Fact-checked by Grok 2 weeks ago

Flux qubit

A flux qubit is a type of superconducting quantum bit (qubit) proposed in 1999, consisting of a micrometer-sized loop of superconducting material interrupted by three or four Josephson junctions, where the two computational states correspond to persistent supercurrents circulating in clockwise and counterclockwise directions around the loop. The quantum information is encoded in the coherent superposition of these current states, which are the two lowest-energy eigenstates of the system, separated by a tunable energy gap. The operating principle of the flux qubit relies on the macroscopic quantum coherence of superconducting circuits, where the Josephson junctions introduce the essential nonlinearity for defining discrete energy levels analogous to an artificial atom. An external threads flux through the loop, tuning the qubit's transition frequency via the flux-dependent formed by the circulating currents; optimal coherence is achieved at the degeneracy point of half a flux quantum (Φ₀/2 ≈ 1.07 × 10⁻¹⁵ Wb), where flux noise sensitivity is minimized. The effective in the phase basis for a three-junction flux qubit approximates H = 4E_C (n - n_g)² - E_J cos(φ) + (Φ₀/2π)² (1/L_J + 1/L) (φ - 2π Φ_ext/Φ₀)², where E_C is the charging energy, E_J the Josephson energy, L_J the junction inductance, L the loop inductance, n the charge number operator, φ the phase difference, and Φ_ext the external flux—capturing the interplay of capacitive, Josephson, and inductive energies. Early demonstrations in the early 2000s confirmed coherent dynamics and Rabi oscillations in flux qubits, marking them as one of the first viable superconducting qubit designs alongside charge and phase qubits. Compared to the charge qubit (sensitive to offset charge noise in a Josephson junction island) and the (based on a current-biased Josephson junction with a cubic potential), the flux qubit offers advantages in gate speed (∼10–50 ns) due to its and insensitivity to tunneling, though it faces challenges from magnetic flux noise. Variants have addressed these limitations: the capacitively shunted flux qubit (CSFQ) reduces flux noise by operating away from degeneracy, while the fluxonium, introduced in 2009, incorporates a large geometric (superinductor) to delocalize the flux variable, achieving coherence times exceeding 1 ms and high for improved single- and two-qubit gate fidelities. Recent advancements include zero-flux-biased flux qubits using ferromagnetic Josephson π-junctions, enabling operation without external flux tuning and potentially simplifying scalable quantum processors. Flux qubits have been integrated into multi-qubit systems for and error-corrected computation, with ongoing research focusing on enhancing connectivity and mitigating decoherence for fault-tolerant .

Introduction

Definition and operating principle

A flux qubit is a type of solid-state superconducting qubit that encodes quantum information in the persistent circulating currents within a micrometer-scale superconducting loop interrupted by a small number of Josephson junctions. The two basis states of the qubit correspond to clockwise and counterclockwise current directions, which generate opposite magnetic fluxes through the loop, on the order of 10^{-3} flux quanta. These persistent current states arise from the quantization of magnetic flux in the superconducting loop, enabling the qubit to store information in a manner analogous to a classical loop current but with quantum coherence. The operating principle relies on the quantum mechanical superposition of the clockwise and counterclockwise current states, forming the qubit's ground and excited states. An external threaded through the loop tunes the relative energies of these states, creating a controllable energy splitting that is minimized at the flux degeneracy point (half a flux quantum), where the system exhibits optimal coherence. This splitting typically ranges from 10 to 20 GHz, corresponding to frequencies suitable for initialization, manipulation, and readout of the . In a standard design, the flux qubit consists of a closed superconducting loop incorporating three Josephson junctions, with two identical larger junctions and one smaller junction to introduce asymmetry. This configuration generates a flux-tunable in the space of fluxoid numbers (integer multiples of the flux quantum), where the qubit states localize in the wells corresponding to the opposite current directions. The asymmetry ensures that the external can effectively tilt the potential, modulating the tunneling between wells and thus the qubit's . Flux qubits differ from other superconducting qubits, such as charge qubits that encode information in charge states on an island or phase qubits that rely on phase differences across a single current-biased , by using the geometry to make the primary control variable. This flux-based encoding provides inherent tunability and reduced susceptibility to charge noise compared to early charge qubits.

Historical development

The flux qubit, also known as the persistent-current qubit, was theoretically proposed in 1999 by a team including T. P. Orlando and J. E. Mooij, who described a superconducting loop interrupted by three Josephson junctions to encode quantum information in circulating persistent currents. This design leveraged flux quantization in superconducting circuits to create a two-level quantum system tunable via external magnetic flux, marking a key advancement in solid-state quantum bits beyond charge-based approaches. The proposal built on earlier observations of macroscopic quantum coherence in superconducting quantum interference devices (SQUIDs) from the 1980s and 1990s. The first experimental realization of coherent dynamics in a flux qubit was achieved in 2003 by I. Chiorescu and colleagues at , who fabricated a niobium-based device and observed Rabi oscillations with coherence times approaching 1 μs, confirming and control in the flux basis. This milestone, using Nb loops and junctions, demonstrated the qubit's potential despite challenges from flux noise. Early spectroscopy efforts further characterized coupled flux qubits, revealing inductive interactions and energy spectra in 2005 experiments that informed multi-qubit designs. By 2006, theoretical work on tunable coupling schemes using intermediary inductors enabled adjustable qubit interactions at the optimal symmetry point, with experimental verification of sign- and magnitude-tunable couplers following in 2007 using rf-SQUID flux qubits. Subsequent evolution focused on enhancing and , with a shift from 2D planar architectures to to minimize decoherence from surface losses and flux trapping. In 2017, researchers at demonstrated flip-chip bonding of flux qubit chips to control layers, achieving improved isolation and times exceeding 10 μs in hybrid modules. Recent variants, such as fluxonium qubits developed initially at Yale in 2009, have pushed beyond 1 ms by 2023 via shunted large inductances, supporting error-corrected operations in small-scale processors. In 2024, a zero-flux-biased flux qubit using ferromagnetic Josephson π-junctions was demonstrated, enabling operation without external tuning and achieving times of about 1.45 μs. These advances underscore the flux qubit's role in hybrid quantum systems, though into large-scale processors remains challenged by noise mitigation.

Theoretical foundations

Superconducting loops and flux quantization

In superconductors, the magnetic flux threading a closed loop is quantized in discrete units known as the flux quantum, \Phi_0 = h / (2e) \approx 2.07 \times 10^{-15} Wb, where h is Planck's constant and e is the . This quantization arises from the coherence of the superconducting order parameter, specifically the phase of the wavefunction, which must be single-valued around the loop, leading to flux values restricted to integer multiples of \Phi_0. In a multiply-connected superconductor, such as a thin-walled cylindrical ring, deviations from this quantization would require breaking the phase coherence, which is energetically unfavorable below the critical temperature. To maintain flux quantization in the presence of an external , persistent supercurrents circulate indefinitely within the superconducting without , screening the interior flux to the nearest multiple of \Phi_0. These currents adjust dynamically to minimize the total energy, resulting in an inductive energy cost for any deviation from the quantized state, given by E_L = (\Phi - n \Phi_0)^2 / (2L), where \Phi is the applied flux, n is an , and L is the . This energy expression reflects the quadratic dependence on flux , analogous to the energy of a in an , and underscores the loop's role in stabilizing macroscopic quantum states. When the external flux is tuned to approximately half a flux quantum (\Phi / \Phi_0 \approx 0.5), the inductive energy landscape forms a symmetric with two degenerate minima corresponding to clockwise and counterclockwise circulating currents of \pm I_p, where I_p = \Phi_0 / (2L). Quantum tunneling between these wells, mediated by the superconducting phase dynamics, allows the system to exist in a superposition of the two states, which forms the basis for information encoding in flux-based quantum devices. The barrier height and well separation are controlled by the loop and , enabling tunable properties. A primary source of decoherence in these loops is low-frequency flux noise, often exhibiting a $1/f spectrum, originating from magnetic moments associated with material defects such as unpaired spins or impurities on the surfaces of the superconducting films. These defects couple to the loop's magnetic field, inducing random flux fluctuations that shift the double-well minima and broaden the energy levels, limiting the achievable superposition times. While the exact microscopic origins remain under investigation, surface preparation techniques have been shown to reduce this noise by minimizing defect density.

Role of Josephson junctions

Josephson junctions serve as the primary nonlinear circuit elements in flux qubits, enabling the required to define well-separated qubit states within the superconducting loop. The governs the behavior of these junctions, permitting a dissipationless supercurrent to flow across a thin insulating barrier between two superconductors without an applied voltage; this supercurrent is given by I = I_c \sin \delta, where I_c is the critical current and \delta is the superconducting phase difference across the junction. This nonlinear current-phase relation contrasts with the linear response of conventional , allowing the junction to function as a tunable, flux-dependent essential for the qubit's . In a typical flux qubit design, three Josephson junctions are integrated into the superconducting loop to balance tunability and nonlinearity. Two of these junctions are usually identical and smaller, forming a dc SQUID that enables control, while the third is larger to ensure the loop's reduced parameter \beta_L = 2\pi L I_c / \Phi_0 > 1, where L is the loop and \Phi_0 is the ; this condition creates a with sufficient to isolate the two lowest energy levels as the states, minimizing leakage to higher excitations. A deliberate , often achieved by a small imbalance in the areas (and thus critical currents) of the two smaller junctions, shifts the optimal to a "sweet spot" at \Phi / \Phi_0 = 0.5, where the qubit energy gap exhibits first-order insensitivity to low-frequency , enhancing . The nonlinear inductance provided by the junctions is captured in the Josephson energy term E_J (1 - \cos \delta), which introduces phase-dependent curvature to the landscape, unlike the of the linear geometric in the loop. This nonlinearity is crucial for the 's operation, as it allows the levels to deviate from behavior, facilitating selective addressing of the qubit . For practical implementation, materials such as aluminum/aluminum oxide/aluminum (Al/AlO_x/Al) are favored for their low-loss tunneling barriers and compatibility with nanoscale fabrication, though niobium/aluminum oxide/niobium (Nb/AlO_x/Nb) junctions are also used to leverage higher critical temperatures and improved uniformity in some designs.

Design and fabrication

Loop geometry and components

The flux qubit features a basic geometry consisting of a closed superconducting loop interrupted by three Josephson junctions connected in series, forming a structure analogous to a direct-current superconducting quantum interference device (dc SQUID). This loop typically has a perimeter on the order of 10-100 μm, enabling the trapping of quanta due to the asymmetry introduced by the junctions, where two are of comparable size and the third is smaller to facilitate states. The design allows for flux quantization within the loop, with the smaller junction often providing the necessary imbalance for stable fluxoid configurations. Key components include the geometric loop inductance L_\text{geo}, derived from the superconducting wire geometry, approximated as L_\text{geo} \approx \frac{\mu_0 l}{2\pi}, where l is the length. Josephson junction areas are typically around 0.1 μm², with variations to achieve the desired critical current asymmetry. On-chip flux bias lines, integrated directly into the , enable precise external threading through the for state control. Design variants encompass both planar and three-dimensional geometries; for instance, configurations using flip-chip bonding have been employed to integrate flux qubits with readout structures while reducing stray . Multi-loop architectures, such as those incorporating additional coupled loops, facilitate tunable inter-qubit strengths by adjusting in shared inductances. As of 2024, zero-flux-biased flux qubits using ferromagnetic Josephson π-junctions have been developed, operating without external flux tuning; these incorporate nitride (NbN) with palladium-nickel (PdNi) for stable π-states. Symmetry considerations in the loop design emphasize an optimal aspect ratio to minimize contributions from kinetic inductance, which arises from the superconductor's finite penetration depth and can degrade performance if dominant over geometric inductance; elongated or square-like ratios are often selected to balance flux sensitivity and inductance uniformity.

Fabrication techniques

Flux qubits are fabricated using thin films of superconducting materials such as aluminum (Al), niobium (Nb), or tantalum (Ta), deposited on insulating substrates like high-resistivity silicon (Si) or sapphire to minimize dielectric losses. The tunnel barriers in Josephson junctions, essential for the qubit's nonlinear inductance, are typically formed by controlled oxidation of an Al layer to create an aluminum oxide (AlO_x) insulator. Electron-beam lithography (EBL) is employed to define sub-micrometer features, enabling precise patterning of the qubit's loop geometry and junctions, while deposition techniques include thermal for Al and sputtering for Nb or Ta films. For Josephson junctions, the Dolan technique—originating in the 1960s but refined in the 2020s for three-dimensional () structures—uses angled shadow to form overlapping superconducting s separated by the oxidized barrier without requiring etching. This method involves a suspended resist that shadows the during sequential evaporations, first at incidence for the bottom , followed by oxidation and a second angled for the top . The fabrication process typically proceeds in key steps: first, the superconducting loop wires are patterned using EBL and lift-off after evaporating the metal film onto a resist-coated substrate; second, the Josephson junctions are formed via the Dolan bridge shadow evaporation with intervening oxidation; third, ground planes and flux bias lines are added through sputtering of Nb or Ta, followed by patterning and lift-off to ensure electrical isolation and shielding. These steps are conducted in a cleanroom environment to avoid contamination, often starting with substrate preparation via piranha cleaning or plasma ashing. Challenges in fabrication include reducing parasitic two-level systems (TLS) that cause decoherence, addressed post-2015 through techniques such as SF_6 or milling to remove surface oxides and contaminants before metal deposition, thereby lowering losses. Advances in integration, including flip-chip bonding of chips to control interposers using or aluminum bumps, have enabled modular architectures while preserving high , with techniques demonstrated in achieving relaxation times exceeding 100 μs in flux-tunable devices. Yield limitations arise from variations in Josephson junction critical currents due to evaporation angle inconsistencies or oxidation nonuniformity, leading to flux offsets that detune the qubit's symmetry point and require individual .

Quantum parameters and Hamiltonian

Energy states and basis

The quantum mechanical description of a flux qubit begins with the quantization of the superconducting , treating the \Phi through the loop and the conjugate Q on the associated as canonical variables. The resulting in the charge-flux representation is H = 4 E_C (n - n_g)^2 - E_J \cos \delta + \frac{(\Phi - \Phi_\text{ext})^2}{2L}, where n = Q / (2e) is the reduced operator, n_g is the dimensionless gate (typically zero in unbiased flux qubits), E_C = e^2 / (2C) is the charging with C, \delta = 2\pi \Phi / \Phi_0 is the phase difference across the Josephson junction with quantum \Phi_0 = h / (2e), E_J is the Josephson , L is the loop , and \Phi_\text{ext} is the externally applied . This form arises from the standard quantization procedure applied to the loop interrupted by Josephson junctions, capturing the inductive , capacitive , and nonlinear Josephson potential. Typical values (e.g., L \approx 10 pH, E_J / h \approx 100 GHz, E_C / h \approx few GHz) date from early designs. For the typical three-junction flux qubit operating near half-flux quantum bias (\Phi_\text{ext} \approx \Phi_0 / 2), the potential term forms a double-well structure due to the interplay of the inductive and Josephson energies, with minima corresponding to and counterclockwise persistent currents \pm I_p. The full many-level spectrum is truncated to the two lowest-energy states to form the , as higher excitations are well-separated. In this two-state approximation, the basis consists of localized current states |L\rangle (left-well, counterclockwise current) and |R\rangle (right-well, current), which are approximate eigenstates far from degeneracy but mix near \Phi_\text{ext} = \Phi_0 / 2. The effective Hamiltonian in the \{|L\rangle, |R\rangle\} basis is H = \frac{\epsilon}{2} \sigma_z + \frac{\Delta_0}{2} \sigma_x, where \epsilon = 2 I_p (\Phi_\text{ext} - \Phi_0 / 2) is the flux-tunable energy bias (longitudinal field term), and \Delta_0 is the tunneling splitting (transverse field term) arising from quantum tunneling between wells. yields symmetric and antisymmetric eigenstates: at the optimal (degeneracy) point \epsilon = 0, the |g\rangle = \frac{1}{\sqrt{2}} (|L\rangle + |R\rangle) and |e\rangle = \frac{1}{\sqrt{2}} (|L\rangle - |R\rangle), with energy splitting \Delta_0; away from optimality, the eigenstates are mixtures, and the full splitting is \Delta \approx \sqrt{\epsilon^2 + \Delta_0^2}. At \epsilon = 0, the simplifies to H = \frac{\Delta_0}{2} \sigma_x, emphasizing the transverse coupling. The spectrum exhibits strong anharmonicity, with higher energy levels spaced by approximately E_J, much larger than the qubit splitting \Delta \ll E_J. This anharmonicity justifies the two-level truncation for qubit operations, as it suppresses leakage to higher states and allows selective microwave addressing of the |g\rangle-|e\rangle (or |0\rangle-|1\rangle) transition without exciting subsequent levels; brief consideration of these higher modes confirms the validity of the approximation but is not required for the qubit subspace.

Key parameters and coherence

Flux qubits are characterized by several key parameters that determine their quantum behavior and performance. The loop inductance L is typically around 10 , the Josephson energy E_J is approximately 100 GHz (in units where \hbar = 1), and the charging E_C is on the order of a few GHz. These values enable the qubit's states and tunability via external . The flux sensitivity, quantified as df/d(\Phi/\Phi_0) (where f is the transition ), is approximately 1 GHz per milliflux quantum (or ~1 THz per \Phi_0) for typical s of 0.1–0.5 μA, reflecting the landscape's response to magnetic variations away from the degeneracy point. The relaxation time T_1 is primarily limited by flux noise, with a S_\Phi \approx 10^{-6} \Phi_0 / \sqrt{\mathrm{Hz}} at 1 Hz, arising from -induced fluctuations and 1/f-type environmental coupling. In optimized designs as of 2016–2023, T_1 reaches 20–60 μs, approaching the limit set by losses and junction imperfections. These improvements stem from refined fabrication on low-loss substrates like and reduced density. As of 2024, zero-flux-biased variants using π-junctions achieve T_1 \approx 1.45 μs without external flux tuning. Dephasing times T_2 are enhanced through echo techniques, achieving up to 100 μs in echo experiments, particularly at the flux sweet spot where first-order flux sensitivity vanishes. Operation at this degeneracy point minimizes sensitivity to 1/f flux noise, which otherwise dominates pure . Dynamical sequences further extend by refocusing low-frequency , pushing T_2 toward the $2T_1 limit. Single-qubit gate fidelities for flux-tunable operations reach ~99.9% using short flux pulses, limited by pulse rise times and residual flux crosstalk. Variability in these fidelities arises from Josephson junction critical current spreads of ±5%, which affect the symmetry parameter \alpha and thus the energy gap tunability. Precise control of junction fabrication uniformity is essential to mitigate this. Recent advancements in 2025 incorporate tantalum-based materials for fluxonium qubits (a flux qubit variant), yielding dephasing times T_2 \approx 300 μs (Ramsey) at optimal bias when combined with dynamic decoupling protocols like CPMG sequences. These Ta-based designs reduce surface losses and 1/f noise, enabling longer coherence for scalable quantum circuits.

Coupling and control

Inter-qubit coupling

Inter-qubit coupling in flux qubits is essential for implementing multi-qubit operations in quantum computing architectures, primarily achieved through inductive or galvanic mechanisms that leverage the qubits' superconducting loop structures. The most common approach is direct inductive coupling, where two flux qubits share a mutual inductance M between their loops. This results in an interaction Hamiltonian H_{\text{int}} = J \sigma_z^i \sigma_z^j, where \sigma_z are Pauli operators in the persistent current basis, and the coupling strength J = M I_p^2 (with I_p the persistent current amplitude). This ZZ-type interaction arises from the magnetic flux threading the loops and enables controlled-phase gates when qubits are detuned. To enable switchable interactions and reduce always-on crosstalk, tunable couplers are integrated between qubits, typically using an additional flux-tunable superconducting quantum interference device (SQUID). This coupler, often a three-junction loop biased by an external flux, mediates the inductive coupling and allows on/off control by tuning the coupler's effective inductance from ferromagnetic to antiferromagnetic states or zero. The coupling rate g / 2\pi can reach up to 100 MHz, providing high on/off ratios (e.g., >10:1) while minimizing unwanted residual interactions. Such designs have been experimentally realized, demonstrating tunable J values spanning tens of mK (corresponding to ~200 MHz) with precise flux control. In contrast, fixed galvanic coupling connects directly via shared Josephson junctions, offering stronger interactions (up to GHz scales) but lacking tunability, which limits flexibility in scalable arrays. This method uses large shared junctions to couple the phase variables across qubit loops, yielding strong coupling through shared phase variables, with effective interaction strengths up to the GHz scale. Galvanic links are less prone to but introduce fabrication challenges for multi-qubit connectivity compared to inductive schemes. Scalability of flux qubit arrays is hindered by crosstalk from stray magnetic fields and mutual inductances, which can induce unwanted terms degrading fidelities. Recent experiments with flux-tunable superconducting arrays (incorporating flux qubit principles) have demonstrated in 16-qubit systems with errors calibrated to <1% using machine learning protocols, paving the way for larger scales. In 2023 demonstrations, arrays approaching 50 qubits in related flux-biased architectures achieved similar low-error through optimized shielding and coupler isolation. In 2024, a tunable inductive coupler for heavy fluxonium qubits demonstrated two-qubit fidelities exceeding 99.9% with strengths around 50 MHz. Additionally, as of 2025, strong between flux qubits and single bismuth donors in silicon has been achieved, enabling coherent transfer of quantum information. Multi-qubit gates, such as iSWAP, are realized by detuning the qubits to resonate the ZZ interaction, enabling photon swapping with exceeding 99% in recent flux-based systems. For instance, strong flux modulation in galvanically coupled setups has produced iSWAP gates with >99.9% , highlighting the potential for high-performance entangling operations. These advances rely on the intrinsic parameters like \Delta and I_p from single-qubit designs to optimize interaction strengths without excessive decoherence.

External control mechanisms

Flux qubits are manipulated externally primarily through precise tuning of the threading the superconducting loop, denoted as \Phi_{\text{ext}}, using on-chip current bias lines that generate localized magnetic fields. These lines, typically superconducting wires inductively coupled to the qubit loop, allow for continuous adjustment of the qubit's , such as biasing at the optimal symmetry point where the energy levels exhibit minimal first-order sensitivity to flux fluctuations. Independent control is achieved via dedicated flux lines for each qubit, enabling scalability in multi-qubit arrays while minimizing . The of these flux bias lines supports rapid tuning up to approximately 1 GHz, facilitated by arbitrary generators (AWGs) that deliver high-fidelity pulses at cryogenic temperatures. This capability is essential for dynamic adjustment of the frequency, which varies from a few GHz at the flux-insensitive point to higher values away from it, allowing operations like adiabatic passage or fast sweeps to avoid noise-sensitive regions. In contrast to inter-qubit schemes that rely on shared flux interactions, single-qubit flux biasing isolates control to individual devices. Microwave drives for flux qubits involve applying RF flux pulses through the same or auxiliary on-chip lines to induce coherent Rabi oscillations between the qubit states. These pulses, resonant with the qubit transition frequency (typically 5-10 GHz), rotate the qubit state on the , with \pi-pulse durations around 20 ns achieving full state flips for operations. The drive amplitude determines the , enabling precise control while higher-power drives can be used to probe noise spectra via oscillation decay. Fast fluxing techniques extend this control to sub-nanosecond timescales, employing high-speed flux lines integrated with low-noise elements like Josephson parametric amplifiers to execute qubit gates with minimal decoherence. Such rapid switching is critical for high-fidelity operations in scalable processors, where flux pulses modulate the Hamiltonian transiently to implement single-qubit rotations or prepare specific states. Feedback control mechanisms enhance stability by adaptively nulling low-frequency flux drifts, particularly the 1/f that limits . Closed-loop systems monitor the qubit frequency in via dispersive readout and adjust the DC bias to maintain the optimal , effectively suppressing in cryogenic environments. Recent implementations in multi-qubit setups demonstrate improved T_2 times by countering environmental flux variations dynamically. To ensure isolated local control, flux qubits incorporate shielding against global magnetic fields, such as enclosures or on-chip gradiometric loop geometries that cancel common-mode noise. These protections reduce sensitivity to ambient fields from components or neighboring qubits, allowing precise local biasing without unintended state perturbations. Designs biased at half- quantum points further enhance first-order insensitivity to both global and local fluctuations.

Readout techniques

Inductive detection

Inductive detection in flux qubits exploits the distinct signatures of the qubit's quantum states, where the persistent currents circulating in the superconducting produce a flux difference of approximately \Phi_0/2 between the clockwise and counterclockwise states. This state-dependent flux is inductively coupled to a pickup or, more typically, a DC superconducting quantum interference device () for measurement. The coupling allows the qubit's flux to modulate the properties of the readout device without direct electrical contact, enabling non-destructive readout in principle. The DC SQUID readout scheme employs a loop containing two Josephson junctions, biased at an external flux of \Phi/\Phi_0 = 0.5 to maximize sensitivity to small flux perturbations. In the switching current mode, the qubit-induced flux shifts the SQUID's critical current, which is detected as a change in the bias current required to induce a voltage across the SQUID. Alternatively, in the inductive mode, the qubit flux alters the SQUID's effective Josephson inductance, measured via an applied AC signal. The measurement circuit integrates the first-stage SQUID directly on-chip with the qubit, coupled through mutual inductance, and employs a flux-locked loop with cryogenic amplification to enhance signal gain while suppressing noise. Dispersive readout protocols, where the qubit is detuned from its optimal point to enhance flux contrast, have been demonstrated for flux qubits. These performance metrics are constrained by back-action effects, in which the readout process can induce qubit transitions via photon exchange or flux noise, limiting the overall efficiency. Noise in the system arises primarily from the amplifier's quantum limit of approximately hf/2, where h is Planck's constant and f the operating frequency, alongside thermal contributions that are minimized through cryogenic operation at temperatures around 20 mK.

Advanced readout methods

Advanced readout methods for flux qubits extend beyond direct inductive detection by incorporating quantum-limited amplification and resonator-based schemes to achieve higher fidelity and scalability. One prominent approach is parametric amplification using Josephson parametric converters, which enable readout with minimal added noise. These devices, such as flux-driven Josephson parametric amplifiers (JPAs), couple to the qubit's flux signal and provide quantum-limited detection, adding less than 1 of noise, as demonstrated in implementations for superconducting qubits including flux types. Dispersive readout represents another key technique, where the flux qubit is coupled to a high-quality-factor —typically via capacitive or inductive means—resulting in a state-dependent shift in the 's . This shift, known as the dispersive \chi \approx g^2 / \Delta, where g is the qubit-resonator strength and \Delta is the detuning, allows the qubit state to be inferred from the 's or without strongly disturbing the qubit. Such methods support multiplexed readout in qubit arrays by assigning distinct frequencies to each qubit, enabling parallel state discrimination with reduced wiring complexity. Quantum non-demolition (QND) measurements further enhance readout by allowing repeated observations of the state without collapsing it, achieved through continuous weak monitoring. In this scheme, a weak drive probes the 's persistent current via a coupled , such as a Josephson , suppressing relaxation and back-action to realize ideal QND conditions across variable bias points. This approach has been theoretically and experimentally validated for qubits, providing high repeatability exceeding 99%. Recent innovations (as of 2025) include all-optical readout techniques that convert signals to the optical domain for detection, leveraging superconducting single-photon detectors (SNSPDs) to achieve faster timescales and reduced . These methods, demonstrated in superconducting systems, support integration with by enabling multiplexed, low-interference measurements. Compared to simple SQUID-based inductive readouts, these advanced techniques offer higher fidelities but introduce greater circuit complexity for .

References

  1. [1]
    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 ...Single-qubit gates · The CPHASE two-qubit gate in... · Two-qubit gates using only...
  2. [2]
    [PDF] Superconducting Qubits and the Physics of Josephson Junctions
    We review here three different ways that these nonlinear resonators can be made, and which are named as phase, flux, or charge qubits. The circuit for the phase ...
  3. [3]
    Low-decoherence flux qubit | Phys. Rev. B
    Apr 27, 2007 · A flux qubit can have a relatively long decoherence time at the degeneracy point, but away from this point the decoherence time is greatly ...
  4. [4]
    Fluxonium: Single Cooper-Pair Circuit Free of Charge Offsets
    A clever piece of quantum circuit engineering that can suppress the effect of the quantum noise and allow the quantum circuit to operate without disturbance.
  5. [5]
    Superconducting flux qubit with ferromagnetic Josephson π-junction ...
    Oct 11, 2024 · We report the realization of a zero-flux-biased flux qubit based on three NbN/AlN/NbN Josephson junctions and a NbN/PdNi/NbN ferromagnetic π-junction.
  6. [6]
    Scalable interconnection using a superconducting flux qubit - Nature
    Jul 16, 2024 · We report an implementation technology to overcome the scaling bottlenecks using a reliable connection qubit with a demonstration of quantum annealing.
  7. [7]
    [cond-mat/9908283] A Superconducting Persistent Current Qubit
    Aug 19, 1999 · Abstract: We present the design of a superconducting qubit that has circulating currents of opposite sign as its two states.
  8. [8]
  9. [9]
    A perspective on superconducting flux qubits - AIP Publishing
    Aug 23, 2021 · In this paper, we discuss the basic properties of flux qubits using the radio frequency superconducting quantum interference device geometry.IV. CHOICE OF QUBIT... · THREE- AND FOUR... · C-shunted flux qubit · Fluxonium
  10. [10]
    Superconducting persistent-current qubit | Phys. Rev. B
    Dec 1, 1999 · We present the design of a superconducting qubit that has circulating currents of opposite sign as its two states.
  11. [11]
    Coherent Quantum Dynamics of a Superconducting Flux Qubit
    Abstract. We have observed coherent time evolution between two quantum states of a superconducting flux qubit comprising three Josephson junctions in a loop.Missing: Chiorescu | Show results with:Chiorescu
  12. [12]
    Spectroscopy on Two Coupled Superconducting Flux Qubits
    Mar 9, 2005 · We have performed spectroscopy measurements on two coupled flux qubits. The qubits are coupled inductively, which results in a ...
  13. [13]
    Tunable coupling scheme for flux qubits at the optimal point
    Mar 15, 2006 · We discuss a practical design for tunably coupling a pair of flux qubits via the quantum inductance of a third high-frequency qubit.
  14. [14]
    and Magnitude-Tunable Coupler for Superconducting Flux Qubits
    Apr 23, 2007 · In this Letter we demonstrate such a coupler between two superconducting flux qubits. The qubit design used is a bistable rf SQUID, magnetically ...
  15. [15]
    3D integrated superconducting qubits | npj Quantum Information
    Oct 9, 2017 · In this work, we use a flip-chip process to bond a chip with superconducting flux qubits to another chip containing structures for qubit readout and control.
  16. [16]
    Signal Crosstalk in a Flip-Chip Quantum Processor
    Our study introduces an on-chip signal-delivery architecture in a flip-chip multiqubit processor with competitive crosstalk performance across different ...
  17. [17]
    Theoretical Considerations Concerning Quantized Magnetic Flux in ...
    Theoretical considerations concerning quantized magnetic flux in superconducting cylinders. N. Byers and CN Yang. Institute of Theoretical Physics.
  18. [18]
    Fluxoid Quantization in a Multiply-Connected Superconductor
    This is a consequence of "fluxoid" quantization which was originally predicted by Fritz London. The basic unit of magnetic flux h c 2 ⁢ 𝑒 is explained in terms ...
  19. [19]
    [cond-mat/0207277] Persistent current in superconducting nanorings
    Jul 11, 2002 · We reduce the problem of low-energy properties of a superconducting nanoring to that of a quantum particle in a sinusoidal potential.
  20. [20]
    Double-well potentials in current qubits
    Oct 7, 2003 · The effective potentials of the rf-SQUID and three-Josephson junction loop with a penetrating external magnetic flux are studied.
  21. [21]
    Origin and Reduction of Magnetic Flux Noise in Superconducting ...
    Oct 18, 2016 · Recent work indicates that the noise is from unpaired magnetic defects on the surfaces of the superconducting devices.
  22. [22]
    The flux qubit revisited to enhance coherence and reproducibility - NIH
    Nov 3, 2016 · We revisit the design and fabrication of the superconducting flux qubit, achieving a planar device with broad-frequency tunability, strong anharmonicity, high ...
  23. [23]
    [PDF] Quantum Bits with Josephson Junctions - Franco Nori
    It is also known as a persistent-current qubit. The flux qubit in its simplest form consists of a superconducting loop interrupted by one Josephson junction.<|control11|><|separator|>
  24. [24]
    [PDF] Fabrication and measurements of hybrid Nb/Al Josephson junctions ...
    May 2, 2014 · While niobium serves as the base material for most of conventional superconducting circuits employing Nb/AlOx/Nb Josephson junctions, quantum ...
  25. [25]
    Quantum theory of three-junction flux qubit with non-negligible loop ...
    May 30, 2006 · The three-junction flux qubit (quantum bit) consists of three Josephson junctions connected in series on a superconducting loop.Article Text · INTRODUCTION · THREE-DIMENSIONAL... · SOLUTION TO THE...
  26. [26]
    (a) Experimental setup. The superconducting flux qubit with a size of...
    The superconducting flux qubit with a size of 10.4 µm × 10.2 µm is surrounded by a SQUID loop connected to a readout line. Nano-diamond particles with a ...Missing: perimeter | Show results with:perimeter
  27. [27]
    [PDF] arXiv:1608.02363v4 [quant-ph] 27 Apr 2018
    Apr 27, 2018 · For example, one may use the inductance formula for coaxial cables if the magnetic flux lines have the tight tube-like shape. Such a shape can ...
  28. [28]
    Improving Josephson junction reproducibility for superconducting ...
    Apr 25, 2023 · ... qubit fabrication, 2.5–6.3% normal resistance variation was reported for 0.01–0.16 μm2 junction area on a 76-mm wafer. In another study ...
  29. [29]
    Flux qubits and readout device with two independent flux lines
    Two on-chip flux bias lines allow independent flux control of any two of the three elements, as illustrated by a two-dimensional qubit flux map. The ...
  30. [30]
    Superconducting `twin' qubit | Phys. Rev. B
    Sep 21, 2020 · We study a flux qubit consisting of a symmetrical pair of superconducting loops, with two Josephson junctions in each, joined by a common ...<|control11|><|separator|>
  31. [31]
    Characterizing and Optimizing Qubit Coherence Based on SQUID ...
    May 29, 2020 · By investigating SQUID loops of varying aspect ratio X / Y , we are able to confirm the linear scaling of A Φ with SQUID perimeter rather than ...
  32. [32]
    [PDF] Material matters in superconducting qubits - arXiv
    Interactions between qubits and their environment, including fabrication and composition, are key. Dielectric loss and two-level systems are also important ...
  33. [33]
    Materials challenges and opportunities for quantum computing ...
    Apr 16, 2021 · We identify key materials challenges that currently limit progress in five quantum computing hardware platforms, propose how to tackle these problems,
  34. [34]
    Simplified Josephson-junction fabrication process for reproducibly ...
    For each qubit, the junction area is based on the design parameters, while ... Junction area (μm2) . f01 (GHz) . T1 (μs) . T 2 * (μs) . Q (106) . S ...
  35. [35]
    Nanoscale direct-write fabrication of superconducting devices for ...
    Mar 10, 2023 · Electron beam lithography is a very popular fabrication technique in quantum technologies and has been applied to fabricate Josephson persistent ...
  36. [36]
    Wafer-scale uniformity improvement of Dolan-bridge Josephson ...
    Aug 1, 2025 · But Dolan bridge technique does not require large evaporation angles, unlike Manhattan junctions, which leads to improved line edge roughness.Missing: flux | Show results with:flux
  37. [37]
    [PDF] arXiv:1101.4576v2 [cond-mat.mes-hall] 2 Feb 2011
    Feb 2, 2011 · We present a novel shadow evaporation technique for the realization of junctions and capacitors. The design by E-beam lithography of ...
  38. [38]
    Stress accommodation in nanoscale dolan bridges designed for ...
    (b) Fabrication process flow for JJs using the shadow evaporation technique. The dark arrows illustrate the angle of the metal evaporation direction with ...
  39. [39]
    [PDF] 1 High Density Fabrication Process for Single Flux Quantum Circuits
    The bumps are fabricated in two steps. First, depending on a qubit design, 5 or 10 μm high aluminum or copper bump cores are deposited by evaporation. Then ...
  40. [40]
    Coherent superconducting qubits from a subtractive junction ...
    Typically, the junctions for qubits are fabricated using shadow evaporation techniques to reduce dielectric loss contributions from the superconducting film ...
  41. [41]
    High-coherence fluxonium qubits manufactured with a wafer-scale ...
    This process eliminates the need for angle evaporation and enables almost 100% qubit yield.
  42. [42]
    Reducing intrinsic loss in superconducting resonators by surface ...
    May 4, 2015 · We use two techniques to reduce losses associated with two-level systems: an additional substrate surface treatment prior to NbTiN deposition ...
  43. [43]
    Effects of surface treatments on flux tunable transmon qubits - Nature
    Oct 29, 2021 · UV light and ion milling can be used to remove unwanted surface contamination and in the case of ion milling even to remove thin surface oxide ...
  44. [44]
    [2303.01481] Fluxonium Qubits in a Flip-Chip Package - arXiv
    Mar 2, 2023 · In this paper, we report work on fluxonium qubits packaged in a flip-chip architecture, where a classical control and readout chip is bump-bonded to the ...Missing: >100 μs
  45. [45]
    [PDF] Signal Crosstalk in a Flip-Chip Quantum Processor
    Sep 12, 2024 · Our flip-chip fabrication process has been demonstrated to be compatible with qubit (transmon) coherence at the. 100-µs level [20]. The ...
  46. [46]
    Improving wafer-scale Josephson junction resistance variation in ...
    Quantum bits, or qubits, are an example of coherent circuits envisioned for next-generation computers and detectors. A robust superconducting qubit with a ...
  47. [47]
    Hamiltonian of a flux qubit-LC oscillator circuit in the deep–strong ...
    Apr 26, 2022 · The derived circuit Hamiltonian consists of terms associated with the LC oscillator, the flux qubit (and its higher energy levels), and the ...
  48. [48]
    The flux qubit revisited to enhance coherence and reproducibility
    Nov 3, 2016 · Here, we revisit the design and fabrication of the superconducting flux qubit, achieving a planar device with broad-frequency tunability, strong ...
  49. [49]
    [PDF] Energy Relaxation, Charge and Flux Noise - arXiv
    associated flux noise is shown to be superOhmic. We discuss how this quasiparticle flux noise can limit T∗. 2 coherence times in flux-tunable qubits. The ...
  50. [50]
    Tunable Superconducting Flux Qubits with Long Coherence Times
    Feb 24, 2023 · In this work, we study a series of tunable flux qubits inductively coupled to a coplanar waveguide resonator fabricated on a sapphire substrate.Abstract · Article Text · INTRODUCTION · TUNABILITY OF THE QUBITS
  51. [51]
    -Gate Operation on a Superconducting Flux Qubit via its Readout ...
    Mar 11, 2015 · In this work, we implement a Z gate by pulsing a current through the qubit's readout dc SQUID. While the dc SQUID acts as a magnetic flux sensor for qubit ...Missing: spread | Show results with:spread
  52. [52]
    Improving Josephson junction reproducibility for superconducting ...
    Apr 25, 2023 · Niemeyer–Dolan bridge technique used in this paper. (a) Scheme of the used shadow evaporation process. The first metal layer was deposited ...
  53. [53]
    [PDF] Master in Physics of Complex Systems Design of Novel Coupling ...
    Oct 25, 2021 · We have analyzed and proposed coupling mechanisms between Three Josephson Junction. Flux Qubits (3JJQ). For this, we have developed a ...
  54. [54]
    [PDF] Superconducting flux qubits for high-connectivity quantum ...
    This dissertation is about superconducting flux qubits for high-connectivity quantum annealing without lossy dielectrics, submitted for a Ph.D. in Physics.
  55. [55]
    None
    ### Summary of Tunable Coupler Using Additional Flux-Tunable SQUID Between Flux Qubits
  56. [56]
    Compound Josephson-junction coupler for flux qubits with minimal ...
    Aug 20, 2009 · Custom tuned qubit CJJ flux offsets provide one means of mitigating this undesirable effect. Alternate qubit designs which contain an in situ ...
  57. [57]
    Mediated tunable coupling of flux qubits - IOP Science
    Nov 7, 2005 · It is sketched how a monostable rf- or dc-SQUID can mediate an inductive coupling between two adjacent flux qubits.
  58. [58]
    Galvanic coupling of flux qubits: simple theory and tunability - arXiv
    May 16, 2006 · Galvanic coupling of small-area (three-junction) flux qubits, using shared large Josephson junctions, has been shown to yield appreciable interaction strengths.
  59. [59]
    Galvanic Phase Coupling of Superconducting Flux Qubits - MDPI
    In this study we derive the exact inductive coupling strength between two flux qubits coupled directly and coupled through a connecting central loop.
  60. [60]
    Learning-Based Calibration of Flux Crosstalk in Transmon Qubit ...
    Aug 28, 2023 · In this work, we introduce a learning-based calibration protocol and demonstrate its experimental performance by calibrating an array of 16 flux-tunable ...Abstract · Article Text · INTRODUCTION · PROTOCOL SCALING
  61. [61]
    [PDF] arXiv:2310.16159v1 [cond-mat.mes-hall] 24 Oct 2023
    Oct 24, 2023 · A major current challenge in solid-state quantum computing is to scale qubit arrays to a larger number of qubits.
  62. [62]
    Fast High-Fidelity Gates for Galvanically-Coupled Fluxonium Qubits ...
    Dec 27, 2022 · We propose a scheme for coupling multiple fluxonium qubits that has two distinct features: first, the coupling is tunable and can explicitly be turned off.
  63. [63]
    DC flux crosstalk reduction with dual flux line - AIP Publishing
    Jun 18, 2024 · We observe a significant reduction in the DC flux crosstalk when utilizing the dual flux line, as supported by electromagnetic field simulations.
  64. [64]
    Flux qubit noise spectroscopy using Rabi oscillations under strong ...
    Jan 9, 2014 · We infer the high-frequency flux noise spectrum in a superconducting flux qubit by studying the decay of Rabi oscillations under strong driving conditions.
  65. [65]
    Parametric Control of a Superconducting Flux Qubit | Phys. Rev. Lett.
    Mar 13, 2006 · Parametric control of a superconducting flux qubit has been achieved by using two-frequency microwave pulses. We have observed Rabi ...
  66. [66]
    Improving qubit coherence using closed-loop feedback - PMC - NIH
    We demonstrate that using the feedback protocol to suppress low-frequency noise enables gate fidelities exceeding 99.9% even far away from the flux sweet spot, ...Missing: T2 | Show results with:T2
  67. [67]
    Evolution of 1 / f Flux Noise in Superconducting Qubits with Weak ...
    Recent progress in superconducting devices for quantum information has highlighted the need to mitigate sources of qubit decoherence, driving a renewed interest ...Missing: 2D 3D
  68. [68]