1 Introduction
The stabilizer formalism was originally introduced to describe quantum error-correcting codes [
1], but now plays many different roles in quantum information and quantum computation. The key idea of this formalism is that quantum states can be represented by operators that stabilize them. A quantum system is fully described by a quantum state, which is viewed as a mathematical object. The description of quantum states is a complicated task, because we need exponentially many parameters in the number of qubits [
2]. The stabilizer formalism is a powerful tool to describe a considerable class of entangled states.
One of the challenges facing quantum computation and other attempts to create specific highly entangled states is decoherence and controlling operational errors. A subclass of quantum codes, the stabilizer codes, has developed based on a group theoretical approach to meet this challenge [
3]. A stabilizer quantum error-correcting code encodes logical qubits into physical qubits. In other words, a stabilizer code appends ancilla qubits to qubits that we want to protect. Stabilizer codes have found a wide range of applications, for instance, decoherence-free stabilizer codes are constructed for open quantum systems [
4] and stabilizer codes are related to Lorentzian lattices and non-chiral CFTs [
5].
An
-qubit stabilizer state is the simultaneous +1 eigenvector of
independent, commuting elements of the Pauli group. A graph state is a special type of multi-qubit state that can be represented by a graph. Stabilizer states and graph states have applications in quantum error correction and form a universal resource for measurement-based quantum computation [
6]. The class of stabilizer states is rich enough to be considered useful in almost all applications of quantum networks. For example, the GHZ-state is seen in many quantum protocols [
7]. Some of these applications include quantum secret sharing [
8], conference key agreement [
9], anonymous transfer [
10] and clock synchronization [
11].
A stabilizer circuit is a quantum circuit in which every gate is a controlled-NOT, Hadamard, phase, or 1-qubit measurement gate. Stabilizer circuits can be applied to create highly entangled states. They also perform the encoding and decoding steps for quantum error-correcting codes. They also play a central role in fault-tolerant circuits [
12]. The stabilizer formalism enables us to encompass most of the paradoxes of quantum mechanics including dense quantum coding [
13], the GHZ experiment [
14] and quantum teleportation [
15]. However, the Knill−Gottesman theorem allows such circuits to be simulated classically with an efficient runtime [
16].
This paper is devoted to the description of the stabilizer formalism and its applications arising in quantum information and quantum computation. This paper is organized as follows. In Section 2, we provide basic definitions and lemmas on stabilizer states, binary representation and Clifford group. In Section 3, we study the measurement process of a Pauli observable by using the projective measurement and give the respective algorithm based on stabilizer generators. In Section 4, we introduce two fundamental concepts of “entanglement detection” and “entanglement witness”. Then it is shown that the stabilizer witness is coarser than the GHZ witness by applying the GHZ basis and methods of operator theory. Furthermore, it is shown that to detect genuine entanglement by stabilizer witness we need a full set of stabilizer generators. In this way, we apply methods of standard linear algebra. In Section 5, we give basic definitions and important instances of stabilizer codes, then a stabilizer code is constructed from a given classical linear code. We discuss Pauli errors, error-correcting codes and the process of syndrome extraction. The symplectic structure of the stabilizer formalism is established and it is shown that any stabilizer code with logical qubits is unitarily equivalent to a -trivial stabilizer code. Two equivalent definitions of graph states are given. Graph codes associated with linear codes are characterized and it is shown that any graph code is a stabilizer code by identifying its stabilizer generators. The cleaning lemma based on independent logical operators is shown and the distance of an embeddable code in a lattice is obtained. Generalization of stabilizer codes and Clifford group is given. In Section 6, stabilizer gates are introduced and their simulating runtime is obtained by applying stabilizer matrices. The Knill−Gottesman theorem, tableau representation and frame representation are discussed. The runtime of resolution of a Pauli group element into production of a stabilizer state and a destabilizer state is obtained. The algorithm for updating global phases is given. In Section 7, we discuss stabilizer channels and their types including Pauli reset channels and Clifford channels. Resolution of a quantum channel into a sum of stabilizer channels is given. The key concept of negativity is discussed. The capacity of the quantum erasure channel is obtained by a class of capacity achieving stabilizer codes. In Section 8, we discuss shadow tomography problem and the framework of classical shadow estimation is given as an algorithm. And finally, in Section 9 we present some research lines.
2 Pauli products
The Pauli operators on a two-level quantum system (qubit) are defined as follows:
where is the Hilbert space on which the operators act. The Pauli group is defined as the group generated by the Pauli operators up to the phase factors , , i.e.,
For qubits, the Pauli group is defined as
Pauli operators on qubit form a basis for the operator space on a single qubit. Products of operators of this sort, called Pauli products, span the space of operators on qubits. If is an operator on qubits, its base is the set of qubits on which it acts nontrivially (, , or ). The weight of is the number of qubits in its base. For example
has base , so is of weight 4. Here, the index (of operators) indicates the Hilbert space (qubit) that the respective operator acts on.
2.1 Stabilizer states
A stabilizer of qubits is an Abelian subgroup of that does not include . Let , where are commuting and independent (i.e., no one can be expressed as a product of other generators, up to a sign) generators. Every element of the Pauli group is either Hermitian operator or anti-Hermitian operator. Any Hermitian operator has order two, i.e., where is the adjoint operator, and any anti-Hermitian operator has order four, i.e., . Therefore, cannot contain anti-Hermitian operators as otherwise . It shows that . When , there is a state that is the common eigenvector of elements of and is called a stabilizer state. For example, the EPR (Einstein−Podolsky−Rosen) state is a stabilizer state, because it is a +1 eigenvector of the stabilizer subgroup . The stabilized subspace is defined as follows
An -qubit GHZ state is defined by . In fact, the GHZ state is a +1 eigenvector of the following Pauli products:
In other words, for . Let be the stabilizer generated by . Then and by letting , we have
We will apply Eq. (1) in Section 4.
2.2 Binary representation
At each qubit we use the following encoding of the Pauli operators as pairs of bits:
We can reconstruct the Pauli operators as follows:
where . By applying Eq. (2) it follows that
where , , , “t” is transposition and computations are done modulo 2. The left side of Eq. (3) is the symplectic inner product of the binary vectors. For example, using this notation, the Pauli product acting on has the following encoding
where is the binary field.
Lemma 1. For every with corresponding binary vectors in encoding ,
i) ,
ii) ,
where ~ denotes equality up to a global phase.
Proof. It is derived from the following representation:
where .□
Remark 1. Property (i) shows that the encoding is a homomorphism of groups, and property (ii) shows that when two Pauli products commute. Throughout this paper, the phrase “up to a global phase” means that we can ignore global phases.
For a given set of Pauli products , the matrix
is called the parity-check matrix corresponding to the Pauli products.
Corollary 1. A set of Pauli products with parity-check matrix are stabilizer generators of a stabilizer state if and only if the following conditions hold
i) ,
ii) has full rank.
Proof. It follows from Lemma 1.□
2.3 Clifford group
The
Clifford group on qubits is defined as follows
, where
is the unitary group of degree
. We take into account that the action of every
is completely determined by the images of
and
, moreover
and
must anti-commute. Then,
can go to any element of
(since
is Hermitian if and only if
is Hermitian) and
can go to any element of
. Therefore, including the global phase
has
elements. In general,
[
17]. We note that qubit
on the Bloch sphere can be written as
. Let
and define
A qubit rotation by angle about the arbitrary axis is written as a similarity transformation, i.e., , where . We can think of as rotations of the Bloch sphere that permute directions.
3 Measurement process
In this section, we study the process of measuring a Pauli observable, i.e., Hermitian Pauli group element, on a stabilizer state . We recall that the projective measurement described by is based on the decomposition , where is the projector onto the eigenvector of with eigenvalue . Upon measuring, the probability of getting result is . Given that outcome occured, the post-measurement state is .
Lemma 2. Adding an independent and commuting generator to the set of stabilizer generators cuts the stabilized subspace dimension in half.
Proof. Given a generator ,
thus is a projector. We note that if and only if , and if and only if , thus is the projector onto , where and
By adding an independent and commuting generator to , we have
and it means that the subspace stabilized by both and is of dimension . By induction on the number of generators of , the proof is completed.□
Let be a set of stabilizer generators for . There are two cases:
Case I. For every , commutes with . From Lemma 1, it follows that cannot be independent of the generators in . Hence, , where . Therefore,
measurement gives result with probability
and the post-measurement state is .
Case II. There is a generator in , such that anti-commutes with . For a given generator , either commutes with or anti-commutes with . In the first case, we do not change , but in the second case we construct a new set of generators as follows. We note that , then is a generator that commutes with . is replaced by and it is also denoted by . The action of on the (un-normalized) post-measurement state is
where is the measurement outcome. Now, we only need one more independent stabilizer generator (instead of ) to characterize the post-measurement state completely. The key idea is that an qubit stabilizer state is uniquely determined by stabilizing operators. is the generator that is needed because, firstly, anti-commutes with , so it is independent of all the generators. Furthermore, stabilizes the post-measurement state as follows:
This argument shows that by obtaining the measurement outcome , we reach stabilizer generators for the post-measurement state to be and . The measurement outcomes are with equal probability, because
or .
From this discussion, it results the following algorithm to measure a Pauli observable in a stabilizer state :
Algorithm 1. (Measurement process)
1) Find a set of stabilizer generators for such that at most one generator anti-commutes with .
2) If all the stabilizer generators commute with , then the outcome will be certain and the post-measurement state will be .
3) If one generator, say , anti-commutes with , then outcome will be random and by replacing with we reach a new set of generators that stabilize the post-measurement state.
In the following section, we present the method of entanglement witnesses as the most common entanglement detection tool.
4 Entanglement detection
Entangled state describes a system made of two separable parts which cannot be written as a tensor product of states and describing separately each of the subsystems, i.e., . A pure state is called biseparable, if it can be written as a tensor product of two multi-partite states , we take into account that each of and can be an entangled state. A mixed state is biseparable, if it can be written as a convex combination of biseparable pure states, i.e., with and , where ’s and ’s are biseparable pure states. A state that is not biseparable is called genuine multipartite entangled.
Theorem 1. ([
18])
A density operator on , where ’s are Hilbert spaces, is entangled if and only if there exists a Hermitian operator such that , and for all separable states , .
is called entanglement witness.
Let be a pure state, be a fixed bipartite splitting, and consider product states . By choosing an orthonormal product basis for partition , we can write
Let be the coefficient matrix, and be the normalized coefficient vectors. Thus
where denotes the singular values of . In other words, are the Schmidt coefficients of with respect to bipartite splitting . Let be the square of the maximal Schmidt coefficient over all possible bipartite partitions of . Therefore, is a witness operator that detects genuine multipartite entanglement of . We can summarize this discussion in the following lemma.
Lemma 3. A witness operator that detects genuine multipartite entanglement of a pure state is given by , where is the identity operator and , where denotes the set of biseparable states.
Corollary 2. For , .
The witness
detects genuine
-qubit entanglement around the GHZ state, which has been applied in an experiment [
19]. By applying Eq. (1), it turns out that
is a stabilizer witness, i.e., constructed by stabilizer generators, and can be measured locally, i.e., can be written as a sum of Pauli products. But the number of measurement settings increases exponentially with the number of qubits [
20], which is not appropriate.
The GHZ state basis is defined as the set of the following states
where and for . is an orthogonal basis and . We represent elements of as .
For a given entanglement witness , let be the set of states which are detected by , i.e.,
For two given entanglement witnesses and , we say that is finer (coarser) than , if ().
Theorem 2. There is a stabilizer witness consisting of all the generators which is coarser than .
Proof. It is shown that is the desired witness. We note that , , and . Let
Given a basis element , there exists a stabilizer generator such that , because case 1: if , then , and case 2: if all of ’s in are not equal, let be the first index for which , then . It follows that , where . Furthermore, . It shows that all of the eigenvalues of are nonnegative, thus . Hence, and then for every state . It means that .□
Theorem 2 shows that detects only states with genuine -qubit entanglement. In fact, uses only two measurement settings: 1) to measure , and 2) to measure ,..., . This is the advantage of using .
Theorem 3. In Theorem 2, we need a full set of generators to detect genuine entanglement. In other words, with less than generators, we cannot detect genuine -qubit entanglement.
Proof. We remove one of the stabilizer generators, say . We claim that there is a Pauli product such that commutes with for , and anticommutes with . Since the parity check matrix has full rank, then the system of equations
has a solution , where and are binary columns. Let , then for rows of , we have
Let be the Pauli product corresponding to , then according to Lemma 1, it turns out that for , and . Let be the witness which is obtained from by removing . Since (), then the expectation values of in and are equal and it means that there are at least two elements of the basis which give the minimum. We note that , and for any basis element , , thus . Therefore, takes the minimal value only for . There is a superposition of and which is biseparable. It follows that the witness takes the minimal value in the biseparable state and then it is not applicable to detect genuine -qubit entanglement.□
In the next section, we study stabilizer codes as a strong tool to correct errors in multiple qubit systems.
5 Stabilizer codes
In classical coding theory, our goal is to encode classical bits (as -bit strings) into binary codewords that are -bit strings, say . This is a linear code, if there exists an full-rank matrix such that . The matrix is called the generator matrix of the code. The code is a -dimensional subspace of and is denoted by . The distance of is defined as , where with . Here, it is denoted by . The dual of is a linear code defined as , where is the standard inner product. The parity check matrix of is the generator matrix of and is denoted by .
Quantum stabilizer codes are quantum analogs of binary linear codes. Given an Abelian subgroup of which does not contain , the (quantum) stabilizer code corresponding to is defined as . When is generated by generators, the subspace has dimension (Lemma 2), it encodes logical qubits to physical qubits and it is denoted by . The distance of the code is defined as , where is the normalizer of . The elements of are called nontrivial logical operators and trivial logical operators are elements of . We denote the set of all logical operators by . Here, the code is denoted by . For example, the 7-qubit code (Steane code) is a code with the following stabilizer
This code encodes 1 qubit to 7 qubits and its codewords are
The Steane code belongs to the family of CSS (Calderbank−Shor−Steane) codes. A CSS code is a stabilizer code whose stabilizer splits into two sets and satisfying: 1) and is the minimal group generated by and . 2) If , then . 3) If , then . Let and be the subgroups of corresponding to and in Lemma 1, respectively. Hence, up to a global phase, we have the following bijection:
where if and only if , and similarly for and . It follows that , thus and . In this example, the distance of the stabilizer code is defined as .
The toric code is an instance of CSS codes. Let be a two-dimensional lattice, i.e., a set of vertices , edges and faces glued together in a certain way. The toric code in two dimensions is defined on a lattice by placing one qubit on every edge, and associating - and -type generators with vertices and faces of as follows:
where and denote Pauli and operators, respectively, acting on qubit placed on the edge (see Fig.1).
The Gottesman−Kitaev−Preskil (GKP) code is another instance of stabilizer codes. The GKP code is an
-mode quantum lattice code with
stabilizers, i.e., constructed using a non-degenerate lattice. This family of codes is quite important in fault-tolerant quantum computation. The GKP code is an important type of bosonic quantum error-correcting codes, which encodes a qubit in an oscillator [
47].
A natural question gives rise: Given a linear code, how can we construct a stabilizer code? Let be an linear code with generator . Let be the rows of . We extend the set of rows to a basis for denoted by . Let be an matrix with rows and be the inverse of . Denote the columns of by . By applying this notation and , where , we define the following generators on qubits:
1) ’s are independent. It follows from the linear independence of ’s, the linear independence of ’s, and the fact that the binary representation of -type operators is linear independent of the binary representation of -type operators.
2) ’s mutually commute. It is clear for the first generators and for the last generators. Between them, the commutation follows from the following identity
and the fact that for , and for the presence of and guarantees commutation. Hence, defines a stabilizer code.
5.1 Quantum error correction
Classical computers use repetition codes, i.e., each bit is encoded in many electrons, and after each time step the error correction is done, i.e., it is turned back to the value held by the majority of the electrons. The simplest example is the following repetition code:
. It will correct the state
to the majority value
. Because of the No−Cloning theorem, quantum computers cannot use a quantum repetition code as follows:
. The No−Cloning theorem says that there is no quantum operation that takes any state
to
[
21].
Quantum error correction comes from meeting error correction in classical information theory and principles of quantum mechanics [
23]. It is concerned with the main problem of communication in the presence of noise. Consider a single qubit in a pure state which interacts with its environment. Without loss of generality, we may assume that the initial state of the environment is pure, say
. The evolution of the qubit and its environment is described by a unitary operator
as follows:
where are states of the environment. For an arbitrary state , we have
The interpretation of this expanding is that one of the following situations occurs to the qubit: trivial (), bit flip (), phase flip (), or both () (in practice, this classification makes sense when four states and of the environment are mutually orthogonal).
In general, for qubits, the action of an arbitrary unitary operator can be described as follows:
where ranges over Pauli products and are the corresponding states of the environment. We aim to identify a subset of correctable errors, i.e., the errors that we wish to be able to correct.
5.1.1 Error-correcting codes
A
quantum error-correcting code is a mapping from the Hilbert space of dimension
corresponding to
logical (encoded) qubits into the Hilbert space of dimension
corresponding to
physical qubits, where
. The image of this mapping is called the
code subspace. Our goal is to protect
encoded qubits from error. In this way,
qubits are added as redundant qubits. Let
be an orthonormal basis for the code subspace,
’s are called
codewords. Now the quantum error correction criteria [
22], i.e., a necessary and sufficient condition in terms of the codewords to preserve the code subspace, is proved.
Theorem 4. Let be a linear space of errors acting on the subspace code . Then recovery is possible if and only if
.
Proof. Suppose that recovery is possible and let be the unitary operator describing the whole recovery operation, then
where ’s are states of the ancilla (it is employed to implement the recovery operation) and the environment. Therefore,
Since the recovery operation must work particularly when , then and Eq. (4) is proven. To prove Eq. (5), we can write as follows:
Similarly, by starting from
, the same result is obtained and then Eq. (5) is proved. Analog of Eq. (5) is not possible in classical coding theory [
42].
Suppose that Eqs. (4) and (5) or equivalently holds. It is clear that is a Hermitian matrix. The action of an error on a basis state and the environment is , where ’s are elements of an orthonormal basis for the environment, and ’s are linear combinations of the Pauli products ’s, satisfying the normalization condition
It follows that . The error can be reversed if there exist operators ’s such that (normalization) and
where is an orthonormal basis for the Hilbert space of the ancilla and the state of environment and ancilla does not depend on . The basis for the environment can be chosen such that the matrix is diagonalized
where follows from Eq. (6). For each with , let , therefore
Equation (8) shows that the ’s do indeed reverse the error. It remains to check the normalization condition. Let
By applying Eq. (7) it follows that
By adding the projector to ’s, we obtain a set of recovery operators (Kraus operators) satisfying the normalization condition.□
Remark 2. The distance of a stabilizer code can be defined as the smallest positive integer such that the quantum error correction criteria is satisfied for all error combinations of weight in the range .
A code is called nondegenerate, if for all . A code that is not nondegenerate is degenerate. For example, consider the 9-qubit code (Shor code) with the following stabilizer:
It is a code with these codewords:
The Shor code is degenerate, because we cannot distinguish and ,
5.1.2 Syndrome extraction
Based on the property that every two Pauli products either commute or anti-commute, we will see how a stabilizer code detects and corrects errors. The corresponding syndrome of a given error is a bit string including which stabilizer generator commutes or anti-commutes with the error. Given an error , its syndrome is denoted by . For example, consider the 5-qubit code with the following stabilizer
The syndrome for the error is , where 0’s indicate that the generators commute with and 1’s indicate that the generators anti-commute with .
Proposition 1. Given a stabilizer code , there exists an error with any given syndrome.
Proof. Let be the stabilizer of . The syndrome of an error is a bit string of length . Therefore, is the syndrome of if and only if for all (Lemma 1). It gives a system of equations with equations and variables which has a nonzero solution.□
Any code allows correction of a particular family of correctable errors. Given a stabilizer code with stabilizer , error operators in are all correctable. If the noise of the system under consideration includes only these errors, the code is called a decoherence free subspace. There is a large family of errors which are correctable by extracting the corresponding syndromes. It is shown that if for every product of two members of , either or for some then is correctable. Let be a general encoded state.
In the first case, and have the same syndrome, so are indistinguishable, but both are correctable because the common syndrome leads us to apply as the corrective operator. If it was which occured, this is well, while if occured, the final state will be which is also correct. This case is related to degenerate codes.
In the second case, either , or . It follows that and have different syndromes, thus they are distinguishable from each other. The noisy state is . Quantum error correction determines the error syndrome with the help of ancilla qubits:
By a projective measurement of the ancilla, the sum will be collapsed to a single term randomly and we will obtain as the measurement result. Since there is only one error with this syndrome, we are able to apply to correct the error. This case is related to nondegenerate codes.
5.2 Symplectic structure
Matrix is called symplectic, if . The set of all symplectic matrices in is denoted by , which is a group under matrix multiplication.
Lemma 4. Up to a global phase, .
Proof. For , where , let
It follows that for every ,
and every satisfies for some functions and . It turns out that
From this equality, we have
Therefore, for some and
for some . Let be the symplectic inner product. We note that and commute (anti-commute) if and only if . Given unitary , conjugating by () preserves commutation (anti-commutation) of and , then its corresponding matrix preserves the symplectic inner product, i.e., for every and . Hence, , i.e., , and which maps onto defines a group homomorphism. It is clear that . Let , then and for all . The sign changes reveal that which of the Pauli operators and is being to qubit by , call this operator . Together these operators we construct the Pauli product which is equal to , because they perform the same thing when conjugating every generator of .□
Given , the -trivial stabilizer code on qubits is defined by
Theorem 5. For any stabilizer code , there exists a unitary operator such that . In other words, and are (Clifford) unitarily equivalent.
Proof. Given code stabilized by . We take into account that the stabilizer of is generated by . We need to find a unitary such that for all . In general, this unitary is not unique since we can redefine it by any which acts trivially on . However, any choice of stabilizer basis of induces a . In fact, a choice of logical basis corresponds to a full completion of the generators and the logical basis is determined by the eigenvalues of . The unitary defined by for is a Clifford operator because the Pauli products on the left and right hand side are fully commuting sets. This unitary transforms input states as follows:
□
In classical coding theory, it is usual to use finite fields larger than . The simplest example is with . It is a field of characteristic 2 with the following two operations: 1) Conjugation: . 2) Trace: . Let be a mapping which sends each to . For and in ,
Here, the trace inner product turns into the symplectic inner product. The Pauli operators are associated with , respectively. Given a linear code over with parity check matrix . Since , then any vector in the dual space of the code can be generated by summing together selected rows of , each multiplied by . It is shown that the 5-qubit code (over ) is derived from the Hamming code over . In this way, we need to find Pauli products associated with the rows of and as follows:
It follows that with , , , and is the stabilizer for the 5-qubit code.
5.3 Graph codes
Let be a graph with a set of of vertices and a collection of edges. We associate the vertices to the numbers . We are only interested in graphs without loops, i.e., there are no edges of the form . The graph state is defined as the stabilizer state stabilized by the following Pauli products:
where . The corresponding parity check matrix for generators of a graph state is in the standard form, i.e., . Let be the controlled phase operator (gate) between qubits and ,
and
, the graph state
can be defined as [
24]
Each graph state corresponds uniquely to a graph, i.e., two different graphs and cannot describe the same graph state, because
where
denotes the symmetric difference of the edge sets, which is a contradiction. Another fact is that for any stabilizer state
, there is a graph state
and a unitary operator
such that
(local Clifford equivalence) [
25].
Given a graph and a classical linear code , a graph code is defined as the linear span of the states , where is any codeword in . The simplest example of a graph code is yielded by a trivial graph with no edges. The distance of this code is 1 (see Remark 2).
Theorem 6. Given a graph and a linear code , let be the associated graph code. Then .
Proof. Suppose that . We take into account that
then ’s commute with ’s. For every , we define , therefore by applying (10)
Since and where , then the set of states of the form constitutes an orthonormal basis of elements in which each qubit is or . As is a unitary operator, it maps one orthonormal basis to another orthonormal basis, therefore and .□
We recall that the support of a vector is the set of indices for which and is denoted by . After this note, we will construct a set of stabilizer generators for a graph code as follows.
Theorem 7. Given a graph and a linear code , let be a minimal weight generating set for the dual code . The graph code is a stabilizer code with the following stabilizer generators
where is defined in Eq. (9).
Proof. Let be the stabilizer for . We show that . Independence of ’s follows from the linear independence of the binary vectors ’s. Since for all and , then ’s commute. It remains to show that for every codeword , the state is stabilized by ’s. For every , we can rewrite as follows:
where , and is a binary vector. Since , then and commute, thus by applying Eq. (11) we have
□
Graph codes play a key role in quantum technologies by yielding considerable protection against qubit loss, a dominant noise mechanism [
26].
5.4 Cleaning lemma
Given an stabilizer code with stabilizer . An operator (Pauli product) is said to be supported on a region (a subset of qubits) if it acts trivially on (the complement of ). The support of an operator is the minimal region on which is supported and is denoted by . According to Lemma 1, we can regard as an -dimensional vector space on . We say a set of independent logical operators to mean one that maps to a linear independent set in the quotient vector space . We recall that two vectors are orthogonal (with respect to the symplectic inner product) if and only if the corresponding operators commute, thus where is the orthogonal complement of in .
Lemma 5. Let and be the number of independent logical operators that can be supported on and , respectively. Then .
Proof. Let be the subspace of which is supported on . We can write as , where and are the subspaces of which are supported on and , respectively and includes whatever remains beyond . Let us define the restriction of a vector to as , where is the standard basis for . Let denote the restriction of to . We take into account that linearly independent vectors in remain linearly independent in , otherwise a nontrivial linear combination of the vectors would have a trivial restriction to , thus would be in . Since is trivially orthogonal to , then is the orthogonal complement of in . Hence,
Similarly by replacing with we have
Since logical operators supported on () are elements of () which commute with and are not contained in , then by applying Eqs. (12) and (13)
It follows that
□
For a logical operator
, we say that
can be cleaned on , if there exists a logically equivalent operator
, i.e.,
where
, such that
can be supported on
. Now we are ready to present the cleaning lemma [
27,
28] as follows.
Corollary 3. Let be a region which does not support any nontrivial logical operator. Then any logical operator can be cleaned on .
Proof. The assumption means that . Hence, by applying Lemma 5, it follows that . Therefore, any logical operator may be supported on . It means that any logical operator can be cleaned on . In other words, for any logical operator , we can choose an element , with and , such that acts trivially on .□
Given a lattice ( is sufficiently large) we say that a stabilizer code with stabilizer is embeddable into , if we can associate the qubits to vertices of such that each is supported only on one hypercube, say of size 1, of .
Theorem 8. Given an embeddable stabilizer code in an -dimensional lattice,
Proof. We will prove the theorem for two-dimensional lattices, the higher dimensional case will be clear from . Suppose that for some
For sufficiently large , we can divide into an even number of vertical strips, say strips, of width at most but still larger than 1. Let be a nontrivial logical operator on . According to Corollary 3, each vertical strip can be cleaned out by applying the generators whose supports overlap with that strip. We note that these supports are located on that strip or adjacent strips, but not on any other strips (Fig.2).
Fig.2 Applying the cleaning lemma. |
Full size|PPT slide
By repeating this process, all even strips will be cleaned out and we obtain a logical operator which is supported only on odd strips , where is a Pauli product which is supported on the -th strip. Since each generator overlaps with at most one odd strip and , then for all . Since , there is at least one odd strip such that which means is a nontrivial logical operator. But which is a contradiction with Eq. (14). Hence, .□
5.5 Generalization
The notion of stabilizer code can be generalized to qudits, i.e., -level systems, where is a prime number. Here, the system has only a finite number of accessible states (degrees of freedom), thus the phase space needs to be discrete. In this case, the Hilbert spaces of constituents are with the standard basis . The shift operator and the clock operator are defined as follows:
and modulo- arithmetic is used. These operators satisfy . We associate with each point in phase space a Weyl operator according . The Weyl operators can be conceived as generalized Pauli products. They satisfy the Weyl commutation relations:
It follows that two Weyl operators and commute if and only if the standard symplectic inner product vanishes
For qudits, we can encounter (position and momentum) coordinates in phase space with . The above inner product gives rise to the following form
where . The generalized Weyl operator is defined by
Let . An isotropic subspace is a subspace on which the inner product of Eq. (15) vanishes. A character is a map from into the circle group . A generalized stabilizer code associated with and is defined as follows
The generalized Clifford group is defined as the set of unitary operators that map Weyl operators onto Weyl operators under conjugation, i.e.,
where belongs to the symplectic group induced by Eq. (15) and is the symbol of proportionality.
Remark 3. We take into account that the direct analog of stabilizer states makes sense in the setting of continuous Weyl systems by describing singular states within an algebraic approach [
29].
In the following section, we study stabilizer circuits as a very important class of quantum circuits. In fact, they are sufficient to implement many quantum algorithms and protocols.
6 Stabilizer circuits
In quantum computation, it is more convenient to work with circuit model rather than matrix equivalent representation. For example, four gates are allowed in the stabilizer formalism, including controlled NOT (CNOT), phase (P), Hadamard (H) and measurement (M) are given below:
where is the addition modulo 2 and
6.1 Simulating the stabilizer gates
Up to a global phase, the Clifford group
is generated by one and two qubit gates in the set
, which are called the
Clifford gates [
30]. Let
and
be two positive functions, we write
if
, and
if
and
.
Theorem 9. Simulating a Clifford gate on an -qubit stabilizer state requires time, while a measurement gate is simulated in time for random outcomes and in time for deterministic outcomes.
Proof. Given an -qubit stabilizer state with stabilizer , we apply the idea of representing by a stabilizer matrix whose rows are a set of generators of . Accordingly, any computational basis state can be represented as follows, where signs denote global phases for the respective generator,
By simulating a Clifford gate on , , the -th row of , is mapped to , hence at most two columns of are updated. Therefore, is simulated in time. We note that each qubit in is either in a state, deterministic outcome, or in an unbiased superposition of both states, random outcome. In the case of random outcome, we flip an unbiased coin to decide the outcome and then update to match the outcome. Let be a row in with an operator in its -th position. The global phase of is set to +1(−1) if . Based on the fact that and anti-commute, if any other rows in anti-commute with , we multiply them by to make them commute with . After that, we replace with . This process needs up to row multiplications, then the overall runtime is . In the case of deterministic outcome, we need to find out whether the qubit is in the state. In other words, whether the qubit is stabilized by . In this way, we can perform Gaussian elimination to convert into a row-echelon form as follows:
where entries in (upper -diagonal) are or , entries in (lower -diagonal) and (upper -diagonal) are or , and entries in (lower -diagonal) are only . This process removes redundant operators from , then it is possible to identify the row as . The phase of this row decides the outcome of the measurement. Since the Gaussian elimination process takes time, thus the runtime is .□
6.2 Knill−Gottesman theorem
Given a quantum circuit with input qubits as one bit string, e.g., , by measuring the qubits, we would read out some sample bits , with probability , where . A classical simulation of a quantum circuit is the process of sampling from the output distribution efficiently using a classical probabilistic computer. By simulating a circuit, we will obtain an outcome that is in accordance with the probability distribution .
Theorem 10. (Knill−Gottesman theorem [
16])
Any quantum circuit that applies only the following elements can be simulated efficiently (in polynomial time) on a classical probabilistic computer.
● Preparation of qubits in computational basis states;
● Quantum gates from ;
● Measurements in the computational basis.
Proof. Without loss of generality, we start with the state
which is a stabilizer state with stabilizer
. Let
be a generator of
and
. It follows that
is a stabilizer state with stabilizer
. Implementing a generator
implies updating the
generators which describe the initial state. This step requires
operations on a classical computer [
30]. If our circuit is a product of
terms, each a generator of
, then the computation can be simulated by a classical computer in
time.□
A circuit satisfying the conditions of Theorem 10 is called a
stabilizer circuit. Stabilizer circuits refer to quantum circuits including all of the four types of stabilizer gates, while
unitary stabilizer circuits consist of only the Clifford gates. By adding any non-Clifford gate to the set of Clifford gates, we obtain a set of universal gates for quantum computation, while stabilizer circuits are probably not even universal for classical computation [
30]. From Theorem 10, it follows that quantum algorithms employing the entanglement created by Clifford gates do not give any advantage over classical computers.
6.3 Tableau representation
Let be an -qubit stabilizer state with stabilizer . We may write , where acts on just as acts on . In other words, we can think of the generators as surrogate operators. The operators ’s are called the canonical Z operators on a set of virtual qubits (syndrome qubits).
A destabilizer (group) D associated with a stabilizer , is a stabilizer of , generated as such that for each ,
Just as stabilizer generators can be thought of as canonical operators, destabilizer generators can be thought of as canonical operators.
Let be a stabilizer of with generators. A full tableau is a pair , where is a corresponding destabilizer.
Proposition 2. Given a full tableau , every can be decomposed as in time, where , and .
Proof. Since is a full tableau, then generators of are canonical operators and up to a global phase, they can together generate . Therefore, such a decomposition exists. Given , is in the factorization of if and only if (). We note that checking each of the anti-commutation relations takes time. After finding and , we compare the phases of and to find , which gives the desired decomposition in time.□
The matrix representation of a full tableau is an matrix as follows:
where if the global phase of the -th generator is 1(−1). The standard initial tableau is the following matrix,
We say that two
-qubit unitary stabilizer circuits
and
are
equivalent, if the final states of them are equal, i.e.,
for all stabilizer states
. By applying linearity, it follows that two equivalent stabilizer circuits have the same action on all states. In addition, there is a one-to-one correspondence between unitary stabilizer circuits and tableaus. That means, given a tableau matrix
, there is a unitary stabilizer circuit
such that
is the final tableau when
is run on
. This fact is based on a key theorem which says any unitary stabilizer circuit has an equivalent circuit in canonical form, i.e., it consists of 11 rounds in the following sequence [
16],
Corollary 4. An -qubit state is a stabilizer state if and only if it can be obtained from by a stabilizer circuit.
Proof. Given a stabilizer state , let be its tableau matrix and be the unitary stabilizer circuit corresponding to . Therefore, . The proof of converse direction is clear.□
Corollary 5. Any -qubit stabilizer state can be transformed into the state by applying a stabilizer circuit.
Proof. According to Corollary 4, there is a unitary stabilizer circuit such that . It suffices to reverse to do the inverse transformation. In this way, we reverse the order of gates and replace every gate with .□
There is a simulation algorithm based on applying tableaus which takes
time for measurements [
16]. Furthermore, tableau representation has a key role in the development of simulation algorithms in the generalized stabilizer formalism [
31].
6.4 Frame representation
Let be the stabilizer matrix of an -qubit stabilizer state . By modifying the leading phases of , we can construct additional orthogonal stabilizer states. These states, together with , form an orthonormal basis. This basis is specified by , and any of the other basis states specifies the same basis.
A set of
stabilizer states
which form an orthonormal basis for a subspace of
, is called an
-qubit
frame [
32]. We represent
by a pair
, where
is a stabilizer matrix and
is a set of distinct phase vectors with
. We represent
by
denoting the ordered assignment of the elements in
as the
-phases of the rows in
. The
size of the frame is denoted by
, which is equal to
. As an example, for stabilizer states
and
, with
and
we have
Consider state which is stabilized by . Conjugating the stabilizer by the phase gate gives . However, in the vector representation we have . Hence, the global phase is not maintained by the stabilizer. This example shows that when simulating single stabilizer states, global phases are unobservable. However, when simulating stabilizer-state superpositions, global phases become relative. To see this, we note that acting a Clifford gate on maps the basis to , where is the global phase of the stabilizer state . In other words, this operation rotates and along this rotation, the superposition maps to . Here, the global phase of each becomes relative with respect to , thus we need to compute such phases.
Stabilizer frames provide a simulation algorithm based on the global-phase maintenance as follows. According to Corollary 4, for any stabilizer state there exists a stabilizer circuit such that . To compute the global phase of , we can keep an account of the global factors generated when each gate in is simulated sequentially. In this way, the global phase of each state in is maintained by using the amplitude vector . Each forms a pair with phase vector , and in the process of simulating Clifford gates , is updated according to the following algorithm.
Algorithm 2. (Updating global phases)
1) Set the -phases of to .
2) Store nonzero amplitude of the basis state .
3) Find state and amplitude by computing .
4) Find by computing and store its amplitude .
5) Update the global factor as .
We note that in step 2, for , it may be necessary to sample a sum of two nonzero (real and imaginary) basis amplitudes.
In order to sample the computational-basis amplitudes and , needs to be in row-echelon form (see Theorem 9). Therefore, simulating with global phase-maintenance takes time. We can improve this runtime by noting that during simulation, the form of is invariant, i.e., remains in row-echelon form. This invariance can be repaired with row multiplications, each of which with runtime. It results that updating takes time. Hence, the overall runtime for simulating a Clifford gate is .
In the next section, we study some important classes of quantum channels, which have a key role in quantum information theory.
7 Quantum channels
Let be a linear map. is called positive, if for any positive semi-definite matrix , is so. Given a positive integer , is called -positive, if is a positive map, where is the identity map. is called completely positive, if for any positive integer , is -positive. is called trace-preserving, if , for every . A completely positive and trace-preserving map is called quantum channel. For example, Pauli channel maps an input state with density operator as follows:
where
and
. In the case of
, the Pauli channel is called the
depolarizing channel, which is one of the most important instances of diagonal channels [
33].
7.1 Stabilizer channels
A quantum channel that maps any stabilizer state to a stabilizer state is called a stabilizer channel. Stabilizer channels include two classes of quantum channels. The first one is the class of Clifford channels, i.e. channels that map to where is a Clifford operator. The second one is the class of nonunitary channels on the space of stabilizer states, which map any state to a determined state.
A Pauli reset channel associated with a Pauli product is an example of stabilizer channels, which sets any given qubit to a +1 eigenvector of . This channel is denoted by and can be written as follows:
where
is the Clifford channel that maps
to
. We note that
is the projector onto the
eigenspace of
. Stabilizer channels give rise to calculate the expectation value for an observable on the final state of a circuit by using the Monte Carlo method [
34].
Theorem 11. Any quantum channel can be decomposed as , where ’s are stabilizer channels and with .
Proof. For Pauli products and , let be the superoperator (generally unphysical channel) that maps to and every other Pauli product to . Since Pauli products span the space of operators on qubits, thus is a basis for the space of qubit channels. For Pauli products and we define when and commute (anti-commute). Since every nonidentity Pauli product commutes with exactly half of the Pauli products and anti-commutes with the remaining elements, then
Let be the channel that maps to , then by applying Eq. (17) and the fact that we have
It shows that can be decomposed as a linear combination of stabilizer channels with real coefficients summing to 1. By applying Eq. (16) we obtain the following decomposition for where ,
For and , we can decompose as follows
where is the Clifford channel that maps to . We take into account that a quantum channel cannot have a superoperator of the form as a component, since this superoperator maps two operators and which have equal traces to operators with nonequal traces. Equations (18)−(20) show that each allowed basis element can be decomposed as desired, so can any quantum channel.□
This decomposition is more important than the decomposition of unitary channels as the complex linear combinations of Pauli channels [
16]. In fact, the importance of this decomposition is that it can represent both unitary and non-unitary channels. As an example, let
,
and
be the channels corresponding to
,
and
gates, respectively. We have the following stabilizer decomposition
where is the identity channel.
We note that the stabilizer decomposition is not unique and
’s may be negative. Each term in this decomposition describes a physical process and only when
for all
, they can be interpreted as probabilities of physical processes. Generally,
’s may be considered as a quasiprobability distribution. Accordingly, we can define
negativity of a decomposition as follows:
. In fact,
indicates nonclassical behaviour and traces back to the development of the Wigner function [
35].
7.2 Capacity of the erasure channel
The capacity of a quantum channel is the highest rate of reliable information transmission through many independent uses of the channel. To determine reliability, we need to compare the input and output states, and , by applying the following functional which is called the fidelity, , where and is typically a mixed state. The quantum capacity of a quantum channel is the largest number such that for any and any , there are block sizes and and a quantum error-correcting code mapping states of qubits into independent uses of the channel with , such that every state can be recovered with fidelity at least at the receiving end.
The encoder and decoder are completely positive and trace preserving maps such that maps qubits into intermediate systems, then each of them is sent through an independent instance of the channel , and finally maps the channel outputs into qubits (Fig.4).
Fig.4 Quantum information transfer using a code has a rate of . |
Full size|PPT slide
A quantum erasure channel with probability is a quantum channel from states in into states in defined by
where is an orthogonal state to and corresponds to a lost qubit. When we apply the erasure channel, each qubit is erased independently with probability . An erased qubit is subjected to a random uniform Pauli error ( or with equal probability ) and we know that this qubit is erased.
For qubits, the characteristic vector of the erased locations, , is defined as an element of which each of its components follows a Bernoulli distribution of probability . When a quantum state is subjected to a random Pauli error , the qubit in location is lost if and only if . And if . We say that erasure covers the error , if , i.e., the support of is included in the set of erased locations.
An encoded state
is transformed by a random error
into a state
for which we know that
covers
. To recover the original state, we need to compute the corresponding syndrome of
and then infer an error
covered by
. If
and
are in the same coset of
, then
and therefore the effect of
is corrected. In this case, the erasure is called
correctable. If
is not equivalent to
, we cannot recover the quantum state in general. In this case, the erasure is called
non-correctable. To see an example of a non-correctable erasure, refer to Ref. [
36].
The capacity of
is upper bounded by
, which follows from the no-cloning theorem, see Ref. [
37]. Hence, it does not rely on the quantum code structure. Therefore, we can find achievable rates of stabilizer codes over
and then the capacity is derived. A family of quantum codes
is called
capacity achieving if for sufficiently large
, we can find a code
with rate arbitrarily close to the capacity. Theorem 12 shows that the family of stabilizer codes is capacity achieving over
, thus
.
Theorem 12. ([
36])
Let be a family of stabilizer codes of rates and achieving vanishing decoding error probability over .
Furthermore, suppose that every code has a set of generators of its stabilizer whose weights are upper bounded by .
Then
Compared to Pauli errors, erasures or errors with known locations are a more desirable error type for quantum error-correcting codes. Therefore, transforming physical noise into erasures can considerably improve the performance of quantum error correction [
38].
8 Tomography
The main task of tomography is to reconstruct a full classical description of an
-qubit state from experimental data. In the process of developing methods of tomography, the following question entitled
shadow tomography problem is raised. Given an
-qubit state
and two-outcome measurements
, estimate
up to error
, for
. This problem is solved by designing a quantum procedure with
copies of state, where
hides factors which are polynomial in
and
[
39].
After this brief introductory note, let us describe the framework of classical shadows. Let be an unknown -qubit state. First we apply a set of unitary evolutions , where is randomly selected from an ensemble and then we measure the rotated state in the computational basis . Here, different ensembles lead to different versions of procedure and in practical scheme, each ensemble unitary should be realizable as an efficient quantum circuit. We suppose that this collection is tomographically complete, i.e., for each there are and such that . The framework of classical shadow estimation is based on the following algorithm.
Algorithm 3. (Constructing classical shadow)
1) Prepare a copy of the unknown quantum state .
2) Sample randomly.
3) Transform .
4) Perform a measurement in the computational basis.
5) Apply the inverse of to the resulting state from 4 for collapsing to , where .
6) Let .
7) Construct a classical snapshot by computing
State is called classical shadow, which has a unit trace, but need not be positive semi-definite. We note that tomographically completeness ensures that exists.
Classical shadow is a useful tool in numerical experiments based on random Clifford measurements [
40]. We know that a stabilizer circuit can be simulated efficiently in
time by applying the algorithm based on tableau representation. This algorithm enables us to simulate systems efficiently with even more than 160 qubits in practice. To test the performance of predicting with classical shadows, we need to implement repeatedly the following efficient protocol.
1) Sample a Clifford unitary
from
by applying the algorithm proposed in Ref. [
41].
2) The action of on - and -type operators is fully characterized by parameters
as follows:
where and and .
3) A parameterized unitary by can be applied on stabilizer states by changing the stabilizer generators and the destabilizer generators.
4) A computational basis measurement can be simulated by the improved algorithm based on tableau representation.
9 Discussion
In this section, we present some research lines as follows.
The family of quantum low-density parity-check (qLDPC) codes is a subclass of CSS codes, which plays a central role in classical error correction [
43]. Existence of qLDPC codes whose minimal distance scales linearly with the number of qubits is one of the major open problems in quantum information [
44]. Furthermore, to see a set of open problems on geometric and combinatorial aspects of qLDPC codes, refer to Ref. [
45].
The
stabilizer formalism is an extension of the stabilizer formalism which includes fractional
rotations around the
axis, where
is an integer. There is an equivalence between
stabilizer states and weighted hypergraph states. To see a set of open problems in the
stabilizer formalism, refer to Ref. [
46].
10 Conclusion
The process of measuring a Pauli observable is discussed. It is shown that there exists a stabilizer witness consisting of a full set of generators, which is coarser than the GHZ witness. Moreover, with less than generators, it is not possible to detect genuine -qubit entanglement. A stabilizer code is constructed from a given linear code. The evolution of a quantum system in the presence of noise is discussed by giving the error-correcting criteria. Moreover, the process of syndrome extraction for correctable errors is discussed. By applying logical bases, it is shown that any stabilizer code is unitarily equivalent to a trivial code. Given a graph code, its stabilizer generators are obtained. By applying the cleaning lemma, the distance of a stabilizer code embeddable in a lattice is given. By applying stabilizer matrices, the runtime of simulating stabilizer gates is obtained. The Knill−Gottesman theorem on classical simulation of stabilizer circuits is discussed. It is shown that given a full tableau, every Pauli product can be decomposed efficiently into a product of a stabilizer and a destabilizer. An algorithm for updating global phases by applying frame representation is given. Resolution of a quantum channel into a sum of stabilizer channels is shown. A family of capacity achieving stabilizer codes is applied to obtain the capacity of the quantum erasure channel. An algorithm for constructing classical shadow is given.
{{custom_sec.title}}
{{custom_sec.title}}
{{custom_sec.content}}