Fact-checked by Grok 2 weeks ago

Superconducting quantum computing

Superconducting quantum computing is a for processing that utilizes superconducting electrical circuits, primarily based on Josephson junctions, to implement qubits—the fundamental units of —that operate at millikelvin temperatures to preserve quantum coherence. These qubits, such as transmons or fluxoniums, function as artificial atoms coupled to microwave resonators via (cQED), enabling precise control of quantum states through microwave pulses and readout of measurement outcomes. This approach has emerged as one of the most mature and scalable methods for building quantum processors, supporting both gate-based and analog quantum computation paradigms. The foundations of superconducting quantum computing trace back to theoretical proposals in the for using superconducting circuits as quantum bits, with experimental breakthroughs in the late 1990s demonstrating coherent quantum oscillations in systems like single Josephson junctions and SQUIDs. Key advancements in the included the development of charge, phase, and flux qubits, followed by the introduction of the robust qubit architecture in 2007, which significantly reduced sensitivity to charge noise and improved coherence times. By the 2010s, multi-qubit devices enabled the implementation of universal quantum gates with fidelities exceeding 99%, paving the way for small-scale quantum algorithms and simulations. A pivotal milestone occurred in 2019 when demonstrated quantum supremacy using a 53-qubit programmable superconducting , performing a specific sampling task infeasible for classical supercomputers. As of 2025, superconducting quantum computing has advanced to processors with over 100 qubits, including a high-performance 105-qubit device that rivals leading architectures in gate fidelity and connectivity. Major contributors, such as , Google Quantum AI, and , have integrated these systems into cloud-accessible platforms, enabling noisy intermediate-scale quantum (NISQ) applications in optimization, chemistry, and . Recent innovations include times exceeding 1 millisecond for individual qubits, enhanced correction via surface codes, and improved scalability through modular architectures and advanced cryogenic engineering. Despite these strides, significant challenges persist, including qubit decoherence due to material defects and two-level systems, between adjacent , and the complexity of control electronics at scale. Efforts to achieve fault-tolerant require integrating thousands of logical protected by error-correcting codes, demanding innovations in , shielding, and hybrid classical-quantum interfaces. Ongoing research emphasizes improvements, such as tantalum-based for longer , and architectural shifts toward 3D integration to mitigate planar limitations.

Fundamentals

Quantum Computing Basics

Quantum computing represents a computational paradigm that harnesses to process information in ways that classical computers cannot efficiently replicate. Unlike classical computing, which relies on bits that exist definitively in one of two states (0 or 1), quantum computing employs quantum bits, or qubits, which are two-level quantum systems capable of representing a superposition of states. This fundamental difference enables quantum computers to explore multiple computational paths simultaneously, potentially offering exponential speedups for certain problems. The idea of quantum computing originated with Richard Feynman's 1982 proposal that quantum systems could simulate physical processes more effectively than classical computers, highlighting the limitations of classical simulations for quantum phenomena. Building on this, David Deutsch formalized the concept of a universal quantum computer in 1985, demonstrating that a could perform any quantum computation efficiently, extending the Church-Turing thesis to the quantum domain. The 1990s saw rapid advancements, including the introduction of the term "" by Benjamin Schumacher in 1995 to describe the basic unit of , analogous to a classical bit but with quantum properties. These early theoretical foundations laid the groundwork for exploring quantum advantages without delving into specific hardware realizations. At the heart of quantum computing are key phenomena like superposition and entanglement. Superposition allows a qubit to exist in a linear combination of its basis states, |0⟩ and |1⟩, such as α|0⟩ + β|1⟩ where α and β are complex amplitudes satisfying |α|² + |β|² = 1, enabling parallel evaluation of possibilities. Entanglement, a non-local correlation between qubits, permits states where the quantum state of one qubit cannot be described independently of others, such as the Bell state (|00⟩ + |11⟩)/√2, which underpins quantum parallelism and correlations unattainable classically. These properties provide the basis for quantum computational advantages. Quantum operations are implemented via quantum gates that manipulate qubit states reversibly. Basic single-qubit gates include the Pauli-X gate, which flips |0⟩ to |1⟩ and vice versa, analogous to a classical NOT gate, and the Hadamard gate, which creates equal superpositions from basis states, such as H|0⟩ = (|0⟩ + |1⟩)/√2. These, combined with two-qubit gates like controlled-NOT (CNOT), form universal sets capable of approximating any quantum computation, as shown by constructions using elementary gates. Superconducting qubits serve as one physical platform for implementing such gates. Prominent quantum algorithms illustrate these principles' power. Shor's algorithm, introduced in 1994, factors large integers in polynomial time on a quantum computer, exponentially faster than the best classical algorithms, threatening current by efficiently solving the problem. Grover's search algorithm, from 1996, quadratically accelerates unstructured database searches, finding a marked item among N entries in O(√N) steps compared to O(N) classically. These algorithms demonstrate quantum computing's potential for specific tasks without requiring hardware details.

Superconductivity Principles

Superconductivity is a quantum mechanical phenomenon observed in certain materials where electrical resistance drops to zero and magnetic fields are expelled (the ) when the material is cooled below a critical temperature T_c. This state enables lossless current flow, which is fundamental for constructing low-dissipation quantum circuits. The arises from the formation of persistent supercurrents that perfectly screen external magnetic fields, distinguishing from simple zero-resistance conductivity. The microscopic mechanism of conventional superconductivity is described by the Bardeen-Cooper-Schrieffer (BCS) theory, which posits that electrons form bound pairs—known as Cooper pairs—due to an attractive interaction mediated by lattice vibrations (phonons). These pairs have opposite spins and momenta, allowing them to condense into a coherent quantum state with a superconducting energy gap \Delta \approx 1.76 k_B T_c at zero temperature, where k_B is the Boltzmann constant. This gap prevents low-energy excitations, ensuring stability of the superconducting state. In this framework, the paired electrons behave collectively as bosons, forming a Bose-Einstein condensate that underlies the macroscopic phase coherence essential for quantum applications. Superconductors are classified into Type I and Type II based on their response to . Type I materials, such as lead, mercury, and aluminum, exhibit complete field expulsion up to a critical field H_c, beyond which is destroyed abruptly. Type II superconductors, such as and , allow to penetrate via quantized vortices between lower (H_{c1}) and upper (H_{c2}) critical fields, enabling higher field tolerance. For quantum devices, materials like (T_c \approx 9.2 K) and aluminum (T_c \approx 1.2 K, particularly in thin films) are preferred due to their ability to maintain in moderate fields while supporting compact, high-density circuits. Key operational limits include the critical T_c, critical H_c, and critical J_c, which define the boundaries of the superconducting phase; for , H_c is approximately 0.2 T at low temperatures. Achieving superconductivity requires cooling to temperatures well below T_c, typically using dilution refrigerators that exploit the phase separation of helium-3 and helium-4 mixtures to reach millikelvin (mK) regimes, often below 20 mK for quantum computing applications. These systems provide the ultra-low thermal noise environment necessary to minimize decoherence in superconducting quantum circuits.

Josephson Junctions

A Josephson junction consists of two superconducting electrodes separated by a thin insulating barrier, typically with a thickness of 1–2 nm, enabling quantum tunneling of Cooper pairs across the barrier. This structure forms a weak link that exhibits macroscopic quantum coherence at cryogenic temperatures, serving as the core nonlinear element in superconducting quantum circuits. The junction's behavior arises from the phase coherence of the superconducting wavefunctions on either side of the insulator. In 1962, Brian Josephson theoretically predicted the tunneling of supercurrents through such a barrier while he was a graduate student at the University of Cambridge, a discovery that earned him the 1973 Nobel Prize in Physics shared with Ivar Giaever and Leo Esaki. His seminal work extended the Bardeen-Cooper-Schrieffer (BCS) theory of superconductivity to tunnel junctions, revealing novel quantum effects. The primary phenomena are the DC and AC Josephson effects. In the DC Josephson effect, a dissipationless supercurrent flows across the junction at zero voltage bias, given by
I = I_c \sin \phi,
where I_c is the critical current and \phi is the superconducting phase difference across the barrier. This current arises from the coherent tunneling of Cooper pairs without energy dissipation, limited by I_c, beyond which a voltage develops. The AC Josephson effect occurs when a DC voltage V is applied, producing an alternating supercurrent at microwave frequencies related by the frequency-voltage relation
f = \frac{2eV}{h},
or equivalently,
\frac{d\phi}{dt} = \frac{2eV}{\hbar},
where e is the electron charge, h is Planck's constant, and \hbar = h / 2\pi. This relation stems from the time evolution of the phase difference \phi, which advances via phase slips driven by the voltage, enabling applications like voltage standards and microwave generation.
Quantum mechanically, the junction is described by a Hamiltonian capturing its inductive nonlinearity:
H = E_J (1 - \cos \phi),
where E_J = \frac{\Phi_0 I_c}{2\pi} is the Josephson energy, \Phi_0 = \frac{h}{2e} is the , and \phi is treated as a quantum operator conjugate to the charge across the junction. The phase \phi and the number of Cooper pairs N satisfy the commutation relation [\phi, 2e N] = 2\pi i, analogous to position and momentum in , allowing the junction to support quantized levels.
The cosine potential in the introduces strong , where the energy spacing between levels decreases nonlinearly, mimicking the discrete spectrum of natural atoms but on a . This , with the first separated from the by roughly E_J while higher levels are more closely spaced, enables selective addressing of states without populating unwanted higher excitations, positioning the Josephson junction as an artificial atom central to superconducting qubits. Fabrication of Josephson junctions typically employs overlap or superconductor-insulator-superconductor (SIS) processes, involving sequential deposition of superconducting thin films (e.g., aluminum or ) via or , followed by oxidation to form the barrier and patterning through . These methods ensure high uniformity and critical current densities on the order of $10^2–$10^3 A/cm², essential for reproducible quantum devices.

Hardware and Fabrication

Superconducting Circuits

Superconducting circuits form the foundational architecture for quantum processors in superconducting quantum computing, integrating classical electrical elements with quantum to realize controllable . These circuits are constructed from superconducting materials, which enable dissipationless transport of pairs at cryogenic temperatures below the critical temperature, typically using materials like or aluminum. The primary components include inductors arising from the of the superconducting film, geometric capacitors formed by parallel plates or interdigital structures, and nonlinear inductors provided by Josephson junctions (JJs). The originates from the inertia of the superconducting electron fluid, contributing significantly to the total in thin films where it can dominate over geometric . Capacitors, in contrast, store energy electrostatically and are linear, while JJs introduce essential nonlinearity through the cosine dependence of the Josephson energy on the phase difference across the junction. To analyze these circuits quantum mechanically, a quantization procedure maps the classical circuit description to a quantum using lumped-element approximations, valid when wavelengths are much larger than circuit dimensions. This involves formulating the from the circuit's kinetic (inductive) and potential (capacitive plus Josephson) energies, followed by to the Hamiltonian, with coordinates as node fluxes and momenta as node charges. For a general , the resulting Hamiltonian is quadratic in bosonic modes for linear elements but anharmonic when JJs are present, leading to multi-level systems that can be truncated to effective qubits. This approach, rooted in the Caldeira-Leggett framework adapted for lossless superconductors, enables precise prediction of circuit dynamics. Superconducting resonators, essential for storing and manipulating microwave photons, are realized as LC circuits or distributed transmission line structures, such as coplanar waveguides, operating at frequencies around 5-10 GHz to match qubit transition energies. In an LC resonator, the inductance L and capacitance C determine the resonance frequency \omega_r = 1/\sqrt{LC}, with quality factors exceeding $10^6 due to low dielectric losses in superconductors. Transmission line resonators extend this to quarter- or half-wavelength modes, providing larger mode volumes for coupling multiple qubits while maintaining high coherence times on the order of microseconds. These resonators act as quantum buses for photon-mediated interactions. The interaction between qubits and resonators is described by (cQED), where the JJ-based qubit serves as an artificial atom coupled to the bosonic field of the resonator. In the simplest model, the Jaynes-Cummings Hamiltonian captures this light-matter coupling: H = \hbar \omega_r a^\dagger a + \frac{\hbar \omega_q}{2} \sigma_z + \hbar g (a^\dagger \sigma_- + a \sigma_+), where \omega_r is the resonator frequency, \omega_q the qubit frequency, a^\dagger (a) creates (annihilates) a photon, \sigma_z, \sigma_-, \sigma_+ are Pauli operators, and g is the vacuum Rabi coupling strength, typically 50-200 MHz in modern devices. This model assumes the rotating-wave approximation and a two-level qubit approximation, valid for weak anharmonicity. A key operating regime in cQED is the dispersive limit, achieved when the qubit-resonator detuning \Delta = \omega_q - \omega_r \gg g, preventing direct energy exchange while enabling indirect readout. Here, the effective Hamiltonian becomes H \approx \hbar \omega_r a^\dagger a + \frac{\hbar \omega_q}{2} \sigma_z + \hbar \chi a^\dagger a \sigma_z, with dispersive shift \chi \approx g^2 / \Delta, causing the resonator frequency to shift by \pm \chi depending on the qubit state. This allows parametric readout via microwave probes without backaction on the qubit during measurement, with \chi / 2\pi on the order of 1-10 MHz. Photon-dependent dephasing and relaxation arise from higher-order terms but are mitigated in optimized designs. Historically, the foundations of quantum behavior in superconducting circuits trace back to studies of in resistively shunted Josephson junctions, as demonstrated by Koch et al. in 1982, who measured spectral densities confirming quantum tunneling effects at low temperatures. This work laid groundwork for macroscopic quantum coherence, evolving through the demonstrations of coherent oscillations in single JJs to the emergence of cQED in the early 2000s, with seminal cavity-qubit coupling experiments realizing the Jaynes-Cummings ladder. Today, cQED architectures support scalable multi-qubit processors with fidelities exceeding 99%.

Manufacturing Processes

The fabrication of superconducting quantum devices begins with the selection of high-purity substrates, typically or , to minimize losses and ensure compatibility with cryogenic environments. These substrates undergo rigorous cleaning via the process, which involves sequential immersion in solutions of , ammonium hydroxide, , and to remove organic contaminants, metal ions, and native oxides, thereby reducing surface defects that could degrade performance. Following substrate preparation, thin films of superconducting materials such as () and () are deposited using techniques like radio-frequency sputtering for layers or electron-beam evaporation for , achieving uniform thicknesses on the order of 50-200 nm. Patterns for circuit elements are then defined using , which provides sub-micron resolution essential for precise geometries, often in combination with resist masks to enable selective deposition. Type II superconductors like are commonly employed in these layers due to their high critical fields and ease of integration. Josephson junctions, critical nonlinear elements in these devices, are fabricated primarily through the Dolan bridge technique or angle evaporation, where Al-AlOx-Al structures are formed by sequential evaporation of aluminum electrodes separated by a thin aluminum barrier created via in-situ oxidation. This process involves shadow masking with suspended bridges or angled vapor deposition to overlap the electrodes precisely, yielding tunnel barriers with critical current densities around 1-10 A/cm². Subsequent steps include lift-off processes to remove excess material from patterned areas and (RIE) using gases like or fluorine-based s to define fine structures such as transmission lines and capacitors, ensuring sharp edges and minimal residue. These etching techniques help achieve aspect ratios suitable for multi-layer interconnects while preserving superconducting properties. All fabrication occurs in ISO 5 cleanrooms to prevent particulate contamination, with devices subsequently tested at millikelvin temperatures in dilution refrigerators to verify electrical properties before packaging. Yield rates for functional qubits typically range from 90-99%, corresponding to defect rates of about 1-10%, which pose significant challenges for as device complexity increases, often requiring redundant designs or advanced error mitigation. Advancements in 2025 have further enhanced fabrication techniques. As of April 2025, the NIST Superconducting and s (SQMS) nanofabrication improved coherence times to over 0.6 milliseconds by optimizing processes to reduce two-level defects at interfaces, enabling higher-fidelity devices through refined surface treatments and purities. In September 2025, a new fabrication approach demonstrated by NYU Tandon-led researchers enables the exploration of a broader range of superconducting beyond traditional aluminum and , potentially improving performance. Additionally, as of November 2025, research on and silicon-based has achieved millisecond-scale times, enhancing for superconducting .

Qubit Types

Transmon Qubits

The is a type of superconducting derived from the charge-sensitive box, modified by adding a large shunt to form a weakly anharmonic , which significantly reduces sensitivity to charge noise. This design operates the qubit in a regime with small charging energy E_C relative to the Josephson energy E_J, with E_J \gg E_C to suppress fluctuations in the offset charge n_g. The of the is given by \hat{H} = 4 E_C (\hat{n} - n_g)^2 - E_J \cos \hat{\phi}, where \hat{n} is the operator, \hat{\phi} is the across the junction, E_C = e^2 / 2C_\Sigma is the charging with total capacitance C_\Sigma, and E_J is the Josephson energy. For charge insensitivity, the ratio E_J / E_C is typically set to 50–100, which exponentially suppresses the dependence of energy levels on n_g while preserving sufficient for operation. The energy spectrum of the approximates that of a for low-lying levels, but with an \alpha \approx -E_C that enables selective addressing of the transition (0–1) without exciting higher levels (1–2). This arises from the nonlinearity of the Josephson potential, allowing the to function effectively as a despite its multi-level nature. Transmon qubits are implemented in two primary geometric forms: three-dimensional () versions housed in macroscopic superconducting cavities for enhanced through reduced surface losses, and two-dimensional (2D) versions integrated into planar structures for improved scalability and on-chip connectivity. In modern devices, such as IBM's 127-qubit Eagle processor (announced 2021), transmon times have reached T_1, T_2 > 400~\mu\text{s} as of 2022, with recent material innovations such as tantalum-based transmons achieving times exceeding 1 ms for individual qubits as of , limited primarily by losses and defects. The design forms the basis for leading superconducting quantum processors, including Google's 53-qubit Sycamore chip, which demonstrated in 2019 by performing a specific sampling task infeasible for classical computers. Similarly, IBM's rely on transmons for their multi-qubit architectures, enabling advances in error-corrected computation. Qubits in these systems are coupled via capacitive or inductive mechanisms to enable multi-qubit interactions.

Fluxonium Qubits

The fluxonium qubit is a flux-tunable superconducting circuit composed of a Josephson junction (JJ) in parallel with a large shunt , typically around 1 μH, forming a superconducting loop. This shunt , often realized as a geometric or a superinductor made from arrays of thin-film nanowires, creates a low-inductive-energy regime that results in a multi-well anharmonic potential for the superconducting phase variable across the JJ. The design renders the qubit insensitive to charge offsets, addressing a key limitation of early charge-based . The effective describing the fluxonium is H = 4 E_C n^2 + \frac{1}{2} E_L \phi^2 - E_J \cos \phi, where n is the (conjugate to the \phi), E_C is the charging energy of the junction's , E_L is the small inductive energy from the shunt (E_L \ll E_J), and E_J is the Josephson energy. The small E_L produces a potential landscape with multiple shallow wells separated by a Josephson barrier, enabling coherent across wells. The energy spectrum of the fluxonium features a high-frequency plasma oscillation mode at \omega_p = \sqrt{8 E_C E_L}/\hbar, typically in the tens of GHz range, while the low-energy qubit transitions occur at frequencies around 0.1–5 GHz, depending on flux bias. This separation allows encoding the qubit in the low-lying states within the potential wells, with the anharmonicity ensuring distinct levels for gate operations. At the flux sweet spot (half flux quantum), the qubit exhibits first-order insensitivity to flux noise, enhancing coherence. Key advantages of the fluxonium include extended times due to suppressed charge and , with relaxation times T_1 exceeding 1 ms demonstrated in optimized devices, such as 1.43 ms at low frequency as of 2023. tunability enables precise control of the frequency over a wide range without inducing significant decay, facilitating applications in quantum simulation and error-corrected computing. Compared to traditional qubits, the large shunt in fluxonium provides stronger suppression of , leading to superior protection. Challenges arise from the complex multi-level spectrum, which can lead to leakage errors if higher states are populated during operations, necessitating careful selection of the computational and advanced for gates. Recent developments include the use of nanowire-based superinductors to achieve values exceeding 1 μH, further reducing E_L and improving noise resilience in scalable architectures.

Flux and Phase Qubits

and qubits represent early implementations of superconducting qubits that relied on the magnetic or across Josephson junctions as the primary degree of freedom for encoding. These designs emerged in the late and early 2000s, leveraging the nonlinear of Josephson junctions to create anharmonic potential wells supporting discrete quantum states. Unlike later qubit types, and qubits exhibited high sensitivity to environmental biases, which limited their times and , ultimately leading to their supersession by more robust architectures. The flux qubit consists of a superconducting quantum interference device (SQUID) loop interrupted by three Josephson junctions (JJs), forming a micrometer-sized circuit where the two computational basis states, denoted |L⟩ and |R⟩, correspond to clockwise and counterclockwise persistent supercurrents of magnitude I_p. The energy bias between these states is given by \varepsilon = I_p (\Phi - \Phi_0/2), where \Phi is the total magnetic flux threading the loop and \Phi_0 = h/2e is the magnetic flux quantum. The qubit's transition frequency is tunable by applying an external magnetic flux \Phi_\text{ext} through the loop, allowing precise control of the symmetry point where the states are degenerate. However, coherence in flux qubits is severely limited by low-frequency $1/f flux noise, arising from defects in the superconducting material or surrounding environment, which couples strongly to the flux-sensitive Hamiltonian and causes rapid dephasing. In contrast, the phase qubit employs a single current-biased Josephson junction operated in the zero-voltage state, where the qubit states are encoded in the phase difference \phi across the junction, treated as a macroscopic quantum variable analogous to a particle in a tilted washboard potential. Quantum operations involve or coherent escape from the metastable zero-voltage well via macroscopic quantum tunneling (MQT), with the tunneling rate \Gamma \propto \exp(-\Delta U / \hbar \omega_p), where \Delta U is the barrier height and \omega_p is the plasma frequency of small oscillations in the well. The bias current tunes the tilt of the potential, setting the energy splitting between the ground and first excited states near the top of the well for optimal . The was first proposed in 1999 by Orlando et al. as a persistent-current device using three nanoscale aluminum JJs in a loop. Experimental demonstration of coherent dynamics followed shortly thereafter. The was experimentally realized in 2002 at NIST, with Rabi oscillations observed in a current-biased JJ, marking a key milestone in superconducting qubit development. Despite these advances, both qubit types suffered from short times T_2 on the order of nanoseconds to microseconds, primarily due to their strong sensitivity to flux or current biases and associated noise sources. This bias sensitivity, coupled with challenges in fabrication uniformity, led to their gradual replacement by charge-insensitive designs in practical efforts. Early variants of superconducting qubits included charge-based devices, such as the Cooper pair box (CPB), which encoded information in the number of s on a superconducting island coupled to a Josephson junction and a gate electrode. The CPB served as a direct precursor to more advanced charge-derived qubits, addressing initial charge noise issues through regime adjustments. and phase qubits have influenced subsequent designs, including shunted variants that enhance while retaining flux tunability.

Quantum Operations

Single-Qubit Gates

Single-qubit gates in superconducting quantum computing are implemented by applying pulses to coherently manipulate the of an individual , enabling rotations on the . These operations are essential for universal quantum computation, as they allow preparation of arbitrary superpositions and form the basis for more complex algorithms when combined with multi-qubit gates. In superconducting systems, such as or fluxonium qubits, the qubits are typically weakly anharmonic oscillators, and control is achieved through resonant driving that induces Rabi oscillations between the computational states |0⟩ and |1⟩. The primary drive method involves applying pulses at the transition frequency ω_q, which couples to the via its , typically through a or . This resonant driving generates Rabi oscillations, where the angle θ on the is given by θ = Ω t, with Ω being the Rabi rate proportional to the drive amplitude and t the pulse duration. For example, a π-pulse (θ = π) implements an X-gate, flipping the state, while shorter pulses achieve partial . To realize arbitrary single- gates, in-phase (I) and (Q) of the is used, allowing over the and angle in the XY-plane of the . Implementation can occur via direct drive on the qubit or through an intermediary in (cQED) architectures, with the latter often preferred to enhance coupling efficiency while minimizing unwanted interactions. Direct drive applies the pulse to the qubit's , but care must be taken to avoid population leakage to higher energy levels due to the qubit's finite , which can degrade gate fidelity. techniques, such as Gaussian or hyperbolic secant envelopes, are employed to suppress such leakage and broaden the operational bandwidth. Calibration of these gates relies on Rabi and Ramsey experiments to characterize the qubit's response. In Rabi experiments, a series of pulses with varying durations or amplitudes is applied from an initial state, followed by measurement, to map the oscillation frequency Ω and optimize pulse parameters for desired rotations. Ramsey interferometry, involving two π/2 pulses separated by free evolution, measures the qubit detuning and dephasing rate, enabling fine-tuning of the drive frequency and phase for coherent control. These methods ensure pulse shaping that compensates for hardware imperfections like drive nonlinearity. State-of-the-art single-qubit gate fidelities as of 2025 exceed 99.99%, with demonstrations achieving average errors below 10^{-5}, such as 99.998% in fluxonium qubits through advanced control techniques. For instance, in qubits, fidelities above 99.99% have been reported using refined microwave pulses. These high fidelities are primarily limited by leakage to higher levels and residual decoherence during the gate time, typically on the order of 20-50 ns. In qubits, which operate in the dispersive regime when coupled to a (with strength g ≪ |ω_q - ω_r|), single-qubit drives target the dressed states to selectively address the |0⟩-|1⟩ transition while avoiding the . The effective drive frequency is shifted by the dispersive interaction χ, where the - in this regime is H ≈ (ω_q + χ σ_z / 2) a†a + (ω_q / 2) σ_z, ensuring the pulse primarily affects the without exciting the . This selectivity is crucial for integrating with readout in scalable architectures. For fluxonium qubits, single-qubit gates can leverage pulses applied to the superconducting loop enclosing the Josephson junction, enabling fast adiabatic manipulations. These pulses tune the bias Φ through the qubit's potential landscape, allowing coherent transitions between states via avoided crossings. Adiabatic protocols, accelerated using counter-diabatic terms, achieve gate times below 20 ns with fidelities over 99.99%, up to 99.998% as of 2025, exploiting the fluxonium's large and long times.

Multi-Qubit Coupling

In superconducting quantum computing, multi-qubit enables the creation of entanglement necessary for quantum , primarily through interactions between qubits mediated by capacitive, inductive, or elements. Fixed coupling schemes provide always-on interactions, while tunable methods allow dynamic control to implement on-demand entangling operations. These approaches leverage the (cQED) framework, where superconducting qubits like transmons interact via shared degrees of freedom. Fixed coupling between qubits arises from direct capacitive or inductive connections, leading to a static interaction strength. For transmon-transmon pairs, capacitive coupling via overlapping electric fields induces a fixed transverse interaction, often resulting in an always-on ZZ term described by the Hamiltonian H_{\text{int}} = \frac{\hbar \chi}{4} \sigma_z^{(1)} \sigma_z^{(2)}, where \chi is the dispersive shift quantifying the conditional phase accumulation. Inductive coupling, common in flux-flux qubit systems, uses mutual inductance through superconducting loops to generate similar ZZ interactions, enabling persistent entanglement but complicating selective addressing. Tunable coupling addresses the limitations of fixed schemes by allowing adjustable interaction strengths, often implemented via auxiliary coupler qubits or flux-tunable Josephson junctions (JJs). Coupler qubits, such as additional transmons, mediate interactions between primary qubits and can be tuned to zero coupling when idle, suppressing unwanted crosstalk. Flux-tunable JJs, typically SQUIDs, enable continuous variation of the coupling g by applying magnetic flux, facilitating rapid activation of entangling gates without residual interactions. Key entangling gate types include the iSWAP and controlled-NOT (CNOT) gates. The iSWAP gate, often resonator-mediated, swaps excitations between s while introducing a \pi/2 shift, achieved by bringing qubit and frequencies into for a controlled duration. In bus-mediated schemes, a shared acts as a quantum bus, connecting multiple s with swap times typically ranging from 20 to 50 ns, depending on the strength g and detuning. The CNOT gate is commonly realized via the cross- (CR) , where the is driven at the 's frequency; the shifts the target's frequency by g^2 / \Delta (with g the and \Delta the detuning), enabling conditional . Alternative methods include the gate, which exploits adiabatic flux modulation in a loop enclosing the qubits to accumulate a conditional Berry phase without dynamic contributions. Heisenberg exchange interactions, manifesting as XX + YY terms in the effective , support gates like \sqrt{\text{iSWAP}} through parametric driving or direct transverse coupling. Reported fidelities for CR-based CNOT and gates as of 2025 exceed 99.9%, with demonstrations reaching 99.93% in systems, highlighting their robustness against certain errors. However, always-on ZZ interactions pose challenges, introducing unwanted conditional phases that degrade gate performance and require compensation via or dynamical .

Qubit Readout

In superconducting quantum computing, qubit readout is primarily achieved through dispersive readout, a technique where the is strongly but off-resonantly coupled to a high-quality-factor , resulting in a state-dependent shift of the 's by the dispersive strength χ. This shift, typically on the order of a few megahertz, allows the state (|0⟩ or |1⟩) to be inferred by probing the without directly exciting the . The method operates within the (cQED) framework, enabling indirect measurement that preserves in multi-qubit systems. The readout process involves sending a short pulse to the , tuned to one of the state-dependent frequencies, and detecting the transmitted or reflected signal using detection. To achieve quantum-limited noise performance essential for , the weak signal is amplified by low-noise devices such as Josephson amplifiers (JPAs) or traveling-wave amplifiers (TWPAs). JPAs, based on nonlinear Josephson junctions pumped at twice the signal frequency, provide high gain in a narrow , while TWPAs offer broader exceeding 1 GHz with near-quantum-limited added noise. Typical measurement times as of 2025 range from 20 to 100 ns, with single-shot assignment fidelities exceeding 99.5% and up to 99.9% when using these amplifiers. A key advantage of dispersive readout is its quantum non-demolition (QND) property, where the measurement projects the onto its energy eigenstates without collapsing in the computational basis, facilitated by the ac-Stark shift induced by off-resonant drive photons. This enables repeated measurements on the same for error detection or control, such as dynamical to extend . Advanced implementations include multiplexed readout, where multiple qubits share a single and readout line, distinguished by frequency-encoded signals, achieving fidelities above 95% for up to 10 qubits simultaneously. However, dispersive readout is susceptible to limitations like Purcell decay, where the qubit relaxes through the resonator channel if the resonator linewidth κ exceeds the qubit's intrinsic decay rate, potentially reducing fidelity unless mitigated by Purcell filters or engineered resonator designs.

Performance and Criteria

DiVincenzo Criteria

The DiVincenzo criteria, proposed in 2000, outline five essential requirements for a to enable scalable processing: a scalable array of well-characterized s, the ability to initialize states to a simple fiducial state, decoherence times much longer than the duration of gate operations, a of quantum gates, and -specific measurement capabilities. These criteria, along with two extensions for quantum communication, have become foundational benchmarks for evaluating platforms, including those based on superconducting s. Superconducting systems, operating at frequencies and cooled to millikelvin temperatures, have progressively demonstrated fulfillment of these criteria, enabling practical demonstrations of quantum algorithms on small scales. The first criterion requires a scalable system of well-characterized qubits, meaning the qubits must be identical, controllable, and fabricable in with consistent properties. In superconducting quantum computing, qubits—nonlinear microwave resonators made from Josephson junctions—serve as the primary implementation, with their frequency and coupling strengths tunable via fabrication and biases. Scalability has advanced significantly, with processors exceeding 1000 qubits demonstrated, such as IBM's processor featuring 1121 superconducting qubits in 2023. These qubits are well-characterized through and routines that map their energy levels and interactions with high precision. The second criterion demands reliable initialization of qubits to a known state, typically the , on demand and with high . Superconducting qubits achieve this through passive thermal relaxation, where the qubit is allowed to dissipate energy to its cold environment (at ~10 ) over the relaxation time T_1, often reaching the ground state with >99% fidelity in microseconds. For faster initialization, active reset protocols use driven pulses to couple the qubit to a lossy or auxiliary circuit, enabling unconditional reset in under 200 ns with fidelities exceeding 99%. The third criterion specifies that the qubit decoherence time T_2—characterizing the loss of phase coherence—must greatly exceed the gate operation time to allow coherent manipulations. In superconducting qubits, T_2 times have reached 100 μs to over 1 ms through improved materials and designs that minimize charge and flux noise, while single-qubit gates typically operate in ~20–50 and two-qubit gates in ~100–300 . This ratio ensures tens of thousands of gate operations before significant decoherence, as demonstrated in systems where T_2 > 1 ms supports gate fidelities above 99.99%. The fourth criterion calls for a universal gate set capable of generating any , typically comprising single-qubit rotations and a two-qubit entangling gate like the controlled-NOT (CNOT). Superconducting s implement single-qubit gates via microwave pulses resonant with the transition, achieving fidelities >99.99% in ~20 ns. The CNOT is realized through the cross-resonance () protocol, where a drive on the control at the target 's frequency induces conditional rotation, yielding entangling gates with ~99.5% fidelity in fixed-frequency arrays. This combination forms a , extensible to larger systems via nearest-neighbor couplings. The fifth criterion requires fast, high-fidelity readout of individual states without disturbing others. In superconducting platforms, dispersive readout couples the to a , shifting its frequency based on the state; a brief pulse then probes the response, enabling . Fidelities exceeding 99% are routine in ~100–200 ns, with advanced amplifiers pushing beyond 99.5% while minimizing . Two additional criteria address quantum communication: the ability to interconvert stationary () and flying () information carriers, and to faithfully transmit flying qubits over distances. Superconducting qubits meet these through on-chip integration with cavities and lines, enabling efficient and of photons as flying qubits; quantum protocols have been demonstrated via qubit-photon entanglement and coupling networks. Superconducting quantum have fulfilled all DiVincenzo criteria on small scales, as exemplified by IBM's 127-qubit in 2021, which integrated scalable transmons with gates, active resets, and dispersive readout to execute multi-qubit circuits. Ongoing advances continue to extend this fulfillment toward fault-tolerant regimes.

Coherence and Fidelity

times quantify the duration over which a superconducting maintains its before decohering due to interactions with the environment. The energy relaxation time, denoted T_1, represents the characteristic time for the qubit to decay from the |1\rangle to the |0\rangle through emission of energy to the bath, typically driven by noise at the qubit's transition frequency \omega_q. The pure time T_\phi captures the loss of phase information in superposition states without energy exchange, often arising from low-frequency noise. The overall time T_2 combines these effects via the relation \frac{1}{T_2} = \frac{1}{2T_1} + \frac{1}{T_\phi}, limiting T_2 \leq 2T_1 in the absence of pure . Additionally, the inhomogeneous time T_2^* accounts for quasistatic fluctuations, such as those from 1/f noise, and is typically shorter than T_2, measured under free evolution conditions. Fidelity metrics assess the accuracy of quantum operations on superconducting qubits. Process fidelity F for a quantum compares the ideal superoperator \chi_\text{ideal} to the experimental one \chi_\text{exp} via F = \frac{\text{Tr}(\chi_\text{ideal} \chi_\text{exp})}{d^2}, where d is the dimension ( d=2 for qubits). The average fidelity, which averages over all input states, relates to process fidelity by F_\text{avg} = \frac{d F + 1}{d+1}, providing a practical measure of performance. These metrics highlight deviations from ideal unitary evolution due to noise and imperfections. Coherence and fidelity are evaluated using techniques like randomized benchmarking (RB), which applies sequences of random to extract average error rates per gate, typically on the order of 0.1% for state-of-the-art single-qubit operations. In RB, the survival probability decays exponentially with sequence length, yielding the error rate r from fits to A p^m + B, where p = 1 - \frac{3r}{2} for qubits, independent of state preparation and errors. Recent implementations have achieved single-qubit gate fidelities exceeding 99.99% using this method. Dielectric loss in superconducting qubits arises primarily from material interfaces and is characterized by the loss tangent \tan \delta \approx 10^{-6}, limiting T_1 through dissipation in capacitor dielectrics. This loss contributes to energy relaxation rates via participation ratios of in lossy regions. Additionally, 1/f , stemming from defects such as trapped charges or spins at interfaces, induces flux and critical current fluctuations that dominate , with power scaling as $1/f^\alpha where \alpha \approx 1. State-of-the-art benchmarks for qubits as of 2025 include T_1 times exceeding 1 ms (up to 1.68 ms), enabled by improved fabrication and materials such as tantalum-on-sapphire to reduce surface losses. Single-qubit gate fidelities routinely surpass 99.99%, with two-qubit gates exceeding 99.9%, as verified by and . For logical qubits, the serves as a , quantifying the largest problem size (in qubits) solvable with 97% accuracy using variational algorithms like QAOA on the Max-Cut problem, indicating potential for quantum advantage. Values above 20 suggest utility for industrially relevant optimization tasks.

Challenges and Advances

Decoherence and Noise

In superconducting quantum computing, decoherence arises primarily from interactions between qubits and their environment, leading to loss of quantum information through energy relaxation and . Key sources include mediated by two-level systems (TLS) in amorphous dielectrics, which couple to the of the qubit and cause . Flux and charge noise, characterized by a 1/f power spectrum, originate from fluctuating magnetic impurities or defects near the Josephson junctions, predominantly affecting flux-sensitive qubits like fluxoniums. Thermal photons from in the environment can also excite or relax the state, particularly at higher operating temperatures around 10-20 mK. Additional decoherence mechanisms involve the Purcell effect, where the qubit couples to a readout resonator, enhancing spontaneous emission into the transmission line and limiting relaxation time T_1 to microseconds in typical designs. Cosmic rays, including muons and gamma rays, generate bursts of quasiparticles by breaking Cooper pairs across the chip, inducing correlated errors and sudden decoherence events in qubit arrays. Quasiparticle dynamics post-fabrication result in nonequilibrium densities around 10^{17} to 10^{18} cm^{-3}, far exceeding thermal equilibrium values, which poison the qubit by tunneling across junctions and reducing T_1 by factors of 10 or more. Mitigation strategies target these noise sources directly. qubits employ charge-insensitive designs with large shunt capacitance to suppress charge noise effects, achieving coherence times up to 100 μs while minimizing sensitivity to 1/f fluctuations. For , normal-metal trapping layers adjacent to superconducting films absorb and recombine excitations, reducing density by over an and extending T_1. Shielding with lead or muon veto detectors further attenuates impacts, lowering quasiparticle burst rates in underground or enclosed setups. Error mitigation techniques like zero-noise extrapolation (ZNE) address residual by executing circuits at amplified noise levels via pulse stretching or folding, then extrapolating to the zero-noise limit using polynomial fits, improving expectation value accuracy by 2-5 times in noisy superconducting processors. Collectively, these sources limit circuit depths to approximately 1000 two-qubit gates without correction, constraining NISQ applications to shallow algorithms.

Scalability and Error Correction

One of the primary scalability hurdles in superconducting quantum computing arises from the need for extensive wiring to control and read out individual qubits, with estimates suggesting that a million-qubit system could require on the order of a million lines, leading to significant challenges in signal and management. between these lines further complicates scaling, as unintended interactions between control signals can introduce errors during operations. Additionally, the cryogenic (I/O) infrastructure at millikelvin temperatures imposes limits on the number of connections feasible without excessive heat load or wiring congestion. To address these wiring and I/O bottlenecks, researchers have pursued modular designs that distribute s across multiple chips, enabling chip-to-chip coupling through tunable couplers such as vacuum-gap capacitors or superconducting indium-bump bonds. These approaches facilitate scalable architectures by allowing independent scaling of qubit arrays while minimizing intra-chip wiring density. Emerging proposals also explore optical links between cryogenic modules to interconnect distant processors, potentially surpassing the performance limits of single-unit systems and supporting larger-scale networks. A key strategy for achieving in superconducting systems is the , a topological quantum error-correcting with an error threshold of approximately 1%, below which logical error rates decrease exponentially with size. Implementing typically requires a distance d \approx 10, which encodes one logical using roughly $2d^2 physical s to suppress errors from noise sources like decoherence. This distance ensures the code can correct multiple errors per cycle, enabling reliable computation as qubit counts grow. Beyond the surface code, adaptations of other error-correcting schemes, such as the —a seven-qubit that corrects single-qubit errors—have been demonstrated in superconducting platforms, often integrated into larger architectures for efficient syndrome measurement. Cat codes, leveraging bosonic modes in superconducting cavities, provide hardware-efficient error correction by stabilizing superpositions of coherent states to mitigate photon-loss errors prevalent in these systems. Effective implementation requires low-latency decoders, such as parallel window or predecoding algorithms, to process syndrome data in without exceeding the limits of superconducting qubits. The progression of qubit counts in superconducting quantum processors has advanced from small-scale demonstrations of around 5 qubits in 2010 to over 1000 qubits by 2025, driven by improvements in fabrication and control techniques. Future projections aim for systems exceeding 4000 qubits in the coming years, reflecting ongoing efforts to bridge the gap to fault-tolerant scales. Hybrid classical-quantum approaches, particularly variational quantum algorithms (VQAs), mitigate scalability issues by reducing circuit depth through iterative optimization on classical hardware, allowing shallower quantum circuits that are more feasible on noisy superconducting devices. These algorithms parameterize quantum states and adjust parameters to approximate solutions, thereby lowering the error accumulation in deeper entangling operations. In support of these scaling efforts, the U.S. Department of Energy allocated $625 million in 2025 to renew its five National Research Centers, focusing on advancements in architectures at national laboratories, including superconducting platforms.

Recent Developments

In 2025, superconducting quantum computing advanced through key algorithmic and hardware milestones, enhancing the pathway to practical applications. Quantum AI demonstrated verifiable quantum advantage using the Quantum Echoes algorithm on its Willow processor, a 105-qubit superconducting chip, by simulating complex physics problems beyond classical capabilities in a computationally verifiable manner. Similarly, D-Wave achieved in a peer-reviewed study by simulating magnetic materials properties on its annealing quantum computer, completing tasks in minutes that would take classical supercomputers nearly one million years. Hardware innovations focused on improving qubit performance and fabrication. In November 2025, Princeton engineers developed a new superconducting design that achieves coherence times over 1 millisecond—three times longer than prior state-of-the-art versions—paving the way for more reliable quantum operations. This builds on qubit architectures with advanced materials and fabrication s. In April 2025, introduced a fabrication using to create partially suspended superconducting structures on , significantly reducing substrate noise and improving qubit robustness for scalable systems. Coherence enhancements were a priority, with the NIST-led Superconducting and Systems (SQMS) Center reporting systematic improvements via advanced nanofabrication, yielding relaxation times (T1) exceeding 500 μs in optimized devices and approaching millisecond scales in prototypes. Leading players drove ecosystem growth. IBM's processor, with 1,121 s, remained a for scale, supporting hybrid quantum-classical workflows, while its updated roadmap targets the 1,386-qubit for 2026 multi-chip integration. Quantum AI continued with Willow-class chips, Rigetti launched commercial sales of its 9-qubit Novera QPU in September 2025 for on-premises research, and efforts from , QuTech, and Quantum Circuits advanced tunable couplers and error mitigation. Funding bolstered these efforts, with the U.S. Department of Energy announcing $625 million in November 2025 to renew five , including SQMS and Q-NEXT, allocating $125 million initially over five years to accelerate and software development. Near-term applications emphasized optimization and simulations, where superconducting systems like D-Wave's provide advantages in materials design, with projections for fault-tolerant computing by 2030 through error-corrected logical qubits. Addressing scalability, researchers updated challenges by exploring 3D integration techniques, such as flip-chip bonding and hybrid packaging, to enable systems with up to 10^6 by reducing interconnect losses and supporting dense qubit arrays.

References

  1. [1]
    A Quantum Engineer's Guide to Superconducting Qubits - arXiv
    Apr 13, 2019 · The aim of this review is to provide quantum engineers with an introductory guide to the central concepts and challenges in the rapidly accelerating field of ...
  2. [2]
    [2006.10433] Superconducting Quantum Computing: A Review - arXiv
    Jun 18, 2020 · In this work, we provide a brief review on the experimental efforts towards building a large-scale superconducting quantum computer.
  3. [3]
    [1905.13641] Superconducting Qubits: Current State of Play - arXiv
    May 31, 2019 · In this brief review, we discuss several of the recent experimental advances in qubit hardware, gate implementations, readout capabilities, ...
  4. [4]
    Quantum supremacy using a programmable superconducting ...
    Oct 23, 2019 · Here we report the use of a processor with programmable superconducting qubits2,3,4,5,6,7 to create quantum states on 53 qubits, corresponding ...Missing: review | Show results with:review
  5. [5]
    Physics - Superconducting Quantum Computing Beyond 100 Qubits
    Mar 3, 2025 · A new high-performance quantum processor boasts 105 superconducting qubits and rivals Google's acclaimed Willow processor.<|separator|>
  6. [6]
    Superconducting quantum computers: who is leading the future?
    Aug 19, 2025 · This review examines the state of superconducting quantum technology, with emphasis on qubit design, processor architecture, scalability, and ...
  7. [7]
  8. [8]
    A Review of Design Concerns in Superconducting Quantum Circuits
    ### Summary of Abstract and Introduction (arXiv:2411.16967)
  9. [9]
    Quantum theory, the Church–Turing principle and the universal ...
    Deutsch David. 1985Quantum theory, the Church–Turing principle and the universal quantum computerProc. R. Soc. Lond. A40097–117http://doi.org/10.1098/rspa ...
  10. [10]
    Simulating physics with computers | International Journal of ...
    Feynman, RP Simulating physics with computers. Int J Theor Phys 21, 467–488 (1982). https://doi.org/10.1007/BF02650179
  11. [11]
    Quantum coding | Phys. Rev. A - Physical Review Link Manager
    Quantum coding. Benjamin Schumacher. Department of Physics, Kenyon College, Gambier, Ohio ... A 51, 2738 – Published 1 April, 1995. DOI: https://doi.org/ ...
  12. [12]
    Elementary gates for quantum computation | Phys. Rev. A
    Nov 1, 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.
  13. [13]
    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 ...
  14. [14]
    A fast quantum mechanical algorithm for database search - arXiv
    Nov 19, 1996 · Authors:Lov K. Grover (Bell Labs, Murray Hill NJ). View a PDF of the paper titled A fast quantum mechanical algorithm for database search, by ...
  15. [15]
    Intrinsic Paramagnetic Meissner Effect Due to $s$-Wave Odd ...
    Nov 6, 2015 · Superconductivity, which occurs in certain materials cooled below a critical temperature, involves the disappearance of electrical resistance ...Missing: definition | Show results with:definition
  16. [16]
    Imaging the Meissner effect and flux trapping of superconductors ...
    The Meissner effect, i.e., the expulsion of an external static magnetic field when a material is cooled below its critical temperature T c [1] , is the ...
  17. [17]
    Theory of Superconductivity | Phys. Rev.
    A theory of superconductivity is presented, based on the fact that the interaction between electrons resulting from virtual exchange of phonons is attractive.Missing: definition | Show results with:definition
  18. [18]
    Microscopic model of critical current noise in Josephson-junction ...
    Sep 18, 2009 · We assume ϵ d = 0 , with the superconducting energy gap dependent on temperature according to Δ = 1.76 k B T c 1 − T / T c for T ≤ ...Missing: T_c | Show results with:T_c
  19. [19]
    When superconductivity crosses over: From BCS to BEC
    May 23, 2024 · This attraction causes fermions to form pairs, called Cooper pairs, that are in some sense “bosonic” ( Schrieffer, 1964 ). Because of this ...
  20. [20]
    [PDF] Characterizing Niobium Nitride Superconducting Microwave ... - arXiv
    Second, type II superconductors have a lower coherence length in comparison with type I superconductors; therefore, energy losses have a higher value at single ...
  21. [21]
    Anisotropic superconductivity of niobium based on its response to ...
    It is tremendously important for applications, and it has the highest superconducting transition temperature, T c = 9 . 3 2 K, of all pure metals. In addition ...Missing: T_c 9.2
  22. [22]
    Superconductivity in an extreme strange metal - Nature
    Jul 21, 2021 · Conventional superconductors such as aluminum and niobium are now key players in the race for realizing the quantum computer. Understanding ...
  23. [23]
    Theoretical modeling of critical temperature increase in ...
    May 24, 2016 · First, let us consider the case of niobium, which has T Cbulk = 9.2 K and θ D = 275 K in the bulk form [8] . Enhancing the superconducting ...Missing: T_c | Show results with:T_c
  24. [24]
    Neutron imaging of an operational dilution refrigerator - Nature
    Jan 21, 2022 · Today, dilution refrigeration is the most widespread technology for accessing the temperature range between about 5 mK and 1 K, and is the only ...
  25. [25]
    Quantum bath suppression in a superconducting circuit by ... - Nature
    Jun 14, 2023 · We demonstrate how to remove this thermal constraint by operating a circuit immersed in liquid 3 He. This allows to efficiently cool the decohering environment.
  26. [26]
    [PDF] Superconducting Qubits and the Physics of Josephson Junctions
    In this paper we focus on the physics of the Josephson junction because, being nonlinear, it is the fundamental circuit element that is needed for the ...<|control11|><|separator|>
  27. [27]
    Superconducting Circuit Companion---an Introduction with Worked ...
    Dec 14, 2021 · We consider primarily three different components of a superconducting circuit: linear capacitors, linear inductors, and nonlinear Josephson ...
  28. [28]
    Circuit quantum electrodynamics | Rev. Mod. Phys.
    May 19, 2021 · Section III is devoted to light-matter coupling in circuit QED, including a discussion of the Jaynes-Cummings model and its dispersive limit.
  29. [29]
    [2005.12667] Circuit Quantum Electrodynamics - arXiv
    May 26, 2020 · Circuit QED allows the study and control of light-matter interaction at the quantum level in unprecedented detail.
  30. [30]
    Dispersive regime of circuit QED: Photon-dependent qubit ...
    Jan 23, 2009 · Measurement thus induces both dephasing and mixing of the qubit, something that can reduce the quantum nondemolition aspect of the readout.
  31. [31]
    Dispersive regime of circuit QED: photon-dependent qubit ... - arXiv
    Oct 8, 2008 · In this paper, we investigate how non-linear corrections to the dispersive regime affect the measurement process.
  32. [32]
    Measurements of quantum noise in resistively shunted Josephson ...
    Jul 1, 1982 · Measurements of quantum noise in resistively shunted Josephson junctions. Roger H. Koch, D. J. Van Harlingen*, and John Clarke.
  33. [33]
    Improved superconducting qubit coherence with high-temperature ...
    Oct 11, 2025 · For Si substrates, an RCA cleaning step [64], [124] is typically performed. High-temperature annealing of sapphire substrates both helps to ...
  34. [34]
    [PDF] arXiv:2304.13257v1 [quant-ph] 26 Apr 2023
    Apr 26, 2023 · the qubits prepared on silicon substrates the process involved first cleaning the substrate with a RCA surface treatment de- tailed ...
  35. [35]
    Challenges and solutions in RF sputtering of superconducting Nb for ...
    Nov 22, 2024 · Here, we give practical and simple solutions for the sputter deposition of high-quality nanostructured superconducting Nb thin films. Our ...
  36. [36]
    Path toward manufacturable superconducting qubits with relaxation ...
    Aug 9, 2022 · These have been exclusively fabricated using aluminum (Al) double-angle evaporation ... sputtering deposition, and advanced optical lithography.
  37. [37]
    Fabrication of Al/AlOx/Al Josephson junctions and superconducting ...
    The Dolan bridge shadow mask evaporation technique we use for fabricating Al/AlOx/Al junctions consists of two steps, the first one is to fabricate a ...
  38. [38]
    Wafer-scale uniformity improvement of Dolan-bridge Josephson ...
    Aug 1, 2025 · Here, we report on a robust wafer-scale Al/AlOx/Al Dolan-bridge Josephson junction (JJ) process using preliminary shadow evaporation bias resist ...
  39. [39]
    Superconducting Quantum Circuits - Oxford Instruments
    Oct 27, 2020 · Traditionally, Josephson tunnel junctions for superconducting qubits are fabricated using lift-off processes, thus suffer from outgassing of the ...
  40. [40]
    Fabrication of metal air bridges for superconducting circuits using ...
    Jul 28, 2025 · To remove the cross-linked resist formed during the argon ion mill step and assist the liftoff process, a reactive ion etch process using oxygen ...
  41. [41]
    Facilities - Enabling Labs and Tools - SQMS Center
    Rigetti Fab-1. Pritzker Nanofabrication Facility. PNF is a 10,000 square-foot, ISO 5 (class 100) cleanroom with state-of-the-art equipment for nano/micro ...
  42. [42]
    Fabricating a superconducting quantum processor in a cleanroom
    Apr 24, 2020 · In this blogpost, we'll discuss the origins and the difficulties of fabricating a superconducting quantum processor.
  43. [43]
    Advanced CMOS manufacturing of superconducting qubits on 300 ...
    Sep 18, 2024 · Here we demonstrate superconducting transmon qubits manufactured in a 300 mm complementary metal–oxide–semiconductor (CMOS) pilot line using industrial ...
  44. [44]
    Quantum Breakthroughs: NIST & SQMS Lead the Way
    Apr 4, 2025 · Recent innovations by the SQMS Nanofabrication Taskforce have led to a systematic improvement in qubit coherence, with the best-performing ...
  45. [45]
    New material platform for superconducting transmon qubits ... - Nature
    Mar 19, 2021 · We fabricate two-dimensional transmon qubits that have both lifetimes and coherence times with dynamical decoupling exceeding 0.3 milliseconds.
  46. [46]
    [0906.0831] Fluxonium: single Cooper pair circuit free of charge offsets
    Jun 4, 2009 · Fluxonium: single Cooper pair circuit free of charge offsets. Authors:Vladimir E. Manucharyan, Jens Koch, Leonid Glazman, Michel Devoret (Yale ...
  47. [47]
    Microwaves in Quantum Computing - PMC - NIH
    (c) Under cw resonant driving, the qubit state undergoes so-called Rabi oscillations, where the probabilities |α0|2 and |α1|2 evolve sinusoidally in time. (d) ...
  48. [48]
    Calibrating single-qubit gates by a two-dimensional Rabi oscillation
    Mar 17, 2023 · We present and analyze a simple scheme to calibrate single–qubit gates. It determines the amplitude and phase difference between a quadrature pair of drives.
  49. [49]
    Single qubit operations using microwave hyperbolic secant pulses
    Oct 30, 2017 · Single-qubit gates are required for quantum computing and simulations. In the case of rotations about an axis in the X Y plane, the control ...
  50. [50]
    Error per single-qubit gate below 10−4 in a superconducting qubit
    Nov 3, 2023 · In this work, we fabricate a transmon qubit with long coherence times and demonstrate single-qubit gates with the average gate error below 10 −4.
  51. [51]
    Calibration of Drive Nonlinearity for Arbitrary-Angle Single-Qubit ...
    Aug 15, 2023 · In this work, we use a measurement based on error amplification to characterize and correct the small single-qubit rotation errors.
  52. [52]
    Charge insensitive qubit design derived from the Cooper pair box
    Feb 28, 2007 · Here, we introduce a new type of superconducting qubit called the "transmon." Unlike the charge qubit, the transmon is designed to operate in a ...
  53. [53]
    Analytic Design of Accelerated Adiabatic Gates in Realistic Qubits
    Jul 9, 2021 · We show in detail how our ideas can be used to analytically design high-fidelity single-qubit “tripod” gates in a realistic superconducting ...FLUXONIUM QUBIT · GATE PERFORMANCE... · FULL GATE PERFORMANCE...
  54. [54]
  55. [55]
  56. [56]
    Approaching Unit Visibility for Control of a Superconducting Qubit ...
    Aug 1, 2005 · Approaching Unit Visibility for Control of a Superconducting Qubit with Dispersive Readout. A. Wallraff, D. I. Schuster, A. Blais, L. Frunzio ...
  57. [57]
    A near–quantum-limited Josephson traveling-wave parametric ...
    Sep 3, 2015 · The amplifier exhibited high gain over a gigahertz-sized bandwidth and was able to perform high-fidelity qubit readout. Because the amplifier ...
  58. [58]
    Rapid High-Fidelity Single-Shot Dispersive Readout of ...
    May 26, 2017 · The state-of-the-art readout of superconducting qubits is based on the dispersive interaction with a readout resonator. Here, we bring this technique to its ...
  59. [59]
    The Physical Implementation of Quantum Computation - arXiv
    Feb 25, 2000 · These five requirements, plus two relating to the communication of quantum information, are extensively explored and related to the many schemes ...
  60. [60]
    Discover the World's Largest Quantum Computer in 2025 - SpinQ
    Jan 6, 2025 · Announced at the IBM Quantum Summit in December 2023, Condor is the world's first superconducting qubit quantum processor to surpass the 1,000- ...
  61. [61]
    A quantum engineer's guide to superconducting qubits
    Jun 17, 2019 · The aim of this review is to provide quantum engineers with an introductory guide to the central concepts and challenges in the rapidly accelerating field of ...
  62. [62]
    Efficient protocol for qubit initialization with a tunable environment
    Jul 14, 2017 · We propose an efficient qubit initialization protocol based on a dissipative environment that can be dynamically adjusted.
  63. [63]
    Active initialization experiment of a superconducting qubit using a ...
    Here, we demonstrate the rapid initialization of a superconducting qubit with a quantum circuit refrigerator (QCR).
  64. [64]
    Native Two-Qubit Gates in Fixed-Coupling, Fixed-Frequency ...
    May 20, 2024 · This work, along with CR and native Z Z interactions, completes all possible native two-qubit interactions for fixed-frequency, fixed-coupling ...
  65. [65]
    Transmon qubit readout fidelity at the threshold for quantum error ...
    Mar 16, 2023 · We demonstrate single-shot readout fidelity up to 99.5% (96.9%) for two-state (three-state) discrimination within 140 ns without using a quantum-limited ...
  66. [66]
    IBM Unveils Breakthrough 127-Qubit Quantum Processor
    Nov 16, 2021 · 'Eagle' is IBM's first quantum processor developed and deployed to contain more than 100 operational and connected qubits.Missing: DiVincenzo | Show results with:DiVincenzo
  67. [67]
    None
    ### Extracted Formulas for Process Fidelity and Average Gate Fidelity
  68. [68]
    None
    ### Summary of Randomized Benchmarking for Gate Error Rates in Superconducting Qubits
  69. [69]
    [PDF] Material matters in superconducting qubits - arXiv
    For Si or sapphire substrates, pSUB is approximately 90%, much higher than any other portion of the superconducting qubit geometry. As discussed in the previous ...
  70. [70]
    [PDF] 1 Model for l/f Flux Noise in SQUIDs and Qubits Roger H ... - arXiv
    We propose a model for 1/f flux noise in superconducting devices (f is frequency). The noise is generated by the magnetic moments of electrons in defect states ...
  71. [71]
    Systematic improvements in transmon qubit coherence enabled by ...
    Apr 26, 2024 · Stabilizing and improving qubit coherence by engineering the noise spectrum of two-level systems. Phys. Rev. Appl. 18, 044026 (2022) ...
  72. [72]
    Atos Announces Q-Score Metrics for Assessing Quantum ... - HPCwire
    Dec 4, 2020 · Q-score can be measured for QPUs with more than 200 qubits. Therefore, it will remain the perfect metrics reference to identify and measure ...
  73. [73]
    Decoherence in Josephson Qubits from Dielectric Loss
    Microwave dielectric loss for materials used in superconducting qubit fabrication. ... This decay rate is equivalent to a dielectric loss tangent δ i = ℏ Γ 1 / E ...
  74. [74]
    [PDF] Extending Coherence in Superconducting Qubits - Schoelkopf Lab
    This thesis follows the evolution of superconducting qubits from coherence times of several microseconds to nearly a millisecond. Page 2. Extending Coherence in ...<|separator|>
  75. [75]
    Superconducting Qubit with Purcell Protection and Tunable Coupling
    Jan 20, 2011 · At g=0, the qubit is in a decoherence-free subspace with respect to spontaneous emission induced by the Purcell effect. Furthermore, we show ...
  76. [76]
    Cosmic-ray-induced correlated errors in superconducting qubit array
    May 20, 2025 · Recent experiments with superconducting qubits indicate that they can arise from quasiparticle (QP) bursts induced by cosmic-ray muons and γ-rays.
  77. [77]
    [PDF] Quasiparticle Poisoning of Superconducting Qubits from Resonant ...
    Observed quasiparticle densities are of order 1 µm−3, tens of orders of magnitude larger than the equilibrium density expected from theory. Nonequilibrium ...Missing: post- | Show results with:post-
  78. [78]
    Normal-metal quasiparticle traps for superconducting qubits
    Sep 20, 2016 · In qubits, quasiparticles interact with the phase degree of freedom, providing an unwanted channel for the qubit energy relaxation [4, 5] .Article Text · THE DIFFUSION AND... · QUASIPARTICLE DYNAMICS... · SUMMARY
  79. [79]
    Quasiparticle Dynamics in a Superconducting Qubit Irradiated by a ...
    Aug 7, 2024 · A known source of decoherence in superconducting qubits is the presence of broken Cooper pairs, or quasiparticles. These can be generated by ...Abstract · Article Text · Results · Discussion and conclusion
  80. [80]
    [PDF] arXiv:2505.04463v1 [quant-ph] 7 May 2025
    May 7, 2025 · This paper presents noise-adaptive folding, a technique that enhances zero-noise extrapolation (ZNE) through the use of adaptive scaling factors.
  81. [81]
    Decoherence benchmarking of superconducting qubits - Nature
    Jun 26, 2019 · We benchmark the decoherence of superconducting transmon qubits to examine the temporal stability of energy relaxation, dephasing, and qubit transition ...
  82. [82]
    Scaling up Superconducting Quantum Computers With Cryogenic ...
    Jan 1, 2024 · We motivate a systematic approach for understanding and mitigating the crosstalk noise when executing near-term quantum programs on ...
  83. [83]
    [PDF] Scalable multi-chip quantum architectures enabled by cryogenic ...
    One of the main challenges for scaling up quantum comput- ers is the wiring between the quantum processor and control electronics [7]. Existing quantum ...
  84. [84]
    Modular superconducting-qubit architecture with a multichip tunable ...
    May 30, 2024 · We demonstrate three different designs of multichip tunable couplers using vacuum-gap capacitors or superconducting indium-bump bonds.
  85. [85]
    Quantum error correction below the surface code threshold - Nature
    Dec 9, 2024 · We present two below-threshold surface code memories on our newest generation of superconducting processors, Willow: a distance-7 code, and a distance-5 code.
  86. [86]
    Scaling and logic in the colour code on a superconducting quantum ...
    May 26, 2025 · The smallest instance of the colour code, the Steane code, has been implemented with trapped ions and neutral atoms, demonstrating logical ...
  87. [87]
    Hardware-efficient quantum error correction via ... - Nature
    Feb 26, 2025 · Here, using a superconducting quantum circuit, we realize a logical qubit memory formed from the concatenation of encoded bosonic cat qubits ...
  88. [88]
    Local Predecoder to Reduce the Bandwidth and Latency of ...
    Mar 15, 2023 · To this end, we propose a local “predecoder,” which makes greedy corrections to reduce the amount of syndrome data sent to a standard matching ...
  89. [89]
    [PDF] Quantum Information Processing with Superconducting Qubits
    In the realization of circuit QED, two superconducting. 'transmon' charge qubits are capacitively coupled to a one-dimensional microwave transmis- sion line ...
  90. [90]
    Reducing Circuit Depth in Adaptive Variational Quantum Algorithms ...
    Jul 6, 2022 · We propose a new scheme to construct the effective Hamiltonian with the transformation in the form of a product of linear combinations of excitation operators.
  91. [91]
  92. [92]
    A verifiable quantum advantage - Google Research
    Oct 22, 2025 · Remarkably, we show that running the Quantum Echoes algorithm on the Willow quantum chip is already in the beyond-classical regime for a set of ...
  93. [93]
    D-Wave First to Demonstrate Quantum Supremacy on Useful, Real ...
    Mar 12, 2025 · Research shows D-Wave annealing quantum computer performs magnetic materials simulation in minutes that would take nearly one million years ...
  94. [94]
  95. [95]
    A Quiet Revolution: New Technique Could Accelerate Noise-Free ...
    Apr 10, 2025 · A new fabrication technique that could improve noise robustness in superconducting qubits, a key technology to enabling large-scale quantum computers.
  96. [96]
    IBM roadmap to quantum-centric supercomputers (Updated 2024)
    May 10, 2022 · In 2025, we'll introduce the 1,386-qubit multi-chip processor, “Kookaburra.” With its communication link support for quantum parallelization, ...
  97. [97]
    Rigetti Announces Purchase Orders for Two Quantum Computing ...
    Sep 30, 2025 · The Novera QPU is manufactured in Rigetti's Fab-1, the industry's first dedicated and integrated quantum device manufacturing facility.Missing: launch | Show results with:launch
  98. [98]
  99. [99]
    2025 is the year of quantum computing | Constellation Research Inc.
    Oct 18, 2025 · 2025 is the year of quantum computing. By Larry Dignan. Quantum ... development and commercialization of superconducting quantum computing.
  100. [100]
    Quantum computing's six most important trends for 2025 - Moody's
    Feb 4, 2025 · Quantum computing's six most important trends for 2025 · More experiments with logical qubits · More specialized hardware/software (as opposed to ...
  101. [101]
    Building scalable, innovative quantum systems - Rigetti Computing
    We combine modern silicon semiconductor and MEMS processing technologies with novel manufacturing methods to produce state-of-the-art superconducting qubits and ...
  102. [102]
    EU Selects SUPREME Consortium to Scale Up Industrial Production ...
    Jul 9, 2025 · The SUPREME consortium will focus on developing technologies including angle evaporated and etched Josephson junctions, 3D integration methods, and hybrid ...