Partially Concatenated Calderbank-Shor-Steane Codes Achieving the Quantum Gilbert-Varshamov Bound Asymptotically
Jihao Fan,
Jun Li,
Ya Wang, Yonghui Li, Min-Hsiu Hsieh, and Jiangfeng Du
This work was supported in part by the National Natural Science Foundation of China (No. 61802175), in part by the National Natural Science Foundation of China (No. 61872184), in part by the Fundamental Research Funds for the Central Universities (No. 30921013104), and in part by Future Network Grant of Provincial Education Board in Jiangsu. (Corresponding
authors: Jun Li; Min-Hsiu Hsieh.)J. Fan is with School of Cyber Science and Engineering, Nanjing University of Science and Technology, Nanjing 210094, China (e-mail: jihao.fan@outlook.com).J. Li is with School of Electronic and Optical Engineering, Nanjing University of Science and Technology, Nanjing 210094, China (e-mail: jun.li@njust.edu.cn). Y. Wang and J. Du are with CAS Key Laboratory of Microscale Magnetic Resonance and Department of Modern Physics, University of Science and Technology of China, Hefei 230026, China (e-mail: {ywustc, qcmr}@ustc.edu.cn).Y. Li is with
School of Electrical and Information Engineering, the University of Sydney, Sydney, NSW 2006, Australia (e-mail: yonghui.li@sydney.edu.au).M.H. Hsieh is with Quantum Computing Research Center, Hon Hai Research Institute, Taipei City 114, Taiwan (e-mail: min-hsiu.hsieh@foxconn.com).
Abstract
In this paper, we utilize a concatenation scheme to construct new families of quantum error correction codes achieving the quantum Gilbert-Varshamov (GV) bound asymptotically. We concatenate alternant codes with any linear code achieving the classical GV bound
to construct Calderbank-Shor-Steane (CSS) codes. We show that the concatenated code can achieve the
quantum GV bound asymptotically and can approach the Hashing bound for
asymmetric Pauli channels. By combing Steane’s enlargement construction of CSS codes, we derive a family of enlarged stabilizer codes achieving the quantum GV bound for enlarged CSS codes asymptotically. As applications, we derive two families of fast encodable and decodable CSS codes with
parameters
and
We show that can be encoded
very efficiently by circuits of size and depth . For an input error syndrome, can correct any adversarial error of weight up to half the
minimum distance bound in time. can also be decoded in parallel in time by using classical processors.
For an input error syndrome, we proved that can correct a linear number of -errors with high probability and an almost linear number of -errors in time. Moreover, can be decoded in parallel in time by using classical processors.
Quantum systems are vulnerable to the environment noise induced by decoherence, which is one of the biggest obstacles in quantum information processing. Similar to classical systems, one feasible solution is to exploit quantum error correction codes (QECCs) to encode the primitive quantum information into a larger quantum state. It is widely believed that QECCs are necessary in realizing long-term quantum communications and in building fault-tolerant quantum computers [1, 2]. The construction and design of QECCs with excellent performance is thus one significant task. QECCs can be constructed from classical linear codes, e.g., by using the stabilizer formalism [3] or the Calderbank-Shor-Steane (CSS) construction [4, 5].
But it is not an easy task to do that straightforwardly since an additional dual-containing constraint is needed.
How to construct asymptotically good QECCs with positive rates and linear distances,
is one of the key problems in quantum coding theory [5, 6]. The quantum Gilbert-Varshamov (GV) bound is an important lower bound that promises the existence of such good quantum codes [5, 7, 8, 9, 10, 11].
Compared with classical codes, it is usually much more challenging to obtain quantum codes that can attain the quantum GV bound due to the dual-containing constraint [3]. Therefore very few types of QECCs can attain the quantum GV bound up to now. In [5], it was shown that quantum codes derived from dual-containing codes can attain the quantum GV bound for CSS codes. In [5], random codes were shown to attain the quantum GV bound for binary stabilizer codes. In [9], stabilizer codes derived from self-orthogonal quaternary codes were shown to attain the quantum GV bound for binary stabilizer codes.
In [12], asymptotically good concatenated quantum codes (CQCs) were constructed. In [13], it was shown that CQCs can attain the quantum GV bound for general stabilizer codes asymptotically. A finite version of the quantum GV bound was given in [14]. In [11], quantum GV bounds for asymmetric quantum codes (AQCs) were given. All the known asymptotic results about quantum GV bound rely on some random code arguments which mean that there is little structure in them.
It should be noted that, a unique phenomenon in quantum coding theory, called error degeneracy, exists, which makes degenerate codes more powerful than nondegenerate ones [15, 16]. In an extreme case, it is possible to construct asymptotically good quantum codes from asymptotically bad classical codes. We refer to such codes as extremely degenerate codes. For example,
the qLDPC conjecture asks whether there exist families of
asymptotically good quantum low-density parity-check (qLDPC) codes [17, 18, 19].
Recently, it was shown that asymptotically good qLDPC codes exist with high probabilities [20, 21] and the resultant qLDPC codes are extremely degenerate. The previous results about quantum GV bounds usually rely on asymptotically good classical linear codes. It was unknown whether there exist extremely degenerate codes derived from bad classical codes can attain the quantum GV bound.
In [22], an enlargement of the CSS construction was proposed to construct more efficient stabilizer codes than the standard CSS codes. In [23], the enlargement was generalized to nonbinary situations. Several asymptotic bounds for enlarged CSS codes were also given in [22, 23]. Then in [24, 25, 26], asymptotically good binary quantum algebraic geometry (AG) codes were constructed by using enlargement of the CSS construction. In [27], asymptotically good concatenated quantum AG codes were constructed by using the CSS construction and its enlargement, and these codes can be decoded efficiently in polynomial time.
However, whether binary enlarged CSS codes can attain the corresponding quantum GV bound is unknown. On the other hand, it should be noted that quantum AG codes over sufficiently large field size can even beat the quantum GV bound [28, 29, 30]. However, for codes over small field size, e.g., binary codes, quantum GV bound is much better than asymptotic bounds from quantum AG codes.
In addition to the construction of QECCs with good parameters, practical QECCs need to equip with both efficient encoders and decoders. If the syndrome decoding is slower than the error accumulations, additional noise will be introduced during the syndrome decoding [31]. However, error degeneracy does not simplify the syndrome decoding of QECCs [32]. In contrast,
it makes most decoding methods that are efficient for classical codes fail to fully decode QECCs, e.g., the belief-propagation (BP) algorithm [33]. How to utilize
error degeneracy to correct more errors is an attractive
and crucial issue, and its importance, in our opinion, is
no less than the construction of good QECCs.
On the other hand, the encoding of QECCs is also significant but it has received much less attention compared to the decoding. Noise not only occurs in the transmission and decoding, but also in the encoding circuit [6]. Large-scale encoding circuit and deep encoding depth make the risk of qubits suffering from faulty gates increase rapidly. Consequently, the qubits might be corrupted even before the transmission and the computation. The encoding circuit of stabilizer codes of length with stabilizer generators generally has gates and depth [34]. But in many practical applications, the encoding circuits need to be with linear size and sublinear depth [35, 36].
In this work, we combine classical alternant codes with some other linear codes to construct QECCs achieving the quantum GV bound asymptotically. We use classical alternant codes as the outer code and use any linear code achieving the classical GV bound as the inner code to do the concatenation. We call the constructed codes as partially concatenated CSS (PC-CSS) codes since we only introduce concatenation in the -stabilizers (or the -stabilizers). We show that PC-CSS codes can attain quantum GV bounds for CSS codes and asymmetric CSS codes asymptotically. Here, our scheme can use any linear code achieving the classical GV bound to do the concatenation. Then PC-CSS codes can achieve the quantum GV bound asymptotically.
In particular, if we use low-density parity-check (LDPC) codes as the component codes, we show that error degeneracy can greatly improve the minimum distance of PC-CSS codes. In such case,
the minimum distance of PC-CSS codes without considering error degeneracy is constant to the block length.
Yet, degenerate PC-CSS codes can achieve the quantum GV bound whose minimum distance is linear to
the block length. Thus PC-CSS codes are extremely degenerate and error degeneracy has large advantages in constructing asymptotically good quantum codes. It should be noted that PC-CSS codes are not qLDPC codes since only -stabilizers of PC-CSS codes can satisfy the qLDPC constraints [37]. It is known that asymptotically good qLDPC codes has been shown to exist with high probabilities [20, 21]. However, the asymptotic bound of qLDPC codes in [20, 21] is much weaker than quantum GV bound. We show that PC-CSS codes can not only attain the quantum GV bound but also can approach the capacity of Pauli channels with large asymmetries.
TABLE I: Summary of Asymptotic Bounds for Different Quantum Codes. Let be a prime and let be an integer. Denote by . For a stabilizer code with parameters , we denote by and . For a nonstabilizer code with parameters , we denote by and . For an AQC with parameters , we denote by , , and .
As a special case, we construct a family of asymmetric PC-CSS codes with parameters
by using expander codes as one of the component codes, where and are integers and .
Let , we can derive
a family of quantum codes with parameters
For an input error syndrome, can correct all errors of weight smaller than half the minimum distance bound in time. can also be decoded in parallel in time by using classical processors. We show that can be fast encoded by a circuit of size and depth . Moreover, can correct a random -error with very high probability provided .
If let , then we can derive a family of AQCs with parameters
For an input error syndrome, can correct any random -error with high probability and correct an almost linear number of -errors in time. Further, we show that can be decoded in parallel in time by using classical processors. While fault-tolerant quantum computation [38] and communication [39, 6] prefer fast encodale and decodable codings, and are thus practical and beneficial for the future use.
We further combine our concatenation scheme with the enlarged CSS construction in [22, 23]. We show that the enlarged quantum codes can attain the quantum GV bound for enlarged CSS codes. In Table I, we give a summary of asymptotic bounds for different quantum codes.
This paper is organized as follows. We begin in Section II with Preliminaries on classical linear codes and quantum stabilizer codes. In Section III, we present the partial concatenation construction of CSS codes and show that PC-CSS codes can attain the quantum GV bound for CSS codes and AQCs asymptotically. In Section IV,
we give a family of fast encodable and decodable PC-CSS codes. In Section V, we combine our concatenation scheme with Steane’s enlarged CSS construction. We show that the enlarged quantum codes can attain the quantum GV bound for enlarged CSS codes asymptotically. The discussions and conclusions are given in Section VI.
II Preliminaries
We shortly introduce some needed knowledge about classical linear codes and quantum stabilizer codes, for more details, see the literatures, e.g., [40, 5, 3, 41, 42]. In this paper, we usually neglect
the index in the code parameters for binary linear codes and
binary stabilizer codes provided there is no ambiguity.
II-1 Classical Alternant Codes
Let be a prime and denote by a prime field. Let be a power of , where is an integer. Let be the Galois field of size and denote by a field extension of
, where is an integer. Define the trace operation from to as , where . We review some known results about classical alternant codes which will be used in the construction of quantum codes.
Let be distinct elements of and let be nonzero elements of , where . Denote by and . Denote the ring of polynomials with coefficients in finite field by .
For any polynomial , the degree of is denoted by . For any ,
the generalized Reed-Solomon (GRS) code is defined by
(1)
The parameters of are given by
and GRS codes are maximum-distance-separable (MDS) codes which can attain the Singleton bound in [40]. Let .
The dual of a GRS code is also a
GRS code, i.e., ,
where and , for . The parity-check
matrix of is given by
(2)
For some given GRS code over , the alternant code over denoted by , is defined as the subfield subcode of , i.e.,
(3)
Since consists of all codewords of over , the parity check matrix of is given by that of in
(2).
The integer is called the degree of the alternant code . In addition, is called the support of , and is called a multiplier of . For the dimension and minimum distance of alternant codes, there is the following result:
Lemma 1
[40]
Let be an alternant code defined in (3). The dimension and minimum distance of can be given by and , respectively.
Let be any nonzero vector over . Let be a fixed vector and let be a varying vector. According to [40, Chap. 12, Theorem 3], we know that the number of GRS codes
containing is at most . Therefore, the number of alternant codes containing is also at most . Here, it should be noted that the estimate of the number of alternant codes that contain does not relate to its Hamming weight.
Lemma 2
[40]
For any nonzero vector , the number of alternant codes in (3) that contain is at most .
On the other hand, the total number of alternant codes is equal to the number of choices for different (or ), which is given by [40].
Thus alternant codes form a large family of linear codes and include many important subclasses, e.g., Goppa codes and Bose-Chaudhuri-Hocquenghem codes. Further, alternant codes play a significant role in the McEliece cryptosystem [43]. Moreover, alternant codes can attain the classical GV bound asymptotically [40]. In the next section, we will show that alternant codes can be used to construct concatenated CSS codes achieving the quantum GV bound asymptotically.
For integers , denote by the Hamming ball of radius . Denote by the -ary entropy function. There exists the following important estimate of the Hamming ball.
Lemma 3
[44]
Let be a power of a prime . For an integer and ,
(4)
II-2 Quantum Stabilizer Codes
Let be a primitive
th root of unity and let . Denote the unitary operators and on by
and , respectively. Denote by the finite group
(5)
A stabilizer code is a -dimensional subspace of the Hilbert space , i.e.,
(6)
where with stabilizer generators () is an Abelian subgroup of . is called the stabilizer group. By measuring the eigenvalues of the stabilizer generators, the syndrome of errors can be revealed. An error is detectable if it anticommutes with some stabilizer generator . A detectable error results in a nonzero syndrome, otherwise the syndrome is zero and the error is undetectable.
The minimum distance of code is the minimum weight of an
undetectable error , which does not belong to the stabilizer group. We denote by . The stabilizer code is called nondegenerate if the stabilizer group does not contain a nontrivial element whose weight is less than , otherwise it is degenerate.
The CSS construction in [5, 45] presents a direct way to construct stabilizer codes from two classical linear codes that satisfy the dual-containing relationship.
Let and be two classical linear codes over such that and . Then there exists a stabilizer code with parameters
where .
In most quantum channels, the dephasing errors (-errors) usually happen much more frequently than the amplitude errors (-errors). Asymmetric quantum codes (AQCs) are thus proposed to deal with such biased quantum noise [46, 30, 47]. Moreover, the CSS construction can be used to derive AQCs with two classical linear codes, one for correcting -errors and
the other for correcting -errors.
Let and be two -ary
linear codes with parameters
and , respectively. If , then there exists a AQC, where
(8)
(9)
If and , then is nondegenerate, otherwise it is degenerate.
III Asymptotic Goodness of Partially Concatenated Calderbank-Shor-Steane Codes
Denote by a linear code with the parity check and generator matrices given by and , respectively. Let be another linear code with the parity check and generator matrices given by and , respectively. We use as the inner code and as the outer code, and we concatenate the generator matrix of with the parity-check matrix of as follows:
(10)
Denote the null space of by and we have , where . Denote by
(11)
It is easy to see that and
. Thus we can derive a partially concatenated CSS (PC-CSS) code with parameters by using the CSS construction.
We show that PC-CSS codes can attain the quantum GV bound asymptotically.
Theorem 1
There exist PC-CSS codes with parameters achieving the quantum GV bound asymptotically. can achieve a quantum rate
(12)
for any block length , where is the relative minimum distance.
Proof:
Let be any classical linear code. Let be an alternant code. We construct a pair of dual-containing codes according to (10) and (11). By using the CSS construction, we can derive a PC-CSS code with
(13)
where , and . Since can be chosen arbitrarily, we let be
an asymptotically good linear code which can attain the classical GV bound, i.e.,
(14)
where is the relative minimum distance. Next we need to determine . Since we know nothing about the distance of the dual code , there
maybe some (very) low weight vectors coming from . If we do not consider error degeneracy, the minimum distance is unknown and maybe very small. Fortunately, vectors come from do not affect the computation of because
they are degenerate. For an integer , let be any nonzero vector with Hamming weight less than . Denote by
(15)
If , then must belong to and
is degenerate when we compute . Therefore we only need to consider the case which makes . Let and denote
by , where is an integer and .
For some given and varying , let be a GRS code over . Let
(16)
be an alternant code, where and , for . By Lemma 1, the dimension of satisfies
(17)
By Lemma 2, we know that the number
of alternant codes that contain does not relate to the Hamming weight of and is at most
(18)
On the other hand, the total number of alternant codes is equal to the number of the choice of , and it is equal to [40]. According to the counting argument in [40], if
(19)
then there exists an alternant code such that there does not exist any nonzero vector with weight less than .
According to the estimate of Hamming ball in Lemma 3 and taking the limit as in (19), we can derive
(20)
Combining (13), (14), and (20) above together, we have
(21)
If we let , where is any constant, then we have
(22)
∎
Suppose that we consider an asymmetry between and in the proof of Theorem 1, and we apply the CSS construction for AQCs in Lemma 6 to PC-CSS codes. Then asymmetric PC-CSS codes can attain the quantum GV bound for AQCs asymptotically.
Corollary 1
There exist asymptotically good asymmetric PC-CSS codes such that
(23)
for any block length , where and are the relative distances, and .
It is known that there exist families of asymptotically good LDPC codes which can achieve the classical GV bound. But the dual of LDPC codes
has a low-density generator matrix and thus the minimum distance is constant. If we use LDPC codes as the inner code , then
the minimum distance of the PC-CSS code without considering error degeneracy is constant to the block length.
While considering error degeneracy, the
low weight codewords in the dual of LDPC codes do not affect the counting argument in the proof of Theorem 1. Therefore, the PC-CSS code is extremely
degenerate and the minimum distance is greatly improved from a constant to be linear with the block length.
By applying the Evra-Kaufman-Zémor distance balancing
construction [17], we show
that our scheme can also lead to qLDPC codes wth non-vanishing rates and minimum distance growing with the square root of the block length [48].
In Corollary 1, PC-CSS codes are shown to attain the quantum GV bound for AQCs [11].
Moreover, we show that asymmetric PC-CSS codes can approach the capacity of asymmetric Pauli channels as the channel asymmetry goes to large.
Suppose that we transmit quantum information over the Pauli channel:
(24)
for an input state , where are the Pauli operators, , and .
Denote the total error probability by . Denote by the asymmetry
for the probabilities of -errors and -errors. Usually, we let , then we have
and .
In Corollary 1, the inner code can be chosen arbitrarily, we let it be an LDPC code, which is used to correct -errors. We use alternant codes to correct -errors. It is known that alternant codes can be decoded up to the GV bound under the bounded minimum distance decoder [40]. LDPC codes can approach the Shannon capacity of the binary symmetric channel under the BP decoding [49, 50]. Then the rate of PC-CSS codes under the asymmetric Pauli channel can approach
(25)
In Fig. 1, we compare the the limit of PC-CSS codes in (25) with the Hashing bound given by over Pauli channels with an asymmetry .
If , then the channel is symmetric with . As the asymmetry grows,
the gap between PC-CSS codes and the Hashing bound becomes increasingly smaller. In Fig. 1, when the asymmetry and , the code rate
gaps are less than and , respectively. Therefore, PC-CSS codes can approach the Hashing bound of Pauli channels as the asymmetry grows. It should be noted that, although our codes are highly degenerate, they can not attain the bound in [15] for very noisy channels .
Figure 1: The comparison of the asymptotic bound of asymmetric PC-CSS codes (solid lines) and the Hashing bound (dashed lines) with different channel asymmetries and . The horizontal axis is the total error probability in the Pauli channel and the vertical axis is the quantum code rate.
The PC-CSS codes in Theorem 1 can also be used to prove the security of the famous BB84 quantum key distribution (QKD) protocol directly [51]. By a simple calculation, we can obtain a total qubit error rate (QBER) less than for the security
of BB84 over Pauli channels. Although the rate is lower than the bound of in [51], our codes are not randomly generated and the LDPC component codes can be efficiently decoded. Since
classical LDPC codes can approach the Shannon bound by using efficient BP decoders, PC-CSS codes are particularly applicable to efficient and practical QKD. Furthermore, if we consider asymmetric errors in Pauli channels, we can further
improve the total QBER for the security of BB84. For example, if we admit an asymmetry between -errors and -errors, then the total QBER is less than . Our codes are competitive with quantum Polar
codes in [52, 53] that can achieve the coherent information. But Ref. [52] needs some preshared entanglement assistance.
IV A Family of Fast Encodable and Decodable PC-CSS Codes
In practice, we need quantum codes to be efficiently encoded and decoded. Let the error syndrome as the input, we show that PC-CSS codes can be fast decoded. Moreover, we also show that PC-CSS codes can be fast encoded.
TABLE II: Parameters and the syndrome decoding complexity of several efficiently decodable quantum codes.
Let and be integers and let . There exists a family of asymmetric PC-CSS codes with parameters . The syndrome decoding of can be done in time. Further, the parallel syndrome decoding of can be done in time by using classical processors. Moreover, can be encoded by a circuit of size and depth .
Proof:
Let be a binary repetition code whose parity check matrix and generator matrix are given by:
(26)
respectively.
Let be a linear code with a parity check matrix and a generator matrix , where is an identity matrix of size .
Let be an asymptotically good expander code in [57] such that
, and , where . Let the parity check and generator matrices of
be and , respectively. Then we can construct an asymmetric PC-CSS code with
(27)
It is not difficult to verify that and .
We consider the syndrome decoding of . For an -error , the error syndrome is given by
(28)
where we denote by . Suppose that ,
then we must have . Thus we can decode
by using the expander code in time [57]. Further, can also be decoded in parallel in time by using classical processors [57]. Let
be the decoded -error, where the s correspond to the redundant qubit positions in . Then we have and is degenerate with respect to .
Next, we use to correct any -error such that
. We divide the parity check matrix into
sub-blocks by columns, and each sub-block corresponds to a diagonal block in . Divide into sub-blocks, i.e., . Then there are at most
erroneous sub-blocks in and each erroneous sub-block has at most errors.
For each erroneous sub-block, we use the one-step majority-logic (OSMLG) decoding method [58] to check the error qubit that corresponds to the last column
in . Let be any sub-block in . Then we can derive the following syndrome for :
(29)
where is an all-ones vector of length . If , then there are at most ones in the vector . Thus the Hamming weight of the syndrome is at most . While if , then there are at most ones in the vector . Then the Hamming weight of the syndrome is at least . Therefore we can use the OSMLG decoding to determine , and then can be derived from (35) directly.
The running time for deriving is determined by the computation of the Hamming weight of the syndrome . The computation time complexity is . Thus the running time for decoding all the erroneous sub-blocks is since there are sub-blocks. Here, the decoding of the sub-blocks can also be carried out in parallel by using classical processors. Thus the parallel syndrome decoding of the -error can be done in time by using classical processors. Overall, the syndrome decoding of can be done in time or can be done in parallel in time by using classical processors.
Next, we consider the encoding of and we use the standard encoding method of CSS codes in [34] to do that. We can encode the pure state as follows:
(30)
where . It is not difficult to verify that the generator matrix of is given by
(31)
where is a zero matrix of size , and is in the standard form. Denote by , where each is a column of and . Then the encoding circuit of the PC-CSS code is given in Fig. LABEL:EncodingCircuitBacon. In Stage I of the encoding circuit, the number of gates is determined by the density of and it is at most . It is easy to see that the depth of Stage I is . While in Stage II, the number of C-NOT gates is exactly and the depth is . Therefore the size of the encoding circuit is and the depth is .
∎
Notice that the last column in in the proof of Theorem 2 is quite dense and then is also dense. However, we can transform into a sparse matrix
by multiplying a series of elementary matrices in the left. Therefore the -stabilizer generators of satisfy the
qLDPC constraint [37] . In the fault tolerant settings, we require quantum codes to correct a linear number of random errors. We consider the correction of random errors by using PC-CSS codes. We assume that the noise model used is the
independent error model
in [59, 38, 60]. Denote by the probability of each -error.
It is shown that the PC-CSS code in Theorem 2 can correct any
adversarial -error of weight less than
half the distance bound.
We count all the uncorrectable errors of weight larger than half the minimum distance bound. It is easy to see that if there is at least one
sub-block that has -errors of weight larger than , then is undetectable. All the uncorrectable
-errors are counted as follows in equations (32) - (34).
(32)
(33)
(34)
We denote by
the probability of uncorrectable -errors and denote by . Let be any constant. If we let and let , then vanishes as . If we let , where is constant, then vanishes when as . Therefore the PC-CSS code in Theorem 2 can correct a linear number of random -errors with high probability in time as long as . However, for the correction of random -errors, we cannot get a non vanishing threshold.
If we let or , we can derive the following two families of fast encodable and decodable quantum codes.
Corollary 2
There exists
a family of quantum codes with parameters
For an input error syndrome, can correct all errors of weight smaller than half the minimum distance bound in time. can also be decoded in parallel in time by using classical processors. We show that can be encoded efficiently by a circuit of size and depth . Moreover, can correct a random -error with very high probability provided .
Corollary 3
There exists a family of quantum codes with parameters
For an input error syndrome, can correct an almost linear number of -errors and can also correct a linear number of -errors with high probability in time. Further, can be decoded in parallel in time by using classical processors.
In [27], families of concatenated quantum AG codes were constructed and decoded efficiently in
polynomial time, while our codes in Theorem 2 can be decoded in linear time. Therefore our codes are more efficient than concatenated quantum AG codes in [27]. In particular, the parallel syndrome decoding of Corollary 2 can be carried out in sublinear time, and that of Corollary 3 can be carried out in logarithmic time. In Table II, we compare the parameters and decoding complexity of several efficiently decodable quantum codes.