm E}\in{\mathcal{E}}_{d-1} and |𝐯1,|𝐯2Q(χ)\ket{{\mathbf{v}}_{1}},\ket{{\mathbf{v}}_{2}}\in Q(\chi) with 𝐯1|𝐯2=0\innerproduct{{\mathbf{v}}_{1}}{{\mathbf{v}}_{2}}=0, then 𝐯1|E1E2|𝐯2=0\bra{{\mathbf{v}}_{1}}{\rm E}_{1}\,{\rm E}_{2}\ket{{\mathbf{v}}_{2}}=0, where E:=E1E2{\rm E}:={\rm E}_{1}\,{\rm E}_{2}. If E¯G¯=C\overline{{\rm E}}\in\overline{G}=C, then 𝐯1|E|𝐯2=χ(E¯)(𝐯1|𝐯2)=0\bra{{\mathbf{v}}_{1}}{\rm E}\ket{{\mathbf{v}}_{2}}=\chi(\overline{{\rm E}})(\innerproduct{{\mathbf{v}}_{1}}{{\mathbf{v}}_{2}})=0. Otherwise, E¯C\overline{{\rm E}}\notin C. From wQ(E¯)=wQ(E)d1w_{Q}(\overline{{\rm E}})=w_{Q}({\rm E})\leq d-1 and the assumption (CsC)¯d1=\left(C^{\perp_{s}}\setminus C\right)\cap\overline{{\mathcal{E}}}_{d-1}=\emptyset, we know E¯Cs\overline{{\rm E}}\notin C^{\perp_{s}}. Hence, there exists E¯G¯\overline{{\rm E}}^{\prime}\in\overline{G} such that EE=EE{\rm E}\,{\rm E}^{\prime}=-{\rm E}^{\prime}\,{\rm E}. Then, for |𝐯2Q(χ)\ket{{\mathbf{v}}_{2}}\in Q(\chi), we have

EE|𝐯2=EE|𝐯2=χ(E)E|𝐯2, with χ(E)χ(E).{\rm E}^{\prime}\,{\rm E}\ket{{\mathbf{v}}_{2}}=-{\rm E}\,{\rm E}^{\prime}\ket{{\mathbf{v}}_{2}}=-\chi({\rm E}^{\prime}){\rm E}\ket{{\mathbf{v}}_{2}}\mbox{, with }-\chi({\rm E}^{\prime})\neq\chi({\rm E}^{\prime}).

Therefore, E|𝐯2Q(χ){\rm E}\ket{{\mathbf{v}}_{2}}\in Q(\chi^{\prime}), with χχ\chi^{\prime}\neq\chi. Since |𝐯1Q(χ)\ket{{\mathbf{v}}_{1}}\in Q(\chi) and Q(χ)Q(\chi) is orthogonal to Q(χ)Q(\chi^{\prime}), we confirm 𝐯1|E|𝐯2=0\bra{{\mathbf{v}}_{1}}{\rm E}\ket{{\mathbf{v}}_{2}}=0. ∎

Example 3.

We exhibit a 5,1,3\llbracket 5,1,3\rrbracket-qubit stabilizer code QQ. Consider a subspace C𝔽210C\subset\mathbb{F}_{2}^{10} with generator matrix

[1100000101011001001000110010010001110100]=[𝐯1𝐯2𝐯3𝐯4].\begin{bmatrix}1&1&0&0&0&0&0&1&0&1\\ 0&1&1&0&0&1&0&0&1&0\\ 0&0&1&1&0&0&1&0&0&1\\ 0&0&0&1&1&1&0&1&0&0\end{bmatrix}=\begin{bmatrix}{\mathbf{v}}_{1}\\ {\mathbf{v}}_{2}\\ {\mathbf{v}}_{3}\\ {\mathbf{v}}_{4}\end{bmatrix}.

One reads 𝐯1=(𝐚|𝐛){\mathbf{v}}_{1}=({\mathbf{a}}|{\mathbf{b}}) as having 𝐚=(1,1,0,0,0){\mathbf{a}}=(1,1,0,0,0) and 𝐛=(0,0,1,0,1){\mathbf{b}}=(0,0,1,0,1). The code CC is symplectic self-orthogonal, with dim𝔽2C=4\dim_{\mathbb{F}_{2}}C=4 and dim𝔽2C=6\dim_{\mathbb{F}_{2}}C^{\perp}=6, i.e., the codimension is 22. To extend the basis for CC to a basis for CsC^{\perp_{s}} we use (0,0,0,0,0, 1,1,1,1,1)(0,0,0,0,0,\,1,1,1,1,1) and (1,1,1,1,1, 0,0,0,0,0)(1,1,1,1,1,\,0,0,0,0,0). Since wQ(C)=4w_{Q}(C)=4 and wQ(Cs)=3w_{Q}(C^{\perp_{s}})=3, one obtains wQ(CsC)=3w_{Q}(C^{\perp_{s}}\setminus C)=3. We can write the 5,1,3\llbracket 5,1,3\rrbracket-code Q=Q(χ0)Q=Q(\chi_{0}) explicitly by using G=g1,g2,g3,g4G=\left\langle g_{1},g_{2},g_{3},g_{4}\right\rangle, with g1=σxσxσzI2σzg_{1}=\sigma_{x}\otimes\sigma_{x}\otimes\sigma_{z}\otimes I_{2}\otimes\sigma_{z}, g2=σzσxσxσzI2g_{2}=\sigma_{z}\otimes\sigma_{x}\otimes\sigma_{x}\otimes\sigma_{z}\otimes I_{2}, g3=I2σzσxσxσzg_{3}=I_{2}\otimes\sigma_{z}\otimes\sigma_{x}\otimes\sigma_{x}\otimes\sigma_{z}, and g4=σzI2σzσxσxg_{4}=\sigma_{z}\otimes I_{2}\otimes\sigma_{z}\otimes\sigma_{x}\otimes\sigma_{x}.

Since k=1k=1 and dimQ=2k=2\dim_{\mathbb{C}}Q=2^{k}=2, two independent vectors in 32\mathbb{C}^{32} form a basis of Q={|𝐯32:g|𝐯=χ0(g)|𝐯gG}Q=\{\ket{{\mathbf{v}}}\in\mathbb{C}^{32}:g\ket{{\mathbf{v}}}=\chi_{0}(g)\ket{{\mathbf{v}}}\,\forall g\in G\}. QQ consists of vectors which are fixed by all gGg\in G. After some computation, we conclude that QQ can be generated by |𝐯0=gGg|00000\ket{{\mathbf{v}}_{0}}=\sum_{g\in G}g\ket{00000} and |𝐯1=gGg|11111\ket{{\mathbf{v}}_{1}}=\sum_{g\in G}g\ket{11111}.  ∎

With minor modifications, the qubit stabilizer formalism extends to the general qudit case. A complete treatment is available in [31]. We outline the main steps here. An nn-qudit system is a nonzero element in (q)nqn\left(\mathbb{C}^{q}\right)^{\otimes n}\cong\mathbb{C}^{q^{n}}. Let 𝐚=(a1,,an)𝔽qn{\mathbf{a}}=(a_{1},\ldots,a_{n})\in\mathbb{F}_{q}^{n}. The standard \mathbb{C}-basis is

{|a1a2an:=|a1|a2|an:𝐚𝔽qn}\{\ket{a_{1}a_{2}\ldots a_{n}}:=\ket{a_{1}}\otimes\ket{a_{2}}\otimes\ldots\otimes\ket{a_{n}}:{\mathbf{a}}\in\mathbb{F}_{q}^{n}\}

and an arbitrary vector in qn\mathbb{C}^{q^{n}} is written |ψ=𝐚𝔽qnc𝐚|𝐚\ket{\psi}=\sum_{{\mathbf{a}}\in\mathbb{F}_{q}^{n}}c_{{\mathbf{a}}}\ket{{\mathbf{a}}}, with c𝐚c_{{\mathbf{a}}}\in\mathbb{C} and qn𝐚𝔽qnc𝐚2=1q^{-n}\sum_{{\mathbf{a}}\in\mathbb{F}_{q}^{n}}\norm{c_{{\mathbf{a}}}}^{2}=1.

Let 𝐚=(a1,,an),𝐛=(b1,,bn)𝔽qn{\mathbf{a}}=(a_{1},\ldots,a_{n}),{\mathbf{b}}=(b_{1},\ldots,b_{n})\in\mathbb{F}_{q}^{n}, and ω:=e2πip\omega:=e^{\frac{2\pi i}{p}}, where q=pmq=p^{m}. The error operators form n:={ωβX(𝐚)Z(𝐛):𝐚,𝐛𝔽qn,β𝔽p}{\mathcal{E}}_{n}:=\{\omega^{\beta}X({\mathbf{a}})Z({\mathbf{b}}):{\mathbf{a}},{\mathbf{b}}\in\mathbb{F}_{q}^{n},\,\beta\in\mathbb{F}_{p}\} of cardinality pq2npq^{2n}. The respective actions of X(𝐚)X({\mathbf{a}}) and Z(𝐛)Z({\mathbf{b}}) on |𝐯qn\ket{{\mathbf{v}}}\in\mathbb{C}^{q^{n}} are X(𝐚)|𝐯=|𝐚+𝐯X({\mathbf{a}})\ket{{\mathbf{v}}}=\ket{{\mathbf{a}}+{\mathbf{v}}} and Z(𝐛)|𝐯=(ω)Tr(𝐛𝐯)|𝐯Z({\mathbf{b}})\ket{{\mathbf{v}}}=(\omega)^{\Tr({\mathbf{b}}\cdot{\mathbf{v}})}\ket{{\mathbf{v}}}. Hence, for E:=ωβX(𝐚)Z(𝐛){\rm E}:=\omega^{\beta}X({\mathbf{a}})Z({\mathbf{b}}) and E:=ωβX(𝐚)Z(𝐛){\rm E}^{\prime}:=\omega^{\beta^{\prime}}X({\mathbf{a}}^{\prime})Z({\mathbf{b}}^{\prime}) in n{\mathcal{E}}_{n}, one gets EE=ωTr(𝐛𝐚𝐛𝐚)EE{\rm E}\,{\rm E}^{\prime}=\omega^{\Tr({\mathbf{b}}\cdot{\mathbf{a}}^{\prime}-{\mathbf{b}}^{\prime}\cdot{\mathbf{a}})}\,{\rm E}^{\prime}\,{\rm E}. The symplectic weight of (𝐚|𝐛)({\mathbf{a}}|{\mathbf{b}}) is the quantum weight of E{\rm E}.

The (trace) symplectic inner product of (𝐚|𝐛)({\mathbf{a}}|{\mathbf{b}}) and (𝐚|𝐛)({\mathbf{a}}^{\prime}|{\mathbf{b}}^{\prime}) in 𝔽q2n\mathbb{F}_{q}^{2n} is

(𝐚|𝐛),(𝐚|𝐛)s=Tr(𝐛𝐚𝐛𝐚).\langle({\mathbf{a}}|{\mathbf{b}}),({\mathbf{a}}^{\prime}|{\mathbf{b}}^{\prime})\rangle_{s}=\Tr({\mathbf{b}}\cdot{\mathbf{a}}^{\prime}-{\mathbf{b}}^{\prime}\cdot{\mathbf{a}}). (6)

The symplectic dual of C𝔽q2nC\subseteq\mathbb{F}_{q}^{2n} is Cs={𝐮𝔽q2n:𝐮,𝐜s=0𝐜C}C^{\perp_{s}}=\{{\mathbf{u}}\in\mathbb{F}_{q}^{2n}:\langle{\mathbf{u}},{\mathbf{c}}\rangle_{s}=0\,\forall\,{\mathbf{c}}\in C\}. As in the qubit case, in the general qudit setup, a subgroup GG of n{\mathcal{E}}_{n} is abelian if and only if G¯\overline{G} is a symplectic self-orthogonal subspace of n¯𝔽q2n\overline{{\mathcal{E}}_{n}}\cong\mathbb{F}_{q}^{2n}. The analogue of Theorem 3.3 follows.

Theorem 3.4.

Let CC be an nkn-k-dimensional self-orthogonal subspace of 𝔽q2n\mathbb{F}_{q}^{2n} under the (trace) symplectic inner product. Let d:=wQ(CsC)=min{wQ(𝐯):𝐯CsC}d:=w_{Q}\left(C^{\perp_{s}}\setminus C\right)=\min\{w_{Q}({\mathbf{v}}):{\mathbf{v}}\in C^{\perp_{s}}\setminus C\}. Then there is an n,k,d\llbracket n,k,d\rrbracket-qudit stabilizer code QQ.

{VF}

“With group and eigenstate, we’ve learned to fix
Your quantum errors with our quantum tricks.”

\VA

Daniel Gottesmanin Quantum Error Correction Sonnet

4 Constructions via Classical Codes

Any stabilizer code QQ is fully characterized by its stabilizer group, that specifies the set of errors that QQ can correct. Any linear combination of the operators in the error set is correctable, allowing QQ to correct a continuous set of operators. For this reason, the best-known qubit codes in the online table [24] maintained by M. Grassl are given in terms of their stabilizer generators. The stabilizer approach has massive advantages over other frameworks, some of which will be mentioned below. It describes a large set of QECs, complete with their encoding and decoding mechanism, in a very compact form.

A valid codeword of QQ is a +1+1 eigenvector of all the stabilizer generators. An error E{\rm E}, expressed as a tensor product of Pauli operators, anticommutes with some of the stabilizer generators and commutes with others. It sends a codeword to an eigenstate of the stabilizer generators. The eigenvalue remains +1+1 for all operators that commute with E{\rm E} but becomes 1-1 for those generators that anticommute with E{\rm E}. From the resulting error syndrome, one knows which Pauli operators acts on which qubits. Applying the respective Pauli operators on the corresponding locations corrects the error. Suppose that the location of error is known, but the type is not, then this is a quantum erasure. By the Knill-Laflamme condition, correcting \ell general errors means correcting 22\ell erasures.

The encoding and syndrome reading circuits can be written using only three quantum gates, namely the Hadamard gate, the phase SS gate, and the CNOT gate, whose respective matrices are

H=12[1111],S=[100i],CNOT=[1000010000010010].{\rm H}=\frac{1}{\sqrt{2}}\begin{bmatrix}1&1\\ 1&-1\end{bmatrix},\quad{\rm S}=\begin{bmatrix}1&0\\ 0&i\end{bmatrix},\quad{\rm CNOT}=\begin{bmatrix}1&0&0&0\\ 0&1&0&0\\ 0&0&0&1\\ 0&0&1&0\end{bmatrix}.

A treatment on the circuit implementations is available, e.g., in [25].

We now look into suitable classical codes that fully describe the set of correctable errors. All constructions are applications of the stabilizer formalism. Since all of the inner products used are nondegenerate, i.e., (C)=C(C^{\perp})^{\perp}=C, one can interchange self-orthogonality and dual-containment, provided that the derived parameters are adjusted accordingly.

First, we consider a generic construction of qq-ary quantum codes via additive (i.e., 𝔽q\mathbb{F}_{q}-linear) codes over 𝔽q2\mathbb{F}_{q^{2}}. Let {1,γ}\{1,\gamma\} be a basis of 𝔽q2\mathbb{F}_{q^{2}} over 𝔽q\mathbb{F}_{q}. The map Φ:n¯𝔽q2n𝔽q2n\Phi:\overline{{\mathcal{E}}_{n}}\cong\mathbb{F}_{q}^{2n}\mapsto\mathbb{F}_{q^{2}}^{n} that sends E¯:=𝐯=(𝐚|𝐛)=(a1,,an|b1,,bn)\overline{{\rm E}}:={\mathbf{v}}=({\mathbf{a}}|{\mathbf{b}})=(a_{1},\ldots,a_{n}|b_{1},\ldots,b_{n}) to (a1+γb1,,an+γbn)(a_{1}+\gamma b_{1},\ldots,a_{n}+\gamma b_{n}) is an isomorphism of 𝔽q\mathbb{F}_{q}-vector spaces. It is also an isometry, since wQ(E)=wH(Φ((𝐚|𝐛)))\operatorname{w_{Q}}({\rm E})=\operatorname{w_{H}}(\Phi(({\mathbf{a}}|{\mathbf{b}}))). For any 𝐮:=(u1,,un){\mathbf{u}}:=(u_{1},\ldots,u_{n}) and 𝐯:=(v1,,vn)𝔽q2n{\mathbf{v}}:=(v_{1},\ldots,v_{n})\in\mathbb{F}_{q^{2}}^{n}, we use 𝐮q{\mathbf{u}}^{q} to denote (u1q,,unq)(u_{1}^{q},\ldots,u_{n}^{q}) and define the trace alternating inner product of 𝐮{\mathbf{u}} and 𝐯{\mathbf{v}} as 𝐮,𝐯alt:=Tr𝔽q2/𝔽q(𝐮𝐯q𝐮q𝐯γγq)\displaystyle{\langle{\mathbf{u}},{\mathbf{v}}\rangle_{\rm alt}:=\Tr_{\mathbb{F}_{q^{2}}/\mathbb{F}_{q}}\left(\frac{{\mathbf{u}}\cdot{\mathbf{v}}^{q}-{\mathbf{u}}^{q}\cdot{\mathbf{v}}}{\gamma-\gamma^{q}}\right)}. When q=2q=2, 𝐮,𝐯alt\langle{\mathbf{u}},{\mathbf{v}}\rangle_{\rm alt} coincides with the trace Hermitian inner product 𝐮,𝐯TrH:=j=1n(ujvj2+uj2vj)\langle{\mathbf{u}},{\mathbf{v}}\rangle_{\Tr_{\mathop{{\rm H}}}}:=\sum_{j=1}^{n}\left(u_{j}v_{j}^{2}+u_{j}^{2}v_{j}\right), since γγ2=1\gamma-\gamma^{2}=1. For any (𝐚,𝐛)({\mathbf{a}},{\mathbf{b}}) and (𝐚,𝐛)({\mathbf{a}}^{\prime},{\mathbf{b}}^{\prime}) in n¯\overline{{\mathcal{E}}_{n}}, we immediately verify that (𝐚,𝐛),(𝐚,𝐛)s=Φ((𝐚,𝐛)),Φ((𝐚,𝐛)alt\langle({\mathbf{a}},{\mathbf{b}}),({\mathbf{a}}^{\prime},{\mathbf{b}}^{\prime})\rangle_{s}=\langle\Phi(({\mathbf{a}},{\mathbf{b}})),\Phi(({\mathbf{a}}^{\prime},{\mathbf{b}}^{\prime})\rangle_{\rm alt}. Hence, a linear code C𝔽q2nC\subseteq\mathbb{F}_{q}^{2n} is symplectic self-orthogonal if and only if the additive code Φ(C)\Phi(C) is trace alternating self-orthogonal. The Hermitian inner product of any 𝐮,𝐯𝔽q2n{\mathbf{u}},{\mathbf{v}}\in\mathbb{F}_{q^{2}}^{n} is 𝐮,𝐯H:=j=1nujvjq\langle{\mathbf{u}},{\mathbf{v}}\rangle_{\mathop{{\rm H}}}:=\sum_{j=1}^{n}u_{j}v_{j}^{q}. If Φ(C)\Phi(C) is 𝔽q2\mathbb{F}_{q^{2}}-linear, instead of being strictly additive, then Φ(C)(Φ(C))alt\Phi(C)\subseteq(\Phi(C))^{\perp_{\rm alt}} if and only if Φ(C)(Φ(C))H\Phi(C)\subseteq(\Phi(C))^{\perp_{\mathop{{\rm H}}}}. Thus, Theorem 3.4 has the following equivalent statement.

Theorem 4.1.

Let 𝒞𝔽q2n\mathcal{C}\subseteq\mathbb{F}_{q^{2}}^{n} be an 𝔽q\mathbb{F}_{q}-additive code such that 𝒞𝒞alt\mathcal{C}\subseteq\mathcal{C}^{\perp_{\rm alt}}, with |𝒞|=qnk\absolutevalue{\mathcal{C}}=q^{n-k}. Then there exists an n,k,dq\llbracket n,k,d\rrbracket_{q} quantum code QQ with

d(Q)=wH(𝒞alt𝒞)=min{wH(𝐯):𝐯𝒞alt𝒞}.d(Q)=\operatorname{w_{H}}(\mathcal{C}^{\perp_{\rm alt}}\setminus\mathcal{C})=\min\{\operatorname{w_{H}}({\mathbf{v}}):{\mathbf{v}}\in\mathcal{C}^{\perp_{\rm alt}}\setminus\mathcal{C}\}.

If 𝒞\mathcal{C} is 𝔽q2\mathbb{F}_{q^{2}}-linear, we can conveniently replace the trace alternating inner product by the Hermitian inner product, which is easier to compute.

If CC is 𝔽4\mathbb{F}_{4}-additive and is even, i.e., wH(𝐜)\operatorname{w_{H}}({\mathbf{c}}) is even for all 𝐜C{\mathbf{c}}\in C, then CC is trace Hermitian self-orthogonal. If CC is trace Hermitian self-orthogonal and CC is 𝔽4\mathbb{F}_{4}-linear, then CC is an even code.

The quantum codes in Theorem 4.1 are modeled after classical codes with an additive structure, but the error operators are in fact multiplicative. An error E{\rm E} may have the same effect as ES{\rm E}{\rm S} where SI{\rm S}\neq I is an element of the stabilizer group. A QEC is degenerate or impure if the set of correctable errors contains degenerate errors. Studies on impure codes has been rather scarce. The existence of two inequivalent 6,1,3\llbracket 6,1,3\rrbracket impure qubit codes was shown in [8, Section IV]. Remarkably, there is no 6,1,3\llbracket 6,1,3\rrbracket pure qubit code. A systematic construction based on duadic codes and further discussion on the advantages of degenerate quantum codes are supplied in [3].

A very popular construction is based on nested classical codes. We denote the Euclidean dual of CC by CEC^{\perp_{\rm{E}}}.

Theorem 4.2 (Calderbank-Shor and Steane (CSS) Construction).

Let CjC_{j} be an [n,kj,dj]q[n,k_{j},d_{j}]_{q}-code for j{1,2}j\in\{1,2\} with C1EC2C_{1}^{\perp_{\mathop{{\rm E}}}}\subseteq C_{2}. Then there is an

n,k1+k2n,min{wH(C2C1E),wH(C1C2E)}q-code Q.\llbracket n,k_{1}+k_{2}-n,\min\{\operatorname{w_{H}}(C_{2}\setminus C_{1}^{\perp_{\mathop{{\rm E}}}}),\operatorname{w_{H}}(C_{1}\setminus C_{2}^{\perp_{\mathop{{\rm E}}}})\}\rrbracket_{q}\mbox{-code }Q.

The code is pure whenever min{wH(C2C1E),wH(C1C2E)}=min{d1,d2}\min\{\operatorname{w_{H}}(C_{2}\setminus C_{1}^{\perp_{\mathop{{\rm E}}}}),\operatorname{w_{H}}(C_{1}\setminus C_{2}^{\perp_{\mathop{{\rm E}}}})\}=\min\{d_{1},d_{2}\}.

Proof.

Let GjG_{j} and HjH_{j} be the generator and parity-check matrices of CjC_{j}. Consider the linear code C𝔽q2nC\subseteq\mathbb{F}_{q}^{2n} with generator matrix [H1𝟎𝟎H2]\displaystyle{\begin{bmatrix}H_{1}&{\mathbf{0}}\\ {\mathbf{0}}&H_{2}\end{bmatrix}}. Since C1EC2C_{1}^{\perp_{\mathop{{\rm E}}}}\subseteq C_{2}, we have H1H2=𝟎H_{1}H_{2}^{\top}={\mathbf{0}}. Similarly, from C2EC1C_{2}^{\perp_{\mathop{{\rm E}}}}\subseteq C_{1}, we know H2H1=𝟎H_{2}H_{1}^{\top}={\mathbf{0}}. Define CsC^{\perp_{s}} to be the code with parity check and generator matrices, respectively, [H2𝟎𝟎H1]\displaystyle{\begin{bmatrix}H_{2}&{\mathbf{0}}\\ {\mathbf{0}}&H_{1}\end{bmatrix}} and [G2𝟎𝟎G1]\displaystyle{\begin{bmatrix}G_{2}&{\mathbf{0}}\\ {\mathbf{0}}&G_{1}\end{bmatrix}}. We verify that CCsC\subseteq C^{\perp_{s}}, with dim𝔽qC=2n(k1+k2)\dim_{\mathbb{F}_{q}}C=2n-(k_{1}+k_{2}). By Theorem 3.4, dimQ=ndim𝔽qC=k1+k2n\dim_{\mathbb{C}}Q=n-\dim_{\mathbb{F}_{q}}C=k_{1}+k_{2}-n. The distance computation is clear. ∎

A special case of the CSS construction comes via a Euclidean dual-containing code CECC^{\perp_{\rm{E}}}\subseteq C. From such an [n,k,d]q[n,k,d]_{q}-code CC, one obtains an n,2kn,dq\llbracket n,2k-n,\geq d\rrbracket_{q}-code QQ. The next method allows for most qubit CSS codes to be enlarged while avoiding a significant drop in the distance. The choice of the extra vectors in the generator matrix of CC^{\prime} is detailed in [48, Section III].

Theorem 4.3 (Steane Enlargement of CSS Codes).

Let CC be an [n,k,d]2[n,k,d]_{2}-code that contains its Euclidean dual CECC^{\perp_{\mathop{{\rm E}}}}\subseteq C. Suppose that CC can be enlarged to C=[n,k>k+1,d]2C^{\prime}=[n,k^{\prime}>k+1,d^{\prime}]_{2}. Then there exists a pure qubit code of parameters n,k+kn,min{d,3d/2}\llbracket n,k+k^{\prime}-n,\min\{d,\lceil{3d^{\prime}/2\rceil}\}\rrbracket.

A generalization to the qudit case was subsequently given in [38], where the distance is min{d,q+1qd}\min\{d,\lceil{\frac{q+1}{q}\,d^{\prime}\rceil}\}. Comparing the minimum distances in the resulting codes, the enlargement offers a better chance of relative gain in the qubit case as compared with the q>2q>2 cases.

Lisoněk and Singh, inspired by the classical Construction X, proposed a modification to qubit stabilizer codes in [39]. The construction generalizes naturally to qudit codes.

Theorem 4.4 (Quantum Construction X).

For an [n,k]q2[n,k]_{q^{2}}-linear code CC, let e:=kdim(CCH)e:=k-\dim(C\cap C^{\perp_{\mathop{{\rm H}}}}). Then there exists an n+e,n2k+e,dq\llbracket n+e,n-2k+e,d\rrbracket_{q}-code QQ, with d:=d(Q)min{d(CH),d(C+CH)+1}d:=d(Q)\geq\min\{d(C^{\perp_{\mathop{{\rm H}}}}),d(C+C^{\perp_{\mathop{{\rm H}}}})+1\}, where C+CH:={𝐮+𝐯:𝐮C,𝐯CH}C+C^{\perp_{\mathop{{\rm H}}}}:=\{{\mathbf{u}}+{\mathbf{v}}:{\mathbf{u}}\in C,{\mathbf{v}}\in C^{\perp_{\mathop{{\rm H}}}}\}.

The case e=0e=0 is the usual stabilizer construction. To prevent a sharp drop in dd, we want small ee, i.e., large Hermitian hull CCHC\cap C^{\perp_{\mathop{{\rm H}}}}.

We shift our attention now to propagation rules and bounds. Most of them are direct consequences of the propagation rules and bounds on the classical codes used as ingredients in the above constructions.

Proposition 4.5 (see [8, Theorem 6] for the binary case).

From an n,k,dq\llbracket n,k,d\rrbracket_{q}-code, the following codes can be derived: an n,k1,dq\llbracket n,k-1,\geq d\rrbracket_{q}-code by subcode construction, an n+1,k,dq\llbracket n+1,k,\geq d\rrbracket_{q}-code by lengthening, and an n1,k,d1q\llbracket n-1,k,\geq d-1\rrbracket_{q}-code by puncturing.

The analogue of shortening is less straightforward. It requires the construction of an auxiliary code and, then, a check on whether this code has codewords of a given length. The details on how to shorten quantum codes are available in [26, Section 4], building upon the initial idea of Rains in [43].

How can we measure the goodness of a QEC? For stabilizer codes, given their classical ingredients and constructions, there are numerous bounds.

Theorem 4.6 (Quantum Hamming Bound, see [8] for the binary case).

Let QQ be a pure n,k,dq\llbracket n,k,d\rrbracket_{q}-code with d2+1d\geq 2\ell+1 and k>0k>0. Then

qnkj=0(q21)j(nj).q^{n-k}\geq\sum_{j=0}^{\ell}(q^{2}-1)^{j}\binom{n}{j}. (7)

QQ is perfect if it meets the bound.

The proof comes from the observation that

qnE¯n¯()dim(E¯Q)=dimQ|n¯()|=qkj=0(q21)j(nj).q^{n}\geq\sum_{\overline{{\rm E}}\in\overline{{\mathcal{E}}_{n}}(\ell)}\dim_{\mathbb{C}}(\overline{{\rm E}}\,Q)=\dim_{\mathbb{C}}Q\cdot\absolutevalue{\overline{{\mathcal{E}}_{n}}(\ell)}=q^{k}\sum_{j=0}^{\ell}(q^{2}-1)^{j}\binom{n}{j}.

The code in Example 3 is perfect. It has 2nk=16=j=013j(5j)=1+152^{n-k}=16=\sum_{j=0}^{1}3^{j}\binom{5}{j}=1+15.

Here is another bound which had been established as a necessary condition for pure stabilizer codes.

Theorem 4.7 (Quantum Gilbert-Varshamov Bound [18]).

Let n>k2n>k\geq 2, d2d\geq 2, and nk(mod2)n\equiv k\pmod{2}. A pure n,k,dq\llbracket n,k,d\rrbracket_{q}-code exists if

qnk+21q21>j=1d1(q21)j1(nj).\frac{q^{n-k+2}-1}{q^{2}-1}>\sum_{j=1}^{d-1}(q^{2}-1)^{j-1}\binom{n}{j}.

An upper bound, which is well-suited for computer search, is the quantum Linear Programming (LP) bound. In the qubit case, the bound is explained in details in [8, Section VII]. The same routine adjusts immediately to the the general qudit case, as was shown in [31, Section VI]. The main tools are the MacWilliams identities [40]. These are linear relations between the weight distribution of a classical code and its dual. They hold for all of the inner products we are concerned with here and have been very useful in ruling out the existence of quantum codes of certain ranges of parameters. Rains supplied a nice proof for the next bound, which is a corollary to the quantum LP bound, using the quantum weight enumerator in [43]. A quantum code that reaches the equality in the bound is said to be quantum MDS (QMDS).

Theorem 4.8 (Quantum Singleton Bound).

An n,k,dq\llbracket n,k,d\rrbracket_{q}-code with k>0k>0 satisfies kn2d+2k\leq n-2d+2.

Nearly all known families of classical codes over finite fields, especially those with well-studied algebraic and combinatorial structures, have been used in each of the constructions above. A partial list, compiled in mid 2005 as [31, Table II], already showed a remarkable breadth. The large family of cyclic-type codes, whose corresponding structures in the rings of polynomials are ideals, has been a rich source of ingredients for QECs with excellent parameters. This includes the BCH, cyclic, constacyclic, quasi-cyclic, and quasi-twisted codes. In the family, the nestedness property, very useful in the CSS construction, comes for free. A great deal is known about their dual codes under numerous applicable inner products. For small qq, the structures allow for extensive computer algebra searchers for reasonable lengths, aided by their minimum distance bounds.

The most comprehensive record for best-known qubit codes is Grassl’s online table [24]. Numerous entries have been certified optimal, while still more entries indicate gaps between the best-possible and the best-known. It is a two-fold challenge to contribute meaningfully to the table. First, for n100n\leq 100, many researchers have attempted exhaustive searches. Better codes are unlikely to be found without additional clever strategies. Second, for n>100n>100, computing the actual distance d(Q)d(Q) tends to be prohibitive. As the length and dimension grow, computing the minimum distances of the relevant classical codes to derive the quantum distance is hard [50]. Improvements remain possible, with targeted searches. Recent examples include the works of Galindo et al.on quasi-cyclic constructions of quantum codes [22], where Steane enlargement is deployed, the search reported in [39] on cyclic codes over 𝔽4\mathbb{F}_{4}, where Construction X is used with e{1,2,3}e\in\{1,2,3\}, and similar random searches on quasi-cyclic codes done in [16] for qubit and qutrit codes.

Less attention has been given to record-holding qutrit codes, for which there is no publicly available database of comparative extense. A table listing numerous qutrit codes is kept by Y. Edel in [13] based on their explicit construction as quantum twisted codes in [4]. Better codes than many of those in the table have since been found.

Attempts to derive new quantum codes by shortening good stabilizer codes motivate closer studies on the weight distribution of the classical auxiliary codes, in particular when the stabilizer codes are QMDS. Shortening is very effective in constructing qudit MDS codes of lengths up to q2+2q^{2}+2 and minimum distances up to q+1q+1.

There has been a large literature on QMDS. All of the above constructions via classical codes as well as the propagation rules have been applied to families of classical MDS codes, particularly the Generalized Reed-Solomon and the constacyclic MDS codes. Since the dual of an MDS code is MDS, the dual distance is evident, leaving only the orthogonality property to investigate. While the theoretical advantages are clearly abundant, there are practical limitations. The length of such codes is bounded above by q2+2q^{2}+2, when qq is even, and by q2+1q^{2}+1, when qq is odd, assuming the MDS conjecture.

For qubit codes, the only nontrivial QMDS are those with parameters 5,1,3\llbracket 5,1,3\rrbracket, 6,0,4\llbracket 6,0,4\rrbracket, and 2m,2m2,2\llbracket 2m,2m-2,2\rrbracket. As qq grows larger, the practical value of QMDS codes quickly diminishes, since controlling qudit systems with q>3q>3 is currently prohibitive. A list for qq-ary QMDS codes, with 2q172\leq q\leq 17, is available in [27]. Another list that covers families of QMDS codes and their references can be found in [10, Table V]. More works in QMDS continue to appear, with detailed analysis on the self-orthogonality conditions supplied from number theoretic and combinatorial tools.

Taking algebraic geometry (AG) codes as the classical ingredients is another route. A wealth of favourable results had already been available prior to the emergence of QECs. Curves with many rational points often lead, via the Goppa construction, to codes with good parameters. Their duals are well-understood, via the residue formula. Their designed distances can be computed from the Riemann-Roch theorem. Chen et al.showed how to combine Steane enlargement and concatenated AG codes to derive excellent qubit codes in [11]. A quantum asymptotic bound was established in [17]. Construction of QECs from AG codes was initially a very active line of inquiry. It had somewhat lessened in the last decade, mostly due to lack of practical values to add to the quest as qq grows.

Using codes over rings to construct QECs have also been tried. This route, however, does not usually lead to parameter improvements over QECs constructed from codes over fields. The absence of a direct connection from codes over rings to QECs necessitates the use of the Gray mapping, which often causes a significant drop in the minimum distance.

5 Going Asymmetric

So far we have been working on the assumption that the bit-flips and the phase-flips are equiprobable. Quantum systems, however, have noise properties that are dynamic and asymmetric. The fault-tolerant threshold is improved when asymmetry is considered [2]. It was Steane who first hinted at the idea of adjusting error-correction to the particular characteristics of the channel in [47]. Designing error control methods to suit the noise profile, which can be hardware-specific, is crucial. The study of asymmetric quantum codes (AQCs) gained traction when the ratios of how often σz\sigma_{z} occurs over the occurrence of σx\sigma_{x} were discussed in [30], with follow-up constructions offered soon after in [44]. Wang et al. established a mathematical model of AQCs in the general qudit system in [51].

As in the symmetric case, n:={ωβX(𝐚)Z(𝐛):𝐚,𝐛𝔽qn,β𝔽p}{\mathcal{E}}_{n}:=\{\omega^{\beta}X({\mathbf{a}})Z({\mathbf{b}}):{\mathbf{a}},{\mathbf{b}}\in\mathbb{F}_{q}^{n},\,\beta\in\mathbb{F}_{p}\}. An error E:=ωβX(𝐚)Z(𝐛)n{\rm E}:=\omega^{\beta}X({\mathbf{a}})Z({\mathbf{b}})\in{\mathcal{E}}_{n} has wtX(E):=wH(𝐚){\rm wt_{X}}({\rm E}):=\operatorname{w_{H}}({\mathbf{a}}) and wtZ(E):=wH(𝐛){\rm wt_{Z}}({\rm E}):=\operatorname{w_{H}}({\mathbf{b}}).

Definition 4 (Asymmetric Quantum Codes).

Let dxd_{x} and dzd_{z} be positive integers. A qudit code QQ with dimension KqK\geq q is called an asymmetric quantum code (AQC) with parameters ((n,K,dz,dx))q((n,K,d_{z},d_{x}))_{q} or n,k,dz,dxq\llbracket n,k,d_{z},d_{x}\rrbracket_{q}, where k=logqKk=\log_{q}K, if QQ detects dx1d_{x}-1 qudits of XX-errors and, at the same time, dz1d_{z}-1 qudits of ZZ-errors. The code QQ is pure if |φ\ket{\varphi} and E|ψ{\rm E}\ket{\psi} are orthogonal for any |φ,|ψQ\ket{\varphi},\ket{\psi}\in Q and any En{\rm E}\in{\mathcal{E}}_{n} such that 1wtX(E)dx11\leq{\rm wt_{X}}({\rm E})\leq d_{x}-1 or 1wtZ(E)dz11\leq{\rm wt_{Z}}({\rm E})\leq d_{z}-1. Any code QQ with K=1K=1 is assumed to be pure.

An n,k,d,dq\llbracket n,k,d,d\rrbracket_{q}-AQC is symmetric, with parameters n,k,dq\llbracket n,k,d\rrbracket_{q}, but the converse is not true since, for En{\rm E}\in{\mathcal{E}}_{n} with wtX(E)d1{\rm wt_{X}}({\rm E})\leq d-1 and wtZ(E)d1{\rm wt_{Z}}({\rm E})\leq d-1, wQ(E)\operatorname{w_{Q}}({\rm E}) may be bigger than d1d-1.

To date, most known families of AQCs come from the asymmetric version of the CSS construction and its generalization in [15].

Theorem 5.1 (CSS-like Constructions for AQCs).

Let CjC_{j} be an [n,kj,dj]q[n,k_{j},d_{j}]_{q}-code for j{1,2}j\in\{1,2\}. Let CjC_{j}^{\perp_{*}} be the dual of CjC_{j} under one of the Euclidean, the trace Euclidean, the Hermitian, and the trace Hermitian inner products. Let C1C2C_{1}^{\perp_{*}}\subseteq C_{2}, with

dz\displaystyle d_{z} :=max{wH(C2C1),wH(C1C2)} and\displaystyle:=\max\{\operatorname{w_{H}}(C_{2}\setminus C_{1}^{\perp_{*}}),\operatorname{w_{H}}(C_{1}\setminus C_{2}^{\perp_{*}})\}\mbox{ and }
dx\displaystyle d_{x} :=min{wH(C2C1),wH(C1C2)}.\displaystyle:=\min\{\operatorname{w_{H}}(C_{2}\setminus C_{1}^{\perp_{*}}),\operatorname{w_{H}}(C_{1}\setminus C_{2}^{\perp_{*}})\}.

Then there exists an n,k1+k2n,dz,dxq\llbracket n,k_{1}+k_{2}-n,d_{z},d_{x}\rrbracket_{q}-code QQ, which is pure whenever {dz,dx}={d1,d2}\{d_{z},d_{x}\}=\{d_{1},d_{2}\}. If we have CCC\subseteq C^{\perp_{*}} where CC is an [n,k,d]q[n,k,d]_{q}-code, then QQ is an n,n2k,d,dq\llbracket n,n-2k,d^{\prime},d^{\prime}\rrbracket_{q}-code, where d=wH(CC)d^{\prime}=\operatorname{w_{H}}(C^{\perp_{*}}\setminus C). The code QQ is pure whenever d=d:=d(C)d^{\prime}=d^{\perp_{*}}:=d(C^{\perp_{*}}).

The propagation rules and bounds for AQCs follow from the relevant rules and bounds on the nested codes and their respective duals. Details on how to derive new AQCs from already known ones were discussed by La Guardia in [28]. The asymmetric version of the quantum Singleton bound reads kn(dx+dx)+2k\leq n-(d_{x}+d_{x})+2. To benchmark codes of large lengths, one can use the the asymmetric versions of the quantum Gilbert-Varshamov bound established by Matsumoto in [41].

Many best-performing AQCs with small dxd_{x} and moderate nn and kk were derived in [15]. The optimal ones reach the upper bounds certified by an improved LP bound, called the triangle bound in [15, Section V]. Recent results, covering also AQCs of large lengths, came from an interesting family of nested codes defined from multivariate polynomials and Cartesian product point sets due to Gallindo et al.in [21].

Most known asymmetric quantum MDS (AQMDS) codes are pure CSS. Assuming the validity of the MDS conjecture, all possible parameters that pure CSS AQMDS codes can have were established in [14].

Theorem 5.2.

Assuming the MDS conjecture, there is a pure CSS AQMDS code with parameters [[n,j,dz,dx]]q[[n,j,d_{z},d_{x}]]_{q}, where {dz,dx}={nkj+1,k+1}\{d_{z},d_{x}\}=\{n-k-j+1,k+1\} if and only if one of the followings holds:

  1. 1.

    qq is arbitrary, n2n\geq 2, k{1,n1}k\in\{1,n-1\}, and j{0,nk}j\in\{0,n-k\}.

  2. 2.

    q=2q=2, nn is even, k=1k=1, and j=n2j=n-2.

  3. 3.

    q3q\geq 3, n2n\geq 2, k=1k=1, and j=n2j=n-2.

  4. 4.

    q3q\geq 3, 2nq2\leq n\leq q, kn1k\leq n-1, and 0jnk0\leq j\leq n-k.

  5. 5.

    q3q\geq 3, n=q+1n=q+1, kn1k\leq n-1, and j{0,2,,nk}j\in\{0,2,\ldots,n-k\}.

  6. 6.

    q=2mq=2^{m}, n=q+1n=q+1, j=1j=1, and k{2,2m2}k\in\{2,2^{m}-2\}.

  7. 7.

    q=2mq=2^{m} where m2m\geq 2, n=q+2n=q+2,

    {k=1, and j{2,2m2},k=3, and j{0,2m4,2m1}, or,k=2m1, and j{0,3}.\left\{\begin{array}[]{l}k=1,\mbox{ and }j\in\{2,2^{m}-2\},\\ k=3,\mbox{ and }j\in\{0,2^{m}-4,2^{m}-1\},\mbox{ or,}\\ k=2^{m}-1,\mbox{ and }j\in\{0,3\}.\end{array}\right.

Going forward, three general challenges can be identified. First, find better AQCs, particularly in qubit systems, than the currently best-known. More tools to determine or to lower bound dzd_{z} and dxd_{x} remain to be explored if we are to improve on the parameters. Second, construct codes with very high dzd_{z} to dxd_{x} ratio, since experimental results suggest that this is typical in qubit channels. Third, find conditions on the nested classical codes that yield impure codes.

6 Other Approaches and a Conclusion

We briefly mention other approaches to quantum error control before concluding.

Successful small-scale hardware implementations often rely on topological codes, first put forward by Kitaev [32]. This family of codes includes surface codes [6] and color codes [5]. Topological codes encode information in, mostly 22-dimensional, lattices. They are CSS codes with a clever design. The lattice, on which the stabilizer generators act locally, has a bounded weight. The extra restrictions make the error syndrome easier to infer.

Instead of block quantum codes, studies have been done on convolutional qubit codes, see, e.g., [20] and subsequent works that cited it. The logical qubits are encoded and transmitted as soon as they arrive in a steady stream. The rate kk over nn is fixed, but the length is not. This type of codes, like their classical counterparts, may be useful in quantum communication.

An approach, that does not require self-orthogonality, constructs entanglement-assisted quantum codes (EA-QECs) [7]. The price to pay is the need for a number of maximally entangled states, called ebits for entangled qubits, to increase either the rate or the ability to handle errors. Producing and maintaining ebits, however, tend to be costly, which offset their efficacy. Pairs of classical codes, whose intersections have some prescribed dimensions, were shown to result in EA-QECs in [29, Section 4]. A formula on the optimal number of ebits that an EA-QEC requires is given in [52].

Theorem 6.1.

Given a linear [n,k,d]q2[n,k,d]_{q^{2}}-code CC with parity check matrix HH, the code CHC^{\perp_{\rm H}} stabilizes an EA-QEC with parameters n,2kn+c,d;cq\llbracket n,2k-n+c,d;c\rrbracket_{q}, where c:=rank(HH)c:=\rank(HH^{\dagger}) is the number of ebits required.

A larger class of QECs that includes all stabilizer codes is the codeword stabilized (CWS) codes. The framework was proposed by Cross et al.in [12] to unify stabilizer (additive) codes and known examples of good nonadditive codes. General constructions for large CWS codes are yet to be devised. Also currently unavailable are efficient encoding and decoding algorithms.

The bridge between classical coding theory and quantum error control was firmly put in place via the stabilizer formalism. Various generalizations and modifications have been studied since, benefitting both the classical and quantum sides of the error-control theory. Well-researched tools and the wealth of results in classical coding theory translate almost effortlessly to the design of good quantum codes, moving far beyond what is currently practical to implement in actual quantum devices. Research problems triggered by error-control issues in the quantum setup revive and expand studies on specific aspects of classical codes, which were previously overlooked or deemed not so interesting. This fruitful cross-pollination between the classical and the quantum, in terms of error control, is set to continue.

References

  • [1] D. Aharonov and M. Ben-Or. Fault-tolerant quantum computation with constant error rate. SIAM J. Comput., 38(4):1207–1282, Jan 2008.
  • [2] P. Aliferis and J. Preskill. Fault-tolerant quantum computation against biased noise. Phys. Rev. A, 78(5), Nov 2008.
  • [3] S. A. Aly, A. Klappenecker, and P. K. Sarvepalli. Remarkable degenerate quantum stabilizer codes derived from duadic codes. In Proc. Int. Symp. Inf. Theory (ISIT). IEEE, Jul 2006.
  • [4] J. Bierbrauer and Y. Edel. Quantum twisted codes. J. Comb. Des., 8(3):174–188, 2000.
  • [5] H. Bombín. Gauge color codes: optimal transversal gates and gauge fixing in topological stabilizer codes. New J. Phys., 17(8):083002, Aug 2015.
  • [6] S. Bravyi, M. Suchara, and A. Vargo. Efficient algorithms for maximum likelihood decoding in the surface code. Phys. Rev. A, 90(3), Sep 2014.
  • [7] T. Brun, I. Devetak, and M.-H. Hsieh. Correcting quantum errors with entanglement. Science, 314(5798):436–439, Oct 2006.
  • [8] A. R. Calderbank, E. M. Rains, P. W. Shor, and N. J. A. Sloane. Quantum error correction via codes over GF(4). IEEE Trans. Inform. Theory, 44(4):1369–1387, Jul 1998.
  • [9] E. T. Campbell, B. M. Terhal, and C. Vuillot. Roads towards fault-tolerant universal quantum computation. Nature, 549(7671):172–179, Sep 2017.
  • [10] B. Chen, S. Ling, and G. Zhang. Application of constacyclic codes to quantum MDS codes. IEEE Trans. Inform. Theory, 61(3):1474–1484, Mar 2015.
  • [11] H. Chen, S. Ling, and C. Xing. Quantum codes from concatenated algebraic-geometric codes. IEEE Trans. Inform. Theory, 51(8):2915–2920, Aug 2005.
  • [12] A. Cross, G. Smith, J. A. Smolin, and B. Zeng. Codeword stabilized quantum codes. IEEE Trans. Inform. Theory, 55(1):433–438, Jan 2009.
  • [13] Yves Edel. Parameters of some GF(3)GF(3)-linear quantum twisted codes. Online available at https://www.mathi.uni-heidelberg.de/~yves/Matritzen/QTBCH/QTBCHTab3.html, 2000. Accessed on 2019-01-17.
  • [14] M. F. Ezerman, S. Jitman, H. M. Kiah, and S. Ling. Pure asymmetric quantum MDS codes from CSS construction: A complete characterization. Int. J. Quantum Inf., 11(03):1350027, Apr 2013.
  • [15] M. F. Ezerman, S. Jitman, S. Ling, and D. V. Pasechnik. CSS-like constructions of asymmetric quantum codes. IEEE Trans. Inform. Theory, 59(10):6732–6754, Oct 2013.
  • [16] M. F. Ezerman, S. Ling, B. Özkaya, and P. Solé. Good stabilizer codes from quasi-cyclic codes over 𝔽4\mathbb{F}_{4} and 𝔽9\mathbb{F}_{9}. In Proc. Int. Symp. Inform. Theory (ISIT). IEEE, Jul 2019.
  • [17] K. Feng, S. Ling, and C. Xing. Asymptotic bounds on quantum codes from algebraic geometry codes. IEEE Trans. Inform. Theory, 52(3):986–991, Mar 2006.
  • [18] K. Feng and Z. Ma. A finite Gilbert–Varshamov bound for pure stabilizer quantum codes. IEEE Trans. Inform. Theory, 50(12):3323–3325, Dec 2004.
  • [19] R. P. Feynman. Simulating physics with computers. Int. J. Theor. Phys., 21(6-7):467–488, Jun 1982.
  • [20] G. D. Forney, M. Grassl, and S. Guha. Convolutional and tail-biting quantum error-correcting codes. IEEE Trans. Inform. Theory, 53(3):865–880, Mar 2007.
  • [21] C. Galindo, O. Geil, F. Hernando, and D. Ruano. Improved constructions of nested code pairs. IEEE Trans. Inform. Theory, 64(4):2444–2459, Apr 2018.
  • [22] C. Galindo, F. Hernando, and R. Matsumoto. Quasi-cyclic constructions of quantum codes. Finite Fields Appl., 52:261–280, Jul 2018.
  • [23] D. E. Gottesman. Stabilizer Codes and Quantum Error Correction. PhD thesis, California Institute of Technology, 1997.
  • [24] M. Grassl. Bounds on the minimum distance of linear codes and quantum codes. Online available at http://www.codetables.de, 2007. Accessed on 2020-02-20.
  • [25] M. Grassl. Variations on encoding circuits for stabilizer quantum codes. In Chee Y.M. et al., editor, Coding and Cryptology. IWCC 2011. Lecture Notes in Comput. Sci., volume 6639, pages 142–158. Springer Berlin Heidelberg, 2011.
  • [26] M. Grassl, T. Beth, and M. Rötteler. On optimal quantum codes. Int. J. Quantum Inf., 02(01):55–64, 2004.
  • [27] M. Grassl and M. Rötteler. Quantum MDS codes over small fields. In Proc. IEEE Int. Symp. Inf. Theory (ISIT). IEEE, Jun 2015.
  • [28] G. G. La Guardia. Asymmetric quantum codes: new codes from old. Quantum Inf. Process., 12(8):2771–2790, Mar 2013.
  • [29] K. Guenda, T. A. Gulliver, S. Jitman, and S. Thipworawimon. Linear \ell-intersection pairs of codes and their applications. Des. Codes and Cryptogr., 88(1):133–152, Jan 2020.
  • [30] L. Ioffe and M. Mézard. Asymmetric quantum error-correcting codes. Phys. Rev. A, 75(3), Mar 2007.
  • [31] A. Ketkar, A. Klappenecker, S. Kumar, and P. K. Sarvepalli. Nonbinary stabilizer codes over finite fields. IEEE Trans. Inform. Theory, 52(11):4892–4914, Nov 2006.
  • [32] A. Yu. Kitaev. Fault-tolerant quantum computation by anyons. Ann. Phys., 303(1):2–30, Jan 2003.
  • [33] E. Knill and R. Laflamme. Theory of quantum error-correcting codes. Phys. Rev. A, 55(2):900–911, Feb 1997.
  • [34] E. Knill, R. Laflamme, and W. H. Zurek. Resilient quantum computation. Science, 279(5349):342–345, 1998.
  • [35] R. Laflamme, C. Miquel, J. P. Paz, and W. H. Zurek. Perfect quantum error correcting code. Phys. Rev. Lett., 77(1):198–201, Jul 1996.
  • [36] D. A. Lidar and T. A. Brun, editors. Quantum Error Correction. Cambridge University Press, 2013.
  • [37] R. Lidl and H. Niederreiter. Finite Fields. Cambridge University Press, Oct 1996.
  • [38] S. Ling, J. Luo, and C. Xing. Generalization of Steane’s enlargement construction of quantum codes and applications. IEEE Trans. Inform. Theory, 56(8):4080–4084, Aug 2010.
  • [39] P. Lisoněk and V. Singh. Quantum codes from nearly self-orthogonal quaternary linear codes. Des. Codes and Cryptogr., 73(2):417–424, Feb 2014.
  • [40] F. J. MacWilliams. A theorem on the distribution of weights in a systematic code. Bell System Tech. J., 42(1):79–94, Jan 1963.
  • [41] R. Matsumoto. Two Gilbert–Varshamov-type existential bounds for asymmetric quantum error-correcting codes. Quantum Inf. Process., 16(12), Oct 2017.
  • [42] J. Preskill. Fault-tolerant quantum computation, 1997.
  • [43] E. M. Rains. Nonbinary quantum codes. IEEE Trans. Inform. Theory, 45(6):1827–1832, Sep 1999.
  • [44] P. K. Sarvepalli, A. Klappenecker, and M. Rötteler. Asymmetric quantum codes: Constructions, bounds and performance. Proc. Roy. Soc. London Ser. A, 465(2105):1645–1672, Mar 2009.
  • [45] B. Schumacher. Quantum coding. Phys. Rev. A, 51(4):2738–2747, Apr 1995.
  • [46] P. W. Shor. Scheme for reducing decoherence in quantum computer memory. Phys. Rev. A, 52(4):R2493–R2496, Oct 1995.
  • [47] A. M. Steane. Multiple-particle interference and quantum error correction. Proc. Roy. Soc. London Ser. A, 452(1954):2551–2577, Nov 1996.
  • [48] A. M. Steane. Enlargement of Calderbank-Shor-Steane quantum codes. IEEE Trans. Inform. Theory, 45(7):2492–2495, 1999.
  • [49] A. M. Steane. Overhead and noise threshold of fault-tolerant quantum error correction. Phys. Rev. A, 68(4), Oct 2003.
  • [50] A. Vardy. The intractability of computing the minimum distance of a code. IEEE Trans. Inform. Theory, 43(6):1757–1766, 1997.
  • [51] L. Wang, K. Feng, S. Ling, and C. Xing. Asymmetric quantum codes: Characterization and constructions. IEEE Trans. Inform. Theory, 56(6):2938–2945, Jun 2010.
  • [52] M. M. Wilde and T. A. Brun. Optimal entanglement formulas for entanglement-assisted quantum coding. Phys. Rev. A, 77(6), Jun 2008.
  • [53] W. K. Wootters and W. H. Zurek. A single quantum cannot be cloned. Nature, 299(5886):802–803, Oct 1982.

Index