Fact-checked by Grok 2 weeks ago

Josephson effect

The Josephson effect is a quantum mechanical phenomenon involving the tunneling of pairs across a thin insulating barrier between two superconductors, enabling a supercurrent to flow without any applied voltage. Predicted theoretically by Brian D. Josephson in 1962 while he was a graduate student at the , the effect manifests in two primary forms: the DC Josephson effect, where a direct supercurrent I = I_c \sin \phi flows across the junction with critical current I_c and superconducting phase difference \phi, and the AC Josephson effect, where an applied DC voltage V causes the current to oscillate at a f = 2eV / h, with e the and h Planck's constant. This prediction extended the Bardeen-Cooper-Schrieffer ( of by incorporating phase coherence across the barrier. Experimental verification of the DC effect came swiftly in 1962 by and John M. Rowell at Bell Laboratories, who observed supercurrents in tin-tin oxide-tin junctions at low temperatures, confirming Josephson's theoretical relations. The AC effect was demonstrated in 1963 by and Karl Megerle, who applied a small voltage and detected radiation at the predicted frequency, providing direct evidence of the phase dynamics in superconducting tunneling. These discoveries earned Josephson the in 1973, shared with and for their work on tunneling phenomena in semiconductors and superconductors. The Josephson effect has profound implications for both fundamental physics and technology, demonstrating macroscopic quantum and enabling precise measurements. In applications, Josephson junctions form the basis of superconducting quantum devices (SQUIDs), which achieve unprecedented in detecting , with flux resolutions better than $10^{-6} \Phi_0 (where \Phi_0 \approx 2 \times 10^{-15} is the ), used in , , and . Additionally, arrays of Josephson junctions serve as primary voltage standards, generating stable voltages in steps of n h f / 2e ( steps) under irradiation, ensuring metrological accuracy traceable to fundamental constants and supporting the international volt definition. More recently, Josephson junctions have become integral to , where their nonlinear realizes superconducting qubits with times exceeding microseconds (as of 2025), advancing scalable processing.

Background and History

Superconducting Junctions

is a quantum mechanical phenomenon observed in certain materials at very low temperatures, characterized by zero electrical resistance and the expulsion of magnetic fields from the interior of the material, known as the . This state was first discovered in 1911 by while studying mercury at temperatures. In superconductors, electrons form bound pairs called Cooper pairs, which carry charge -2e and behave as composite bosons due to their even spin and effective integer spin statistics. These pairs condense into a single macroscopic , enabling macroscopic quantum coherence where the entire material exhibits wave-like behavior as a coherent entity. A Josephson junction serves as a weak link between two superconducting regions, allowing the quantum tunneling of pairs across the link while maintaining the overall superconducting properties. The most common configuration is the tunnel junction, consisting of two superconductors separated by a thin insulating barrier, typically 1-2 nm thick, such as aluminum oxide (AlOx). This barrier prevents classical flow between the superconductors but permits phase-coherent quantum tunneling of the bosonic pairs, coupling the macroscopic wavefunctions of the two sides. Other types of Josephson junctions include weak links, such as point contacts where a narrow constriction or sharpened tip connects two bulk superconductors, and superconductor-normal metal-superconductor () junctions featuring a thin normal metal interlayer instead of an . Fabrication methods vary by type but often involve thin-film techniques; for example, tunnel junctions are commonly made by depositing a superconducting film like aluminum, exposing it to controlled oxidation to form the barrier, and then depositing the second superconducting layer, typically using electron-beam or in a environment. These structures are essential for realizing weak-link behavior in superconducting devices.

Prediction and Discovery

In 1962, , a graduate student at the , theoretically predicted the existence of a supercurrent tunneling through an insulating barrier between two superconductors, provided the barrier was thinner than the superconducting coherence length. This prediction extended the Bardeen-Cooper-Schrieffer ( of to weakly coupled superconducting systems, emphasizing the role of quantum phase coherence across the barrier. Josephson specifically forecasted that the critical current I_c of this supercurrent would be proportional to the product of the superconducting energy gaps \Delta_1 and \Delta_2 on either side of the junction. Josephson's proposal initially faced significant skepticism within the physics community, particularly from Nobel laureate , who questioned the validity of applying quantum tunneling to macroscopic superconducting pairs and argued that the predicted zero-voltage current contradicted established principles. This debate highlighted broader theoretical uncertainties about macroscopic , including whether phase coherence could be maintained over distances spanning an insulating barrier in superconducting systems. Despite the controversy, the counterintuitive nature of a dissipationless current at zero voltage spurred rapid experimental efforts to test the predictions. The first experimental verification came in 1963 from and John M. Rowell at Bell Laboratories, who observed an anomalous zero-voltage supercurrent in tin-tin oxide-tin junctions, along with constant-voltage steps induced by microwave radiation, confirming both the DC and AC aspects of the effect. Subsequent measurements in lead-based junctions achieved critical currents up to several microamperes, aligning closely with Josephson's theoretical estimates. These findings resolved the initial doubts, establishing the Josephson effect as a cornerstone of superconducting physics. In recognition of this breakthrough, Josephson shared the 1973 with and for their collective contributions to tunneling phenomena in solids.

Fundamental Equations

The First Josephson Relation

The first Josephson relation establishes the time-independent relationship between the supercurrent and the phase difference across a superconducting tunnel junction, arising from the quantum tunneling of Cooper pairs through an insulating barrier separating two superconductors. In Brian Josephson's seminal derivation, the superconducting state on each side of the barrier is described by a macroscopic wavefunction \psi = |\psi| e^{i\phi}, where \phi is the phase of the order parameter. The tunneling of Cooper pairs, which carry charge $2e, leads to a supercurrent due to the coherent overlap of these wavefunctions, with the amplitude determined by the phase difference \phi = \phi_L - \phi_R + (2e/\hbar) \int_L^R \mathbf{A} \cdot d\mathbf{l}, the gauge-invariant phase shift incorporating vector potential effects. This relation is expressed as I = I_c \sin \phi, where I is the supercurrent, I_c is the critical current (the maximum supercurrent occurring at \phi = \pi/2), and \phi is the gauge-invariant phase difference. The equation predicts a dissipationless supercurrent up to the critical value, beyond which the junction switches to a voltage state. Physically, \phi quantifies the relative quantum mechanical phase shift between the superconducting order parameters, driving the directional flow of Cooper pairs; the \sin \phi form ensures the current's $2\pi periodicity, consistent with the single-valued nature of the wavefunction under phase transformations. This periodicity arises because a phase change of $2\pi corresponds to an integer number of Cooper pairs tunneling, preserving coherence. The critical current I_c is given by the Ambegaokar-Baratoff relation, I_c = \frac{\pi \Delta}{2 e R_N} \tanh\left(\frac{\Delta}{2 k_B T}\right), where \Delta is the superconducting energy gap, R_N is the normal-state resistance of the junction, e is the , k_B is Boltzmann's constant, and T is the ; at T = 0, this simplifies to I_c = \pi \Delta / (2 e R_N). Thus, I_c depends on the barrier thickness via R_N, as thinner barriers enhance tunneling probability and reduce R_N, increasing I_c. Material properties influence \Delta; for instance, niobium-based junctions achieve higher I_c than aluminum-based ones due to niobium's larger \Delta stemming from its higher critical temperature (T_c \approx 9.2 K versus $1.2 K for aluminum). Near T_c, \Delta vanishes, causing I_c to decrease to zero. This static current-phase relation underpins the dynamics of \phi, where time evolution incorporates dissipation and quantum effects, as described in the Caldeira-Leggett framework for macroscopic quantum tunneling of the phase in underdamped .

The Second Josephson Relation

The second Josephson relation describes the dynamic evolution of the superconducting phase difference across a Josephson under an applied voltage, originating from the quantum mechanical treatment of the coupled wavefunctions in the adjacent superconductors. This relation emerges from applying the time-dependent to the macroscopic wavefunctions of the superconductors, treating the junction as a tunneling barrier that couples the two sides. The voltage introduces an energy shift between the superconducting condensates, driving a relative phase rotation. The core equation of the second relation is \frac{d\phi}{dt} = \frac{2eV}{\hbar}, where \phi is the phase difference between the wavefunctions, V is the voltage across the junction, e is the elementary charge, and \hbar is the reduced Planck's constant. This follows directly from the phase dynamics in the Schrödinger equation, where the voltage V corresponds to an electrochemical potential difference that accelerates the relative phase accumulation at a rate proportional to $2eV/\hbar. Physically, the equation implies that a constant voltage V results in a linear time dependence of [\phi](/page/Phi), \phi(t) = \phi_0 + (2eV/[\hbar](/page/H-bar)) t, representing a continuous "phase slippage" between the superconductors. At zero voltage (V = 0), d\phi/dt = 0, so [\phi](/page/Phi) remains constant, enabling steady supercurrent flow without dissipation. For finite V, this phase evolution underpins frequency-dependent phenomena by linking electrical and temporal domains. A key consequence is the Josephson frequency f_J, defined as the rate of phase cycles divided by $2\pi, f_J = \frac{2eV}{h} \approx 483.6 \, \mathrm{MHz}/\mu\mathrm{V}, where h is Planck's ; this universal precisely converts voltage to and is exploited in for voltage standards traceable to fundamental constants. The relation's validity extends to environments with external fields through the use of the gauge-invariant difference, \phi = \phi_2 - \phi_1 - \frac{2e}{\hbar} \int_1^2 \mathbf{A} \cdot d\mathbf{l}, where \phi_1 and \phi_2 are the phases in the respective superconductors, and the line integral of the vector potential \mathbf{A} from side 1 to 2 ensures invariance under electromagnetic gauge transformations, preserving the form of the equation even in nonzero magnetic fields.

Key Effects

DC Josephson Effect

The DC Josephson effect manifests as the flow of a dissipationless supercurrent across a Josephson junction when no voltage is applied, with the current magnitude limited by a critical value I_c determined by the junction properties. This supercurrent arises from the coherent quantum tunneling of Cooper pairs through the insulating barrier, enabled by a fixed phase difference between the superconducting wave functions on either side of the junction. The relation between the current and the superconducting phase difference \phi underscores this coherence, where the current is proportional to \sin\phi for bias currents below I_c. Observation of the DC Josephson effect is characterized by zero-voltage transport in the current-voltage (I-V) curve for applied currents up to I_c, often accompanied by in underdamped junctions due to the barrier associated with slippage. The dissipationless nature of this transport is confirmed experimentally by the lack of , as no power is dissipated at zero voltage, distinguishing it from normal resistive flow. The effect requires coherent locking across the junction, with \phi remaining constant under bias currents less than I_c; exceeding I_c causes the to evolve dynamically, initiating a finite voltage. First experimentally verified in late 1962 (published 1963) by P. W. Anderson and J. M. Rowell using junctions with tin oxide barriers between superconducting tin and lead films (Sn-SnO-Pb), these early measurements demonstrated supercurrents with critical current densities up to approximately $10^5 A/cm² at low temperatures. This phenomenon serves as direct evidence of quantum superposition in macroscopic systems, as the sustained supercurrent reflects the collective wave function of billions of Cooper pairs maintaining a definite phase relation across the barrier, embodying macroscopic quantum coherence.

AC Josephson Effect

When a constant voltage V is applied across a Josephson junction, the phase difference \phi between the superconducting wave functions on either side evolves linearly with time according to the second Josephson relation: \frac{d\phi}{dt} = \frac{2eV}{\hbar}, where e is the and \hbar is the reduced Planck's constant. This linear phase winding causes the supercurrent through the junction, given by the first Josephson relation I = I_c \sin \phi (with I_c the critical current), to oscillate as I(t) = I_c \sin(\phi(t)), producing an at a single f = \frac{2eV}{h}, where h is Planck's constant; this is the core mechanism of the AC Josephson effect. The resulting AC current is purely oscillatory with no net DC component, enabling direct voltage-to-frequency conversion that underpins precise metrological applications. For typical biases in the microvolt range (e.g., 20–200 μV), the emission frequency falls in the microwave band of 10–100 GHz, with radiated power reaching up to several nanowatts, though higher values near 1 μW are possible in optimized setups. Due to the intrinsic nonlinearity of the \sin \phi term, the emitted spectrum includes harmonics at integer multiples of the fundamental frequency, with power in higher harmonics typically 1–2 orders of magnitude lower than the primary signal. Experimentally, the AC Josephson effect was confirmed through observations of constant-voltage steps ( steps) in the current-voltage characteristics of junctions under irradiation, where the applied locks the intrinsic , leading to quantized voltage plateaus at V_n = n \frac{h f}{2e} (n ). These steps provide of phase-locking between the external field and the junction's AC response. The linewidth of the emitted is determined by the junction's quality factor , which reflects and ; high-quality junctions achieve > 10^5, yielding narrow linewidths essential for coherent applications.

Inverse AC Josephson Effect

The inverse AC Josephson effect manifests as a series of constant-voltage steps in the current-voltage (I-V) characteristic of a Josephson exposed to external alternating-current () electromagnetic , such as microwaves at f. These steps, termed Shapiro steps, appear as horizontal plateaus where the voltage locks to discrete values V_n = n \frac{h f}{2e}, with n an , h Planck's constant, and e the , allowing a range of currents at fixed voltage. This phenomenon arises from phase synchronization between the external drive and the intrinsic dynamics of the superconducting phase difference across the . Predicted theoretically by in 1962 as a consequence of the AC component of the supercurrent under combined and bias, it demonstrates the quantum mechanical locking of the junction's oscillatory behavior to the incident . The underlying mechanism involves the time evolution of the gauge-invariant phase difference \phi(t) across the junction, governed by the second Josephson relation modified for the applied voltages: \phi(t) = \phi_0 + \frac{2e}{\hbar} V t + \frac{2e V_\mathrm{rf}}{\hbar} \sin(\omega t), where V is the DC voltage, V_\mathrm{rf} is the amplitude of the RF voltage induced by the external drive, \omega = 2\pi f, and \hbar is the reduced Planck's constant. The resulting supercurrent is I(t) = I_c \sin[\phi(t)], with I_c the critical current. Expanding this using the Jacobi-Anger identity yields I(t) = I_c \sum_{n=-\infty}^\infty J_n\left( \frac{2e V_\mathrm{rf}}{\hbar \omega} \right) \sin\left[ \left( \frac{2e V}{\hbar} + n \omega \right) t + \psi \right], where J_n are Bessel functions of the first kind. Phase locking occurs when \frac{2e V}{\hbar} = -m \omega for integer m, making the n = -m term time-independent and enabling a DC current component over a finite voltage range, thus forming the step at V_m = m \frac{h f}{2e}. The height of each step is proportional to I_c |J_n(\alpha)|, with \alpha = \frac{2e V_\mathrm{rf}}{\hbar \omega}, and can approach I_c for low-order steps under optimal drive amplitudes, while higher-order steps diminish due to the oscillatory nature of the Bessel functions. First observed experimentally by Sidney Shapiro in 1963 using microwave irradiation (9–24 GHz) on Al-Al_2O_3-Pb and Sn-SnO-Sn tunnel junctions at 1.2–4.2 K, the steps appeared as zero-current intercepts spaced by ΔV ≈ 2 μV/GHz, with widths up to several μA and heights comparable to I_c for the fundamental step. Integer-order Shapiro steps (n = 0, 1, 2, \ldots) dominate in standard sinusoidal current-phase relation junctions, but fractional steps (e.g., at V = (p/q) h f / 2e, with integers p, q) emerge in systems with higher harmonics in the current-phase relation, such as ferromagnetic or high-transparency junctions. These steps have been observed up to high orders, with n \approx 1000 in optimized tunnel junctions under strong drives, confirming the scalability of the effect. This effect unequivocally confirms the quantum interference nature of Cooper pair tunneling and precisely verifies the universal frequency-voltage relation f = \frac{2e V}{h}, measured to accuracies better than 1 part in $10^{18} through comparisons with cesium frequency standards. The step characteristics are highly sensitive to via flux quantization, as an applied field \mathbf{B} modulates the through the Aharonov-Bohm-like \Delta \phi = \frac{2e}{\hbar} \int \mathbf{A} \cdot d\mathbf{l}, equivalent to flux \Phi = B A (with junction area A) altering the effective I_c via the pattern and shifting step positions, enabling sub-pT detection in SQUID-like configurations.

Theoretical Models and Properties

Resistively and Capacitively Shunted Junction (RCSJ) Model

The resistively and capacitively shunted junction (RCSJ) model provides a phenomenological framework for describing the dynamics of a Josephson junction by treating it as a parallel combination of an ideal Josephson element, a , and a . The ideal Josephson element carries a supercurrent I_c \sin \phi, where I_c is the critical current and \phi is the gauge-invariant across the junction. The shunt R accounts for the normal current I_N = V / R, with V denoting the voltage, while the C contributes the I_C = C \, dV/dt. This circuit analogy captures the interplay between superconducting, dissipative, and inertial effects in the junction's response to bias currents. The total current I through the junction is the sum of these components: I = I_c \sin \phi + \frac{V}{R} + C \frac{dV}{dt}. Using the second Josephson relation d\phi/dt = 2eV / \hbar, where e is the elementary charge and \hbar is the reduced Planck's constant, the voltage V can be eliminated to yield a second-order differential equation for the phase \phi: \frac{\hbar C}{2e} \frac{d^2 \phi}{dt^2} + \frac{\hbar}{2e R} \frac{d\phi}{dt} + I_c \sin \phi = I. This equation resembles the motion of a damped, driven pendulum, with the phase \phi playing the role of the angular displacement. To analyze the dynamics, the equations are often normalized. Introducing the characteristic time \tau_J = \hbar / (2e I_c R) and normalizing the current by I_c, the equation becomes \beta_c \frac{d^2 \phi}{d\tau^2} + \frac{d\phi}{d\tau} + \sin \phi = i, where \tau = t / \tau_J is the normalized time and i = I / I_c is the normalized bias current. The Stewart-McCumber parameter \beta_c = (2e I_c / \hbar) C R^2 quantifies the damping: large \beta_c indicates underdamped (hysteretic) behavior, while small \beta_c corresponds to overdamped dynamics. Additionally, the plasma frequency \omega_p = \sqrt{2e I_c / (\hbar C)} sets the scale for small-amplitude oscillations around the zero-voltage state, representing the natural frequency of the junction's "plasma" modes. The RCSJ model elucidates key dynamical features, such as the hysteresis in current-voltage (I-V) characteristics for underdamped junctions, where the junction switches abruptly from the zero-voltage state to a finite-voltage running state upon exceeding I_c, but retraps at a lower current during decreasing bias. It also explains phase diffusion in the thermal regime, where random phase fluctuations broaden the zero-voltage state and suppress the supercurrent, as well as the probabilistic switching from the metastable zero-voltage state due to thermal activation over the washboard potential barrier. The model is valid for temperatures much below the critical temperature (T \ll T_c) to ensure negligible thermal smearing of the superconducting gap and for small junctions where the phase is uniform across the device, avoiding fluxon effects. Extensions incorporate thermal noise via Langevin terms and quantum tunneling for ultra-low temperatures, addressing limitations in the classical description.

Josephson Inductance

The effective inductance of a Josephson junction arises from its nonlinear current-phase relation, where the supercurrent is I = I_c \sin \phi, with I_c the critical current and \phi the superconducting phase difference across the junction. This relation implies that the junction behaves as a phase-dependent nonlinear inductor in superconducting circuits, distinct from linear inductors due to its dependence on the operating bias point. To derive the inductance, consider the second Josephson relation, which links voltage V to the phase evolution: V = \frac{\Phi_0}{2\pi} \frac{d\phi}{dt}, where \Phi_0 = \frac{h}{2e} \approx 2.07 \times 10^{-15} is the . The current change rate is \frac{dI}{dt} = \frac{dI}{d\phi} \frac{d\phi}{dt} = I_c \cos \phi \cdot \frac{d\phi}{dt}. Substituting into the voltage expression yields V = \frac{\Phi_0}{2\pi I_c \cos \phi} \frac{dI}{dt}, defining the effective Josephson inductance as L_J(\phi) = \frac{\Phi_0}{2\pi I_c \cos \phi}. This derivation highlights the inductive response in the zero-voltage state, treating the junction as an for small perturbations around a bias \phi. The inductance L_J(\phi) exhibits key properties: it reaches a minimum value L_J(0) = \frac{\Phi_0}{2\pi I_c} at \phi = 0, typically ranging from 30 to 300 for I_c in the 1–10 μA range common in quantum circuits. As \phi approaches \pi/2, \cos \phi \to 0, causing L_J to diverge, which limits the junction's operation near the critical current. For small-signal applications biased near \phi = 0, an average or small-signal inductance \langle L_J \rangle \approx L_J(0) (1 + \frac{\phi^2}{2}) is often used to approximate linear behavior. In practice, this nonlinear serves as a tunable in superconducting , enabling compact on-chip inductors far smaller than geometric ones. In superconducting quantum devices (SQUIDs), threading modulates I_c and thus L_J, providing sensitive control over circuit parameters. The inherent nonlinearity supports advanced functionalities, such as parametric amplification through phase-dependent gain and dynamical bifurcations under drive, where the inductor's response switches between stable states.

Josephson Energy

The Josephson energy arises from the phase-dependent across the junction and serves as the for the superconducting phase difference \phi. It is expressed as E_J(\phi) = E_J (1 - \cos \phi), where E_J = \frac{\hbar I_c}{2e} = \frac{\Phi_0 I_c}{2\pi} defines the characteristic energy scale, with I_c the critical current, \hbar the reduced Planck's constant, e the , and \Phi_0 = h/2e the . For typical junctions used in classical applications with I_c \approx 1 mA, E_J \sim 10^{-19} J. This energy form is derived from the fundamental Josephson relations: the supercurrent I = I_c \sin \phi and the voltage-phase relation V = \frac{\hbar}{2e} \frac{d\phi}{dt}. The instantaneous power dissipated or stored in the junction is P = I V = I_c \sin \phi \cdot \frac{\hbar}{2e} \frac{d\phi}{dt}. Integrating over the phase evolution yields the change in stored energy: dE = \frac{\hbar I_c}{2e} \sin \phi \, d\phi = -E_J \, d(\cos \phi), resulting in E_J(\phi) = -E_J \cos \phi + \mathrm{const}, conventionally shifted to E_J (1 - \cos \phi) so that the minimum energy occurs at \phi = 0. The E_J(\phi) profile resembles a tilted washboard potential for the phase variable \phi, with periodic minima separated by barriers of height up to $2E_J. In the presence of bias , the potential tilts, creating metastable states at the wells from which the phase can escape thermally or quantum mechanically over the barriers, influencing junction switching dynamics. In quantum regimes, E_J establishes the primary scale for dynamics, setting the frequency of plasma oscillations and thus the timescale for quantum , typically on the order of E_J / \hbar \sim 10^{10}–$10^{12} rad/s. The ratio E_J / E_C, where E_C = e^2 / 2C is the charging with junction C, determines the classical-to-quantum crossover: large E_J / E_C \gg 1 yields classical, phase-dominated behavior, while E_J / E_C \sim 1–$100 enables quantum fluctuations and coherent superposition states. The anharmonic nature of the potential, stemming from the nonlinear \cos \phi dependence, is crucial for quantum applications; it produces unequally spaced energy levels in the phase-basis Hamiltonian H = 4E_C n^2 + E_J (1 - \cos \phi) (with n the operator), allowing selective microwave-driven transitions between adjacent levels without populating higher ones in operations. In the resistively and capacitively shunted junction model, modifies escape rates from this landscape but preserves the underlying .

Josephson Penetration Depth

The Josephson penetration depth, denoted as \lambda_J, represents the characteristic length scale over which and supercurrents extend into a Josephson junction, particularly in extended or large-area configurations. It quantifies the balance between the coherence of the superconducting order parameter and the screening of applied or self-generated . The expression for \lambda_J is \lambda_J = \sqrt{\frac{\Phi_0}{2\pi \mu_0 J_c d}}, where \Phi_0 = h / 2e is the , \mu_0 is the permeability of free space, J_c is the , and d is the effective magnetic thickness of , approximated as d \approx 2\lambda_L + t with \lambda_L the of the superconducting electrodes and t the thickness of the insulating barrier. This length scale emerges from the static electrodynamics of the junction, derived by considering the spatial variation of the gauge-invariant phase difference \phi(x) across the junction in the presence of a magnetic field. For extended junctions, the phase satisfies the one-dimensional sine-Gordon equation in the stationary case: \frac{\partial^2 \phi}{\partial x^2} = \frac{1}{\lambda_J^2} \sin \phi, which balances the curvature of the phase (related to field screening) against the nonlinear Josephson current-phase relation. The derivation assumes a uniform critical current density and incorporates Ampère's law for the magnetic field generated by the supercurrent, leading to an exponential decay of the field penetration similar to the Meissner effect in bulk superconductors. In typical tunnel junctions fabricated with materials like or lead, \lambda_J takes values between 10 and 100 \mum, depending on J_c (often 100–1000 A/cm²) and the /barrier parameters. Fluxons—topological defects corresponding to 2\pi phase slips across the junction—manifest as solutions to the sine-Gordon equation and propagate along the junction at the Swihart velocity \bar{c} = c / \sqrt{\epsilon_r}, where c is the in and \epsilon_r is the relative of the barrier (typically 4–10 for insulators). For junctions much smaller than \lambda_J (e.g., submicron overlap areas), the supercurrent flows uniformly, enabling simple point-like models. In larger junctions where the dimensions exceed \lambda_J, the current density becomes nonuniform due to self-field effects, resulting in Meissner-like screening currents at the edges and quantized entry of in units of \Phi_0. The value of \lambda_J imposes fundamental limits on device scaling, as junctions larger than this scale can trap multiple flux quanta, degrading uniformity and performance in applications like SQUIDs or microwave circuits. In long junctions (length \gg \lambda_J), it facilitates dynamic effects such as Fiske modes, which appear as constant-voltage steps in the current-voltage characteristics due to resonant coupling between the AC Josephson current and electromagnetic cavity modes propagating at \bar{c}.

Applications and Significance

Metrological Standards

The AC Josephson effect underpins the primary voltage standard in electrical , providing an exact relationship between an applied and the resulting voltage across an array of Josephson junctions. In a (JVS), arrays of thousands of junctions biased with microwaves of f generate a quantized voltage V = n \frac{h f}{2e}, where n is the number of junctions, h is Planck's constant, and e is the . This frequency-voltage relation enables voltage calibration with direct traceability to the second through precise . Practical JVS systems employ series arrays of 10,000 to 100,000 niobium-based tunnel junctions on a single chip to produce stable output voltages ranging from 1 V to 10 V at frequencies around 70 GHz, achieving uncertainties below $10^{-9}. The first operational JVS devices appeared in the early , initially limited to millivolt outputs from single junctions, with series-array designs enabling 1 V standards by 1985 and broader adoption thereafter. An international agreement in 1990 established a conventional value for the Josephson constant K_J = 2e/[h](/page/H+) = 483597.9 GHz/V, aligning national voltage references worldwide and superseding earlier Zener diode-based standards. The 2019 redefinition of the SI units fixed the numerical values of h and e exactly, eliminating conventional constants like K_J and making the Josephson relation intrinsically precise without reliance on agreed-upon values. This ties the volt directly to fundamental physical constants, enhancing global uniformity in electrical measurements. The DC Josephson effect supports realizations of the ampere through series arrays of junctions that sustain supercurrents without voltage drop, though practical implementations remain uncommon due to thermal dissipation and noise challenges in maintaining quantization. Modern programmable JVS incorporate superconductor-normal metal-superconductor (SNS) junctions, such as Nb/PdAu/Nb configurations with up to 32,768 elements per chip, enabling binary-coded rapid switching to discrete voltage steps at lower frequencies (10–20 GHz) for automated, high-speed calibrations. These systems integrate with quantum Hall resistance standards to form the quantum metrology triangle, closing the circuit for ampere and ohm realizations via I = V / R.

Sensing and Detection Devices

The phase of Josephson junctions enables highly precise detection of and through devices that exploit quantum effects. Superconducting Quantum Interference Devices (SQUIDs) represent the primary class of such sensors, leveraging the periodic dependence of the junction's critical current on the enclosed to achieve unprecedented . These devices operate at cryogenic temperatures to maintain , with performance ultimately limited by that introduce noise in the flux measurement. The DC SQUID consists of two Josephson junctions connected in parallel within a superconducting loop, forming a closed sensitive to external . The total critical current I_c through the device is modulated by the applied \Phi according to I_c(\Phi) \propto \cos(\pi \Phi / \Phi_0), where \Phi_0 = h/(2e) \approx 2.07 \times 10^{-15} Wb is the ; this interference arises directly from the phase coherence across the junctions. By biasing the SQUID with a slightly above I_c and measuring the resulting voltage, small changes in produce detectable shifts in the voltage output, enabling magnetometry applications down to femtotesla () levels when coupled to appropriate pickup coils. Typical noise sensitivities reach \delta \Phi \sim 10^{-6} \Phi_0 / \sqrt{\mathrm{Hz}} at 1 Hz, corresponding to sensitivities below 1 /\sqrt{\mathrm{Hz}} in optimized configurations. RF SQUIDs, employing a single Josephson junction in the loop and coupled to a resonant tank circuit, extend detection capabilities to low-frequency magnetic signals where SQUIDs may be more susceptible to environmental . The RF SQUID modulates the quality factor or frequency of the tank circuit in response to , allowing indirect readout via changes in reflected RF ; this configuration has been particularly useful for broadband, low-frequency magnetometry in noisy environments. While generally less sensitive than SQUIDs (with flux noise around 10^{-5} to 10^{-6} \Phi_0 / \sqrt{\mathrm{Hz}}), RF variants simplify fabrication and have facilitated early advancements in practical sensing systems. Beyond magnetometry, the AC Josephson effect underpins bolometric detectors for sub-millimeter wave radiation, where incident photons generate Shapiro steps in the junction's current-voltage characteristics, leading to measurable heating or resistance changes. These Josephson bolometers, often based on one-dimensional arrays or graphene-integrated junctions, offer high responsivity in the 100 GHz to 1 THz range, with noise equivalent powers below 10^{-16} W/\sqrt{\mathrm{Hz}} for terahertz detection; they convert absorbed radiation into temperature rises via the junction's nonlinear dynamics, enabling applications in astrophysics and spectroscopy. SQUIDs find widespread use in , such as (MEG) for non-invasive brain activity mapping, where multi-channel arrays detect neural currents at picotesla levels; geophysical surveys for mineral exploration and earthquake precursor monitoring; and of materials for subsurface defects in components. Commercial SQUID systems emerged in the , initially for laboratory magnetometry, evolving into integrated platforms for these fields by the with multichannel capabilities up to 300 sensors. Low-temperature SQUIDs typically require cooling at 4 , but high-T_c variants using materials like YBa_2Cu_3O_7 enable operation at 77 with , reducing cryogenic costs while maintaining sensitivities above 10 fT/\sqrt{\mathrm{Hz}}; thermal noise from Johnson-Nyquist effects sets the fundamental limit, often mitigated by flux locking and shielding.

Quantum Technologies

Josephson junctions serve as the core nonlinear elements in superconducting quantum processors, enabling the realization of for processing. These devices exploit the quantum mechanical properties of superconducting circuits to encode quantum bits in discrete energy states, facilitating operations essential for and . The provided by the Josephson potential allows selective addressing of two-level systems, distinguishing them from oscillators and enabling precise control over quantum states. Superconducting qubits based on Josephson junctions include several prominent types, each leveraging different . The Cooper pair box, an early charge-based qubit, operates by tuning the charge imbalance across a Josephson junction to select discrete charge states as the qubit basis. Its variant, the , reduces sensitivity to charge noise by increasing the Josephson energy relative to the charging energy, achieving greater . Phase qubits utilize the anharmonic energy levels in the phase across a current-biased Josephson junction, while flux qubits encode information in the circulating current states of a superconducting loop interrupted by Josephson junctions. Modern designs routinely achieve coherence times exceeding 100 μs, a significant driven by material and fabrication advances. The operational mechanism of these qubits relies on the inherent of the Josephson energy landscape, E_J(\phi) = -E_J \cos \phi, where \phi is the difference across the junction, which deviates from a simple harmonic potential and supports well-separated two-level systems for the ground and first excited states. For charge-based designs like the box, discrete charge states provide the anharmonicity, while in and flux qubits, it arises from the nonlinear or flux dynamics. Qubit control is achieved through pulses to drive transitions between these levels or flux pulses to tune the splitting, allowing for single- and two-qubit gates in architectures. The Josephson E_J sets the typical scale for qubit transition frequencies in the 4–8 GHz range. The first experimental demonstration of a Josephson junction-based occurred in 1999 with the coherent manipulation of charge states in a box, marking the advent of solid-state quantum bits. By 2025, these s have scaled to large integrated circuits, with processors from and featuring over 50 qubits—such as IBM's with 1,121 qubits and Google's chip supporting advanced error-corrected operations—enabling demonstrations of quantum advantage in specific tasks. In hybrid quantum systems, Josephson junctions function as tunable couplers for interconnecting qubits in quantum networks, facilitating entanglement distribution across distant nodes. Voltage-tunable Josephson junctions enable parametric gates, where nonlinear amplification or mixing of signals generates controlled interactions, such as cross-resonance or parametric-resonance entangling operations between transmons. These elements support modular architectures for scalable quantum computing. Despite progress, challenges persist in superconducting qubit technologies, particularly decoherence from and material losses, which limit operational . Advances in , including surface code implementations that suppress logical errors below physical thresholds, mitigate these effects and enable fault-tolerant computation. Scalability remains constrained, with current systems approaching 1,000 qubits but requiring innovations in wiring, cryogenic infrastructure, and fabrication uniformity to reach millions for practical applications.

References

  1. [1]
    Possible new effects in superconductive tunnelling - ScienceDirect
    Possible new effects in superconductive tunnelling☆. Author links open overlay panel B.D. Josephson ... Shirkov. A new method in the theory of superconductivity.
  2. [2]
    Josephson Effects and Josephson Voltage Standards - PTB.de
    Original paper of B.D. Josephson. B.D. Josephson, "Possible new effects in superconductive tunnelling", Phys. Lett. 1 (1962) 251-253.
  3. [3]
    [PDF] Lecture 11: Basic Josephson Junctions - MIT
    Oct 9, 2003 · The principle of the dc Volt: Put 5000 Josephson junctions in series, and apply a fixed frequency, which can be done very accurately, and ...
  4. [4]
    [PDF] Superconducting Qubits and the Physics of Josephson Junctions
    Josephson junctions are used for qubits due to low dissipation and long coherence times. The Josephson effect describes the supercurrent flowing through the ...
  5. [5]
    Superconductivity | CERN
    In 1911, while studying the properties of matter at very low temperature, the Dutch physicist Heike Kamerlingh Onnes and his team discovered that the ...
  6. [6]
    Discovery of superconductivity | Physics Today - AIP Publishing
    Apr 8, 2015 · On this day in 1911 Heike Kamerlingh Onnes dipped a wire made of solid mercury into liquid helium and lowered the temperature.
  7. [7]
    [PDF] SOLID STATE PHYSICS PART IV Superconducting Properties of ...
    2.1 The Cooper Pair. Most of the distinctive properties of superconductivity are explained by the macroscopic quantum description of superconductivity. The ...
  8. [8]
    Josephson Junction - an overview | ScienceDirect Topics
    Josephson junctions are defined as weak links between two superconductors that enable the transfer of Cooper pairs and the coupling of their macroscopic wave ...Missing: contacts | Show results with:contacts
  9. [9]
    Simulating the fabrication of aluminium oxide tunnel junctions - Nature
    Jan 28, 2021 · In this paper, we use molecular dynamics to develop models of Al–AlO x –Al junctions by iteratively growing the structures with sequential calculations.Missing: insulating | Show results with:insulating
  10. [10]
    [PDF] Controlling the thickness of Josephson tunnel barriers with atomic ...
    Aug 12, 2014 · aluminum oxide layer is not thick enough to act as a diffusion barrier (h). Page 3. 4EPo1C-04. For JJ tunnel barriers, less than 10 ALD cycles ...
  11. [11]
    [PDF] Superconductor-normal-superconductbr (SNS)
    Because of difficulties in fabricating high T, tunnel junctions, SNS weak liriks may be the most promising Josephson de- vices for applications at ...
  12. [12]
    [PDF] Brian D. Josephson - Nobel Lecture
    His lecture course in. Cambridge introduced the new concept of 'broken symmetry' in supercon- ductors, (1) which was already inherent in his 1958 pseudospin ...
  13. [13]
    The Nobel Prize in Physics 1973 - NobelPrize.org
    The Nobel Prize in Physics 1973 was divided, one half jointly to Leo Esaki and Ivar Giaever for their experimental discoveries regarding tunneling phenomena.
  14. [14]
  15. [15]
    Broadband Josephson Voltage Standards | NIST
    May 20, 2001 · A Josephson junction is a perfect frequency to voltage converter, that is, V=f/Kj where Kj = 483597.9 GHz/V.
  16. [16]
    Experimental Test of the Josephson Frequency-Voltage Relation
    The measured frequency-voltage ratio in this case was equal to the currently accepted value of to within an experimental accuracy of 0.006% (60 ppm). The ...
  17. [17]
  18. [18]
    None
    Nothing is retrieved...<|separator|>
  19. [19]
    A survey of classical and quantum interpretations of experiments on ...
    Feb 17, 2016 · We address two questions: (1) How successful is the non-quantum theory (RCSJ model) in replicating those experiments? (2) How strong is the ...Missing: original | Show results with:original
  20. [20]
    Superconducting Qubits and the Physics of Josephson Junctions
    Feb 16, 2004 · The Josephson equations are derived using standard first and second order perturbation theory that describe quasiparticle and Cooper-pair ...
  21. [21]
    Quantum-state engineering with Josephson-junction devices
    May 8, 2001 · The article reviews the properties of these devices and the practical and fundamental obstacles to their use. Two kinds of device have been ...
  22. [22]
    [PDF] arXiv:2103.01225v3 [quant-ph] 10 Nov 2023
    Nov 10, 2023 · L(Φ) = ∂I. ∂Φ. −1. = LJ cos. 2π. Φ. Φ0. ,. (14) where we define the Josephson inductance LJ = Φ0/2πIc. Since the inductance is associated with ...Missing: L_J | Show results with:L_J
  23. [23]
    Tunable ohmic environment using Josephson junction chains
    May 17, 2018 · The effective Josephson junctions are implemented using SQUIDs threaded by a magnetic flux Φ B to achieve a tunable Josephson inductance L J = Φ ...
  24. [24]
    Bolometric detection of Josephson inductance in a highly resistive ...
    Dec 1, 2023 · The derivation relies on the hypothesis of very weak Josephson coupling, 2EJ < kBTS,D. This condition is not well satisfied for both samples ...
  25. [25]
    Self-Field Limiting of Josephson Tunneling of Superconducting ...
    Self-Field Limiting of Josephson Tunneling of Superconducting Electron Pairs. Richard A. Ferrell and Richard E. ... R. E. Prange, Phys. Rev. 129, 2495 (1963).Missing: penetration depth
  26. [26]
    [PDF] Applied Superconductivity: Josephson Effects and ... - wmi.badw.de
    We will introduce the. Josephson coupling energy and the Josephson inductance as well as mechanical analogs for lumped. Josephson junctions. In sections 2.2 ...
  27. [27]
    [PDF] Josephson penetration depth in coplanar junctions based on 2D ...
    We derive a new expression for the Josephson penetration depth of such junctions and verify our assumptions by numerical simulations. Page 2. 2. I. Introduction.Missing: seminal | Show results with:seminal
  28. [28]
    Josephson penetration depth in coplanar junctions based on 2D ...
    We have shown that the Josephson penetration depth of this sort of junction is proportional to 1 / t and is around tens of micrometers for typical Josephson ...Missing: seminal | Show results with:seminal
  29. [29]
    Josephson junctions, superconducting circuits, and qubit for ... - arXiv
    May 31, 2024 · In its simplest form, the Josephson effect involves the flow of supercurrent (a current without resistance) between two superconductors ...
  30. [30]
    Fiske modes in Josephson junctions | Phys. Rev. B
    Oct 5, 2006 · The self-resonance modes, known as Fiske steps, are calculated by assuming the presence of an arbitrary number of 0 − 𝜋 singularities in the ...
  31. [31]
    Fiske and flux-flow modes of the intrinsic Josephson junctions in
    Apr 27, 1998 · The characteristic velocities of the intrinsic Josephson junctions are estimated to be m/s, which corresponds to the highest velocity mode, ...
  32. [32]
    [PDF] Application of the Josephson Effect to Voltage Metrology
    The Josephson effect, relating frequency and voltage, is used to improve the accuracy of dc voltage standards, relating the volt to the second.
  33. [33]
    The Josephson Effect and Voltage Standards - ResearchGate
    Aug 6, 2025 · Josephson voltage standards with series arrays of up to 20 000 tunnel junctions can generate stable reference voltages at arbitrary values ...
  34. [34]
    [PDF] New internationally adopted reference standards of voltage and ...
    New standards for voltage and resistance are based on the Josephson effect and the quantum Hall effect, respectively, to improve uniformity of electrical ...
  35. [35]
    Ampere: The Quantum Metrology Triangle | NIST
    May 15, 2018 · Because radiation frequency can be controlled with extraordinary precision, a Josephson junction can produce exact levels of voltage, with ...
  36. [36]
  37. [37]
    [PDF] REVIEW ARTICLE Superconducting quantum interference device ...
    SQUID uses Josephson effect phenomena to measure ex- tremely small variations in magnetic flux. Typically, a. SQUID is a ring of superconductor interrupted by ...
  38. [38]
    [PDF] Superconducting Quantum Interference Devices: State of the Art and ...
    SQUIDs are sensitive magnetic flux detectors, consisting of a superconducting loop with Josephson junctions. They measure magnetic fields, voltage, and ...
  39. [39]
    [PDF] Investigations of 1/f flux noise in superconducting quantum circuits ...
    Typical flux gains are in the range of 10-100 and provide a noise floor of. < 1 µΦ0/. √. Hz referred to the input of the first stage SQUID. A final interesting ...
  40. [40]
    [PDF] Superconducting Quantum Interference Devices for the Detection of ...
    Mar 26, 2015 · The RF SQUID consists of one Josephson junction placed in a superconducting ring ... “Quantum interference effects in Josephson tunneling”.
  41. [41]
    Fully Superconducting Josephson Bolometers for Gigahertz ... - MDPI
    Jan 14, 2021 · Here, we review the design and experimental realization of two bias-current-tunable sensors based on one dimensional fully superconducting Josephson junctions.2. Sensors Structure And... · 3.2. Jes Bolometer... · 4.3. Bias Current Control Of...<|separator|>
  42. [42]
    Bolometric detection of Josephson radiation - arXiv
    Feb 14, 2024 · Here we design and build an engineered on-chip reservoir that acts as an efficient bolometer for detecting the Josephson radiation under non- ...
  43. [43]
    [PDF] SQUIDS - Vanderbilt University
    The rf SQUID is coupled to an rf tank circuit. The Josephson junc- tions are each shunted by a resistor. The physical capacitance of the junction is shown as ...Missing: effect | Show results with:effect
  44. [44]
    The Impact of High-Temperature Superconductivity on SQUID ...
    SQUIDs fabricated from high T(c) superconductors such as YBa(2)Cu(3)O(7) and operated in liquid nitrogen at 77 K offer a greater convenience in operation.Missing: 4K | Show results with:4K
  45. [45]
    A quantum engineer's guide to superconducting qubits
    Jun 17, 2019 · Therefore, the Josephson junction is the key ingredient that makes the oscillator anharmonic and thus allows us to identify a uniquely ...
  46. [46]
    [PDF] Quantum Bits with Josephson Junctions - Franco Nori
    Adding the extra capacitance decreases the charging energy EC in the circuit. By changing the EJ/EC ratio from EJ/EC ≈ 10−1 to EJ/EC ≈ 102, the charge-qubit ...Missing: E_J / E_C
  47. [47]
    Charge-insensitive qubit design derived from the Cooper pair box
    Oct 12, 2007 · In the transmon, it is the natural anharmonicity of the cosine potential which allows qubit operations, whereas in the phase qubit, the E J ∕ E ...
  48. [48]
    Methods to achieve near-millisecond energy relaxation and ... - arXiv
    Feb 3, 2025 · Short coherence times have historically been a disadvantage of superconducting qubits ... Transmon qubits with coherence times above 100 µs ...
  49. [49]
    [PDF] Microwaves in Quantum Computing - Time and Frequency Division
    Jan 7, 2021 · The quantization of microwave energy describes how microwave photons can interact resonantly with other quantum objects, such as the quantum two ...<|control11|><|separator|>
  50. [50]
    Superconducting quantum computers: who is leading the future?
    Aug 19, 2025 · Performance metrics include relaxation time ( T 1 ) and coherence time ( T 2 ), both measured in microseconds (μs), as well as qubit frequencies ...
  51. [51]
    Parametric-Resonance Entangling Gates with a Tunable Coupler
    Aug 27, 2021 · The Josephson junctions in the superconducting quantum interference device (SQUID) loops of the transmon qubits are then fabricated through ...
  52. [52]
    Four-body coupler for superconducting qubits based on Josephson ...
    Jun 10, 2025 · We theoretically propose a circuit of the four-body coupler for superconducting qubits based on Josephson parametric oscillators (JPOs).
  53. [53]
    Quantum error correction below the surface code threshold - Nature
    Dec 9, 2024 · Equipped with below-threshold logical qubits, we can now probe the sensitivity of logical error to various error mechanisms in this new regime.
  54. [54]
    Quantum Computing: Vision and Challenges - arXiv
    Sep 6, 2024 · Reducing the probability of decoherence and creating effective error correction ... 1000 qubits may increase the effectiveness and scalability ...