Fact-checked by Grok 2 weeks ago

Quantum logic gate

A quantum logic gate, also known as a quantum gate, is a fundamental operation in that acts on one or more s to manipulate their quantum states, represented mathematically as a that preserves the norm of the . These gates are the building blocks of quantum circuits, analogous to gates in digital electronics, but they exploit quantum phenomena like superposition and entanglement to enable computations that process multiple states simultaneously. Unlike classical gates, which operate on definite bit values (0 or 1), quantum gates apply reversible transformations to superpositions of qubit states, allowing for parallel processing of information across exponentially many possibilities. Quantum gates must be unitary operators, satisfying U^\dagger U = I, where U^\dagger is the and I is the ; this ensures reversibility and of , preventing irreversible as seen in some classical operations like the . Single-qubit gates, such as the Pauli-X gate (which flips |0\rangle to |1\rangle and vice versa, represented as \begin{pmatrix} 0 & 1 \\ 1 & 0 \end{pmatrix}), the Hadamard gate (creating equal superpositions, H = \frac{1}{\sqrt{2}} \begin{pmatrix} 1 & 1 \\ 1 & -1 \end{pmatrix}), and phase gates like Pauli-Z (\begin{pmatrix} 1 & 0 \\ 0 & -1 \end{pmatrix}), form the basis for more complex operations. Multi-qubit gates, including the controlled-NOT (CNOT) gate—which flips the target qubit only if the control qubit is in |1\rangle—introduce entanglement, linking qubits such that the state of one instantaneously influences another, a key resource for quantum algorithms. Sets of quantum gates, such as those comprising the Clifford group (including Hadamard, Pauli, and CNOT) augmented with non-Clifford gates like the T-gate, are universal, meaning they can approximate any unitary operation on an arbitrary number of qubits to desired precision, enabling the of general quantum computations. In practice, quantum gates are implemented physically using technologies like superconducting circuits, trapped ions, or photons, where operations correspond to controlled interactions such as microwave pulses or laser beams. Their development has been pivotal in advancing processing, supporting applications in , optimization, and of quantum systems that outperform classical computers.

Fundamentals

Definition and basic principles

A quantum logic gate is a that acts on one or more s, the basic units of . Qubits represent two-level quantum systems, such as the spin states of an or the polarization states of a , with their states described by vectors in a two-dimensional complex . The standard computational basis for a qubit consists of the orthonormal states denoted as |0\rangle and |1\rangle, allowing a general qubit state to be expressed as a linear superposition |\psi\rangle = \alpha |0\rangle + \beta |1\rangle, where \alpha and \beta are complex amplitudes satisfying the normalization condition |\alpha|^2 + |\beta|^2 = 1. This superposition enables qubits to encode more information than classical bits, which are restricted to definite 0 or 1 values. The action of a quantum logic gate U on a state |\psi\rangle produces a new state U |\psi\rangle = |\phi\rangle, where U is a satisfying U^\dagger U = I (with U^\dagger as the Hermitian conjugate and I the identity). Unitarity ensures that the operation preserves the norm of the , thereby maintaining the probabilities associated with measurement outcomes and the inner products between states. Consequently, quantum logic gates inherently preserve key quantum features such as superposition—where the output state remains a coherent combination of basis states—and entanglement, the between multiple qubits that cannot be described independently, as unitaries act linearly on the of multiple qubits. Unlike many gates, which can be irreversible (e.g., mapping multiple inputs to the same output and losing ), quantum logic gates are always reversible due to the invertibility of unitary operators (U^{-1} = U^\dagger). This reversibility is essential for quantum computation, as it allows the exact recovery of input states from outputs without loss. Quantum circuits, which form the practical framework for quantum algorithms, consist of sequences of such gates applied successively to one or more s, often visualized as diagrams with horizontal lines representing qubit wires and symbols denoting the gates. These circuits enable the composition of complex operations while respecting the unitary evolution of closed quantum systems.

Comparison to classical logic gates

Classical logic gates, such as , and NOT, function as operations on classical bits, which exist definitively in one of two states: or , and produce deterministic outputs without preserving all input information. For instance, an yields an output of only when both inputs are , but the original inputs cannot be uniquely recovered from this single-bit result, leading to inherent information loss. Quantum logic gates extend this framework by operating on qubits, which can occupy superpositions of states, thereby enabling computations across multiple possibilities in parallel. A quantum NOT gate applied to the superposition |+ \rangle = \frac{|0 \rangle + |1 \rangle}{\sqrt{2}} produces \frac{|1 \rangle + |0 \rangle}{\sqrt{2}}, maintaining the superposition while flipping its components, unlike classical operations that would require sequential evaluation of individual states. This parallelism arises from the linear nature of , allowing a single gate to process an exponential number of input combinations simultaneously. A fundamental distinction lies in reversibility: classical gates like AND and OR are typically irreversible, dissipating in accordance with law of , whereas quantum gates must be unitary operators, ensuring full reversibility and conservation of to preserve quantum . This unitarity prevents the information loss common in classical computing and aligns with the principles of quantum evolution, where every operation is invertible. In computational models, quantum gates underpin the model, which parallels classical gate arrays but surpasses the efficiency of classical Turing machines for specific tasks by leveraging quantum effects. While both models are universal—capable of simulating any computation within their domains—the quantum gate model facilitates exponential speedups in problems intractable for classical Turing machines. These capabilities enable landmark algorithms, such as Shor's, which uses quantum gates to create superpositions for parallel in , and Grover's, which exploits entanglement via gates to achieve in unstructured search problems.

Historical Development

Origins in

The conceptual foundations of quantum logic gates trace back to the early formalisms of in the 1930s, which provided the mathematical language for describing quantum states and operations. introduced the bra-ket notation in 1939, a compact system for representing quantum states as vectors in an abstract space and operations as transformations between them, emphasizing the abstract, basis-independent nature of quantum descriptions. This notation facilitated the manipulation of quantum amplitudes and laid groundwork for later unitary transformations central to gate operations. Concurrently, John von Neumann's 1932 work formalized within , portraying physical observables as operators and state evolutions as unitary mappings that preserve probabilities, thus establishing the essential for reversible quantum processes. In the mid-20th century, advancements in further developed ideas of controlled quantum manipulations akin to gate-like operations. During the and 1960s, researchers explored coherent states—eigenstates of the annihilation operator that minimize in and for the —enabling precise control over quantum light fields through unitary evolutions. Roy Glauber formalized these states in 1963, demonstrating their role in describing light as classical-like superpositions of number states, which influenced subsequent views of processing via optical elements. These developments shifted focus from purely analog wave descriptions to discrete, manipulable quantum entities, bridging continuous quantum dynamics with potential computational applications. The late 1970s and early 1980s marked a pivotal transition toward discrete models of quantum computation, moving from analog simulations of physical systems to structured, gate-based frameworks. Feynman's highlighted the inefficiency of classical computers in simulating and advocated for quantum devices using unitary operations to mimic , inspiring the idea of programmable quantum simulators. Building on this, Paul Benioff's 1980 model of a demonstrated how discrete quantum processes, governed by time-independent Hamiltonians, could perform universal computation reversibly, paving the way for the gate model by discretizing quantum evolutions into sequential steps. This era formalized the shift from continuous analog to a discrete paradigm, where unitary operators act as building blocks for information processing.

Key milestones and contributors

The development of quantum logic gates gained momentum in the 1980s with foundational theoretical work that emphasized reversibility and universality in quantum operations. In 1984, Charles Bennett and Gilles Brassard introduced quantum key distribution in their protocol, which relied on reversible quantum measurements and highlighted the need for unitary transformations akin to logic gates to ensure information security without decoherence. This work influenced the design of reversible quantum gates by demonstrating practical applications of quantum reversibility. The following year, David Deutsch proposed the universal quantum computer model, formalizing quantum computation through arrays of unitary gates that could simulate any quantum process, establishing the gate-based paradigm central to modern quantum computing. The 1990s marked a surge in algorithmic and experimental progress, driving the adoption of gate-based frameworks for quantum tasks. In 1991, developed entanglement-based quantum cryptography protocols using , which employed controlled quantum gates to generate and measure entangled states for secure . This advanced the role of multi-qubit gates in harnessing quantum correlations. Peter Shor's 1994 algorithm for and discrete logarithms further motivated gate decompositions, showing that a universal set of quantum gates could efficiently solve problems intractable for classical computers, spurring hardware development. Experimentally, the first realization of a controlled-NOT (CNOT) gate was achieved in 1995 by and colleagues using trapped ions, demonstrating two-qubit entanglement and control with fidelity sufficient for basic quantum operations. Into the 2010s, industry leaders like advanced practical gate implementations through scalable superconducting systems. 's 2016 launch of the Quantum Experience platform enabled cloud access to a 5- , followed by demonstrations of 16- and 20- chips by 2017, showcasing gate fidelities above 90% for single- and two- operations and paving the way for hybrid quantum-classical algorithms. These milestones highlighted the feasibility of gate arrays for near-term applications despite noise limitations. Recent advancements from 2019 to 2025 have focused on fault-tolerant gates integrated with error correction, transitioning toward scalable quantum computing. Google's 2019 Sycamore processor achieved quantum supremacy by executing random circuit sampling with 53 qubits and over 1 million two-qubit gates in 200 seconds—a task estimated to take classical supercomputers 10,000 years—using gate circuits to benchmark programmable quantum advantage. Building on this, progress in error-corrected gates includes Quantinuum's 2025 demonstration of a universal fault-tolerant gate set with logical qubits achieving below-threshold error rates via surface code protocols, enabling repeatable error suppression. Similarly, a 2024 experiment reported quantum error correction below the surface code threshold using a 105-qubit processor, sustaining logical gate operations with error rates under 0.1% through active syndrome measurement and decoding. These developments underscore the maturation of gate-based error correction for practical fault tolerance.

Mathematical Representation

Unitary matrix formalism

In the formalism of quantum logic gates, operations are described by unitary matrices acting on the state vectors of qubits. A unitary matrix U satisfies the condition U^\dagger U = I, where U^\dagger denotes the conjugate transpose of U and I is the identity matrix. This property ensures that quantum gates are reversible, meaning the original state can be recovered by applying the inverse operation U^{-1} = U^\dagger, and that the Euclidean norm of the state vector is preserved, upholding the unitarity of quantum probabilities. For a single qubit, whose state is a superposition in the two-dimensional computational basis \{ |0\rangle, |1\rangle \}, quantum gates are 2×2 unitary matrices. These matrices transform the qubit state vector |\psi\rangle = \alpha |0\rangle + \beta |1\rangle (with |\alpha|^2 + |\beta|^2 = 1) into another valid while maintaining of the basis. The general form of such a single-qubit gate admits an parameterization, for example using the ZYZ decomposition U = R_z(\psi) R_y(\theta) R_z(\phi), where R_z(\alpha) = \begin{pmatrix} e^{-i \alpha / 2} & 0 \\ 0 & e^{i \alpha / 2} \end{pmatrix} and R_y(\beta) = \begin{pmatrix} \cos(\beta / 2) & -\sin(\beta / 2) \\ \sin(\beta / 2) & \cos(\beta / 2) \end{pmatrix}, with \theta \in [0, \pi] and \psi, \phi \in [0, 2\pi) the real parameters that capture the three essential degrees of freedom of the special unitary group SU(2), up to a physically irrelevant global phase. This representation allows any single-qubit evolution to be expressed compactly. Geometrically, single-qubit gates correspond to rotations in the Bloch sphere representation, where a pure qubit state is depicted as a vector \vec{r} on the unit sphere in three-dimensional real space, with components related to the expectation values of the Pauli operators: r_x = \langle \sigma_x \rangle, r_y = \langle \sigma_y \rangle, r_z = \langle \sigma_z \rangle. Applying a unitary U rotates this vector by an angle determined by the gate parameters around an axis specified by the equivalent rotation vector, providing an intuitive visualization of how gates manipulate qubit coherence without changing the overall purity of the state. For systems of multiple qubits, the formalism extends to the tensor product space, where the total Hilbert space dimension is $2^n for n qubits, and gates are correspondingly $2^n \times 2^n unitary matrices. Independent operations on separate qubits are constructed using the of their individual unitary matrices, for example, U \otimes V for two qubits, which acts diagonally across the tensor factors while preserving the overall unitarity. This tensor structure facilitates the composition of larger circuits from basic gates, though entangled multi-qubit gates involve non-separable unitaries.

Connection to time evolution operators

In , the time evolution of a closed quantum system is governed by the time-dependent , which yields the unitary time evolution operator U(t) = e^{-i H t / \hbar}, where H is the of the system and \hbar is the reduced Planck's constant. This describes the continuous propagation of the quantum state |\psi(t)\rangle = U(t) |\psi(0)\rangle over time t, preserving the norm and ensuring unitarity for reversible dynamics. In the context of , this framework underpins the ideal behavior of quantum logic gates, which are discrete unitary transformations applied instantaneously to qubits. Quantum logic gates can be viewed as idealized, discrete approximations to short-time evolutions under specific Hamiltonians, often implemented via Trotterization or exact methods for infinitesimal time steps. For small t \to 0, the evolution operator simplifies to U(t) \approx I - i H t / \hbar, where I is the identity operator, allowing gates to mimic perturbative dynamics. Trotterization decomposes the exponential of a sum of non-commuting Hamiltonians into a product of exponentials, enabling the simulation of complex evolutions through sequences of elementary gates, with error bounds scaling favorably for higher-order approximations. This connection bridges continuous quantum dynamics to the discrete gate model essential for quantum circuits. Physically, quantum gates are realized by engineering time-dependent Hamiltonians in platforms, such as applying shaped optical pulses to trapped ions or pulses to superconducting circuits, which approximate the desired unitaries over finite durations. In superconducting qubits, for instance, resonant drives induce rotations around specific axes in the , effectively implementing single-qubit gates with gate times on the order of nanoseconds. Similarly, optical implementations use pulses to hyperfine transitions in qubits, achieving high-fidelity approximations to ideal unitaries. These methods rely on precise of strengths and durations to align the induced evolution with the target gate. However, real implementations deviate from ideal unitarity due to environmental interactions, primarily decoherence, which introduces non-unitary effects like amplitude damping and , causing loss of quantum coherence during gate execution. Decoherence arises from coupling to external baths, such as phonons in solids or stray electromagnetic fields, leading to mixed states and reduced gate performance, particularly for longer-duration operations. To quantify these deviations, the average gate serves as a key metric, defined as F = \int |\langle \psi | U^\dagger U_{\text{ideal}} | \psi \rangle|^2 \, d\psi, averaged over all pure input states |\psi\rangle on the (or higher-dimensional equivalents for multi-level systems). This , ranging from 0 to 1, encapsulates both coherent errors (e.g., over-rotation) and incoherent noise, with values above 99.9% typically required for fault-tolerant .

Single-Qubit Gates

Pauli gates (X, Y, Z)

The Pauli gates, consisting of the X, Y, and Z gates, are fundamental single-qubit operations in quantum computing that correspond to 180-degree rotations around the respective axes of the Bloch sphere. These gates are unitary and Hermitian, making them both reversible transformations and observable operators for quantum measurements. The Pauli X gate, also known as the bit-flip gate, is represented by the matrix \begin{pmatrix} 0 & 1 \\ 1 & 0 \end{pmatrix}. It swaps the computational basis states |0\rangle and |1\rangle, effectively flipping the qubit's bit value. The Pauli Y gate, which performs both a bit flip and a phase flip, has the matrix \begin{pmatrix} 0 & -i \\ i & 0 \end{pmatrix} and rotates the qubit state by \pi radians around the Y-axis of the Bloch sphere. The Pauli Z gate, or phase-flip gate, is given by \begin{pmatrix} 1 & 0 \\ 0 & -1 \end{pmatrix} and applies a relative phase of -1 to the |1\rangle state while leaving |0\rangle unchanged. As observables, the Pauli gates have eigenvalues \pm 1, with corresponding eigenstates forming the bases for measurements along the X, Y, and Z directions. For the X gate, the eigenstates are |+\rangle = \frac{1}{\sqrt{2}}(|0\rangle + |1\rangle) (eigenvalue +1) and |-\rangle = \frac{1}{\sqrt{2}}(|0\rangle - |1\rangle) (eigenvalue -1); for the Y gate, they are |+i\rangle = \frac{1}{\sqrt{2}}(|0\rangle + i|1\rangle) (+1) and |-i\rangle = \frac{1}{\sqrt{2}}(|0\rangle - i|1\rangle) (-1); and for the Z gate, |0\rangle (+1) and |1\rangle (-1). Measuring a in one of these bases projects it onto the corresponding eigenstate, yielding the eigenvalue as the outcome, which is crucial for tomography and error detection. In applications, the Pauli gates form the basis for stabilizer codes in , where the code space is defined as the +1 eigenspace of a commuting set of Pauli operators that stabilize the encoded state. Errors are detected by measuring these stabilizers, as deviations from +1 indicate Pauli-type faults correctable via decoding. Additionally, the Pauli gates generate the , which underlies the Clifford group—the normalizer of the under conjugation—enabling efficient simulation of Clifford circuits via the Gottesman-Knill theorem.

Hadamard gate

The Hadamard gate, denoted as H, is a fundamental single-qubit quantum gate that plays a crucial role in generating superpositions, enabling the exploration of quantum parallelism in algorithms. Its in the computational basis is given by H = \frac{1}{\sqrt{2}} \begin{pmatrix} 1 & 1 \\ 1 & -1 \end{pmatrix}, which is unitary and Hermitian. Applying H to the basis state |0\rangle yields the equal superposition |+\rangle = \frac{1}{\sqrt{2}} (|0\rangle + |1\rangle), while H |1\rangle = |-\rangle = \frac{1}{\sqrt{2}} (|0\rangle - |1\rangle). This transformation shifts the qubit from the Z-basis to the X-basis, creating balanced superpositions that allow quantum states to interfere constructively or destructively in subsequent operations. Unlike the Pauli X gate, which simply flips the basis states without superposition, the Hadamard gate introduces coherent overlap between them. The gate's effect on general superpositions can be understood through its linearity: for an input state \alpha |0\rangle + \beta |1\rangle, [H](/page/H+) produces \frac{\alpha + \beta}{\sqrt{2}} |0\rangle + \frac{\alpha - \beta}{\sqrt{2}} |1\rangle, preserving the norm while redistributing amplitudes. A key property is its self-inverse nature, satisfying [H](/page/H+)^2 = I, where I is the identity ; thus, applying [H](/page/H+) twice returns any state to its original form, making it reversible and efficient for basis changes. On the Bloch sphere, the Hadamard gate corresponds to a 180° rotation around the axis \frac{\hat{x} + \hat{z}}{\sqrt{2}}, which interchanges the positive and negative Z-poles with the X-equator points, visually capturing its role in basis transformation. In quantum algorithms, the Hadamard gate is essential for initializing superpositions in the (QFT), where it is applied to each to prepare the input state for phase encoding and interference, as utilized in Shor's factoring algorithm. Similarly, in Grover's search algorithm, Hadamard gates create the uniform superposition for the initial state, and they form part of the diffusion operator $2|s\rangle\langle s| - I (with |s\rangle the equal superposition), which amplifies the amplitude of the target state through inversion about the mean.

Phase shift and rotation gates

Phase shift gates introduce a relative between the computational basis states of a single without altering the probabilities of measurement outcomes. The general phase shift gate, often denoted as S(\theta) or P(\theta), is defined by the S(\theta) = \begin{pmatrix} 1 & 0 \\ 0 & e^{i\theta} \end{pmatrix}, which applies a e^{i\theta} to the |1\rangle state while leaving |0\rangle unchanged. This operation corresponds to a rotation around the z-axis of the by an angle \theta. Specific instances of the phase shift gate include the S gate, where \theta = \pi/2, resulting in the matrix \begin{pmatrix} 1 & 0 \\ 0 & i \end{pmatrix}, and the T gate, with \theta = \pi/4, given by \begin{pmatrix} 1 & 0 \\ 0 & e^{i\pi/4} \end{pmatrix}. The T gate is particularly significant as it is a non-Clifford gate, meaning it cannot be generated by Clifford operations alone, and its inclusion with Clifford gates enables universal quantum computation for single qubits. More generally, rotation gates allow arbitrary manipulations of a single qubit's state on the . The general rotation gate R_{\mathbf{n}}(\phi), where \mathbf{n} is a specifying the and \phi is the angle, is expressed as R_{\mathbf{n}}(\phi) = e^{-i \phi \mathbf{n} \cdot \boldsymbol{\sigma} / 2}, with \boldsymbol{\sigma} = (X, Y, Z) denoting the vector of . Phase shift gates are special cases of these rotations along the z-axis, where \mathbf{n} = (0, 0, 1). According to for rotations, any single-qubit unitary operation, which corresponds to a rotation in the three-dimensional SU(2) represented on the , can be decomposed into a sequence of three rotations around fixed axes, typically expressed in terms of (\alpha, \beta, \gamma) as R_z(\alpha) R_y(\beta) R_z(\gamma). This decomposition provides a constructive method for implementing arbitrary single-qubit gates using standard rotation primitives.

Identity gate

The identity gate, denoted as I, is a single-qubit quantum gate represented by the 2×2 I = \begin{pmatrix} 1 & 0 \\ 0 & 1 \end{pmatrix}. This gate applies no transformation to the state, satisfying I |\psi\rangle = |\psi\rangle for any |\psi\rangle, thereby preserving the input exactly. As a , the identity gate fulfills I^\dagger I = I, ensuring it conserves the and inner products of quantum states, with all eigenvalues equal to 1 and every state serving as an eigenvector. It constitutes the trivial element of the SU(2), corresponding to a zero-rotation transformation in the representation. In practical quantum circuits, the identity gate functions as a "do-nothing" operation to introduce delays, enabling synchronization of qubit timings across parallel branches or compensating for hardware latencies during gate execution. It also plays a key role in theoretical proofs of reversibility, exemplifying how unitary operations allow perfect recovery of prior states without information loss, as I^{-1} = I. A global phase factor applied to the identity gate, such as e^{i\phi} I for \phi \neq 0 \mod 2\pi, yields an equivalent operation indistinguishable from I in quantum measurements, since global phases do not affect observable probabilities or interference patterns.

Multi-Qubit Gates

Controlled-NOT and generalizations

The controlled-NOT (CNOT) gate is a fundamental two-qubit quantum gate in which the qubit determines whether a Pauli X operation (bit flip) is applied to the target qubit: if the is in the state |1⟩, the target is flipped; otherwise, the target remains unchanged. This operation can be expressed as |x y⟩ → |x, y ⊕ x⟩ for x, y ∈ {0, 1}, where ⊕ denotes the XOR function. In the standard computational basis {|00⟩, |01⟩, |10⟩, |11⟩}, the unitary matrix representation of the CNOT gate is \begin{pmatrix} 1 & 0 & 0 & 0 \\ 0 & 1 & 0 & 0 \\ 0 & 0 & 0 & 1 \\ 0 & 0 & 1 & 0 \end{pmatrix}. The CNOT gate is a specific instance of the more general controlled-U gate family, where U is any single-qubit . The action of a controlled-U gate on control and target qubits is given by |0⟩_c ⟨0| ⊗ I + |1⟩_c ⟨1| ⊗ U, leaving the target unchanged when the control is |0⟩ and applying U when the control is |1⟩. This generalization enables a wide range of conditional operations essential for quantum algorithms. Controlled gates like the CNOT play a key role in generating , a resource central to advantages. For example, starting from the |00⟩, applying a Hadamard gate to the first yields (|0⟩ + |1⟩)|0⟩ / √2, and subsequent application of the CNOT (with the first qubit as control) produces the entangled (|00⟩ + |11⟩) / √2.

Toffoli (CCNOT) gate

The , also known as the controlled-controlled-NOT (CCNOT) gate, is a three- quantum gate that applies an X (NOT) operation to a target qubit conditional on both control qubits being in the state |1⟩. In the computational basis, it permutes the basis states such that |000⟩ through |101⟩ remain unchanged, while |110⟩ maps to |111⟩ and |111⟩ maps to |110⟩, effectively implementing a reversible classical AND operation on the control bits to decide whether to flip the target. This behavior is captured by its 8×8 , which is the identity on the first six basis states and exchanges the last two via a 2×2 X block: \text{CCNOT} = \begin{pmatrix} I_6 & 0 \\ 0 & X \end{pmatrix}, where I_6 is the 6×6 and X = \begin{pmatrix} 0 & 1 \\ 1 & 0 \end{pmatrix}. As a , the is inherently reversible, preserving the information content of the . It is self-inverse, meaning applying the gate twice returns the original state, up to any required ancilla qubits in implementations, which aligns with its role in extending classical reversible logic to quantum settings. The can be decomposed using elementary gates, including six controlled-NOT (CNOT) gates along with single-qubit rotations such as Hadamard (H) and eighth-root-of-NOT (T) gates. In the Clifford + T gate set, relevant for fault-tolerant , any exact decomposition requires at least seven T gates, establishing a lower bound on the non-Clifford resource cost. The is essential for constructing reversible arithmetic operations in quantum algorithms, such as in-place circuits that enable efficient quantum simulations of classical computations without information loss. For example, multi-bit adders rely on networks of to compute carries reversibly, supporting algorithms like quantum transforms and error-corrected arithmetic.

Swap gate

The Swap gate is a two-qubit quantum logic gate that exchanges the quantum states of the two s on which it acts, mapping the joint state |ab⟩ to |ba⟩ where a and b are the states of the first and second , respectively. This operation is unitary and reversible, preserving the overall without requiring ancillary qubits. In the standard computational basis {|00⟩, |01⟩, |10⟩, |11⟩}, the matrix representation of the Swap gate is \begin{pmatrix} 1 & 0 & 0 & 0 \\ 0 & 0 & 1 & 0 \\ 0 & 1 & 0 & 0 \\ 0 & 0 & 0 & 1 \end{pmatrix}. The Swap gate can be decomposed into a sequence of three controlled-NOT (CNOT) gates, where the first CNOT has the first as and the second as , followed by a CNOT with the roles reversed, and then a third CNOT identical to the first. This construction leverages the CNOT's ability to add or subtract basis states modulo 2, effectively permuting the joint basis states. The decomposition is optimal, as at least three CNOT gates are required to implement the Swap operation. In quantum circuits, the Swap gate is crucial for rearranging positions to enable non-local operations on architectures with sparse qubit connectivity, such as superconducting processors where direct interactions are limited to neighboring s. It supports routing by moving logical s to desired physical locations, thereby optimizing circuit execution and reducing latency in multi- algorithms. The Swap gate preserves quantum entanglement, as it is a local unitary operation on the two qubits that merely interchanges their states without introducing or removing correlations. For instance, applying the Swap gate to a maximally entangled , such as \frac{1}{\sqrt{2}}(|00\rangle + |11\rangle), yields the identical state, maintaining the full entanglement. Similarly, for \frac{1}{\sqrt{2}}(|01\rangle + |10\rangle), the output is the same , while for \frac{1}{\sqrt{2}}(|01\rangle - |10\rangle), it results in a global but equivalent entanglement.

Universal Gate Sets

Two-qubit universal gates

In quantum computing, a set of gates is considered universal for n qubits if the group it generates forms a dense subgroup of the Lie group SU(2^n), enabling the approximation of any n-qubit unitary operation to arbitrary precision using finite sequences from the set. This density criterion ensures that the generated operations can simulate any quantum evolution required for universal quantum computation, up to global phases irrelevant for most algorithms. A foundational result demonstrates that combining all single-qubit gates, which form the group U(2), with a single entangling two-qubit gate like the controlled-NOT (CNOT) achieves universality for arbitrary n. More generally, Barenco et al. proved that almost any two-qubit gate capable of generating entanglement, paired with single-qubit rotations, suffices to generate a dense of SU(4) that extends to multi-qubit universality via tensor products and compositions. This highlights the minimal role of two-qubit interactions: they introduce the necessary non-local correlations absent in single-qubit operations alone. For practical implementations with discrete gate sets, the CNOT gate combined with single-qubit Hadamard (), phase (S), and π/8-rotation (T) gates forms a . Here, and S generate Clifford operations alongside CNOT, but the Clifford group alone is insufficient for universality, as circuits using only these gates can be efficiently simulated classically per the Gottesman-Knill theorem. The T gate, being non-Clifford, breaks this limitation by enabling approximations outside the stabilizer formalism, thus achieving density in SU(2^n). To compile arbitrary unitaries from such finite sets, the Solovay-Kitaev algorithm efficiently decomposes single-qubit gates into sequences from a dense-generating subset, requiring O(\log^{3.97}(1/\epsilon)) gates to achieve approximation error \epsilon. This method incurs only polylogarithmic overhead compared to exact implementations, facilitating the extension to two-qubit and multi-qubit universality through standard decompositions like the Euler-angle parameterization of SU(4).

Deutsch gate

The Deutsch gate is a three-qubit quantum logic gate introduced by in 1989 as part of his work on quantum computational networks. It generalizes the classical Toffoli (CCNOT) gate and, when combined with single-qubit gates, forms a for quantum computation, capable of approximating any unitary operation on an arbitrary number of qubits. The gate applies a conditional to the target qubit based on the states of two control qubits, introducing the entanglement necessary for universal quantum processing. The Deutsch gate, denoted D(\theta), acts on the computational basis states \{|000\rangle, |001\rangle, \dots, |111\rangle\} as the on all states except |111\rangle, where it applies a e^{i\theta} to the . Its is an 8×8 with 1's everywhere except the (8,8) entry, which is e^{i\theta}. Equivalently, it can be expressed as D(\theta) = |00\rangle\langle 00| \otimes I + |01\rangle\langle 01| \otimes I + |10\rangle\langle 10| \otimes I + |11\rangle\langle 11| \otimes R_z(\theta), where R_z(\theta) = \begin{pmatrix} 1 & 0 \\ 0 & e^{i\theta} \end{pmatrix} is a Z-axis . This structure enables the implementation of conditional shifts essential for multi-qubit interactions. For appropriate choices of \theta (e.g., \theta = \pi/4), sequences of Deutsch gates with single-qubit operations generate a dense of SU(8), extending to full universality. Deutsch showed that the Deutsch gate, augmented with single-qubit unitary gates, allows the construction of any multi-qubit unitary up to global phase through compositions, with efficient decompositions requiring a small number of gate applications. This establishes the gate's role as a primitive for quantum circuits. Historically, the Deutsch gate built on the framework of universal quantum computers, demonstrating that quantum networks can simulate any physical process, thereby affirming the quantum extension of the Church-Turing and inspiring modern models and algorithms.

Common universal sets (e.g., Clifford + T)

The Clifford group forms a finite of the on n qubits, generated by the Hadamard (H), phase (S), and controlled-NOT (CNOT) gates. These operations preserve the under conjugation, meaning they map products of to other Pauli products, which enables efficient classical simulation of Clifford circuits via the stabilizer formalism. However, the Clifford group alone is not universal, as it cannot approximate arbitrary unitaries beyond its discrete set, limiting its expressive power for general quantum algorithms. A widely adopted extension to achieve universality is the Clifford + T gate set, which augments the Clifford generators with the non-Clifford T gate—a π/4 around the Z-axis that introduces the necessary "magic" to break structure. This set allows approximation of any single-qubit unitary to precision ε using the Solovay-Kitaev algorithm, requiring O(log^{3.97}(1/ε)) T gates and a total of roughly 10^4 gates for typical high-precision targets (e.g., ε ≈ 10^{-12}). In multi-qubit settings, combining Clifford + T with CNOT enables fault-tolerant computation, particularly in architectures relying on to implement T gates with low error rates. The set {H, T, CNOT} serves as a minimal explicit variant, avoiding the full Clifford group while still supporting efficient . Hardware-specific universal sets prioritize native operations to reduce compilation depth and error accumulation in noisy intermediate-scale quantum (NISQ) devices. In trapped-ion systems, native gates include arbitrary single-qubit rotations and the Mølmer-Sørensen (MS) entangling gate, which generates a through combinations that approximate standard two-qubit interactions like CNOT. Similarly, Google's employs the fSim gate—a continuously tunable two-qubit interaction derived from cross-resonance coupling—alongside single-qubit Pauli rotations, forming a hardware-native that achieves lower circuit depths than Clifford + T decompositions. Practical trade-offs in selecting sets balance gate fidelity, overhead, and . Clifford + T excels in fault-tolerant due to the low overhead of Clifford operations and mature protocols for T-gate protection, but it incurs higher gate counts in NISQ regimes where non-native T implementations amplify noise. Native sets like fSim or gates minimize physical operations and leverage hardware strengths for faster execution, though they often require platform-specific calibration and may complicate cross-device portability. These considerations guide implementations in both NISQ experiments and long-term scalable architectures.

Quantum Circuit Composition

Serial composition of gates

Serial composition refers to the sequential application of quantum gates on the same set of qubits, where each subsequent gate acts on the output state of the previous one. In this arrangement, the overall transformation of the quantum state is given by the composition of unitary operators U = U_n \circ \cdots \circ U_1, such that the final state is |\psi_{\text{final}}\rangle = U_n \cdots U_1 |\psi_{\text{initial}}\rangle. This right-to-left ordering reflects the temporal sequence of gate applications, with the first gate U_1 applied closest to the initial state. Mathematically, the total unitary operator for serial composition is obtained through : U_{\text{total}} = U_n U_{n-1} \cdots U_1, where each U_i is a corresponding to the respective gate. Since the product of unitary matrices remains unitary, the composed preserves the quantum state's norm and reversibility. This underpins the of quantum circuits as sequences of such operations, enabling the implementation of complex algorithms through layered gate applications. The circuit depth, defined as the number of sequential layers in the composition (i.e., the longest path of dependent gates), is a critical metric in serial arrangements. Deeper circuits increase the total execution time, which can exceed coherence times—the duration over which remains intact before decohering due to environmental interactions. Minimizing depth is thus essential for fault-tolerant , as it reduces error accumulation in noisy intermediate-scale quantum (NISQ) devices. A representative example of serial composition is the of an arbitrary single-qubit unitary into a sequence of , such as U = R_x(\theta_1) R_z(\phi) R_x(\theta_2), where R_x and R_z are around the x- and z-axes, respectively. This Euler-like allows any single-qubit gate to be synthesized from three basic , facilitating practical on with native capabilities.

Parallel composition and tensor products

In quantum computing, parallel composition of logic gates on disjoint sets of qubits is achieved through the tensor product operation, which allows multiple gates to act simultaneously and independently on their respective subsystems. For two single-qubit gates U_a acting on qubit a and U_b on qubit b, the combined operation is represented as the tensor product U = U_a \otimes U_b, resulting in a two-qubit unitary that applies U_a to the state of qubit a and U_b to qubit b without interference between them. This parallel application preserves the structure of separable multi-qubit states. Consider an initial product state |\psi\rangle \otimes |\phi\rangle, where |\psi\rangle is the state of the first subsystem and |\phi\rangle of the second; under the parallel gate U_a \otimes U_b, it evolves to U_a |\psi\rangle \otimes U_b |\phi\rangle, maintaining the separable form. On entangled states, such as a , the tensor product still acts locally on each , applying the respective unitaries without introducing additional entanglement beyond what was already present, as the operation respects the independence of the subsystems. Representing general n-qubit unitaries via tensor products highlights their computational complexity, as the Hilbert space dimension scales as $2^n, leading to an O(2^n) size for the full matrix description, though implementations of sparse or local gates—common in parallel compositions—remain efficient with polynomial resource requirements in practice. A prominent example of parallel composition is the application of the Hadamard gate H across n qubits, forming H^{\otimes n}, which creates a uniform superposition over the computational basis and serves as the initial step in the quantum Fourier transform for algorithms like Shor's.

Gate exponents and inversion

In quantum circuits, the exponent of a gate refers to the repeated serial application of a U, denoted U^k, which composes the gate with itself k times. For integer k \geq 0, this is defined as U^k = U \circ U \circ \cdots \circ U (k times), with U^0 = I the identity operator. Since quantum gates are unitary, U^k remains unitary, preserving the of quantum states during . This repeated application is fundamental for implementing parameterized operations, such as phase accumulations or rotations in algorithms like phase estimation. A key example arises with rotation gates, which are exponentials of Pauli operators: the z-rotation R_z(\phi) = e^{-i \phi Z / 2}, where Z is the Pauli-Z matrix. The exponent property simplifies to R_z(\phi)^k = R_z(k \phi), scaling the rotation angle linearly with repetitions. This holds generally for rotations around a fixed axis \hat{n}: R_{\hat{n}}(\phi)^k = R_{\hat{n}}(k \phi). Fractional exponents, such as U^{1/m} for integer m, correspond to roots of the gate and are used in approximations via theorems like Solovay-Kitaev, enabling synthesis of arbitrary unitaries from a discrete gate set with polynomial overhead. For instance, the eighth-root of the identity, related to the T gate (R_z(\pi/4)), facilitates non-Clifford operations in fault-tolerant computing. Inversion of a quantum gate undoes its effect through the inverse operator U^{-1}, which for unitaries is the adjoint U^\dagger, satisfying U U^\dagger = I and U^\dagger U = I. The adjoint is the conjugate transpose of the matrix representation, ensuring reversibility essential for coherent quantum evolution. This property is crucial in circuit design for uncomputing auxiliary qubits (ancillas), where applying U^\dagger after U restores the ancilla to its initial state, preventing garbage accumulation in reversible computations like those using Toffoli gates. For rotation gates, inversion corresponds to negation: R_z(\phi)^{-1} = R_z(-\phi) = R_z(\phi)^\dagger. The Toffoli (CCNOT) gate, a controlled-controlled-NOT, is self-inverse up to its structure, meaning \text{CCNOT}^\dagger = \text{CCNOT} and \text{CCNOT}^2 = I, as it is a real acting as an on the computational basis. This self-inversion simplifies circuit optimization, allowing the same gate to both compute and uncompute the target bit without additional resources, which is vital in multi-qubit and fault-tolerant implementations.

Measurement in Quantum Circuits

Projective measurement basics

In quantum circuits, projective measurements represent a fundamental non-unitary operation that collapses the of a system into one of the eigenstates of the measured observable, fundamentally altering the evolution unlike reversible unitary gates. This process is governed by the projection postulate of , which dictates that upon , the state is projected onto the corresponding to the observed outcome, with the probability determined by the . For a single qubit, the standard projective measurement is performed in the computational Z-basis, defined by the Pauli-Z operator with eigenstates |0\rangle and |1\rangle. If the qubit is in a general state |\psi\rangle = \alpha |0\rangle + \beta |1\rangle (where |\alpha|^2 + |\beta|^2 = 1), the measurement projects the state to |0\rangle with probability |\alpha|^2 = |\langle 0 | \psi \rangle|^2 or to |1\rangle with probability |\beta|^2 = |\langle 1 | \psi \rangle|^2. The post-measurement state is then the normalized projection: for outcome |0\rangle, it becomes |0\rangle \langle 0 | \psi \rangle / \sqrt{|\langle 0 | \psi \rangle|^2} = |0\rangle, and similarly for |1\rangle. Projective measurements assume orthogonal outcomes, but quantum mechanics allows more general measurements described by positive operator-valued measures (POVMs), which extend projective measurements to non-orthogonal bases through a set of positive semi-definite operators \{E_m\} satisfying \sum_m E_m = I, where the probability of outcome m is \langle \psi | E_m | \psi \rangle. In quantum circuits, projective measurements (or their POVM generalizations) are typically applied at the circuit's end to extract classical information, but they can also occur mid-circuit to enable adaptive computations based on measurement outcomes, influencing subsequent gate applications.

Effects on single-qubit states

In , measuring a single in the computational basis collapses its superposition into one of the basis states. For instance, consider a prepared in the equal superposition state |+\rangle = \frac{1}{\sqrt{2}} (|0\rangle + |1\rangle). Upon projective in the \{|0\rangle, |1\rangle\} basis, the outcome is |0\rangle with probability \frac{1}{2} or |1\rangle with probability \frac{1}{2}, and the post-measurement state is the corresponding eigenstate, thereby destroying the between the basis states. This collapse can be described using the density operator formalism, where the pre-measurement \rho evolves to a classical after . Specifically, the post-measurement becomes \rho' = \sum_i p_i |i\rangle\langle i|, which is diagonal in the measurement basis and encodes only the probabilities p_i = \langle i | \rho | i \rangle without off-diagonal terms. This diagonal form represents a mixed , reflecting the irreversible loss of inherent in the process. The destructive nature of measurement underscores the no-cloning theorem, which prohibits perfect copying of an unknown quantum state. Extracting classical information via measurement provides probabilistic knowledge about the qubit but at the cost of rendering the original state unusable for further coherent operations or cloning attempts, as the collapsed state no longer retains the full superposition. In practical quantum devices, measurement errors can be modeled approximately by a depolarizing channel, which with probability p replaces the qubit state with the maximally mixed state \frac{I}{2}, mimicking readout inaccuracies that partially collapse or randomize the state. This model captures the average effect of imperfect projective measurements without resolving the full error mechanisms, aiding in error mitigation strategies for single-qubit operations.

Impacts on entangled multi-qubit states

In entangled multi-qubit systems, projective on a of qubits induces a that propagates non-locally, disrupting correlations across the entire . For instance, in a such as |\Phi^+\rangle = \frac{1}{\sqrt{2}} (|00\rangle + |11\rangle), measuring the first in the computational basis yields outcome with probability 1/2, collapsing the joint to |00\rangle, or outcome , collapsing it to |11\rangle; this instantly determines the second 's without direct , destroying the superposition while preserving perfect correlation. The operation formalizes the effect on the unmeasured subsystem after . When measuring a of qubits in a multi-qubit operator \rho, the reduced for the remaining qubits is obtained by tracing over the measured , yielding a mixed that captures the post-measurement ensemble; for example, tracing over one qubit in the maximally entangled results in \rho_A = \frac{1}{2} I, a maximally mixed indicating complete loss of local purity due to entanglement. This reduction reflects how erases quantum in the traced-out subsystem while imprinting classical correlations on the observer's record. Entanglement swapping demonstrates measurement's role in redistributing correlations without direct interaction. In the protocol proposed by Żukowski et al., two independent entangled pairs—say, qubits A and B in |\Phi^+\rangle_{AB}, and C and D in |\Phi^+\rangle_{CD}—undergo a Bell state measurement on B and C; the outcome projects A and D into an entangled state (e.g., |\Phi^+\rangle_{AD} for certain s), transferring entanglement across the unmeasured parties and enabling applications like quantum repeaters. This process highlights measurement's constructive potential amid destructive , as the joint creates new non-local links. In registers featuring pairwise entanglement, such as a of Bell pairs forming a larger entangled , measurement on one triggers propagation of the collapse, fracturing global coherence. For a linear where adjacent qubits are pairwise entangled (e.g., via controlled-NOT gates on a product ), measuring an interior collapses its partner immediately and, through correlated updates, decoheres the entire register into a classical-like , breaking multipartite entanglement while leaving separable remnants; this cascading underscores measurement's incompatibility with maintaining extended quantum superpositions in multi-qubit architectures.

Gate Synthesis and Applications

Synthesizing quantum logic functions

The primary goal of synthesizing functions is to decompose a target unitary operation into a of elementary from a universal gate set, enabling practical implementation on quantum hardware. This decomposition ensures that complex quantum operations, represented as unitary matrices, can be realized using a finite library of basic gates such as single-qubit rotations and controlled-NOT (CNOT) gates. For instance, an arbitrary two-qubit unitary can be exactly decomposed into at most 23 elementary gates, providing a constructive method for realization. Exact synthesis methods achieve precise decompositions without approximation errors for specific classes of operations. For Clifford circuits, which form a generated by Hadamard, phase, and CNOT gates, graph-state based approaches represent the target as a symmetric matrix and apply elementary Clifford operations to reduce it to an form, yielding circuits with a two-qubit gate depth of $7n - 2 for n-qubit operators on linear nearest-neighbor architectures. This method leverages properties to optimize gate count via decoding and depth via matching, outperforming prior techniques on random instances and benchmarks. For reversible functions, Toffoli networks provide exact synthesis by constructing cascades of Toffoli gates, with techniques such as iterative Reed-Muller spectra selection and reducing the average gate count to 6.38 for all 40,320 three-qubit functions, while resynthesis further minimizes larger circuits by replacing subnetworks with optimized equivalents. Approximate synthesis is essential for non-Clifford operations in universal gate sets, where exact decompositions may be inefficient or impossible. The Solovay-Kitaev algorithm approximates an arbitrary single-qubit unitary to accuracy \epsilon using a sequence of O(\log^{3.97}(1/\epsilon)) gates from a dense , via recursive decomposition of approximation errors into balanced group commutators, with a classical runtime of O(\log^{2.71}(1/\epsilon)) after compression. In the noisy intermediate-scale quantum (NISQ) , variational methods optimize parametrized circuits by minimizing the distance to the target unitary through gradient-based search, incorporating coherent multi-start strategies with controlled-phase gates to reduce CNOT counts—for example, achieving 8 CNOTs for a three-qubit Toffoli on nearest-neighbor topologies. These approaches address local minima in , making them suitable for hardware-constrained devices. Synthesis quality is evaluated using metrics that balance computational efficiency and error resilience, particularly for . Gate count measures total elementary operations, minimizing accumulation in NISQ devices and resource overhead in fault-tolerant regimes. depth quantifies sequential layers, directly impacting and decoherence limits, with optimizations targeting parallelism to enhance . The CNOT count, or more broadly two-qubit count, is critical due to their higher error rates (up to $10^{-2} versus $10^{-4} for single-qubit gates), while in Clifford+T libraries, T-count and T-depth non-Clifford usage, as each T incurs significant overhead (e.g., O(d^3) with constants 160–310), prioritizing for overall .

Universality proofs and approximations

A fundamental result in theory demonstrates that the Clifford group, generated by gates such as the Hadamard (H), (S), and controlled-NOT (CNOT), is insufficient for universal quantum computation due to its finite nature and efficient classical simulability via the Gottesman-Knill theorem. However, augmenting the Clifford group with a single non-Clifford gate, such as the T gate (a π/4 rotation), generates a dense in SU(2^n) for an n-qubit system. This density arises because the Clifford operations conjugate the non-Clifford gate to produce a set of unitaries whose phases form a dense of group, enabling approximation of any special unitary operation. This construction, rooted in the developed by Gottesman and the error-correction frameworks of Knill and Laflamme, establishes the theoretical basis for universality in fault-tolerant . The Solovay-Kitaev theorem provides a rigorous approximation bound for such universal gate sets, stating that any single-qubit unitary U ∈ SU(2) can be approximated to within error ε using O(log^{3.97}(1/ε)) gates from a finite set generating a dense subgroup of SU(2), with the constant improved to approximately 1 in modern variants. Extending to multi-qubit systems, the set {H, T, CNOT}—where H and CNOT are Clifford, and T is non-Clifford—allows any n-qubit unitary to be approximated to precision ε using a polynomial in n and log(1/ε) number of gates, ensuring efficient circuit synthesis for universal computation. This theorem not only proves the possibility of universality but also offers a constructive algorithm for gate decomposition, critical for practical implementations. DiVincenzo's criteria formalize the physical requirements for , including the need for a " of quantum gates" capable of generating, up to global phase, any unitary evolution on the computational . This criterion ensures that the underlying physical system supports the dense approximations guaranteed by gate set universality, bridging theoretical proofs to experimental realizations across platforms like superconducting qubits or trapped ions.48:9/11<771::AID-PROP771>3.0.CO;2-E) In practical applications, these universality proofs underpin algorithms like the Quantum Approximate Optimization Algorithm (QAOA), where universal gate sets approximate the unitary time evolutions e^{-iHt} for problem Hamiltonians H via Trotterization or direct synthesis, enabling optimization of combinatorial problems on near-term quantum devices. Such approximations leverage the polynomial-depth circuits from Solovay-Kitaev decompositions to simulate continuous-time with controlled error, highlighting the impact of theoretical universality on algorithmic design.

References

  1. [1]
    [PDF] Quantum Computation and Quantum Information - Michael Nielsen
    In this section we describe some simple quantum gates, and present several example circuits illustrating their application, including a circuit which teleports.
  2. [2]
    Quantum Logic Gates | NIST
    Mar 21, 2018 · Quantum logic gates use energy levels or motion of an ion to represent 0 or 1, and can process multiple possibilities simultaneously, unlike ...
  3. [3]
    [quant-ph/9703032] Programmable quantum gate arrays - arXiv
    Mar 18, 1997 · The universal quantum gate array we construct requires an exponentially smaller number of gates than a classical universal gate array.
  4. [4]
    A new notation for quantum mechanics | Mathematical Proceedings ...
    Oct 24, 2008 · A new notation for quantum mechanics. Published online by Cambridge University Press: 24 October 2008. P. A. M. Dirac.
  5. [5]
  6. [6]
    Coherent and Incoherent States of the Radiation Field | Phys. Rev.
    Methods are developed for discussing the photon statistics of arbitrary fields in fully quantum-mechanical terms.Missing: optics | Show results with:optics
  7. [7]
    Simulating physics with computers | International Journal of ...
    Download PDF · International Journal of ... Cite this article. Feynman, R.P. Simulating physics with computers. Int J Theor Phys 21, 467–488 (1982).
  8. [8]
    Quantum cryptography: Public key distribution and coin tossing - arXiv
    Mar 14, 2020 · Title:Quantum cryptography: Public key distribution and coin tossing. Authors:Charles H. Bennett, Gilles Brassard. View a PDF of the paper ...
  9. [9]
    Quantum theory, the Church–Turing principle and the universal ...
    It is shown that quantum theory and the 'universal quantum computer' are compatible with the principle.
  10. [10]
    Quantum cryptography based on Bell's theorem | Phys. Rev. Lett.
    Quantum cryptography based on Bell's theorem. Artur K. Ekert. Merton College and Physics Department, ... 67, 661 – Published 5 August, 1991. DOI: ...Missing: entanglement | Show results with:entanglement
  11. [11]
    Algorithms for quantum computation: discrete logarithms and factoring
    This paper gives Las Vegas algorithms for finding discrete logarithms and factoring integers on a quantum computer that take a number of steps which is ...
  12. [12]
    IBM's roadmap for scaling quantum technology
    Sep 15, 2020 · IBM has been exploring superconducting qubits since the mid-2000s, increasing coherence times and decreasing errors to enable multi-qubit ...Missing: demonstrations | Show results with:demonstrations
  13. [13]
    Quantinuum Crosses Key Quantum Error Correction Threshold ...
    Jun 27, 2025 · Quantinuum reports it has achieved the first universal, fully fault-tolerant quantum gate set with repeatable error correction.
  14. [14]
    Quantum error correction below the surface code threshold - Nature
    Dec 9, 2024 · Our error-corrected processors also demonstrate other key advances towards fault-tolerant quantum computing. We achieve repeatable performance ...
  15. [15]
    [PDF] Quantum Information and Computation Chapter 5 - John Preskill
    But our quantum logic gates will be unitary transformations, and hence will be invertible, while classical logic gates like the AND gate are not invertible.
  16. [16]
    A simple formula for the average gate fidelity of a quantum ... - arXiv
    May 7, 2002 · This note presents a simple formula for the average fidelity between a unitary quantum gate and a general quantum operation on a qudit.Missing: original | Show results with:original
  17. [17]
  18. [18]
    [quant-ph/9705052] Stabilizer Codes and Quantum Error Correction
    May 28, 1997 · A group-theoretical structure and associated subclass of quantum codes, the stabilizer codes, has proved particularly fruitful in producing codes.Missing: Pauli operators
  19. [19]
    [PDF] quantum-computation-and-quantum-information-nielsen-chuang.pdf
    This comprehensive textbook describes such remarkable effects as fast quantum algorithms, quantum teleportation, quantum cryptography, and quantum error- ...
  20. [20]
    [PDF] Quantum circuits of CNOT gates - arXiv
    Dec 16, 2020 · The controlled Pauli-X gate, also called the CNOT gate, is a very common and usefull gate in quantum circuits. This gate involves two qubits i ...
  21. [21]
    Training of quantum circuits on a hybrid quantum computer - Science
    Oct 18, 2019 · Here, we implement a data-driven quantum circuit training algorithm on the canonical Bars-and-Stripes dataset using a quantum-classical hybrid machine.Training Of Quantum Circuits... · Introduction · Materials And Methods
  22. [22]
    [PDF] On the CNOT-cost of TOFFOLI gates - Rinton Press
    Mar 22, 2008 · The magic decomposition is a two-qubit phenomenon,cbut the cosine-sine and demultiplexing decompositions hold for n-qubit operators are ...
  23. [23]
    [0803.2316] On the CNOT-cost of TOFFOLI gates - arXiv
    Mar 15, 2008 · In physical implementations, however, TOFFOLI gates are decomposed into six CNOT gates and several one-qubit gates. Though this decomposition ...Missing: CNOTs | Show results with:CNOTs
  24. [24]
    [1308.4134] An algorithm for the T-count - arXiv
    Aug 19, 2013 · We implemented our algorithm and used it to show that any Clifford+T circuit for the Toffoli or the Fredkin gate requires at least 7 T gates.
  25. [25]
    [quant-ph/0408173] Reversible addition circuit using one ancillary ...
    Aug 28, 2004 · In this paper I give a network of O(n^3) Toffoli gates for reversibly performing in-place addition with only a single ancillary bit.
  26. [26]
    SwapGate (latest version) | IBM Quantum Documentation
    The SWAP gate. This is a symmetric and Clifford gate. Can be applied to a QuantumCircuit with the swap() method.
  27. [27]
    1-3. Multiqubit representation - Quantum Native Dojo!
    The action of the three CNOT gates Λ(X)1 ... We see that this is the gate which swaps two qubits. (See Nielsen-Chuang 1.3.2 Multiple qbit gates for detail) ...
  28. [28]
    Optimal quantum circuits for general two-qubit gates | Phys. Rev. A
    Mar 22, 2004 · To compute the SWAP at least three CNOT gates are needed. Proof. We construct a proof by contradiction. Suppose that there is a circuit ...
  29. [29]
    What is a SWAP gate? - PennyLane
    The SWAP gate is a gate in quantum computing that swaps the states of two qubits. The diagram below shows how a SWAP gate is represented in quantum circuits.
  30. [30]
    Programmable Swap Gate for Quantum Computing Applications
    The SWAP gate facilitates the exchange of quantum states between two qubits and plays a significant role in qubit routing and circuit optimization. The ...
  31. [31]
    [PDF] Quantum circuits 1) The swap gate cannot create entanglement
    The swap gate cannot create entanglement; it merely interchanges the state of two qubits. Single-qubit operations also cannot create entanglement.
  32. [32]
    [quant-ph/9503016] Elementary gates for quantum computation - arXiv
    Mar 23, 1995 · We derive upper and lower bounds on the exact number of elementary gates required to build up a variety of two-and three-bit quantum gates.
  33. [33]
    [quant-ph/0505030] The Solovay-Kitaev algorithm - arXiv
    May 6, 2005 · This pedagogical review presents the proof of the Solovay-Kitaev theorem in the form of an efficient classical algorithm for compiling an arbitrary single- ...
  34. [34]
    [PDF] The Classification of Clifford Gates over Qubits - arXiv
    Most non-degenerate gate sets generate the Pauli group, which alone suffices to set the phase bits of the tableau arbitrarily by applying gates at the beginning ...
  35. [35]
    Efficient Simulation of Clifford Circuits | PennyLane Demos
    Apr 11, 2024 · More importantly, they can be efficiently simulated classically, according to the Gottesman-Knill theorem, which states that any n -qubit ...
  36. [36]
    6-qubit optimal Clifford circuits | npj Quantum Information - Nature
    Jul 5, 2022 · We ran a script to calculate the distribution of the number of Clifford group elements across optimal CNOT gate costs. Given the database, it ...Missing: HS | Show results with:HS
  37. [37]
    (Clifford + T) Gate Set | PennyLane Quantum Compilation
    The (Clifford + T) gate set contains S, H, CNOT, and T gates and is typically the target gate set in FTQC (fault-tolerant quantum computing).
  38. [38]
    [PDF] efficient clifford+t approximation of single-qubit operators
    By contrast, the. Solovay-Kitaev algorithm achieves T-count O(logc(1/ε)), where c is approximately 3.97. Keywords: circuit synthesis, Clifford+T, efficient ...
  39. [39]
    Universal quantum computation with ideal Clifford gates and noisy ...
    More specifically, the Gottesman-Knill theorem states that by operations from O ideal one can only obtain quantum states of a very special form called ...
  40. [40]
    [2202.09235] Qutrit metaplectic gates are a subset of Clifford+T - arXiv
    Feb 18, 2022 · A popular universal gate set for quantum computing with qubits is Clifford+T, as this can be readily implemented on many fault-tolerant ...
  41. [41]
    Native Gates - IonQ Quantum Cloud Documentation
    The native gateset is the set of quantum gates that are physically executed on IonQ hardware by addressing ions with resonant lasers via stimulated Raman ...When to use native gates · Introducing the native gates · MS gates · ZZ gates
  42. [42]
    Demonstrating a Continuous Set of Two-Qubit Gates for Near-Term ...
    Sep 15, 2020 · We demonstrate a continuous two-qubit gate set that can provide a threefold reduction in circuit depth as compared to a standard decomposition.
  43. [43]
    [PDF] to implement fault-tolerantly on the QEC codes such as - arXiv
    A special protocol called magic state distillation is employed to implement a non-Clifford T ... ×10 4 d=3 d=5 d=7 d=9. 2p/15. 0.00002 0.00004 0.00006 ...
  44. [44]
    Boundaries of quantum supremacy via random circuit sampling
    Apr 11, 2023 · Google's quantum supremacy experiment heralded a transition point where quantum computers can evaluate a computational task, random circuit ...Missing: paper | Show results with:paper
  45. [45]
    [PDF] Synthesis of Quantum-Logic Circuits
    Generic gates used in this paper are limited to the following. A generic unitary gate. An Rz gate without a specified angular parame- ter; conventions for Rx, ...Missing: seminal | Show results with:seminal
  46. [46]
    Quantum circuit optimization with AlphaTensor - Nature
    Mar 20, 2025 · Composition of quantum gates is achieved via matrix multiplication (serial) and Kronecker product (parallel). Signature tensor. Given an N ...
  47. [47]
    [2503.16208] Constant-Depth Quantum Circuits for Arbitrary ... - arXiv
    Mar 20, 2025 · The optimization of quantum circuit depth is crucial for practical quantum computing, as limited coherence times and error-prone operations ...
  48. [48]
    [PDF] Circuit Construction for General n-Qubit Gates Based on Block ZXZ ...
    Apr 3, 2024 · One-qubit gates do not require any CNOTs and can be decomposed into a sequence of three rotation gates [3]. Arbitrary two-qubit gates can be ...
  49. [49]
    [PDF] With a Few Square Roots, Quantum Computing is as Easy as - arXiv
    Oct 21, 2023 · sequentially, and by tensor product when gates are composed in parallel. For example, the controlled. 5. Page 6. gates used in the circuit ...
  50. [50]
    [PDF] Quantum Computing - IFIS | Institute of Information Systems
    ISBN 978-1-84628-887-6. - Nielsen, Michael A.; Chuang, Isaac (2010). Quantum Computation and. Quantum Information. Cambridge: Cambridge University ...
  51. [51]
    [PDF] Lecture 2: Quantum Algorithms 1 Tensor Products - People @EECS
    An extreme case of this phenomenon occurs when we consider an n qubit quantum system. ... The second Hadamard gate cancels out the first, since H2 = I. If ...
  52. [52]
    [PDF] arXiv:2310.11288v3 [cs.LO] 29 Jan 2024
    Jan 29, 2024 · monoidal categories is that the parallel composition of each choice does not correspond to the tensor product. In a way, we also subsume ZX ...
  53. [53]
    [PDF] Quantum Computation and Quantum Information
    is because unitary quantum logic gates are inherently reversible, whereas many classical logic gates such as the gate are inherently irreversible. Any ...
  54. [54]
    [PDF] GENERATORS AND ROOTS OF THE QUANTUM LOGIC GATES
    May 1, 2019 · To al ulate generators and roots of di erent logi gates we need to build up some general relations, using well known methods of matrix linear ...
  55. [55]
    18.435 - Quantum Computation - Lecture 6 - MIT Mathematics
    The second part of the lecture went over the basics of the quantum circuit model. ... projective measurement in a higher dimensional space. Suppose we have ...
  56. [56]
  57. [57]
    Measurement of Quantum Objects — QuTiP 4.6 Documentation
    Feb 8, 2022 · Performing a basic measurement (Projective)​​ The probabilities and respective output state are calculated for each projection operator. Now, ...
  58. [58]
    [PDF] Notes: Measurement and POVMs 1 Positive Operator-Valued ...
    It is important to note that measurements induced by POVMs, while generalizing projective measurements, don't introduce anything fundamentally new to quantum ...
  59. [59]
    [PDF] The Quantum Density Matrix and Its Many Uses - arXiv
    Aug 16, 2023 · A diagonal density matrix corresponds to a classical probability distribution, and deco- herence provides a means to understand how a quantum ...
  60. [60]
    A single quantum cannot be cloned - Nature
    Oct 28, 1982 · We show here that the linearity of quantum mechanics forbids such replication and that this conclusion holds for all quantum systems.
  61. [61]
    Simple Mitigation of Global Depolarizing Errors in Quantum ... - arXiv
    Jan 5, 2021 · By measuring the errors directly on the device, we use an error model ansatz to infer error-free results from noisy data.
  62. [62]
    [PDF] Chapter 4 Quantum Entanglement - John Preskill
    ... entangled pure state of two qubits violates some Bell inequality. It is not hard to generalize the argument to an arbitrary bipartite pure state. For ...
  63. [63]
    [PDF] Chapter 3: Entanglement, Density Matrices, and Decoherence
    May 18, 2016 · 6.2 Partial measurement and partial trace. Density matrices were introduced by the fact that measuring one part of a larger system leaves the ...
  64. [64]
    Event-ready-detectors'' Bell experiment via entanglement swapping
    Dec 27, 1993 · Our proposal involves two parametric down-converters. Subcoherence-time monitoring of the idlers provides a noninteractive quantum measurement entangling.
  65. [65]
    [PDF] A graph-state based synthesis framework for Clifford isometries
    Jan 14, 2025 · We tackle the problem of Clifford isometry compilation, i.e, how to syn- thesize a Clifford isometry into an executable quantum circuit.
  66. [66]
    [PDF] Techniques for the Synthesis of Reversible Toffoli Networks
    We present certain new techniques for the synthesis of reversible networks of Toffoli gates, as well as improvements to previous methods.
  67. [67]
  68. [68]
    None
    ### Summary of Variational Synthesis of Quantum Circuits in NISQ Context
  69. [69]