Fact-checked by Grok 2 weeks ago

Stabilizer code

Stabilizer codes are a class of quantum error-correcting codes defined by an abelian subgroup S of the n- Pauli group, where the code consists of the joint +1 eigenspace of all elements in S, enabling the protection of logical against errors induced by decoherence and noise in . These codes are parameterized as [[n, k, d]], where n is the number of physical qubits, k is the number of encoded logical qubits (with dimension $2^k = 2^{n-r} where r is the number of independent generators, so |S| = 2^r), and d is the code representing the minimum weight of any nontrivial logical . The d determines the error-correcting capability, allowing detection of up to \lfloor (d-1)/2 \rfloor errors and correction of up to \lfloor (d-1)/2 \rfloor errors via measurements that reveal error patterns without disturbing the logical state. Introduced independently in by Daniel Gottesman in his Ph.D. thesis and by Calderbank, Rains, Shor, and Sloane in their work on quantum stabilizer codes derived from classical linear codes, stabilizer codes provide a unified group-theoretic framework for constructing and analyzing a wide variety of quantum error-correcting schemes. This generalizes classical linear error correction to the quantum setting, where stabilizer generators—products of [X, Y](/page/X&Y), Z—must commute to ensure simultaneous measurability, and the code space is stabilized by these s acting as the identity on valid states. Notable subclasses include CSS codes, which are built from pairs of classical linear codes and admit transversal gates for fault-tolerant operations, and surface codes, which achieve high thresholds for error rates due to their topological structure on a . Stabilizer codes form the foundation for fault-tolerant quantum computation, enabling the implementation of universal quantum gates through Clifford operations and measurements while bounding the overhead required to suppress error rates below thresholds like 1% for practical scalability. Key examples encompass the 5-qubit code (perfect for single-qubit errors with d=3), the 7-qubit (derived from the ), and the 9-qubit Shor code (a concatenated construction achieving d=3), all of which demonstrate how stabilizer simplifies error decoding and logical identification.

Introduction

Overview

Quantum systems are highly susceptible to decoherence and , which rapidly degrade the fragile encoded in qubits, making reliable quantum computation challenging without protective measures. These errors arise from interactions with the surrounding environment, leading to loss of and in quantum states. codes address this need through a that defines a protected codespace as the simultaneous +1 eigenspace of an abelian of the , enabling systematic detection and correction of errors. This approach encodes k logical qubits into n physical qubits, allowing the code to correct up to t errors while preserving the logical information. The performance of stabilizer codes is characterized by the parameters [[n, k, d]], where n denotes the number of physical qubits, k the number of encoded logical qubits, and d the minimum distance, which determines the code's error-correcting via t = \lfloor (d-1)/2 \rfloor. This has become foundational for fault-tolerant by providing a structured to mitigate without direct measurement of the .

Historical context

The foundations of quantum error correction were laid in 1995 when Peter Shor proposed the first explicit quantum error-correcting code, a nine-qubit scheme capable of correcting arbitrary single-qubit errors by encoding one logical qubit into a subspace protected against decoherence. This work demonstrated that quantum information could be safeguarded against noise, inspiring further developments in constructing more efficient codes. Building on this, the Calderbank-Shor-Steane (CSS) codes emerged in 1996, adapting classical linear error-correcting codes over GF(4) to quantum settings, allowing simultaneous correction of bit-flip and phase-flip errors while leveraging dual classical codes for syndrome measurement. In his 1997 PhD thesis, Daniel Gottesman introduced the formalism as a general framework for , extending CSS codes by defining the code subspace as the simultaneous +1 eigenspace of an Abelian subgroup of the , known as the stabilizer group. Independently, Calderbank, Rains, Shor, and Sloane introduced the stabilizer formalism in a 1997 deriving quantum codes from classical linear s. This abstraction unified previous constructions and enabled the design of diverse codes with efficient extraction via Clifford operations. Gottesman's work was further disseminated through his 1997 preprint and subsequent publication in 1998, where he outlined fault-tolerant operations on stabilizer codes, establishing their role in scalable quantum computation. By the early 2000s, stabilizer codes became integral to fault-tolerant protocols, with Gottesman's theory providing the basis for transversal gates and error s, as explored in concurrent works on concatenated codes and theorems. A significant occurred in 1997 when proposed the surface code, a topological stabilizer code on a two-dimensional that achieves high error s through local stabilizer measurements, linking abstract to physically realizable architectures with anyonic excitations. This connection spurred research into -based implementations, influencing designs throughout the 2000s. In the 2010s, stabilizer codes, particularly surface and color codes, were adopted in proposed quantum computing architectures due to their low overhead and compatibility with nearest-neighbor interactions in superconducting and ion-trap systems. By 2023, experimental demonstrations advanced significantly, with Google's implementation of a distance-5 surface code logical demonstrating suppression of logical s with a reduced logical error rate as the physical count scaled from 17 to 49, marking a milestone toward practical error-corrected computation. As of 2025, stabilizer codes underpin widespread quantum hardware simulations and the realization of error-corrected logical s in leading platforms, enabling longer coherence times and rudimentary fault-tolerant gates in noisy intermediate-scale quantum devices.

Mathematical Foundations

Pauli group

The Pauli group on n qubits, denoted G_n, consists of all unitary operators of the form g = i^k P, where k = 0, 1, 2, 3 and P is a of the single-qubit Pauli operators \{I, [X, Y](/page/X&Y), Z\} acting on the n qubits. This generates a under , serving as the foundational for defining stabilizer operators in codes. The group G_n has $4^{n+1}. Its Z(G_n) = \{ I, iI, -I, -iI \} has 4, and the G_n / Z(G_n) is an of rank $2n over \mathbb{Z}_2. The comprises all scalar multiples of the , which commute with every in G_n. Commutation within G_n is determined by the Pauli strings: two i^{k_1} P_1 and i^{k_2} P_2 commute P_1 P_2 = P_2 P_1, since phase factors commute with all operators. For the Pauli strings P_1 and P_2, they commute the symplectic inner product of their binary vectors (defined in the next subsection) is zero 2; otherwise, they anticommute. These relations are essential for identifying abelian s suitable for stabilizers, as all in such a must pairwise commute. The inclusion of global phase factors i^k is critical, as they influence measurement outcomes and the unitarity of operators; for example, stabilizer generators are typically selected from the Hermitian subgroup with phases \pm 1 to ensure eigenvalues of \pm 1 upon . In quantum error correction, these phases distinguish distinct error operators within G_n, though equivalent errors differing only by a phase produce identical syndrome measurements. For constructing and analyzing quantum codes, the projective representation of G_n modulo phases—effectively quotienting by the center—simplifies the structure to a vector space over \mathbb{F}_2, enabling efficient computational models. In error models, elements of G_n represent correctable Pauli errors, with syndrome extraction relying on their commutation properties.

Binary symplectic representation

The binary symplectic representation provides an isomorphism between the n-qubit Pauli group (modulo phases) and the vector space (\mathbb{Z}/2\mathbb{Z})^{2n}, facilitating the algebraic treatment of stabilizer codes through classical linear algebra over the finite field GF(2). Each Pauli operator, which is a tensor product of single-qubit Paulis I, X, Y, Z on n qubits, maps to a unique 2n-bit binary vector (x_1 \dots x_n \mid z_1 \dots z_n), where the X-part vector (x_1, \dots, x_n) has x_i = 1 if the i-th qubit has X or Y (and 0 otherwise), and the Z-part vector (z_1, \dots, z_n) has z_i = 1 if the i-th qubit has Z or Y (and 0 otherwise). This mapping ignores overall phase factors of the form i^k (where k = 0,1,2,3), which are handled separately to preserve the full group structure. The commutation relations between Pauli operators are captured by the symplectic inner product on these binary vectors. For two vectors \mathbf{u} = (x_u \mid z_u) and \mathbf{v} = (x_v \mid z_v), the symplectic product is defined as \langle \mathbf{u}, \mathbf{v} \rangle = x_u \cdot z_v + z_u \cdot x_v \pmod{2}, where \cdot denotes the standard over GF(2). Two Pauli operators corresponding to \mathbf{u} and \mathbf{v} commute \langle \mathbf{u}, \mathbf{v} \rangle = 0, and anticommute otherwise; this bilinear form is alternating and nondegenerate, endowing the space with a structure. Phases arising from products of Paulis (e.g., XY = iZ) are tracked independently in this representation, as the binary vectors alone do not encode them. While often omitted during code design for simplicity—focusing on the error-correcting —the phases must be accounted for in full simulations or implementations to ensure gate fidelity, particularly when composing operators. This representation transforms the problem of finding stabilizer generators into solving systems of linear equations over GF(2), enabling efficient computation of code parameters and error syndromes using standard operations. For example, the single-qubit Pauli X maps to (1 \mid 0), Z to (0 \mid 1), and Y to (1 \mid 1), with the symplectic product \langle X, Z \rangle = 1 \pmod{2} reflecting their anticommutation.

Core Concepts

Definition of stabilizer codes

Stabilizer codes are a class of quantum error-correcting codes defined using an abelian subgroup S of the n-qubit Pauli group \mathcal{G}_n, which consists of all tensor products of the identity I and Pauli matrices X, Y, Z on n s, multiplied by phases \pm 1, \pm i. The group S must satisfy two key conditions: it is abelian, meaning all elements commute, and it does not contain the element -I, ensuring the existence of a non-trivial fixed . Typically, S is generated by n-k independent Pauli operators g_1, \dots, g_{n-k}, so S = \langle g_1, \dots, g_{n-k} \rangle, where |S| = 2^{n-k} and the generators form a basis for S. The code subspace C(S), or simply the code space, is the set of all states in the n-qubit Hilbert space that are fixed by every operator in S: C(S) = \left\{ |\psi\rangle \in (\mathbb{C}^2)^{\otimes n} \;\middle|\; g |\psi\rangle = |\psi\rangle \ \forall \, g \in S \right\}. This subspace consists of the simultaneous +1 eigenspace of all stabilizers in S, and its dimension is $2^k, allowing the encoding of k logical qubits into n physical qubits. In this framework, encoding logical involves preparing states within C(S), where the logical states are those under the action of all stabilizers. The of the generators ensures that S has the correct size |S| = 2^{n-k}, so that the code space has dimension $2^k. For the trivial code, S = \{ I \}, which encodes all n qubits without any protection, corresponding to k = n. These definitions can be efficiently represented using the binary symplectic formalism, where generators are encoded as rows in a (n-k) \times 2n binary matrix.

Code space and logical operators

The code space of a stabilizer code, denoted C(S), is the $2^k-dimensional subspace of the $2^n-dimensional Hilbert space consisting of all states |\psi\rangle that are fixed by every stabilizer generator, i.e., s|\psi\rangle = |\psi\rangle for all s \in S, where S is the abelian stabilizer group of order $2^{n-k} generated by n-k independent Pauli operators. This subspace encodes k logical qubits and forms the eigenspace with eigenvalue +1 for the projector P_S = \frac{1}{|S|} \sum_{s \in S} s, which maps any state onto C(S). States in C(S) are simultaneous +1 eigenstates of the stabilizer generators, ensuring that the code space is invariant under the action of S. Logical Pauli operators \bar{X}_j and \bar{Z}_j for j = 1, \dots, k are representatives from the N(S)/S, where N(S) is the centralizer (or normalizer) of S in the n- Pauli group G_n, defined as N(S) = \{ g \in G_n \mid gs = sg \ \forall s \in S \}. These operators commute with every element of S (so they preserve C(S)) but act non-trivially on the logical states, satisfying the canonical anticommutation relations \{ \bar{X}_j, \bar{Z}_j \} = 0 and [\bar{X}_j, \bar{Z}_l] = [\bar{X}_j, \bar{X}_l] = [\bar{Z}_j, \bar{Z}_l] = 0 for j \neq l. The group N(S) has order $4 \cdot 2^{n+k} and contains S as a of index $4^k, with the logical operators generating the Pauli algebra on the encoded qubits up to stabilizers. For , the minimum of any non-identity element in N(S)/S must be at least the code distance d, ensuring that low-weight errors cannot mimic logical operations. An encoding unitary U maps an unencoded k-qubit state |c_1 \dots c_k \rangle_L (e.g., in the computational basis) to a state in C(S) via U |c_1 \dots c_k \rangle_L |0\rangle^{\otimes (n-k)} = \left( \sum_{M \in S} M \right) \bar{X}_1^{c_1} \cdots \bar{X}_k^{c_k} |0\rangle^{\otimes n}, where the ancillary qubits are initialized to |0\rangle and the logical X operators \bar{X}_j are applied first. This unitary can be constructed using at most n(n-k) two-qubit gates, preserving the stabilizer structure such that U s U^\dagger \in S for all s \in S. Stabilizer codes exhibit degeneracy when multiple distinct errors E_a and E_b (with a \neq b) belong to the same of N(S), meaning they produce the same and act equivalently on C(S) up to a , i.e., E_a^\dagger E_b \in S. This property allows correction of error sets without distinguishing individual errors, potentially improving efficiency in fault-tolerant schemes, as seen in codes like the nine-qubit Shor code where paired phase flips are degenerate.

Error Correction

Syndrome extraction

In stabilizer codes, the syndrome is a binary vector s \in \mathrm{GF}(2)^{n-k}, where n is the number of physical qubits and k the number of logical qubits, with each component s_i corresponding to the eigenvalue of the i-th generator g_i: s_i = 0 for eigenvalue +1 (commutation with the error) and s_i = 1 for eigenvalue -1 (anticommutation). This vector encodes the pattern without revealing the logical state, as measurements project the system onto eigenspaces of the stabilizers while preserving the code space. Syndrome extraction involves measuring each stabilizer generator g_i using ancilla qubits to avoid disturbing the encoded information. For a Z-type stabilizer like g_i = Z_1 Z_2 \cdots Z_m, an ancilla qubit is prepared in the state |+\rangle = \frac{|0\rangle + |1\rangle}{\sqrt{2}}, followed by controlled-NOT (CNOT) gates with the relevant data qubits as controls and the ancilla as target; the ancilla is then measured in the X-basis, yielding +1 (syndrome bit 0) or -1 (syndrome bit 1). For enhanced fault tolerance, multiple ancilla qubits can be used in cat states such as |0\cdots0\rangle + |1\cdots1\rangle, with CNOTs propagating the parity to the ancillas, followed by majority voting on their measurements to reduce error rates to second order in the physical error probability. Similar circuits apply to X-type stabilizers by conjugating with Hadamard gates. These "gadgets" ensure the measurement is indirect and preserves the logical superposition. The measurements are non-demolition because the stabilizer generators commute—forming an Abelian subgroup of the —allowing repeated extractions without altering the code space in the absence of errors; any detected persists until correction. In the binary representation, the bits correspond to the inner product between the generator matrix rows and the error vector, facilitating efficient computation. Error identification proceeds via lookup: the vector s is matched to the most probable Pauli E such that s_i = \langle \mathrm{row}_i, \mathrm{vec}(E) \rangle over \mathrm{GF}(2) for each generator row, using precomputed lookup tables for correctable errors in non-degenerate codes. codes exhibit closure under the Clifford group, meaning Clifford operations map the stabilizer group to itself, enabling transversal implementations of like Hadamard and CNOT that preserve the code structure during extraction and correction.

Error recovery conditions

In stabilizer codes, an error operator E is correctable if, for every other correctable error F, either E^\dagger F belongs to the stabilizer group S or the syndromes of E and F are distinct. This condition ensures that errors with the same syndrome do not interfere destructively on the code space, allowing unambiguous recovery by identifying the error coset in the . The stabilizer formalism partitions the error space into cosets based on syndromes, where correction succeeds if applying any representative from the coset to a codeword returns it to the code space up to a stabilizer element. The general mathematical framework for quantum error correction is provided by the Knill-Laflamme conditions, which state that a code corrects a set of errors \{E_i\} if \langle \psi | E_i^\dagger E_j | \phi \rangle = \delta_{\psi \phi} c_{ij} for all logical states |\psi\rangle, |\phi\rangle in the code space, where c_{ij} is independent of the logical states. In the context of stabilizer codes, these conditions are satisfied for Pauli errors whose syndromes uniquely identify correctable cosets, ensuring that the logical information remains intact after correction. This criterion generalizes classical error correction by accounting for the non-orthogonality of quantum errors. The error-correcting capability of a stabilizer code is quantified by its d, defined as the minimum (number of non-identity Pauli factors) of any non-trivial logical that stabilizes the code space. A code of d can correct any error of up to t = \lfloor (d-1)/2 \rfloor, as such errors produce distinct syndromes and do not include undetectable logical operations. For instance, errors below this threshold are guaranteed to be recoverable without logical flips, establishing a fundamental limit on the code's robustness against adversarial or noisy channels. Stabilizer codes exhibit degeneracy, where multiple distinct Pauli errors can share the same syndrome, corresponding to the same coset in the normalizer of S. Correction is still possible in degenerate cases because all errors in a given coset act equivalently on the code space (differing by stabilizers, which act trivially), allowing the decoder to choose any representative (typically minimum-weight) whose inverse restores the code space without altering the logical state. This property enhances efficiency, as it permits correcting higher-weight errors without additional overhead, a feature particularly useful in fault-tolerant quantum computing protocols. Stabilizer codes are subject to the quantum Singleton bound, which states that for an [[n, k, d]] code with n physical qubits encoding k logical qubits and d, the d \leq n - k + 1 holds. Codes achieving equality are termed maximum distance separable (MDS) quantum codes, providing optimal trade-offs between encoding rate and error correction; examples include certain CSS codes derived from classical MDS codes. This bound underscores the inherent limitations of , mirroring classical bounds but adapted to the and superposition principles.

Examples

Repetition code

The three-qubit repetition code is the simplest non-trivial example of a stabilizer code, encoding a single logical into three physical s to protect against bit-flip errors. It is constructed using the generators S_1 = Z \otimes Z \otimes I and S_2 = I \otimes Z \otimes Z, which generate the group consisting of the identity and all even-weight products of these operators. The code space is the +1 eigenspace of both generators, a two-dimensional spanned by the basis states |0_L\rangle = |000\rangle and |1_L\rangle = |111\rangle, where the subscripts denote the logical basis. This ensures that the logical states have even in the Z basis, allowing detection of discrepancies caused by bit-flips. The logical Pauli operators for this code are \bar{X} = X \otimes X \otimes X, which flips between |0_L\rangle and |1_L\rangle, and \bar{Z} = Z \otimes I \otimes I (or equivalently Z \otimes Z \otimes Z, as the two differ by the stabilizer S_2), which applies a phase flip to |1_L\rangle. These operators commute with the stabilizers but are not in the stabilizer group, preserving the code space while acting non-trivially on the logical information. The minimum-weight representatives yield a code distance of 1 in the general Pauli error model due to the weight-1 \bar{Z}, but the code achieves an effective distance of 3 for bit-flip (X-type) errors alone, enabling correction of any single X error. Error correction proceeds by extracting the syndrome via non-destructive measurements of the stabilizers using ancilla qubits. For instance, a single X error on the first qubit anticommutes with S_1, yielding syndrome outcomes (-1, +1) when measuring the eigenvalues of S_1 and S_2; errors on the second or third qubit produce distinct syndromes (+1, -1) and (-1, -1), respectively, while no error gives (+1, +1). Correction involves applying an X operator to the erroneous qubit based on the syndrome, akin to a majority vote but implemented quantum mechanically. Syndrome extraction circuits typically initialize an ancilla in |0\rangle, apply controlled operations (such as CNOTs after Hadamards for parity checks), and measure the ancilla in the Z basis to obtain the eigenvalue without disturbing the code space. The encoding circuit projects an input state \alpha |0\rangle + \beta |1\rangle onto the code space by appending two ancillary s in |0\rangle, followed by CNOT gates with the first as control to the second and third s, resulting in \alpha |000\rangle + \beta |111\rangle. This entangles the s into the +1 eigenspace of the stabilizers. However, the only corrects bit-flip errors and leaves errors undetectable, as a single error acts equivalently to a logical \bar{Z}. It generalizes to an [[n,1,n-1]] repetition for odd n, using n-1 generators Z_i \otimes Z_{i+1} \otimes I^{\otimes (n-2)} for i=1 to n-1, capable of correcting up to (n-1)/2 bit-flip errors via repeated measurements and majority decoding.

Steane code

The Steane code is a [[7,1,3]] stabilizer code that encodes one logical into seven physical qubits and can correct any single-qubit Pauli error. It belongs to the CSS subclass of stabilizer codes, constructed from the classical binary [7,4,3] , whose minimum of 3 enables the quantum code to achieve a distance of 3. This makes it capable of detecting up to two errors and correcting one, as the code space is the simultaneous +1 eigenspace of six independent stabilizer generators—three of X-type and three of Z-type. The code was introduced as an efficient quantum error-correcting scheme, demonstrating how classical perfect codes can yield quantum analogs with comparable efficiency. The stabilizers are derived from the 3×7 parity-check matrix H of the classical Hamming code, with the same matrix used for both X- and Z-type generators due to the code's self-orthogonality properties (ensuring the rows of H are pairwise orthogonal modulo 2, so the stabilizers commute). A standard form of H is H = \begin{pmatrix} 0 & 0 & 0 & 1 & 1 & 1 & 1 \\ 0 & 1 & 1 & 0 & 0 & 1 & 1 \\ 1 & 0 & 1 & 0 & 1 & 0 & 1 \end{pmatrix}, where the columns correspond to qubits 1 through 7. The Z-type stabilizers are then S_i^Z = \prod_{j: H_{i j}=1} Z_j for i=1,2,3, yielding explicit operators such as S_1^Z = Z_4 Z_5 Z_6 Z_7, S_2^Z = Z_2 Z_3 Z_6 Z_7, and S_3^Z = Z_1 Z_3 Z_5 Z_7. The X-type stabilizers S_i^X = \prod_{j: H_{i j}=1} X_j have identical supports, e.g., S_1^X = X_4 X_5 X_6 X_7. These six generators define the 2^{7-6}=2-dimensional code space, with logical states |\bar{0}\rangle = \frac{1}{\sqrt{8}} \sum_{\substack{c \in C \\ \mathrm{wt}(c) \ even}} |c\rangle and |\bar{1}\rangle = X^{\mathbf{1}} |\bar{0}\rangle = \frac{1}{\sqrt{8}} \sum_{\substack{c \in C \\ \mathrm{wt}(c) \ odd}} |c\rangle, where C is the classical Hamming codewords and \mathbf{1} is the all-ones vector (a valid logical operator). Logical operators, such as the bit-flip \bar{X}, can be constructed as the product of Pauli-X operators over the supports of the coset leaders of the classical code (the minimum-weight representatives for each syndrome, including the identity and single-bit errors), resulting in the weight-7 operator \bar{X} = X_1 X_2 X_3 X_4 X_5 X_6 X_7; a similar construction applies to the phase-flip \bar{Z}. Error correction proceeds separately for X and Z components due to the CSS structure: Z-stabilizer measurements yield a syndrome decoded via the classical Hamming decoder to identify and correct X (or Y) errors, while X-stabilizer measurements correct Z errors analogously. This leverages the classical decoder's ability to correct single errors from the 3-bit syndrome. The Steane code is a perfect code in the sense that its underlying classical Hamming code saturates the classical Hamming bound, achieving optimal single-error correction efficiency (with 7 physical qubits supporting 2^4 classical codewords for the encoding subspace). It also supports transversal implementation of the full Clifford group, allowing fault-tolerant logical Clifford gates via single-qubit operations on corresponding physical qubits (e.g., logical Hadamard as transversal Hadamard), a property stemming from the code's CSS structure and the automorphism group of the Hamming code. This fault-tolerance feature has made it a foundational example for scalable quantum computing architectures.

References

  1. [1]
    [quant-ph/9705052] Stabilizer Codes and Quantum Error Correction
    May 28, 1997 · I will give an overview of the field of quantum error correction and the formalism of stabilizer codes. In the context of stabilizer codes, I ...Missing: original | Show results with:original
  2. [2]
    Qubit stabilizer code - Error Correction Zoo
    An qubit stabilizer code is denoted as [ [ n , k ] ] or , where d is the code's distance. Logical subspace is the joint eigenspace of commuting Pauli operators.
  3. [3]
    Stabilizer codes | IBM Quantum Learning
    ### Summary of Stabilizer Codes from IBM Quantum Learning
  4. [4]
    Quantum stabilizer codes and classical linear codes | Phys. Rev. A
    We show that within any quantum stabilizer code there lurks a classical binary linear code with similar error-correcting capabilities.
  5. [5]
    Scheme for reducing decoherence in quantum computer memory
    Quantum Milestones, 1995: Correcting Quantum Computer Errors. Published 14 ... P. W. Shor, in Proceedings of the 35th Annual Symposium on Foundations of ...
  6. [6]
    Good quantum error-correcting codes exist | Phys. Rev. A
    Good quantum error-correcting codes exist. A. R. Calderbank and Peter W. Shor. AT&T Research, 600 Mountain Avenue, Murray Hill, New Jersey 07974. PDF Share. X ...
  7. [7]
    Error Correcting Codes in Quantum Theory | Phys. Rev. Lett.
    Calderbank and P. W. Shor, Phys. Rev. A (to be published). A. M. Steane, Proc. R. Soc. London A (to be published). P. W. Shor, Phys. Rev. A 52, R2493 (1995) ...
  8. [8]
    Theory of fault-tolerant quantum computation | Phys. Rev. A
    Jan 1, 1998 · I present a theory of fault-tolerant operations on stabilizer codes based on symmetries of the code stabilizer.
  9. [9]
    [quant-ph/9707021] Fault-tolerant quantum computation by anyons
    Jul 9, 1997 · Abstract: A two-dimensional quantum system with anyonic excitations can be considered as a quantum computer. Unitary transformations can be ...
  10. [10]
    Suppressing quantum errors by scaling a surface code logical qubit
    Feb 22, 2023 · Quantum error correction offers a path to algorithmically relevant error rates by encoding logical qubits within many physical qubits, for which ...
  11. [11]
    [PDF] Chapter 7 Quantum Error Correction
    NM. We will use the Pauli group to characterize a QECC in the following way: Let S denote an abelian subgroup of the n-qubit Pauli group Gn. Thus all.
  12. [12]
    [quant-ph/9604034] A Theory of Quantum Error-Correcting Codes
    Apr 26, 1996 · We develop a general theory of quantum error correction based on encoding states into larger Hilbert spaces subject to known interactions.
  13. [13]
  14. [14]
    Quantum repetition code | Error Correction Zoo
    Both phase- and bit-flip codes for 31 qubits and their stabilizer measurements have been realized by Quantinuum [35]. Multiple rounds of Steane error correction ...
  15. [15]
    [quant-ph/9605021] Simple Quantum Error Correcting Codes - arXiv
    May 15, 1996 · Access Paper: View a PDF of the paper titled Simple Quantum Error Correcting Codes, by Andrew Steane (Clarendon Laboratory and 1 other authors.