Fact-checked by Grok 2 weeks ago

Quantum circuit

A quantum circuit is a in that represents quantum algorithms as a sequence of unitary quantum gates applied to qubits, the basic units of analogous to classical bits but capable of existing in superpositions and entangled states. This model extends classical circuit architectures by leveraging principles of quantum mechanics, such as superposition and entanglement, to perform reversible transformations on quantum states within a , ultimately yielding probabilistic classical outputs upon . Quantum circuits are visualized as diagrams with horizontal wires representing qubits evolving from left to right through time, where each gate acts as a local —either on a single qubit (e.g., Hadamard or Pauli gates) or multiple qubits (e.g., controlled-NOT or Toffoli gates)—to manipulate the overall . A finite of such gates, such as single-qubit rotations combined with the controlled-NOT gate, suffices to approximate any desired , enabling the of powerful quantum algorithms like Shor's factoring or Grover's search. Circuits typically begin with qubits initialized in a state, such as |0⟩, and conclude with measurements that project the final state onto classical bit strings, with outcome probabilities determined by the squared magnitudes of the quantum amplitudes. The model's power stems from its ability to exploit quantum parallelism, where operations on superposed states effectively compute on multiple inputs simultaneously, though decoherence and noise in physical implementations necessitate fault-tolerant designs using error-correcting codes and ancillary qubits. Quantum circuits form the backbone of quantum programming frameworks like and Cirq, facilitating simulations and executions on noisy intermediate-scale quantum (NISQ) devices, and continue to drive advancements in demonstrations and scalable architectures.

Classical Foundations

Reversible Classical Logic Gates

Reversible logic gates are computing elements that implement bijective functions, ensuring a mapping between inputs and outputs, thereby preserving all without . This reversibility means that the gate's operation can be inverted to recover the original inputs from the outputs, contrasting with irreversible gates like OR that discard . In the context of classical bits, these gates permute the $2^n possible input states for n bits, represented mathematically as matrices—orthogonal matrices with exactly one 1 in each row and column. The concept of reversible computation, foundational to these gates, was formalized by Charles Bennett in 1973 to address energy dissipation in computing devices, showing that logical reversibility allows arbitrary computations with minimal thermodynamic cost by avoiding information loss. Bennett demonstrated that any irreversible computation can be simulated reversibly with modest overhead in space and time, laying the groundwork for low-power classical and later . A fundamental example is the NOT gate, which operates on a single bit by inverting its value: 0 becomes 1, and 1 becomes 0. Its is:
InputOutput
01
10
The NOT gate is depicted in diagrams as a triangle with a small circle at the output, symbolizing inversion, connected to a single input wire. Mathematically, it corresponds to the \begin{pmatrix} 0 & 1 \\ 1 & 0 \end{pmatrix}, which swaps the basis states |0\rangle and |1\rangle in the classical . Another key is the controlled-NOT (CNOT), a two-bit with one bit and one bit; it flips the target only if the control is 1, otherwise leaving it unchanged, effectively computing the XOR of control and target on the target bit. Its truth table is:
ControlTargetOutput ControlOutput Target
0000
0101
1011
1110
In circuit notation, the CNOT is shown with a solid dot or open circle on the control wire and a standard NOT symbol (triangle with circle) on the target wire, with lines connecting the two bits. Its permutation matrix in the computational basis is \begin{pmatrix} 1 & 0 & 0 & 0 \\ 0 & 1 & 0 & 0 \\ 0 & 0 & 0 & 1 \\ 0 & 0 & 1 & 0 \end{pmatrix}. The , also known as CCNOT, extends this to three bits with two control bits and one target; it inverts the target only if both controls are 1, enabling reversible implementations of nonlinear functions like AND. Introduced by Tommaso Toffoli in , it is for classical reversible computation when combined with simpler gates. Its , for inputs a, b, c (where a and b are controls, c the target), is unchanged unless both controls are 1 (i.e., flips the third bit for inputs 110 to 111 and 111 to 110):
abcOutput aOutput bOutput c
000000
001001
010010
011011
100100
101101
110111
111110
The features two symbols (dots) on the first two wires connected to a NOT on the third wire. Its $8 \times 8 has 1s on the diagonal except swapping the last two rows and columns to reflect the flip on input 111. The classical NOT finds a direct quantum analog in the Pauli-X gate.

Reversible Classical Circuits

Reversible classical circuits are constructed by interconnecting reversible logic gates in sequences or configurations to perform deterministic computations while preserving bijectively. Gates are wired such that outputs of one gate serve as inputs to subsequent gates, with wiring allowing simultaneous operations on of bits to minimize . To ensure reversibility for functions that would otherwise be irreversible, ancillary bits—initialized to a constant value like zero—are introduced as auxiliary lines; these must be restored to their initial state at the circuit's end to maintain overall bijectivity. This construction enables the simulation of arbitrary classical computations in a reversible manner, with the circuit's behavior fully determined by the input and independent of timing. A prominent example is the reversible full , which computes the and carry bits from three input bits (two operands and a carry-in) using Toffoli gates to avoid information loss. The circuit typically employs two Toffoli gates: one to generate the carry-out by controlled-NOT on the target bit if both operands are 1, and another to compute the by XOR operations facilitated by Feynman gates (classical CNOT equivalents), with an ancillary bit to temporarily store intermediate results before uncomputation. This design requires four bits total (three inputs plus one ancillary) and produces the on one output line while restoring the ancillary. Another application is reversible networks, where bit strings are sorted by swapping elements based on comparisons implemented via Toffoli gates. For instance, the Bit String Swapping (BSSSN) iteratively swaps adjacent bit strings differing by a single bit ( 1) using n×n Toffoli gates, optimizing by removing redundant pairs and matching templates to reduce gate count; an example for sorting {1,0,3,2,5,7,4,6} uses five such gates to permute the inputs into ascending order. Key properties of these circuits include their as permutations of the 2^n possible input states, ensuring every output has a predecessor and thus preserving the input cardinality—a classical analog to unitarity. is quantified by , the total number of , and depth, the length of the longest gate sequence affecting any output bit, which together determine and parallelism potential. For even permutations (those with even ), no ancillary bits are needed using controlled-NOT and Toffoli (CNT) gates, while odd permutations require at most one ancillary bit; optimal synthesis for n=3 wires yields circuits of 1 to 7 across all 40,320 permutations. Despite these advantages, limitations arise in scaling to general classical functions, where naive methods—such as direct embedding of irreversible like AND via Toffoli expansions—can demand exponential growth in ancillary bits to maintain reversibility without excessive garbage outputs. Fredkin and Toffoli's 1982 demonstration of a universal gate set using the conservative Fredkin (controlled-SWAP) and Toffoli highlights this , as arbitrary permutations require careful ancillary management to avoid prohibitive resource overhead. Advanced techniques mitigate this by achieving asymptotically optimal gate counts without ancillas, but the fundamental trade-off underscores the overhead in bridging classical irreversibility to reversible models.

Quantum Primitives

Qubits

A , or quantum bit, serves as the basic unit of in quantum circuits, extending the classical bit concept by leveraging quantum mechanical principles. Unlike a classical bit, which is strictly in one of two states (0 or 1), a is realized as a two-level quantum system capable of existing in a linear superposition of basis states. In Dirac notation, the state of a single is expressed as |\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 representation allows the qubit to encode more than a classical bit, as measurement yields 0 with probability |\alpha|^2 and 1 with probability |\beta|^2. The of a qubit can be geometrically visualized using the , a unit sphere in three-dimensional real space where the corresponds to |0\rangle, the to |1\rangle, and equatorial points represent equal superpositions. Pure qubit states on the sphere's surface, parameterized by \theta and \phi such that |\psi\rangle = \cos(\theta/2) |0\rangle + e^{i\phi} \sin(\theta/2) |1\rangle, providing an intuitive tool for understanding state evolution under quantum operations. The concept of the qubit emerged from Feynman's proposal to simulate quantum physical systems efficiently using a quantum mechanical computer, highlighting the need for units that natively handle quantum interference and superposition. Physically, qubits are implemented in diverse platforms to encode the two-level system using stable quantum . Superconducting qubits employ Josephson junctions in superconducting loops to create anharmonic oscillators, where the lowest two energy levels serve as the qubit states. Trapped-ion qubits use hyperfine or optical transitions in ionized atoms confined by electromagnetic fields, offering long times due to the isolation of atomic internal states. Photonic qubits information in properties such as or spatial modes, enabling room-temperature operation and compatibility with optical networks, though challenged by loss and detection efficiency. These realizations prioritize , , and , with ongoing advancements in each modality. Central to the qubit's utility are its intrinsic quantum properties: superposition, entanglement, and the implications of the . Superposition permits a qubit to represent a probabilistic combination of , enabling parallel computation across multiple configurations until collapses the . Entanglement arises when multiple qubits share a joint that cannot be factored into individual descriptions, exemplified by Bell states such as the maximally entangled \frac{1}{\sqrt{2}} (|00\rangle + |11\rangle), which exhibit perfect correlations defying classical intuition. The prohibits the creation of perfect copies of an arbitrary unknown , a consequence of in quantum that underpins and distinguishes from classical. These features collectively form the foundation for quantum circuit operations, where qubits are manipulated to exploit quantum advantages.

Quantum Logic Gates

Quantum logic gates are unitary operators that act on the of qubits to manipulate quantum states while preserving the norm of the . These gates are represented by unitary matrices U satisfying U^\dagger U = I, ensuring reversibility and probability conservation in quantum computations. Single-qubit gates operate on individual qubits and form the basis for more complex operations. The Pauli gates—, and —are fundamental Hermitian and unitary operators with explicit matrix representations in the computational basis: the X gate (bit ) is \begin{pmatrix} 0 & 1 \\ 1 & 0 \end{pmatrix}, the Y gate is \begin{pmatrix} 0 & -i \\ i & 0 \end{pmatrix}, and the Z gate ( ) is \begin{pmatrix} 1 & 0 \\ 0 & -1 \end{pmatrix}. The Hadamard gate H creates superposition and is given by H = \frac{1}{\sqrt{2}} \begin{pmatrix} 1 & 1 \\ 1 & -1 \end{pmatrix}. Phase gates include the S gate, which applies a \pi/2 phase shift (S = \begin{pmatrix} 1 & 0 \\ 0 & i \end{pmatrix}), and the T gate, which applies a \pi/4 phase shift (T = \begin{pmatrix} 1 & 0 \\ 0 & e^{i\pi/4} \end{pmatrix}). Multi-qubit gates entangle or correlate s across the system. The controlled-NOT (CNOT) gate flips the target if the control is in state |1\rangle, with \begin{pmatrix} 1 & 0 & 0 & 0 \\ 0 & 1 & 0 & 0 \\ 0 & 0 & 0 & 1 \\ 0 & 0 & 1 & 0 \end{pmatrix} in the two- basis. The controlled-Z () gate applies a \pi phase to the target if the control is |1\rangle, represented as \begin{pmatrix} 1 & 0 & 0 & 0 \\ 0 & 1 & 0 & 0 \\ 0 & 0 & 1 & 0 \\ 0 & 0 & 0 & -1 \end{pmatrix}. The SWAP gate exchanges the states of two s, with matrix \begin{pmatrix} 1 & 0 & 0 & 0 \\ 0 & 0 & 1 & 0 \\ 0 & 1 & 0 & 0 \\ 0 & 0 & 0 & 1 \end{pmatrix}. General controlled-U gates, where U is an arbitrary single- unitary, can be constructed using two CNOTs and modified U operations on the target and ancillary s. Certain finite sets of quantum gates are universal, meaning they can approximate any unitary operation to arbitrary precision. The Solovay-Kitaev theorem establishes that any finite set of single-qubit gates generating a dense subgroup of SU(2), combined with CNOT, can approximate any unitary on n qubits with a circuit of size O(\log^c (1/\epsilon)) for precision \epsilon and constant c \approx 3.97. The Clifford+T set, comprising Clifford gates (generated by H, S, and CNOT) and the T gate, is a widely used universal gate set for fault-tolerant quantum computing. Unlike classical logic gates, which may be irreversible (e.g., discards information), quantum gates are inherently reversible due to unitarity, though introduces irreversibility by collapsing superpositions. This reversibility aligns quantum gates with reversible classical gates like Toffoli but enables superposition and interference absent in classical permutations.

Circuit Model and Operations

Quantum Circuit Representation

Quantum circuits are typically represented using a diagrammatic notation that visually depicts the evolution of qubits through a sequence of . In this standard , horizontal lines, referred to as wires, represent the qubits, with time progressing from left to right along these lines. Single-qubit are shown as symbols or boxes placed directly on the corresponding wire, while multi-qubit , such as , are illustrated with vertical lines connecting the and qubits to indicate entanglement or conditional application. This notation, popularized in tools like IBM's , facilitates intuitive understanding of circuit structure and parallelism, where non-overlapping can be applied simultaneously in a single time step. Formally, a quantum circuit models the computation as an initial state |\psi\rangle transformed by a sequence of unitary operators applied in succession. For an n-qubit system, the state is represented in the tensor product space \mathcal{H} = (\mathbb{C}^2)^{\otimes n}, and the circuit applies unitaries U_k for k = 1 to m, yielding the final state U_m U_{m-1} \cdots U_1 |\psi\rangle. These unitaries are composed from elementary quantum gates, ensuring the overall transformation remains unitary to preserve . This model assumes reversible operations, with the circuit defining a specific ordering that captures the temporal dynamics of the . Key metrics quantify the and of a quantum . The width is defined as the number of s involved, determining the dimensionality of the state space. Circuit depth measures the number of sequential layers of , corresponding to the longest path through the diagram and influencing time requirements on hardware. The total gate count tallies all used, while parallelism is assessed by the extent to which in different layers can be executed concurrently without qubit overlap. These parameters guide optimization, as deeper circuits are more susceptible to error accumulation. Variants of the representation account for practical implementations. Circuits often begin with initialization of qubits to the |0\rangle , depicted as a symbol at the wire's start. Measurement gates, which perform projective measurements onto the computational basis, are typically placed at the circuit's end, represented by meter-like symbols on the wires; these introduce non-unitarity by collapsing the . Closed circuits refer to purely unitary evolutions without measurements, suitable for theoretical analysis, whereas open circuits incorporate measurements or noise channels, reflecting real-world quantum processes. The conventions for quantum circuit representation were standardized in foundational texts, notably Nielsen and Chuang's 2000 textbook, which established the diagrammatic and formal frameworks widely adopted in the field. This standardization ensures consistency across theoretical developments, software libraries, and experimental platforms, enabling reproducible design.

Quantum Computations and Algorithms

Quantum circuits execute computations through a sequence of unitary operations applied to an initial state, typically prepared in the computational basis, followed by to extract classical outcomes. The overall evolution is governed by a U composed from the individual gate unitaries, transforming the initial state |\psi_0\rangle to a final state |\psi\rangle = U |\psi_0\rangle. in the computational basis then yields an outcome |k\rangle with probability p_k = |\langle k | \psi \rangle|^2, where the squared modulus reflects the for probabilistic collapse of the . This process leverages and entanglement during the unitary phase to explore multiple computational paths simultaneously, with interference effects determining the measurement probabilities. A cornerstone of quantum computations is the (QFT), a unitary transformation that maps a computational basis state to its Fourier coefficients, enabling efficient frequency analysis in quantum algorithms. The QFT circuit for n s consists of Hadamard gates on each qubit combined with controlled-phase rotations R_k = \begin{pmatrix} 1 & 0 \\ 0 & e^{2\pi i / 2^k} \end{pmatrix} applied in a ladder structure, requiring O(n^2) elementary gates. In contrast, the classical (FFT) computes the DFT on $2^n inputs in O(2^n n) time, which is exponential in n, whereas the QFT achieves this in O(n^2) time using n qubits. This efficiency highlights the QFT's role in providing polynomial-time implementations for problems intractable classically. The QFT forms the basis for phase estimation subroutines in many algorithms. Prominent quantum algorithms exploit these primitives to achieve speedups over classical counterparts. , introduced in 1994, factors an integer N in polynomial time by reducing the problem to period-finding via and applying the QFT to extract the period from superposition states, enabling efficient solutions to the and factoring problems believed to be hard classically. Similarly, from 1996 provides a quadratic speedup for unstructured search problems, using an oracle circuit to mark target states and iterative —via diffusion operators—to boost the probability of measuring the solution, requiring O(\sqrt{N}) queries for a database of size N. These algorithms demonstrate how quantum circuits harness to amplify desired amplitudes while suppressing others. Superposition in quantum circuits enables massive parallelism, evaluating exponentially many inputs in a single run, with constructive and destructive providing the speedup mechanism. For instance, the Deutsch-Jozsa algorithm of 1992 determines whether a is constant or balanced with a single query using Hadamard gates to create superposition and a final to detect global , offering an exponential advantage over classical deterministic methods that require up to $2^{n-1} + 1 queries for n bits. Such examples illustrate the potential for quantum circuits to solve decision problems intractable for classical computers. From a complexity-theoretic , quantum computations define the class (bounded-error quantum time), consisting of decision problems solvable by a quantum circuit of size with success probability at least $2/3. Introduced in 1993, contains and is contained in , with problems like integer factoring in but their exact placement relative to remaining open, underscoring the hypothesized power of quantum parallelism beyond classical nondeterminism.

Implementations and Extensions

Quantum Hardware Realizations

Quantum hardware realizations involve implementing quantum circuits on physical platforms that encode qubits and execute gates while contending with and architectural constraints. These platforms leverage diverse physical systems to create controllable quantum states, with ongoing efforts focused on improving , , and . Major approaches include superconducting circuits, trapped ions, neutral atoms, and photonic systems, each offering unique advantages in qubit stability and gate operations but facing distinct engineering challenges such as cryogenic requirements or optical complexity. Superconducting qubits, particularly designs, form the basis of processors from companies like and , operating at millikelvin temperatures to maintain . 's 53-qubit , demonstrated in 2019, achieved a milestone by performing a random sampling task in 200 seconds that would take a 10,000 years, highlighting the platform's potential for complex computations despite . 's systems, such as the 127-qubit processor released in 2021 and subsequent advancements, emphasize modular architectures with tunable couplers for high-fidelity gates, achieving two-qubit gate fidelities exceeding 99.5% in recent iterations. In November 2025, unveiled the 120-qubit processor with a square topology and 5,000 two-qubit gates, enabling 30% more complex calculations than previous generations. These platforms excel in fast gate speeds (on the order of nanoseconds) but suffer from short times due to material imperfections and . Trapped-ion systems use electromagnetic fields to confine charged atoms, enabling all-to-all qubit connectivity without fixed nearest-neighbor limitations, which simplifies circuit implementation. IonQ's Forte system, with 36 qubits, demonstrated algorithmic qubits equivalent to 35 in 2023, while their 2025 Tempo platform reached 100 physical qubits and an #AQ 64 benchmark for application performance. Quantinuum's H2 system achieved 56 qubits with two-qubit gate fidelities above 99.9% in 2024, leveraging ion shuttling for scalable chains. In November 2025, Quantinuum introduced the Helios system with 98 fully connected physical qubits, achieving single-qubit gate fidelity of 99.9975% and two-qubit gate fidelity of 99.921%. These platforms benefit from long coherence times (seconds) and identical qubit properties but require vacuum chambers and laser control, limiting gate speeds to microseconds. Neutral-atom arrays, as pursued by QuEra, trap laser-cooled atoms in , allowing dynamic reconfiguration of positions for flexible connectivity. QuEra's system, with 256 s since 2022, supports parallel operations across reconfigurable lattices, achieving single- fidelities over 99.9% and demonstrating error-corrected logical s in small-scale codes by 2025. Their roadmap targets 10 logical s in 2024 and 100 by 2026, emphasizing room-temperature operation and scalability via atom loading from reservoirs. This approach mitigates decoherence through atomic ground-state encoding but faces challenges in precise trap addressing and readout fidelity. Photonic quantum hardware, exemplified by Xanadu's systems, encodes qubits in or path, enabling via fiber optics and room-temperature operation. Xanadu's , a 12-qubit universal photonic processor unveiled in 2025, integrates squeezed-light sources for fault-tolerant gates, while their on-chip generation of Gottesman-Kitaev-Preskill (GKP) states marked a milestone for error-resistant qubits. These systems achieve high-speed operations (gigahertz) and natural through but struggle with loss and probabilistic gates, requiring heralding techniques for reliability. Implementing quantum circuits on these hardware platforms requires mapping logical qubits to physical ones, accounting for device topologies like linear chains in ions or 2D grids in superconductors. Nearest-neighbor constraints in fixed architectures, such as IBM's heavy-hex , necessitate swap insertions to route non-adjacent qubits, increasing circuit depth and accumulation; optimization algorithms reduce these overheads by up to 50% through initial layout permutations. In reconfigurable systems like neutral atoms, dynamic repositioning minimizes swaps, enhancing efficiency for deep circuits. Key error sources in quantum hardware include decoherence, quantified by relaxation time T1 (energy loss to environment) and dephasing time (phase randomization), typically ranging from 10-100 microseconds in superconductors and milliseconds in ions as of 2024. Gate fidelities, measuring operation accuracy, have surpassed 99% for single qubits and 99.5% for two-qubit gates across platforms, with reporting 99.99% two-qubit fidelity in 2025 via improved laser calibration. Control errors from pulse imperfections and further degrade performance, necessitating dynamical decoupling pulses to extend . Current quantum hardware operates in the Noisy Intermediate-Scale Quantum (NISQ) era, with systems scaling to around 100-1000 qubits but limited by error rates of 0.1-1% per gate, restricting circuit depths to hundreds of operations. Scalability paths emphasize fault-tolerant architectures, particularly surface codes, which encode logical qubits in a lattice of physical qubits to suppress errors below thresholds (around 1%) using syndrome measurements and corrections. Demonstrations on Google's Willow processor in 2024 achieved surface code performance below the error threshold with 105 qubits, paving the way for million-qubit machines by encoding one logical qubit per thousand physical ones.

Simulations and Accelerations

Simulating quantum circuits on classical computers faces significant challenges due to the exponential growth of the state space. For a system of n qubits, the full quantum state is represented by a vector in a $2^n-dimensional complex Hilbert space, requiring storage and computation that scale exponentially with n, making exact simulations infeasible for large n beyond a few dozen qubits on standard hardware. To address this, tensor network methods such as matrix product states (MPS) and projected entangled pair states (PEPS) exploit low-entanglement structures in certain circuits, approximating the state evolution with polynomial resources for circuits where entanglement remains bounded, as demonstrated in simulations of quantum chemistry problems. Classical simulation algorithms vary in approach to mitigate these challenges. State-vector simulation maintains the full wavefunction and applies unitary via matrix-vector multiplications, achieving for small-scale circuits but limited to around 30-40 qubits on high-end classical machines. methods extend this to handle mixed states and noise, useful for fault-tolerant simulations, though they incur a overhead in memory due to the (2^n)^2 elements. For specific gate sets, simulations efficiently track Pauli operators for Clifford circuits—those composed of Hadamard, , CNOT, and similar —using linear over the group, enabling simulations of up to 50+ qubits in time, as shown in benchmarks for random Clifford circuits. Hardware accelerations enhance simulation scalability through parallelism and customization. Graphics processing units (GPUs) parallelize the dense operations in state-vector simulations, achieving speedups of orders of magnitude; for instance, 's cuQuantum library simulates 40-qubit circuits in seconds on A100 GPUs, compared to hours on CPUs. Recent benchmarks include simulating the 53-qubit Sycamore processor in under five minutes on the and a full quantum chip simulation using 6,724 GPUs on Perlmutter in November 2025. Field-programmable gate arrays (FPGAs) offer reconfigurable logic for gate , with studies on UltraScale+ FPGAs demonstrating simulations of over 40 qubits for circuits by optimizing tensor contractions. Hybrid approaches integrate quantum simulation into frameworks for variational and noisy intermediate-scale quantum (NISQ) applications. TensorFlow Quantum combines Cirq circuits with 's automatic differentiation, enabling gradient-based optimization of parameterized quantum circuits on classical hardware, as applied to quantum neural networks. Similarly, Cirq's simulation backend supports backends for efficient handling of low-depth circuits in hybrid quantum-classical algorithms. The practical limits of classical simulations highlight the need for approximations in large-scale studies. As of November 2025, state-vector methods on supercomputers like reach thresholds of about 50 qubits for exact before constraints dominate, with the first full of a 50-qubit universal quantum computer achieved on , beyond which heuristic or approximate methods such as or path-integral Monte Carlo become essential for exploring circuit behaviors.

References

  1. [1]
    [PDF] quantum-computation-and-quantum-information-nielsen-chuang.pdf
    ... quantum circuit model of computation. 202. 4.7 Simulation of quantum systems. 204. 4.7.1 Simulation in action. 204. 4.7.2 The quantum simulation algorithm. 206.
  2. [2]
    [PDF] Lecture 1: Introduction to the Quantum Circuit Model
    Sep 9, 2015 · The goal for this first lecture is to give a lightning-fast, as-barebones-as-possible definition of the quantum circuit model of computation.
  3. [3]
    [PDF] Lecture 5 1 Quantum Circuits - People @EECS
    Jan 31, 2007 · A quantum circuit implements a unitary operator in a Hilbert space C2n. , given as primitive a (usually finite).
  4. [4]
    [PDF] A Review on Reversible Logic Gates and their Implementation
    Reversible logic gates have equal inputs and outputs, with a one-to-one correspondence, and allow for unique input recovery from outputs.
  5. [5]
    Synthesis and optimization of reversible circuits—a survey
    In this survey, we review algorithmic paradigms—search based, cycle based, transformation based, and BDD based—as well as specific algorithms for reversible ...
  6. [6]
    [PDF] Logical Reversibility of Computation* - UCSD Math
    Abstract: The usual general-purpose computing automaton (e.g.. a Turing machine) is logically irreversible- its transition function.
  7. [7]
    [PDF] reversible computing - CSAIL Publications - MIT
    REVERSIBLE COMPUTING. Tommaso Toffoli. February 1980. Page 2. REVERSIBLE COMPUTING*. Tommaso Toffoli. MIT Laboratory for Computer Science. 545 Technology Sq ...
  8. [8]
    [PDF] Reversible Logic Circuit Synthesis - University of Michigan
    The 2-CNOT is called a TOFFOLI gate. Together, the NOT, CNOT, and TOFFOLI form a universal set of gates for classical reversible circuits [16] (we will be ...
  9. [9]
    [PDF] Sorting Network for Reversible Logic Synthesis - arXiv
    In this paper, we have introduced an algorithm to implement a sorting network for reversible logic synthesis based on swapping bit strings.
  10. [10]
    Asymptotically optimal synthesis of reversible circuits - ScienceDirect
    However, this trivial idea has the defect that exponential ancillary bits are required. In this paper, we will propose a reversible logic synthesis approach ...
  11. [11]
    Conservative logic | International Journal of Theoretical Physics
    Fredkin, E., Toffoli, T. Conservative logic. Int J Theor Phys 21, 219–253 (1982). https://doi.org/10.1007/BF01857727. Download citation. Received: 06 May 1981.
  12. [12]
    [PDF] Simulating Physics with Computers - s2.SMU
    International Journal of Theoretical Physics, VoL 21, Nos. 6/7, 1982. Simulating Physics with Computers. Richard P. Feynman. Department of Physics, California ...
  13. [13]
    An elementary review on basic principles and developments of ...
    Mar 18, 2024 · Recently, the experimental realization of logical qubits has been demonstrated in leading quantum platforms, such as superconducting qubits [59] ...Quantum States And Their... · Defining Qubit States · Quantum Error Correction For...
  14. [14]
    [PDF] Chapter 4 Quantum Entanglement - John Preskill
    Entangled states are interesting because they exhibit correlations that have no classical analog. We will begin the study of these correlations in this chapter.
  15. [15]
    [PDF] arXiv:2504.00995v2 [quant-ph] 8 May 2025
    May 8, 2025 · We introduce fundamental concepts such as basis states, quantum gates, and tensor products, illustrat- ing how these form the building blocks of ...
  16. [16]
  17. [17]
    Quantum Gates | SpringerLink
    Mar 23, 2021 · The Hadamard gate has the following matrix representation: (6.3). Using matrix multiplication we can show that application of the Hadamard gate ...
  18. [18]
    [PDF] The Geometry of Quantum Computing - arXiv
    Dec 22, 2023 · Figure 5: X, Y , Z are the Pauli matrices gates, H the Hadamard gate. The. S gate is eiπ/2Z, the T gate is eiπ/4Z, † as usual indicates the ...Missing: XYZ | Show results with:XYZ
  19. [19]
    [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 ...
  20. [20]
    [PDF] arXiv:2305.18128v3 [quant-ph] 27 Feb 2024
    Feb 27, 2024 · The controlled-SWAP and controlled-controlled-NOT gates are at the heart of the original proposal of reversible classical computation by Fredkin ...
  21. [21]
    Optimal quantum circuit synthesis from Controlled-U gates - arXiv
    Aug 29, 2003 · From a geometric approach, we derive the minimum number of applications needed for an arbitrary Controlled-Unitary gate to construct a universal quantum ...
  22. [22]
    [PDF] arXiv:quant-ph/0505030v2 23 Aug 2005
    Aug 23, 2005 · The Solovay-Kitaev (SK) theorem is one of the most important fundamental results in the theory of quantum computation. In its simplest form the ...
  23. [23]
    [PDF] Synthesis of unitaries with Clifford+T circuits - arXiv
    We describe a new method for approximating an arbitrary n qubit unitary with precision ε using a Clifford and T circuit with O(4nn(log(1/ε) + n)) gates.
  24. [24]
    Quantum computation is the unique reversible circuit model ... - Nature
    Jan 22, 2019 · ... reversible transformations consists entirely of single-bit gates, and not even classical computation is possible. In this sense, qubit quantum ...
  25. [25]
    Quantum Circuit Diagram Conventions - Microsoft Learn
    Jan 16, 2025 · This article explains how to read quantum circuit diagrams and their conventions. For more information, see How to visualize quantum circuits diagrams.Reading quantum circuit... · Multi-qubit operations
  26. [26]
    Circuits | IBM Quantum Learning
    Quantum circuits. In the quantum circuit model, wires represent qubits and gates represent operations on these qubits. We'll focus for now on operations we ...
  27. [27]
    [PDF] 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.
  28. [28]
    [PDF] A fast quantum mechanical algorithm for database search - arXiv
    The quantum algorithm uses superposition to examine multiple items simultaneously, finding a desired item in an unsorted database in only steps, faster than ...
  29. [29]
    [PDF] Rapid Solution of Problems by Quantum Computation - cs.Princeton
    A (1992). Rapid solution of problems by quantum computation. BY DAVID DEUTSCH AND RICHARD JOZSA. Wolfson College, Oxford OX2 6UD, U.K.. St Edmund Hall, Oxford ...
  30. [30]
    how the main modalities stack up for HPC leaders - QuEra Computing
    Oct 9, 2025 · Today's serious candidates fall into five buckets: superconducting circuits, trapped ions, neutral atoms, photonic systems, and semiconductor ...Missing: transmon Sycamore 2019 2023
  31. [31]
    Quantum supremacy using a programmable superconducting ...
    Oct 23, 2019 · ... Quantum supremacy is demonstrated using a programmable superconducting processor known as Sycamore, taking approximately 200 seconds to ...
  32. [32]
    IBM quantum computers: evolution, performance, and future directions
    Apr 1, 2025 · This paper explores IBM Quantum's journey in quantum computing, highlighting key technological achievements, current challenges, and future prospects.Missing: realizations | Show results with:realizations
  33. [33]
    IonQ's Accelerated Roadmap: Turning Quantum Ambition into Reality
    Jun 13, 2025 · IonQ's trapped ion technology underpins our architectural advantage. Unlike alternative approaches, ions are naturally identical and stable, ...
  34. [34]
    9 Types of Qubits Driving Quantum Computing Forward [2025] - SpinQ
    Mar 28, 2025 · In this article, we will explore 9 different types of qubits, each with its unique set of features, advantages, and challenges in quantum computing.
  35. [35]
    Roadmap for Advanced Error-Corrected Quantum Computers
    QuEra's roadmap includes 10 logical qubits in 2024, 30 in 2025, and 100 in 2026, with a three-phase release.
  36. [36]
  37. [37]
    On the optimality of quantum circuit initial mapping using ...
    Mar 13, 2024 · Initial logical-to-physical mapping permutes logical qubits while fixing physical ones. According to Qiskit [24], selecting a logical-to- ...
  38. [38]
    [PDF] Quantum Index Report 2025 - QIR - MIT
    Jun 2, 2025 · In particular trapped-ion systems have demonstrated the highest fidelity operations and qubit connectivity and have set ambitious goals for ...Missing: realizations | Show results with:realizations
  39. [39]
    IonQ Achieves Landmark Result, Setting New World Record in ...
    IonQ Achieves Landmark Result, Setting New World Record in Quantum Computing Performance. October 21, 2025. World record 99.99% two-qubit gate performance ...
  40. [40]
    [PDF] Quantum Computing in the NISQ era and beyond - arXiv
    Noisy Intermediate-Scale Quantum (NISQ) technology will be available in the near future. Quantum computers with 50-100 qubits may be able to perform.
  41. [41]
    (PDF) Notes on Google Willow (2024) aka "Quantum Error ...
    Dec 31, 2024 · qubit coherence times and gate fidelities due to new fabrication techniques and design optimizations. Two surface codes were implemented: a ...