Fact-checked by Grok 2 weeks ago

Superconductivity

Superconductivity is a quantum mechanical phenomenon observed in certain materials that, when cooled below a characteristic critical temperature (Tc), exhibit zero electrical resistance to the flow of () and the complete expulsion of magnetic fields from their interior, a property known as the . This behavior allows for the conduction of electrical current without energy loss, fundamentally distinguishing superconductors from ordinary conductors. The discovery of superconductivity occurred in 1911 when Dutch physicist observed the sudden drop in electrical resistance to zero in mercury cooled to approximately 4.2 using , a finding that earned him the 1913 . Initially empirical, the phenomenon puzzled scientists until 1933, when and Robert Ochsenfeld identified the magnetic field expulsion, solidifying its unique quantum nature. In 1957, , , and developed the , which explains superconductivity in conventional materials through the formation of Cooper pairs—bound electrons mediated by lattice vibrations (phonons)—leading to a coherent that enables resistance-free flow; this work was recognized with the 1972 . Superconductors are classified into two main types: low-temperature (conventional) ones, which require cooling to near (e.g., niobium-titanium alloys with Tc around 9-10 K), and high-temperature (unconventional) superconductors, first discovered in 1986 by J. Georg Bednorz and in a copper-oxide with Tc of 30 K, later advancing to over 90 K in (YBCO) by 1987, with the current record for cuprates reaching about 134 K in mercury-based compounds, operable with cooling. The mechanism for high-Tc materials, often involving d-wave pairing and doping effects in cuprates and other unconventional families such as iron-based and nickelates, remains an active area of research beyond the phonon-based BCS framework. Superconductivity is also limited by critical parameters: a critical (Hc) beyond which the fails, and a critical (Jc) above which resistance reappears. Practical applications of superconductivity leverage these properties for efficient energy use and advanced technologies, including superconducting magnets in magnetic resonance imaging (MRI) machines, particle accelerators like the Large Hadron Collider, and nuclear magnetic resonance (NMR) spectrometers, where zero resistance enables strong, stable fields without power loss. Emerging uses span power transmission lines to reduce energy dissipation in grids, magnetically levitated (maglev) trains for frictionless high-speed transport, and quantum computing components exploiting the quantum coherence of superconducting states. Five Nobel Prizes have been awarded for superconductivity-related discoveries (1913, 1972, 1973, 1987, 2003), underscoring its profound impact on physics and engineering.

Discovery and Basic Properties

Historical Discovery

In 1911, Dutch physicist Heike Kamerlingh Onnes and his team at the University of Leiden discovered superconductivity while investigating the electrical resistivity of pure mercury at cryogenic temperatures. On April 8, 1911, they observed an abrupt drop in resistance to apparently zero as the temperature reached approximately 4.2 K, achieved using liquid helium that Onnes had pioneered the liquefaction of in 1908. This unexpected result, detailed in Onnes's seminal paper "The Resistance of Pure Mercury at Helium Temperatures," marked the first identification of a material exhibiting zero electrical resistance below a critical temperature. The breakthrough relied heavily on prior advancements in low-temperature physics, particularly the development of the Dewar flask by Scottish physicist in 1892. This vacuum-insulated vessel allowed for the efficient storage and transfer of liquefied gases like air and hydrogen, enabling Onnes to pre-cool gas effectively before its and maintain the ultra-low temperatures required for his experiments. Without such innovations, reaching and sustaining temperatures near would have been impractical, underscoring how incremental progress in facilitated the discovery. Onnes initially viewed the phenomenon as a novel , coining the term "supraconductivity" in (later standardized as "superconductivity") to describe the complete disappearance of . He speculated it might align with contemporary theories of behavior in metals, such as reduced at low temperatures, but subsequent checks revealed no dissipation even after prolonged current flow, defying existing models and prompting recognition of it as a distinct physical effect. Building on the mercury findings, Onnes's group extended observations in 1912 to other pure metals, noting zero resistance in lead at about 7.2 and in tin at roughly 3.7 , establishing superconductivity as a property shared by multiple elements under similar cryogenic conditions. These early experiments highlighted the need for high-purity samples, as impurities had initially suggested gradual resistance decreases rather than the sharp transition observed in refined materials.

Zero Electrical Resistance

One of the defining characteristics of superconductivity is the complete disappearance of electrical resistance below a critical , enabling infinite conductivity. This phenomenon was first observed in 1911 by , who measured the resistivity of mercury and found it to drop abruptly to zero at approximately 4.2 K when cooled using . Subsequent experiments confirmed this zero-resistivity state in other materials, such as lead and tin, establishing it as a of superconductors under appropriate conditions. The absence of permits s to circulate indefinitely in closed superconducting loops without energy dissipation. These currents arise from induced electromagnetic forces and maintain themselves due to the lack of ohmic losses, allowing practical applications like superconducting magnets. Experimental verification includes observations of such currents in lead cylinders, where a persistent current persisted for over two years without measurable decay, limited only by an external interruption in cooling. Unlike an ideal classical , which might sustain persistent currents through inertia but permit penetration, a superconductor achieves zero resistance while expelling internal s entirely—a complementary property known as the . In the phenomenological framework of the London theory, this zero-resistance behavior for fields is captured by the implication of the first London equation in : \mathbf{E} = 0 inside the superconductor, where \mathbf{E} is the , ensuring no and constant current flow.

Meissner Effect

The is the expulsion of a from the interior of a superconductor upon cooling below its critical temperature T_c, resulting in perfect . This phenomenon was discovered in by German physicists and Robert Ochsenfeld through experiments on lead and tin samples. Their work revealed that, unlike the normal state where s penetrate materials, the superconducting state actively excludes internal fields, distinguishing it from mere zero electrical observed earlier. In their experimental setup, Meissner and Ochsenfeld used cylindrical samples of polycrystalline lead and single-crystal tin, approximately 140 mm long and 3 mm in diameter, placed parallel and separated by 1.5 mm, within a uniform external of about 5 gauss generated by an . They measured changes using a small search , roughly 10 mm long, connected to a , positioned either between the cylinders or inside a hollow lead tube (130 mm long, 3 mm outer diameter, 2 mm inner diameter) for internal field assessment. Upon cooling the samples below T_c (around 7.2 K for lead and 3.7 K for tin) via , the registered deflections indicating a sudden expulsion of flux from the interior, with the external field lines compressing around the sample surfaces as if the material had zero permeability. If the field was applied after achieving the superconducting state, no penetration occurred, confirming the effect's thermodynamic nature. This discovery had profound implications, establishing superconductivity as a true thermodynamic to an equilibrium state rather than a metastable condition tied solely to dissipationless current flow. Prior understanding of zero resistance, found in mercury by in 1911, suggested persistent currents but not field expulsion; the clarified that the superconducting state minimizes magnetic energy through flux exclusion, enabling thermodynamic treatments like specific heat measurements and phase diagrams. Theoretically, the effect is captured by the condition that the magnetic induction \mathbf{B} = 0 inside the superconductor in the absence of currents, arising from \nabla \cdot \mathbf{B} = 0 and \nabla \times \mathbf{H} = \mathbf{J} with \mathbf{J} = 0 in the bulk superconducting region, implying perfect screening. This idealization holds for fields below the , underscoring the superconductor's role as a perfect diamagnet with \chi = -1.

Phase Transition and London Moment

Superconductivity emerges as a second-order at the critical T_c, below which the material abruptly loses electrical and expels . In the Ehrenfest classification, second-order transitions are defined by continuous first derivatives of the but discontinuities in higher-order derivatives, such as the specific heat at constant volume or pressure. For superconductors, this manifests as a sharp jump in the electronic at T_c, reflecting the onset of coherent pairing among electrons that forms the superconducting state. The value of T_c serves as the defining transition point and varies significantly across materials, influenced by factors like atomic structure, electron-phonon , and isotopic mass. Conventional low-temperature superconductors, such as elemental metals like or lead, exhibit T_c values typically below 10 , while high-temperature superconductors like cuprates can reach T_c exceeding 90 under . This material dependence underscores the diversity in superconducting mechanisms, with T_c marking the boundary where disrupts the quantum of paired electrons. A notable consequence of superconductivity in rotating systems is the London moment, a phenomenon where a spinning superconductor generates an internal aligned with its rotation axis. This arises from the conservation of for the Cooper pairs, which carry zero orbital in their ; rotation induces a uniform supercurrent that screens the mechanical , producing a proportional to the . Predicted by in his molecular theory of superconductivity, the effect highlights the rigid, quantized nature of the superconducting wavefunction. Experimental verification of the London moment has been achieved using superconducting gyroscopes, where the induced precisely tracks the spin axis, enabling ultra-stable orientation control. In the mission, niobium-coated quartz spheres served as gyroscopes, with their London moments monitored via superconducting quantum interference devices (SQUIDs) to detect minute drifts in , confirming the effect's reliability for high-precision measurements. This application demonstrates the practical utility of the London moment in inertial navigation and .

Theoretical Frameworks

Phenomenological Theories

Phenomenological theories of superconductivity provide macroscopic descriptions of the material's electromagnetic behavior near the critical temperature, without delving into microscopic mechanisms. These models treat superconductivity as a thermodynamic , using empirical relations to capture observable phenomena such as perfect and zero resistance. The foundational phenomenological approach was developed by Fritz and Heinz London in 1935, who proposed two constitutive equations relating the supercurrent density \mathbf{J} to the electromagnetic fields. The first London equation describes the acceleration of the supercurrent in response to an : \frac{\partial \mathbf{J}}{\partial t} = \frac{n_s e^2}{m} \mathbf{E}, where n_s is the of superconducting s, e and m are the charge and , and \mathbf{E} is the . In the static , this implies \mathbf{E} = 0 for steady currents, explaining the persistence of supercurrents without dissipation. The second London equation relates the curl of the current to the : \nabla \times \mathbf{J} = -\frac{n_s e^2}{m} \mathbf{B}, where \mathbf{B} is the . Combined with , this leads to the magnetic field penetrating the superconductor over a characteristic length, the London penetration depth \lambda_L = \sqrt{\frac{m}{\mu_0 n_s e^2}}, beyond which the field decays exponentially. A more comprehensive phenomenological framework was introduced by and in 1950, extending the London theory to include spatial variations in the superconducting order parameter. In Ginzburg-Landau theory, the superconducting state is described by a complex order parameter \psi, where |\psi|^2 represents the density of Cooper pairs (or superconducting electrons). The theory is formulated through a functional minimized to find states: F = \int \left[ \alpha |\psi|^2 + \frac{\beta}{2} |\psi|^4 + \frac{1}{2m^*} \left| \left( -i \hbar \nabla - 2e \mathbf{A} \right) \psi \right|^2 + \frac{|\mathbf{B}|^2}{2\mu_0} \right] dV, where \alpha = \alpha' (T - T_c) (with \alpha' > 0) changes sign at the critical temperature T_c, \beta > 0, m^* is the effective mass of Cooper pairs, \mathbf{A} is the , and the last term is the energy (in units). Minimizing this functional yields the Ginzburg-Landau equations, which generalize the London equations and introduce a \xi = \sqrt{\frac{\hbar^2}{2m^* |\alpha|}} over which \psi varies spatially. The Ginzburg-Landau framework has key applications in describing inhomogeneous superconducting . For type-II superconductors, where the Ginzburg-Landau parameter \kappa = \lambda_L / \xi > 1/\sqrt{2}, the predicts a mixed consisting of quantized lines or vortices, each carrying a flux quantum \Phi_0 = h / 2e. This vortex lattice structure was theoretically established by Alexei Abrikosov in 1957 using the Ginzburg-Landau equations. Additionally, the enables calculations of critical currents, such as the depairing J_c \propto (T_c - T)^{3/2}, above which the superconducting order parameter is suppressed, leading to a transition to the normal . These applications highlight the 's utility in modeling practical superconducting behaviors under magnetic fields and currents.

Microscopic BCS Theory

The Bardeen–Cooper–Schrieffer () theory, formulated in 1957, offers a microscopic quantum mechanical description of superconductivity in conventional materials, explaining the phenomenon as arising from the formation of pairs bound by interactions. In this framework, conduction s near the experience an attractive potential mediated by phonons—quantized vibrations—that overcomes their Coulomb repulsion at low temperatures. This attraction leads to the creation of Cooper pairs, composite bosons consisting of two s with opposite momenta and spins, which condense into a coherent ground state with long-range order. The pairing instability is treated using a mean-field approximation, where the many-body is decoupled into a form resembling a non-interacting Bogoliubov spectrum, enabling analytical solutions for thermodynamic and transport properties. Central to BCS theory is the superconducting energy gap \Delta, which represents the binding energy of the Cooper pairs and suppresses single-particle excitations below the critical temperature T_c. In the weak-coupling limit, the zero-temperature gap is related to the critical temperature by \Delta(0) \approx 1.76 k_B T_c. The gap equation at zero temperature, derived from the self-consistent condition for pair formation, yields the approximate relation for T_c: k_B T_c \approx 1.14 \hbar \omega_D \exp\left(-\frac{1}{N(0)V}\right), where \hbar \omega_D is the Debye energy scale setting the cutoff for phonon-mediated interactions, N(0) is the single-spin at the , and V is the effective potential. This highlights how even a weak (N(0)V \ll 1) can produce a finite gap due to the exponential sensitivity to the interaction strength. Near T_c, the gap vanishes, providing a direct link between observable T_c and microscopic parameters. A key experimental validation of the phonon-mediated pairing in is the isotope effect, where T_c scales inversely with the square root of the ionic mass M as T_c \propto M^{-1/2}, stemming from the mass dependence of the frequency \omega_D \propto M^{-1/2}. This relation was first observed in mercury isotopes in 1950, predating the full theory but confirming the role of electron-phonon coupling over purely electronic mechanisms. BCS theory also accounts for the thermodynamic signature of the superconducting transition, predicting a sharp discontinuity in the electronic specific heat at T_c. The jump is quantified as \Delta C = 1.43 \gamma T_c, where \gamma is the normal-state electronic specific heat coefficient, reflecting the abrupt opening of the energy gap and the associated entropy change in the paired state. On the electromagnetic front, the theory derives the superfluid density and acceleration equation for Cooper pairs, yielding a microscopic justification for the and perfect , consistent with macroscopic phenomenological models.

Extensions and Unconventional Theories

While the provides a foundational microscopic description for conventional superconductors mediated by interactions in the weak-coupling limit, extensions are necessary to account for stronger electron- couplings where retardation effects become significant. Eliashberg theory, developed as a strong-coupling , incorporates the full frequency dependence of the propagator and the resulting time retardation in the electron pairing interaction, leading to more accurate predictions of the superconducting transition temperature T_c and the gap function. This framework replaces the BCS energy-gap equation with a set of coupled integral equations for the renormalized Green's functions, capturing effects such as the isotope coefficient deviations observed in materials like lead and mercury. In unconventional superconductors, the pairing symmetry deviates from the isotropic s-wave state of BCS, often involving higher angular momentum states mediated by non-phononic mechanisms such as spin fluctuations. For instance, d-wave , characterized by a gap function \Delta(\mathbf{k}) \propto \cos k_x - \cos k_y, has been identified in through phase-sensitive tunneling and neutron scattering experiments, where antiferromagnetic spin fluctuations provide the dominant attractive in the pseudogap . Similarly, p-wave with odd-parity appears in heavy-fermion systems, where spin-triplet states are stabilized by ferromagnetic or spin-fluctuation exchanges, as evidenced by the helical in materials exhibiting chiral superconductivity. These unconventional symmetries arise from the sign-changing , which suppresses impurity scattering and leads to distinctive thermodynamic properties like power-law behaviors in specific heat at low temperatures. Theoretical modeling of these systems often relies on the , a minimal framework for strongly correlated electrons on a , where onsite repulsion U competes with kinetic hopping t to drive Mott insulation and antiferromagnetic order. In the strong-coupling limit, the effective t-J model derived from the Hubbard Hamiltonian highlights the role of interactions in fostering d-wave pairing, with antiferromagnetism providing the spin-fluctuation glue that binds Cooper pairs near half-filling. Numerical methods like and variational simulations confirm that doping away from the antiferromagnetic parent state can induce superconductivity, though the precise mechanism remains tied to the interplay of charge and . Despite these advances, a unified microscopic encompassing all superconductors remains elusive, with ongoing debates centering on the universality of pairing mechanisms across diverse classes like cuprates, iron pnictides, and organic materials. Key challenges include reconciling the absence of a single dominant mediator—phonons for conventional cases versus spin or orbital fluctuations for unconventional ones—and addressing discrepancies between and experiments in pseudogap phases or under . While Eliashberg-like extensions handle strong coupling effectively, adapting them to repulsive interactions or multi-orbital systems highlights the need for hybrid approaches that integrate beyond-mean-field correlations.

Classification of Superconductors

By Magnetic Response

Superconductors are classified into Type I and Type II based on their distinct responses to applied magnetic fields, particularly how they handle field penetration while maintaining the below critical thresholds. Type I superconductors exhibit complete expulsion of magnetic fields from their interior, achieving perfect up to a single critical field H_c, beyond which superconductivity abruptly ceases and the material transitions to . This behavior is characteristic of pure elemental metals such as aluminum and lead, where the Ginzburg-Landau parameter \kappa = \lambda / \xi < 1/\sqrt{2}, with \lambda as the penetration depth and \xi as the coherence length, favoring a positive interface energy that prevents partial field penetration. In Type I superconductors subjected to fields exceeding H_c in geometries like slabs or cylinders, an intermediate state can emerge to minimize free energy, consisting of macroscopic domains of alternating superconducting and normal regions threaded by magnetic flux. These domains form branched or lamellar structures, allowing partial accommodation of the external field without fully destroying superconductivity until the normal phase dominates. This state contrasts with the sharp transition in ideal conditions and highlights the role of demagnetization effects in real samples. Type II superconductors, defined by \kappa > 1/\sqrt{2}, display a more nuanced magnetic response with two critical fields: a lower critical field H_{c1} and an upper critical field H_{c2}. Below H_{c1}, they fully expel fields via the , similar to Type I materials. Between H_{c1} and H_{c2}, magnetic penetrates in a mixed state composed of quantized vortex lines, each carrying a flux quantum \Phi_0 = h/(2e), arranged in an ordered hexagonal Abrikosov lattice to stabilize the superconducting order parameter around the cores. This vortex structure, predicted theoretically and observed experimentally, enables Type II superconductors to sustain higher fields, making them essential for applications like superconducting magnets. The upper critical field H_{c2} marks the point where the vortex density becomes so high that the superconducting state collapses, given by the relation H_{c2} = \frac{\Phi_0}{2\pi \xi^2}, derived from the Ginzburg-Landau theory in the limit of linearized equations near the normal transition. This formula underscores the inverse dependence on the coherence length, explaining why materials with shorter \xi support stronger fields before losing superconductivity.

By Critical Parameters

Superconductors are characterized by three primary critical parameters that delineate the boundaries of the superconducting state: the critical temperature T_c, the critical H_c, and the critical j_c. These parameters define the thermodynamic and transport limits beyond which the material reverts to its normal resistive state. The critical temperature T_c is the maximum temperature at which superconductivity persists, corresponding to the onset of zero electrical resistance and perfect as the becomes insufficient to disrupt the Cooper pairs binding electrons. Across known superconductors, T_c spans a wide range, from millikelvin scales in exotic systems such as heavy-fermion compounds to around 130 in high-temperature variants, highlighting the diversity in pairing mechanisms and material properties. The critical magnetic field H_c(T) quantifies the strength of an external at temperature T that destroys superconductivity by providing energy comparable to the superconducting condensation energy. For type I superconductors, this is the thermodynamic critical field, while in type II materials, it relates to the upper critical field H_{c2} where the mixed state transitions to normal. Its temperature dependence is approximately parabolic, described by the empirical relation H_c(T) \approx H_c(0) \left[1 - \left( \frac{T}{T_c} \right)^2 \right], where H_c(0) is the value at ; this form arises from the balance between magnetic and condensation energies near the . This dependence ensures that superconductivity is suppressed more readily at higher temperatures, as weaken the pairing. The critical current density j_c is the highest density of supercurrent that can flow without inducing , beyond which normal emerges due to the motion of lines under Lorentz forces. In type II superconductors, j_c is particularly limited by flux flow, where unpinned vortices migrate under current drive, generating and resistivity analogous to viscous drag. Pinning centers, such as defects, enhance j_c by immobilizing vortices, enabling practical applications in high-current devices. Phase diagrams in the temperature-magnetic (H-T) map the interplay of these parameters, with the H_c(T) forming a parabolic enclosing the superconducting below T_c. Incorporating introduces additional complexity, as j_c varies with both T and H, often decreasing with increasing due to enhanced vortex mobility; this three-dimensional parameter space guides the design of superconducting magnets and wires by revealing operational limits where or thermal activation destabilizes the state.

By Material Type

Superconductors are categorized by their material composition—ranging from elemental metals to complex compounds—and the symmetry and mechanism of electron pairing, which determine their critical temperatures (Tc) and other properties. This classification distinguishes conventional materials, which adhere to the Bardeen-Cooper-Schrieffer (BCS) theory of phonon-mediated s-wave pairing, from unconventional ones involving anisotropic pairing often driven by magnetic fluctuations or other non-phonon mechanisms. Conventional superconductors primarily include elemental metals, alloys, and compounds with Tc values generally below 40 K, where Cooper pairs form in an isotropic s-wave state via electron-phonon interactions. Notable elemental examples are (Nb, Tc = 9.25 K) and (V, Tc = 5.4 K), which exhibit zero electrical resistance and the at these temperatures. A key compound example is (MgB2, Tc = 39 K), discovered in 2001, representing the highest Tc for conventional superconductors at ambient pressure. Alloys such as niobium-titanium (NbTi, Tc ≈ 9.5 K) and niobium-tin (Nb3Sn, Tc = 18.3 K) are widely used in practical applications like superconducting magnets due to their high critical fields and mechanical properties, while maintaining the s-wave BCS pairing symmetry. A key hallmark of these materials is the presence of the isotope effect, where Tc scales inversely with the square root of the atomic mass (isotope coefficient α ≈ 0.5), confirming the role of lattice vibrations (phonons) in mediating pairing. Unconventional superconductors, often layered compounds, feature higher and pairing symmetries that break rotational invariance, such as d-wave or p-wave, leading to nodes in the gap function and anisotropic properties. Cuprates like lanthanum-barium-copper oxide (La_{2-x}Ba_xCuO_4, up to 35 ) exemplify this class, with their superconductivity arising from strongly correlated electrons rather than phonons. Iron pnictides, such as LaFeAsO_{1-x}F_x ( up to 26 , with some variants reaching 55 ), also show anisotropic pairing, often s± symmetry, and are linked to spin-fluctuation mechanisms near antiferromagnetic instabilities. Unlike conventional superconductors, these materials generally exhibit a weak or absent isotope effect (α ≈ 0), underscoring non-phonon pairing origins. Organic superconductors, based on molecular charge-transfer salts, offer low but unique tunable properties in low-dimensional structures. A representative example is κ-(BEDT-TTF)_2Cu(NCS)_2, where BEDT-TTF denotes bis(ethylenedithio)tetrathiafulvalene, achieving = 10.4 K under and displaying exotic behaviors like potential Fulde-Ferrell-Larkin-Ovchinnikov (FFLO) states in high magnetic fields. These materials often involve quasi-two-dimensional conduction planes and unconventional pairing symmetries. Heavy fermion superconductors, characterized by f-electron systems with large effective masses, exhibit low Tc alongside complex . (CeCoIn_5, Tc = 2.3 K) is a prototypical case, featuring d_{x^2 - y^2}-wave pairing, nodal quasiparticles, and proximity to an antiferromagnetic , which enhances its unconventional nature. Like other unconventional types, it shows minimal effect and anisotropic superfluid density. This material-based classification emphasizes mechanistic distinctions: conventional superconductors align with BCS predictions including the isotope effect, while unconventional, , and heavy types deviate, often enabling higher or exotic states, with high-temperature cuprates serving as prime examples of the latter (detailed further in the High-Temperature Superconductors section).

Historical Development

Early Experiments and Theories

The discovery of superconductivity occurred in 1911 when and his team at the University of Leiden observed that the electrical resistance of pure mercury suddenly vanished below 4.2 K upon cooling with , a phenomenon they termed "superconductivity." This abrupt drop in resistance, plotted as a function of temperature, marked a sharp transition rather than a gradual decrease, prompting further investigations into the behavior of other metals under similar cryogenic conditions. From 1911 to 1933, Onnes' laboratory expanded these studies, mapping resistance-temperature curves for numerous pure elements and alloys. They identified superconductivity in materials like tin (at 3.7 K), lead (at 7.2 K), and (at 2.4 K), revealing that the transition temperature varied by element but consistently involved a complete loss of resistivity below a critical point. In 1950, experiments on mercury isotopes revealed the isotope effect, where Tc varied inversely with the of the atomic mass, suggesting involvement of lattice vibrations (phonons) in the pairing mechanism. These experiments established superconductivity as a reproducible quantum effect limited to low temperatures, though the underlying mechanism remained elusive. A pivotal advancement came in 1933 when and Robert Ochsenfeld at the Physikalisch-Technische Reichsanstalt in demonstrated that superconductors not only exhibit zero resistance but also expel applied magnetic fields from their interior upon entering the superconducting state—a property now known as the . Their measurements on lead and tin samples showed this flux expulsion occurs regardless of whether the field was present during cooling, distinguishing superconductivity from mere perfect conductivity and highlighting its diamagnetic nature. In response to the Meissner effect, Fritz and Heinz London, working in after fleeing , developed the first phenomenological theory of superconductivity in 1935. Their described superconductors as having a characteristic , beyond which magnetic fields decay exponentially, providing a macroscopic framework to explain both zero and perfect diamagnetism without invoking microscopic details. During the 1930s, experiments by J.G. Daunt and K. Mendelssohn at on the thermal properties of superconductors, including specific heat measurements near the temperature, provided early hints of involvement. Their observations of anomalies in suggested that electron-phonon interactions, tied to the ionic vibrations, played a role in the superconducting state, challenging purely electronic models. Early theoretical efforts, such as applications of band theory from , largely failed to account for superconductivity's key features, including the sharp and magnetic expulsion, as they predicted gradual changes or no such effect at low temperatures. By the , superconductivity had been confirmed in approximately 25 elements, primarily soft metals and some alloys, underscoring the phenomenon's specificity and the need for a more robust explanatory framework.

Mid-20th Century Advances

In the early 1950s, the phenomenological provided a framework for understanding superconductivity near the critical temperature, introducing a complex order parameter to describe the superconducting state and enabling predictions of vortex structures in magnetic fields. This theory, developed by and in 1950, extended earlier by incorporating spatial variations in the order parameter, which allowed for the classification of superconductors into Type I and Type II categories based on their response to magnetic fields. Specifically, it predicted the existence of Type II superconductors, where penetrates in quantized vortices rather than being completely expelled, a phenomenon later confirmed experimentally. The mid-1950s also saw significant advances in superconducting materials, particularly with the discovery of niobium-tin (Nb₃Sn) in 1954, which exhibited a critical of approximately 18 , higher than previously known elemental superconductors. This A15 compound, identified by Bernd T. Matthias and colleagues at Bell Laboratories, demonstrated superior performance in high magnetic fields due to its elevated upper critical field (H_{c2}), with early measurements in the early showing values exceeding 10 T at 4.2 . These properties made Nb₃Sn suitable for practical magnet applications, leading to the first superconducting electromagnets in the late 1950s and early commercial wire production by the . A major theoretical milestone came in 1957 with the publication of the Bardeen-Cooper-Schrieffer (BCS) theory, which provided the first microscopic explanation of superconductivity as arising from an attractive electron-phonon interaction forming Cooper pairs. , , and Robert Schrieffer's model quantitatively predicted key observables, such as the energy gap and critical temperature dependence on isotope mass, aligning closely with experimental data and earning widespread acceptance within the by the early 1960s. This theory not only resolved long-standing puzzles from earlier phenomenological approaches but also laid the groundwork for understanding conventional superconductivity in metals and alloys. Building on BCS, predicted in 1962 the tunneling of supercurrents across a thin insulating barrier between two superconductors, known as the , which demonstrated the macroscopic quantum coherence of the superconducting state. This DC Josephson effect, along with the AC variant under irradiation, was experimentally verified shortly thereafter in 1962-1963 by teams using lead-based junctions, confirming the phase-dependent current flow without energy dissipation. These findings highlighted the quantum nature of superconductivity at larger scales and spurred further research into weakly coupled superconducting systems.

High-Tc Breakthroughs

In 1986, J. Georg Bednorz and at IBM's Research Laboratory reported the observation of superconductivity in a barium-doped lanthanum copper oxide (La-Ba-Cu-O) system with a critical temperature (Tc) of 35 K, marking the first breakthrough in beyond the limits of conventional materials. This discovery, published in a seminal paper, demonstrated a resistive transition indicating superconductivity at temperatures significantly higher than the previous record of 23 K for metallic superconductors, and it earned them the 1987 for opening the field of ceramic high-Tc superconductors. The following year, in 1987, Ching-Wu Chu's group at the University of Houston advanced this work by synthesizing yttrium barium copper oxide (YBCO, YBa2Cu3O7-x), achieving a Tc of 92 K under ambient pressure, which surpassed the boiling point of liquid nitrogen (77 K) and enabled practical cooling without expensive liquid helium. This material's orthorhombic perovskite structure with CuO2 planes was confirmed to exhibit bulk superconductivity through zero resistance and the Meissner effect, revolutionizing potential applications by making high-Tc systems more accessible. Rapid material development followed, with bismuth-based cuprates (Bi-Sr-Ca-Cu-O, or BSCCO) reported in 1988 by Hiroaki Maeda's team, with the 2212 reaching Tc ≈ 85 , and the Pb-doped 2223 achieving onset above 110 and zero resistance near 105 . Shortly thereafter, thallium-based cuprates (Tl-Ba-Ca-Cu-O, or TBCCO) were discovered by Z. Z. Sheng and Allen M. Hermann, achieving a record Tc of 125 , further expanding the family of high-Tc oxides with multiple homologous containing varying numbers of CuO2 layers. These breakthroughs ignited a global research surge, with the 1987 American Physical Society March Meeting—dubbed the "Woodstock of Physics"—drawing over 1,800 presentations on high-Tc materials and spawning thousands of publications within months, as laboratories worldwide raced to replicate and extend the findings. However, initial efforts faced challenges in due to multiphase samples, precise oxygen requirements, and to conditions, leading to early skepticism and false alarms until confirmatory measurements solidified the results across institutions. This period highlighted the unconventional nature of pairing, though detailed mechanisms remained elusive at the time.

Unconventional Superconductivity

High-Temperature Superconductors

High-temperature superconductors are materials exhibiting superconductivity at temperatures above approximately 30 K, significantly higher than traditional low-temperature superconductors cooled by . These materials, first discovered in the with the cuprates, have revolutionized the field by enabling potential applications at more accessible temperatures using . Despite their promise, the microscopic pairing mechanisms remain unconventional and not fully understood, distinguishing them from . Cuprates, the archetypal high-temperature superconductors, feature a layered perovskite-like composed of conducting O₂ planes separated by insulating charge reservoir layers. Superconductivity emerges upon doping these antiferromagnetic Mott insulators with holes, typically via chemical substitution, to achieve an optimal carrier concentration around 0.16 holes per atom, where the critical temperature (T_c) reaches maxima, such as 92 K in YBa₂₃O₇. This optimal doping suppresses the and enhances , though the exact mechanism involves strong correlations and possibly d-wave . Iron-based superconductors, discovered in , include pnictides like LaFeAsO and chalcogenides such as FeSe, with T_c values up to 55 K under ambient conditions. These materials exhibit a multi-orbital, multi-band electronic structure, where superconductivity arises from interband pairing mediated by fluctuations, leading to s± with sign-changing parameters between and pockets. Unlike single-band cuprates, the multi-band nature allows for complex vortex dynamics and higher robustness to disorder in some cases. Under extreme pressures, hydrogen-rich compounds like LaH₁₀ have achieved record T_c values, with superconductivity confirmed at approximately 250 K in 2019 at ~170 GPa. This discovery validated prior theoretical predictions based on electron-phonon coupling in metallic hydrides, where light hydrogen atoms enable strong phonon-mediated pairing within conventional , though at pressures far exceeding practical use. Despite their high T_c, high-temperature superconductors face significant challenges in practical implementation. Their layered structures impart strong in superconducting properties, with critical currents and fields varying markedly along in-plane versus out-of-plane directions, complicating wire fabrication. , essential for maintaining high currents in magnetic fields, is often weak and requires engineered defects to enhance vortex immobilization. Additionally, grain boundaries in polycrystalline samples act as weak links, severely limiting intergranular critical currents due to suppressed pairing across misoriented interfaces, a persistent issue in both cuprates and iron-based materials.

Exotic Superconductors

Exotic superconductors encompass a diverse class of materials where the pairing mechanism deviates significantly from the conventional electron-phonon interactions described by , often involving strong electron correlations, spin fluctuations, or . These systems typically exhibit low critical temperatures but reveal profound insights into quantum many-body physics, including potential applications in . Key examples include heavy-fermion compounds, materials, and non-centrosymmetric structures that support chiral or topological pairing symmetries. Heavy-fermion superconductors, such as URu₂Si₂, represent a paradigm of unconventional superconductivity mediated by spin fluctuations in strongly correlated f-electron systems. In URu₂Si₂, superconductivity emerges at a critical Tc ≈ 1.5 , below a mysterious "hidden order" at 17.5 , where the order parameter remains unidentified despite extensive study. The superconducting state is unconventional, characterized by a nodal gap structure and evidence of spin-fluctuation-driven pairing, as indicated by colossal Nernst signals arising from superconducting fluctuations near the hidden-order phase. This material highlights how antiferromagnetic spin fluctuations can suppress conventional pairing and favor exotic order, with the heavy effective masses (up to 100 times the bare ) arising from Kondo lattice interactions. Organic superconductors, exemplified by κ-(BEDT-TTF)₂Cu(NCS)₂, demonstrate superconductivity in molecular crystals tuned close to a Mott insulating state, where electron correlations dominate. This compound achieves a relatively high Tc of approximately 10 K under ambient pressure, making it one of the highest-Tc organic superconductors. The proximity to the Mott insulator, achieved by varying pressure or chemical substitution, leads to a pseudogap phase above Tc, attributed to precursor pairing fluctuations influenced by antiferromagnetic spin correlations. In this layered system, the BEDT-TTF molecules form dimers that support a quasi-two-dimensional band structure, where the metal-insulator transition at the Mott boundary enhances superconducting instability through enhanced density of states at the Fermi level. Topological superconductors introduce spatial structure to the order parameter, potentially hosting protected edge states and non-Abelian anyons. Strontium ruthenate, Sr₂RuO₄, is a prominent candidate for chiral p-wave pairing, with superconductivity at Tc ≈ 1.5 K in a layered perovskite structure. Experimental evidence for p-wave triplet pairing includes muon spin relaxation measurements indicating time-reversal symmetry breaking and Kerr effect observations consistent with chiral currents, though recent phase-sensitive tests have sparked debate on the exact symmetry. In hybrid systems combining conventional superconductors with topological insulators or semiconductors, proximity-induced superconductivity can realize one-dimensional topological phases hosting Majorana zero modes at the ends or vortices. These zero-energy modes, predicted in semiconductor-superconductor nanowires under magnetic fields, enable braiding operations for fault-tolerant quantum computation, with signatures observed in tunneling conductance plateaus. Non-centrosymmetric superconductors, lacking inversion , permit mixed-parity due to antisymmetric spin-orbit , leading to unique vortex structures like half-quantum vortices and chiral . In such materials, the Rashba-type spin-orbit interaction splits the into spin-helical bands, allowing singlet and triplet components to mix and stabilize half-flux quanta (h/2e) in vortex cores, half the usual flux quantum. Examples include β-Bi₂Pd, where mesoscopic rings exhibit flux quantization at Tc ≈ 1.8 K, direct evidence of spin-triplet . Similarly, α-BiPd shows half-quantum flux in Little-Parks oscillations, confirming a helical p-wave state with chiral Majorana modes bound to vortex edges. These chiral states propagate unidirectionally along surfaces, protected by , and can form closed loops with half-quantum vortices in multicomponent systems.

Recent Advances

2D and Nanostructured Materials

Superconductivity in two-dimensional (2D) and nanostructured materials arises from quantum confinement effects that alter electronic band structures, enhancing electron-electron interactions and pairing mechanisms compared to bulk counterparts. In these low-dimensional systems, reduced dimensionality flattens energy bands, increases density of states near the Fermi level, and promotes correlated states, leading to unconventional superconducting phases at low temperatures. These effects are particularly pronounced in van der Waals heterostructures, where stacking and gating enable precise tuning of carrier density and interlayer coupling. In magic-angle twisted trilayer (MATTG), discovered around 2021-2022, superconductivity emerges in flat bands formed by moiré superlattices, with critical temperatures () reaching approximately 2-3 under displacement fields that break layer . These flat bands, arising from interlayer hybridization at twist angles near 1.6°, enhance pairing by slowing charge carriers and amplifying interactions, enabling tunable superconducting domes in the . Experimental transport measurements reveal robust zero-resistance states, highlighting how confinement in the moiré potential mimics strong-coupling superconductivity. In November 2025, researchers reported direct evidence of unconventional superconductivity in MATTG, observing a V-shaped indicative of non-phononic pairing mechanisms. Recent observations in 2025 have identified signatures of chiral superconductivity in rhombohedral tetralayer and pentalayer , without moiré patterns, exhibiting up to 300 in gate-tuned flat conduction bands. This phase, embedded within a - and valley-polarized quarter-metal state, shows magnetic hysteresis in resistivity under out-of-plane fields up to 1.4 T, indicating time-reversal and chiral pairing consistent with confinement-enhanced . The robustness against in-plane fields underscores the role of dimensionality in stabilizing exotic pairing symmetries. In transition metal dichalcogenides (TMDs), such as MoS₂, electrostatic gating induces superconductivity by doping carriers into the conduction band, achieving Tc around 2-4 K at densities of ~0.55 × 10¹⁴ cm⁻², with softening of acoustic modes driving the pairing. Confinement in the limits interlayer screening, enhancing electron- coupling and enabling a Berezinskii-Kosterlitz-Thouless indicative of vortex physics. Similar gate-induced effects occur in other TMDs like NbSe₂, where proximity to substrates modulates waves to favor superconducting order. Nanostructured systems, including 2D films and nanowires, exhibit enhanced through proximity effects or , where interfaces with conventional superconductors induce in semiconductors or topological materials. For instance, in hybrid Ge-Si nanowires, proximity coupling yields hard superconducting gaps up to magnetic fields of 250 mT, while in thin films of materials like YBa₂Cu₃O₇-δ suppresses competing orders to boost by several . These structures also host 2D Josephson junctions, such as van der Waals heterostructures with h-BN barriers, enabling supercurrent modulation via gating and revealing confinement-driven phase coherence over micrometer scales.

New Material Discoveries (2020-2025)

In 2024, researchers identified three novel superconducting materials, marking significant progress in the quest for higher critical temperatures under ambient conditions. Among these, atom-thin transition metal dichalcogenide (TMD) sheets, such as (InSe₂)ₓNbSe₂—a layered structure of indium selenide intercalated in niobium diselenide—exhibited unexpected superconducting properties at interfaces, challenging conventional models of pairing in two-dimensional systems. These discoveries highlight the potential of layered materials to enable superconductivity without extreme cooling, with implications for compact quantum devices. Additionally, Yale University experiments provided evidence for a novel type of superconductor, potentially involving chiral states or unconventional electron pairing mechanisms, observed through advanced imaging techniques on iron selenide materials doped with sulfur. Advancing into 2025, a team at achieved a breakthrough by stabilizing a new class of high-temperature superconductors at room pressure, retaining properties previously requiring megabar pressures and pushing critical temperatures closer to practical thresholds. This stabilization involved nickelate-based structures, building on prior high-pressure findings in high-Tc nickelates. In October 2025, scientists demonstrated superconductivity in for the first time using industry-compatible hyperdoping methods with , achieving zero-resistance states at low temperatures and opening pathways for integrating superconductors into silicon-based electronics. The HTSC-2025 dataset, released in mid-2025, compiled ambient-pressure high-temperature superconductors, featuring promising families like X₂YH₆ and perovskite-type MXH₃ structures, which exhibit enhanced electron-phonon coupling for elevated critical temperatures. Complementing this, MIT's tool, introduced in September 2025, leverages generative AI to predict superconducting materials by enforcing physical constraints, accelerating the discovery of candidates with exotic quantum properties. Amid these advances, the 2023 LK-99 claim of room-temperature superconductivity in a copper-substituted lead apatite was thoroughly debunked through replication efforts, revealing the observed effects as diamagnetism from impurities rather than true zero-resistance flow. Pursuit of room-temperature superconductivity continues via clathrate structures, with theoretical predictions suggesting hexagonal boron-rich variants could achieve high critical temperatures at ambient pressure through optimized phonon-mediated pairing.

Applications

Magnets and Levitation

Superconducting magnets leverage the persistent currents in superconductors to generate strong, stable without energy loss, enabling applications that require high-field homogeneity and reliability. These magnets typically employ Type II superconductors, which allow penetration in a controlled manner via , supporting fields far exceeding those of conventional electromagnets. In , niobium-titanium (NbTi) coils are widely used in (MRI) systems, producing fields from 1.5 (T) in standard clinical scanners to 7 T in research-grade units, providing exceptional due to the uniform fields sustained by zero-resistance loops. For even higher fields, niobium-tin (Nb3Sn) windings are incorporated, as seen in advanced prototypes, though NbTi remains dominant for most operational MRI due to its and cost-effectiveness. In , superconducting magnets form the backbone of large-scale accelerators like the (LHC) at , where NbTi dipole coils achieve an operational field of 8.3 T to guide proton beams in a 27-kilometer ring. These magnets operate in persistent mode, circulating currents indefinitely to maintain the field with minimal power input, a feat enabled by cryogenic cooling to 1.9 K. Upgrades for the High-Luminosity LHC incorporate Nb3Sn coils to reach 11 T in select dipoles, enhancing collision rates while preserving stability. Magnetic exploits the and in Type II superconductors for frictionless suspension. The Superconducting () train exemplifies this, using onboard NbTi superconducting magnets cooled to 4 K to induce repulsive forces via () against aluminum guideway coils, enabling up to 10 cm and operational speeds of 500-600 km/h with minimal . This repulsion arises from eddy currents generated in the guideway, interacting with the persistent fields to provide both and lateral stability, as demonstrated in test runs achieving 603 km/h. Superconducting magnetic energy storage (SMES) systems store energy in the magnetic fields of persistent-mode coils, releasing it rapidly for grid stabilization or power quality improvement. These devices use NbTi or high-temperature superconductors (HTS) like YBCO, cycled between charge and discharge states with efficiencies over 95%, and are projected to grow from a 2023 market value of USD 75 million to USD 168 million by 2030, driven by renewable integration needs. Flux pumping techniques enhance trapped fields in HTS bulks for compact, high-performance bearings by iteratively inducing currents to amplify flux density without direct electrical contacts. In YBCO-based trapped-field magnets, fields up to 1.35 T have been achieved at 77 K via multi-coil pumping arrangements, enabling repulsive forces for or rotor suspensions with load capacities exceeding 1 ton and rotational speeds over 10,000 rpm. These bearings benefit from the permanent trapping of flux lines in melt-textured HTS, providing passive stability and low losses for applications like rotors.

Electronics and Sensing

Superconducting electronics leverage quantum interference and tunneling phenomena to enable devices with exceptional performance in speed, sensitivity, and energy efficiency. Central to these applications are Josephson junctions, thin insulating barriers between two superconductors that allow pairs to tunnel coherently, leading to dissipationless supercurrents. Predicted in 1962, these junctions exhibit two key effects: the DC Josephson effect, where a supercurrent flows across the junction without applied voltage, and the AC Josephson effect, where an applied DC voltage induces an alternating supercurrent at a proportional to the voltage. The AC effect underpins precise voltage-to-frequency conversion, governed by the relation $2eV = h f, where e is the , V is the voltage across the junction, h is Planck's constant, and f is the oscillation ; this relation has been experimentally verified and forms the basis for superconducting voltage standards. A primary application of Josephson junctions is in superconducting quantum interference devices (SQUIDs), which exploit quantum interference in a loop containing one or two junctions to detect minute . DC SQUIDs, using two junctions, achieve ultrahigh sensitivity by measuring flux changes as small as a fraction of the \Phi_0 = h / 2e. These devices reach sensitivities down to $10^{-15} T \mathrm{Hz}^{-1/2}, enabling non-invasive brain imaging via (MEG), where they map neural activity with spatiotemporal resolution superior to . In , SQUIDs facilitate mineral and crustal studies by detecting subtle geomagnetic anomalies in magnetotelluric surveys, outperforming conventional magnetometers in low-signal environments. For digital computing, (RSFQ) uses Josephson junctions to process information via single quanta, enabling clock speeds exceeding 100 GHz with picowatt-level dissipation per gate—orders of magnitude faster and more efficient than counterparts. In RSFQ circuits, states are encoded as the presence or absence of a flux quantum within a short timing window, with junctions clocking operations through phase slips; prototypes have demonstrated shift registers operating at 20 GHz and arithmetic units at 50 GHz, positioning RSFQ for applications in high-performance and hybrid supercomputing. Superconducting qubits for quantum computing also rely on Josephson junctions to create anharmonic energy levels for quantum state manipulation. The qubit, a charge-insensitive , shunts a Josephson junction with a large to suppress charge noise, achieving times up to milliseconds and gate fidelities over 99.9%; it operates in the regime where the Josephson greatly exceeds the charging , enabling robust two-level system behavior. Similarly, the fluxonium qubit incorporates a Josephson junction in series with a superinductance, providing exponential suppression of noise and times exceeding 1 ms, with a multilevel tunable via external for improved scalability in quantum processors. These junction-based qubits form the core of leading superconducting quantum computers, demonstrating multi-qubit entanglement and error-corrected operations essential for fault-tolerant quantum information processing.

Energy Systems

Superconductivity plays a pivotal role in advancing energy systems by enabling lossless and enhancing the of and technologies. In electrical grids, superconducting materials eliminate ohmic losses, allowing for higher power densities and reduced needs. This is particularly valuable for urban and applications, where space and efficiency constraints are significant. Superconducting cables represent a breakthrough in , functioning as zero-resistance lines that minimize energy dissipation over long distances. The AmpaCity project in Essen, , exemplifies this technology, featuring a 1 km-long, 10 cable with a 40 MW that connects two city-center stations. Installed in 2014, it replaces a conventional 110 line, transporting five times more power than traditional s with negligible losses, thereby simplifying grid structures and reducing urban . The cable's high-temperature superconducting (HTS) design, cooled by , demonstrates economic viability for inner-city distribution, with potential to cut costs and enhance reliability. Fault current limiters (FCLs) based on superconductivity provide critical in grids by rapidly responding to short-circuit faults. These devices use HTS tapes that maintain zero resistance under normal operation at cryogenic temperatures, such as 70 K in . Upon detecting a fault, the surge in current heats the superconductor, inducing a rapid quench that transitions it to a high-resistance state, thereby limiting the fault current to safe levels—often reducing it by factors of 10 or more—without interrupting flow. This allows existing to isolate the fault quickly while enabling faster recovery as the material recools and regains superconductivity. In applications like high-voltage and grids, resistive SFCLs (RSFCLs) relieve , support meshing for better reconfigurability, and avoid costly substation upgrades, as demonstrated in prototypes tested near and in EU-funded initiatives like FASTGRID. In generation, high-temperature superconductors enable lighter and more efficient s by replacing windings with HTS materials like ceramic-metallic tapes, which operate at temperatures around -196°C using cryogenic cooling. This reduces generator weight by up to 40% and electrical losses, allowing for compact designs that produce higher power outputs—such as over 3 MW per unit—while using far less rare-earth materials than permanent magnet alternatives. A landmark example is the ECOSWING project, which in 2018 installed the world's first full-scale HTS off Denmark's coast in Thyborøn, powering approximately 1,000 homes with enhanced energy flow densities. These advancements promise 20-30% cost reductions for offshore installations greater than 10 MW, though challenges like wire affordability and cooling reliability persist. Superconductivity is essential for large-scale production in fusion reactors, where strong confine for sustained reactions. The International Thermonuclear Experimental Reactor () incorporates Nb₃Sn superconductors in its central , a key component comprising six independent coil packs that induce and sustain a 15 MA current for 300-500 seconds. Each module uses approximately 6,000 meters of cable-in-conduit Nb₃Sn , heat-treated to form the superconducting , enabling a maximum of and storing 6.4 GJ of . This vertical stack, weighing 1,000 tonnes and standing meters tall, is crucial for shaping and , marking a high-impact application of low-temperature superconductors in clean generation.

Recognition

Nobel Prizes

The discovery of superconductivity was recognized early in the Nobel Prizes in Physics. In 1913, was awarded the prize "for his investigation of the properties of matter at low temperatures, which led, inter alia, to the production of ." His work at the University of enabled the liquefaction of , allowing experiments at temperatures near , where he observed zero electrical in mercury in 1911, marking the first identification of superconductivity. Theoretical understanding advanced significantly with the 1972 Nobel Prize in Physics, shared by , Leon N. Cooper, and "for their jointly developed theory of superconductivity, usually called the ." Developed in the mid-1950s at the University of Illinois, the BCS theory explained superconductivity as arising from the formation of Cooper pairs of electrons mediated by lattice vibrations (phonons), providing a microscopic quantum mechanical framework that predicted key properties like the energy gap and critical temperature. Experimental insights into quantum tunneling in superconductors earned the 1973 Nobel Prize in Physics. Half was awarded jointly to and "for their experimental discoveries regarding tunneling phenomena in semiconductors and superconductors, respectively," while the other half went to Brian D. Josephson "for his theoretical predictions of the properties of a supercurrent through a tunnel barrier (Josephson effects)." Esaki's work at in the late demonstrated tunneling in p-n junctions, leading to the , while Giaever at observed tunneling in superconductor-insulator-superconductor structures, revealing the superconducting energy gap. Josephson, then a graduate student, predicted in 1962 that a supercurrent could flow without voltage across a thin insulating barrier between superconductors and that an AC current would arise at a finite voltage, enabling applications like SQUIDs for precise measurements. A breakthrough in came with the 1987 , awarded to J. and K. Müller "for their important break-through in the discovery of superconductivity in ceramic materials." Working at IBM's Zurich Research Laboratory, they reported in 1986 the observation of superconductivity at 35 in a barium-doped , the first above temperature (77 ) and in a non-metallic ceramic, challenging the prior focus on elemental metals and alloys and sparking global research into high-temperature . Further theoretical contributions were honored in 2003 with the Nobel Prize in Physics. Vitaly L. Ginzburg and Alexei A. Abrikosov shared half "for their pioneering contributions to the theory of superconductors and superfluids," while Anthony J. Leggett received the other half "for his pioneering contributions to the theory of superfluids." Ginzburg and Lev Landau's 1950 phenomenological theory described superconductivity using macroscopic wave functions, predicting behaviors near critical temperatures. Abrikosov extended this in the 1950s to type-II superconductors, introducing vortex lattices that allow magnetic flux penetration, essential for high-field applications like MRI magnets. Leggett's work on superfluidity, including macroscopic quantum coherence, paralleled and informed superconductivity theories.

Other Milestones

In 1961, researchers at Bell Laboratories, led by J. E. Kunzler, developed the first practical Nb₃Sn superconducting wire by filling niobium tubes with powders of and tin, then drawing and heat-treating them into ribbons capable of sustaining high above 8 T, paving the way for advanced high-field applications. This breakthrough marked a significant advancement over bulk Nb₃Sn samples, enabling the production of flexible wires with improved current-carrying capacity for electromagnets. The discovery of accelerated dramatically in the late 1980s, with the critical temperature (T_c) record reaching 92 K in (YBCO) in 1987, achieved by replacing with in cuprate structures, allowing operation above temperatures. This progression from the initial 35 K in lanthanum-based cuprates in 1986 represented a pivotal shift, making practical cooling feasible and spurring global research into ceramic superconductors. In 2015, (H₃S) set a new benchmark for conventional superconductivity with a T_c of 203 K under of 155 GPa, confirmed through resistivity and measurements, demonstrating phonon-mediated pairing in hydrides. This record, the highest for any superconductor at the time, highlighted the potential of compressed hydrogen-rich compounds to approach room-temperature superconductivity, though practical applications remain challenged by extreme pressures. Advancing toward ambient conditions, in February 2025, researchers at and achieved the first stabilization of a new class of high-T_c superconductors—nickelate-based materials—at room pressure, using thin-film epitaxial growth techniques that apply lateral compressive strain from substrates to induce the infinite-layer structure without high-pressure synthesis. This milestone enables further exploration of infinite-layer nickelates without diamond anvil cells, bridging the gap to pressure-free high-T_c systems. Complementing these efforts, in October 2025, scientists at demonstrated the first superconducting germanium using industry-compatible fabrication methods, such as with gallium hyperdoping on semiconductor-grade wafers, achieving T_c of 3.5 K while enabling direct integration with existing silicon- electronics for hybrid quantum devices. This integration breakthrough transforms germanium from a conventional into a platform for scalable superconducting circuits, potentially revolutionizing architectures.

References

  1. [1]
    DOE Explains...Superconductivity - Department of Energy
    Superconductivity is the property of certain materials to conduct direct current (DC) electricity without energy loss when they are cooled below a critical ...
  2. [2]
    Superconductivity
    A superconductor is a material which exhibits the following two properties: Fig. 1(a) Vanishing of electrical resistivity below a critical temperature Tc, ...Missing: definition | Show results with:definition
  3. [3]
    The discovery of superconductivity - Physics Today
    Sep 1, 2010 · A few months earlier, the Leiden team had discovered that lead and tin were also superconductors, with transition temperatures near 6 K and 4 K, ...
  4. [4]
    The discovery of superconductivity - Europhysics News
    One hundred years ago, on April 8, 1911, Heike Kamerlingh Onnes and his staff at the Leiden cryogenic laboratory were the first to observe superconductivity [1] ...
  5. [5]
    [PDF] Onnes 1911 - Physics
    In the thing paper, using a specially built cryostat with Itz leads, the change in p was found to abrupto. HEIKE KAMERLINGH ONNES good. 0035% gast. (. 94360.
  6. [6]
    [PDF] Story of Superconductivity - Indian Academy of Sciences
    Kamerlingh Onnes pre-cooled helium to this low temperature using liquid hydrogen. ... James Dewar invented the vacuum-insulated, silver-plated glass Dewar flask ...
  7. [7]
    This Month in Physics History | American Physical Society
    However, the story of superconductivity begins in 1911, when Heike Kamerlingh Onnes first discovered the phenomenon. ... He published a second paper in May, and ...
  8. [8]
    [PDF] landmark experiments in superconductivity
    Great must have been the excitement with the discovery, in late 1912, of superconductivity in lead and tin (and in not ... Alas, Kamerlingh Onnes's hope was soon ...
  9. [9]
    The electromagnetic equations of the supraconductor - Journals
    (1967) An extension of the London equations of superconductivity, Physica, 10.1016/0031-8914(67)90090-0, 37:4, (504-538), Online publication date: 1-Jan-1967.
  10. [10]
    [PDF] Meissner and Ochsenfeld revisited - Physics Courses
    The paper by Meissner and Ochsenfeld which ap- pears below in translation was first published in Die. Naturwissenschaften in November 1933. The dis- covery ...Missing: discovery Walther Robert
  11. [11]
    The second-order phase transition and the Ehrenfest relations
    of the second-order phase transition of the type envisaged by Ehrenfest are described, one a super- conducting and the other a ferroelectric phase transition.
  12. [12]
    Phys. Rev. B 106, 174518 (2022) - Heat capacity double transitions ...
    Nov 29, 2022 · Standard superconductors display a ubiquitous discontinuous jump in the electronic specific heat at the critical superconducting transition ...
  13. [13]
    [PDF] Survey of superconductive materials and critical evaluation of ...
    Oct 15, 2009 · This publication includes all data on superconductive materials intercepted through March 1975. Dala on the bulk elements have been ...<|control11|><|separator|>
  14. [14]
    Specific heat in superconductors
    ... specific heat which is defined as C = −T (∂ S/∂ T)H = T (∂ 2g/∂ T 2)H has a discontinuous jump, therefore the thermodynamics transition is second order.
  15. [15]
    Nonlinearity of the field induced by a rotating superconducting shell
    Dec 10, 2004 · A rotating superconductor generates a magnetic field. This field is usually known as the “London moment” since a good approximation for it ...
  16. [16]
    [PDF] Observation of the London Moment and Trapped Flux in Precision ...
    The spherical gyroscopes are made of fused quartz and are sputter coated with a Nb film. The superconducting Nb film generates the. London magnetic dipole ...
  17. [17]
    Observation of the London moment and trapped flux in precision ...
    The London-moment readout has been observed in flight quality gyroscopes and it has been demonstrated that it is possible to reduce magnetic field trapped ...
  18. [18]
    On the Theory of superconductivity - Inspire HEP
    ) 1950. 19 pages. Part of Collected Papers of L.D. Landau, 546-568. Published ... In the present paper a theory is constructed which is free from these faults.
  19. [19]
    On the Magnetic Properties of Superconductors of the Second Group
    On the Magnetic Properties of Superconductors of the Second Group. A.A. Abrikosov(. Moscow, Inst. Phys. Problems. ) Nov, 1956. 9 pages. Published in: Sov.
  20. [20]
    Theory of Superconductivity | Phys. Rev.
    A theory of superconductivity is presented, based on the fact that the interaction between electrons resulting from virtual exchange of phonons is attractive.Missing: equations | Show results with:equations
  21. [21]
    [1911.05065] Eliashberg Theory: a short review - arXiv
    Nov 12, 2019 · Eliashberg theory is a theory of superconductivity that describes the role of phonons in providing the attractive interaction between two electrons.
  22. [22]
    A Spin Fluctuation Model for D-wave Superconductivity - arXiv
    Jan 9, 2002 · In this review we demonstrate that superconducting pairing, mediated by the exchange of spin fluctuations, is a viable alternative to conventional phonon- ...
  23. [23]
    [1106.1213] The Challenge of Unconventional Superconductivity
    Jun 6, 2011 · Developing a rigorous theory for any of these classes of materials has proven to be a difficult challenge, and will continue to be one of the major problems in ...
  24. [24]
    [PDF] SUPERCONDUCTIVITY - Portland State University
    property of complete disappearance of electrical resistance in solids when they are cooled below a characteristic temperature. This temperature is called.Missing: key | Show results with:key
  25. [25]
    None
    ### Summary of Type I vs Type II Superconductors, Magnetic Fields, Hc1, Hc2, Vortices, Intermediate State, Examples
  26. [26]
    Equilibrium Topology of the Intermediate State in Type-I ...
    Jun 21, 2007 · Flux tubes represent the equilibrium topology of the intermediate state in type-I superconductors, as concluded from the study.
  27. [27]
    The Intermediate State of Supraconductors - Nature
    A body in the so-called 'intermediate' state consists of alternating layers of the supraconducting and normal phases.<|separator|>
  28. [28]
    Alexei Abrikosov – Nobel Lecture - NobelPrize.org
    Abrikosov shows how the theory of type II superconductors, where the magnetic field penetrates, was developed. All unusual properties of the high-temperature ...<|separator|>
  29. [29]
    Nobel Lecture: Type-II superconductors and the vortex lattice
    Dec 2, 2004 · Now they are called type-II superconductors. I published my derivation in the Russian journal Doklady Akademii Nauk SSSR in 1952 ( Abrikosov, ...
  30. [30]
    [PDF] Ginzburg-Landau Theory: Simple Applications
    (3) in the same limit, the “upper critical field” at which superconductivity is com- pletely destroyed is ∼ φ0/ξ2, more precisely. Hc2 = φ0/2πξ2(T). (50) ...
  31. [31]
    Superconducting Critical Temperature - ScienceDirect.com
    Tc, superconducting critical temperature, can be defined as the temperature at which a material exhibits superconductivity, which is often calculated using ...
  32. [32]
    Using ultra-low temperatures to understand high ... - TU Wien
    Jul 21, 2021 · " In the case of high-temperature superconductors ... temperatures of only one millikelvin, the strange metal turns into a superconductor.
  33. [33]
    The maximum T c of conventional superconductors at ambient ...
    Sep 10, 2025 · It is true that physical laws do not restrict Tc to go beyond 100–120 K, but in practice our data show that the experimental realization of a ...
  34. [34]
    Standards for Measurement of the Critical Fields of Superconductors
    The critical field is temperature dependent, with an approximately parabolic dependence on T/Tc. Below the critical field the flux density is zero in the bulk ...
  35. [35]
    [PDF] Critical current densities in superconducting materials
    Critical current density (JC) is the current density at which a superconductor loses its zero resistance property, and the material develops a 'flux-flow ...
  36. [36]
    [PDF] The critical current density in type-II superconducting bulk materials
    Mar 1, 2012 · In this paper we investigate the critical current density c. J for type-II superconducting bulk materials based on the quantum theory of vortex ...
  37. [37]
    Phase diagram of a type I superconductor in the H-T plane. The...
    Phase diagram of a type I superconductor in the H-T plane. The transition is first order for finite field. The blue lines show the 4 steps in the Gorter cycle ...
  38. [38]
    Current distribution across type II superconducting films - Nature
    Jan 29, 2018 · With increasing current the flux-free region narrows and the critical state moves in towards the centre. ... critical current density of flat ...
  39. [39]
  40. [40]
    [PDF] Thermodynamic Properties of κ-(BEDT-TTF)2X Salts
    Jun 29, 2012 · The well known high-Tc superconductors among organic systems are κ-(BEDT-TTF)2Cu(NCS)2 (Tc = 10.4 K) and κ-(BEDT-TTF)2Cu[N(CN)2]Br (Tc = 11.6 K) ...<|separator|>
  41. [41]
    [PDF] Heavy-fermion superconductivity in CeCoIn5 at 2.3 K - arXiv
    We report the observation of heavy-fermion superconducitivity in CeCoIn5 at Tc=2.3 K. When compared to the pressure-induced Tc of its cubic relative CeIn3 ...
  42. [42]
    The isotope effect in superconductivity | Physics Today
    Ever since Kamerlingh‐Onnes discovered in 1911 that the electrical resistance of mercury abruptly vanished at a temperature just a few degrees above the ...
  43. [43]
    [PDF] Failed theories of superconductivity - arXiv
    Of the “failures” to explain superconductivity during the early post quantum mechanics days, two theories are noteworthy for their elegance and the ...
  44. [44]
    [PDF] Ginzburg-Landau Equations - ResearchGate
    Ginzburg and Landau start their remarkable paper of 1950 by stating that London's existing phenomeno- logical theory, despite successfully describing several as ...
  45. [45]
    Superconductivity of Sn | Phys. Rev. - Physical Review Link Manager
    Superconductivity of N b 3 Sn. B. T. Matthias, T. H. Geballe, S. Geller, and ... The superconducting transition temperature of N b 3 Sn at 18°K is the ...Missing: original paper discovery
  46. [46]
    High-field superconductivity
    Only after Eugene. Kunzler's 1958 discovery of high-field superconductivity at 8.8 T in Nb3Sn was the dream of constructing high- field magnets rekindled. Much ...
  47. [47]
    [PDF] A History of Nb3Sn and Related A15 Wires - Florida State University
    The first strand carried 50 times higher Jc than the bulk sample,1 and their best PIT strand (with 10% more Sn than was required to fully react the. Nb powder ...
  48. [48]
    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.
  49. [49]
    Possible highT c superconductivity in the Ba−La−Cu−O system
    Cite this article. Bednorz, J.G., Müller, K.A. Possible highT c superconductivity in the Ba−La−Cu−O system. Z. Physik B - Condensed Matter 64, 189–193 (1986).Missing: Nobel Prize
  50. [50]
    Press release: The 1987 Nobel Prize in Physics - NobelPrize.org
    Last year, 1986, Bednorz and Müller reported finding superconductivity in an oxide material at a temperature 12°C higher than previously known. This was the ...
  51. [51]
    April 1986: Bednorz and Müller Trigger Avalanche of High ...
    Mar 16, 2023 · April 1986: Bednorz and Müller Trigger Avalanche of High-Temperature Superconductivity Research. The duo's work energized the 1987 APS meeting.
  52. [52]
    Transport and optical studies of single crystals of the 80-K Bi–Sr–Ca ...
    Mar 17, 1988 · Recently, Maeda et al.4 announced the discovery of new superconducting oxides in the system Bi–Sr–Ca–Cu–O, with Tc above 100 ...
  53. [53]
    Bulk superconductivity at 120 K in the Tl–Ca/Ba–Cu–O system | Nature
    Mar 10, 1988 · Here we report stable and reproducible bulk superconductivity with an onset at 120 K and zero resistance above 100 K in the Tl–Ca/Ba–Cu–O system.
  54. [54]
    [PDF] J. Georg Bednorz and K. Alex Müller - Nobel Lecture
    had investigated a Ba-La-Cu oxide with perovskite structure exhibiting metal- lic conductivity in the temperature range between 300 and - 100º C [1.16]. The.
  55. [55]
    Feshbach hypothesis of high-Tc superconductivity in cuprates - Nature
    Jan 2, 2025 · High-temperature superconductivity in the cuprate compounds has a long and storied history and has remained an active field of research since ...
  56. [56]
    [PDF] High Temperature Superconductivity Abstract - arXiv
    The high Tc cuprate superconductors have layered perovskite-like crystal structures which consist of conducting CuO2 planes separated by layers comprised of ...
  57. [57]
    Quasiparticle mass enhancement approaching optimal doping in a ...
    ... superconductivity near optimal doping in the cuprates has impeded access to the metallic ground state. The Fermi surface in underdoped cuprates is known to ...
  58. [58]
    Interplay between hole superconductivity and quantum critical ...
    Mar 20, 2025 · Antiferromagnetic spin fluctuations are the most promising candidate as the pairing glue of high critical temperature (T c ) superconductivity in cuprates.Missing: Hc jc
  59. [59]
    The Puzzle of High Temperature Superconductivity in Layered Iron ...
    May 24, 2010 · In short order, other Fe-based superconductors with the same or related crystal structures were discovered with Tc up to 56 K. Many experiments ...
  60. [60]
    Pairing mechanism in multiband superconductors | Scientific Reports
    May 4, 2020 · In this Article, we investigate the pairing mechanism in multiband iron-based superconductors by RG analysis. ... pairing in a four-band model of ...
  61. [61]
    Role of the orbital degree of freedom in iron-based superconductors
    Oct 18, 2017 · In this review, we summarize angle-resolved photoemission spectroscopy (ARPES) measurements across various FeSC families that have been reported in literature.
  62. [62]
    [PDF] Conventional superconductivity at 203 K at high pressures - arXiv
    The high Tc superconductivity most likely is due to H3S which is formed from H2S under its decomposition under pressure. Even higher Tc, room temperature ...
  63. [63]
    Pressure and high-Tc superconductivity in sulfur hydrides - Nature
    May 11, 2016 · Recently, TC ≈ 203 K was confirmed in H3S formed in the decomposition of H2S under pressure. The present work is mainly concerned with the ...
  64. [64]
    [PDF] arXiv:0808.3408v1 [cond-mat.supr-con] 25 Aug 2008
    Aug 25, 2008 · The dependence of their basal-plane critical current properties on orientation in a magnetic field can pose important design problems. Here we ...
  65. [65]
    Enhanced flux pinning isotropy by tuned nanosized defect network ...
    Oct 28, 2019 · Materials science challenges for high-temperature superconducting wire. ... pinning in high temperature superconducting nanocomposite wires ...
  66. [66]
    High tolerance of the superconducting current to large grain ... - Nature
    Aug 23, 2024 · However, the grain boundaries (GBs) with larger misorientation angles in HTSs act as weak links, leading to a fatal reduction in Jc. In the case ...
  67. [67]
    The behavior of grain boundaries in the Fe-based superconductors
    Oct 4, 2011 · We conclude that grain boundaries in FBS are weak linked in a qualitatively similar way to grain boundaries in the cuprate superconductors.
  68. [68]
    Colossal thermomagnetic response in the exotic superconductor ...
    Dec 1, 2014 · Here we report on the observation of a colossal Nernst signal due to the superconducting fluctuations in the heavy-fermion superconductor URu 2 Si 2.
  69. [69]
    Unfolding the physics of URu2Si2 through silicon to ... - Nature
    Feb 19, 2016 · The heavy fermion intermetallic compound URu 2 Si 2 exhibits a hidden-order phase below the temperature of 17.5 K, which supports both anomalous metallic ...Missing: discovery | Show results with:discovery
  70. [70]
    Superconducting fluctuations in organic molecular metals enhanced ...
    Dec 2, 2013 · We reveal how the Nernst effect, a sensitive probe of superconducting phase fluctuations, evolves in the regime of extreme Mott criticality.
  71. [71]
    Theory of the evolution of superconductivity in Sr 2 RuO 4 under ...
    Mar 2, 2017 · Sr2RuO4 is a leading candidate for chiral p-wave superconductivity. The detailed mechanism of superconductivity in this material is still ...
  72. [72]
    [1210.2992] Chiral P-Wave Order in Sr_2RuO_4 - arXiv
    Oct 10, 2012 · In this paper, I review a broad range of experiments on Sr2RuO4 that are sensitive to p-wave pairing, triplet superconductivity and time-reversal symmetry ...
  73. [73]
    Atomic-scale interface engineering of Majorana edge modes in a 2D ...
    Jul 26, 2019 · A hallmark of 2D topological superconductors is the existence of dispersive, in-gap Majorana modes that are spatially located along the edges of ...<|separator|>
  74. [74]
    Majorana modes with side features in magnet-superconductor ...
    Dec 19, 2022 · Magnet-superconductor hybrid (MSH) systems represent promising platforms to host Majorana zero modes (MZMs), the elemental building blocks for fault-tolerant ...
  75. [75]
    Topological order and non-Abelian statistics in noncentrosymmetric $s
    Jun 5, 2008 · Also, note that the Majorana fermions in vortices of the NC superconductors do not require the existence of half quantum vortices, i.e., ...
  76. [76]
    Observation of half-quantum flux in the unconventional ... - Science
    Oct 11, 2019 · We report the observation of half-integer magnetic flux quantization in mesoscopic rings of superconducting β-Bi 2 Pd thin films.
  77. [77]
    Spin-Triplet Pairing State Evidenced by Half-Quantum Flux in a ...
    Apr 22, 2020 · In this work, we perform a phase-sensitive experiment to examine the pairing state of noncentrosymmetric superconductor -BiPd.
  78. [78]
    Topological Phase Transitions in Multicomponent Superconductors
    Nov 3, 2017 · The chiral Majorana modes bound to the half-quantum vortices and the surfaces form a closed contour, shown in Fig. 3 . This chiral Majorana mode ...
  79. [79]
    Signatures of chiral superconductivity in rhombohedral graphene
    May 22, 2025 · Here we report the discovery of robust unconventional superconductivity in rhombohedral tetralayer and pentalayer graphene without moiré superlattice effects.
  80. [80]
    Promotion of superconductivity in magic-angle graphene multilayers
    Sep 29, 2022 · Graphene bilayers and trilayers consisting of monolayers twisted at just the right angle have been shown to be superconducting.
  81. [81]
    Unconventional Superconductivity in Magic-Angle Twisted Trilayer ...
    Apr 20, 2021 · Magic-angle twisted trilayer graphene (MATTG) recently emerged as a highly tunable platform for studying correlated phases of matter, such as ...
  82. [82]
    Gated tuned superconductivity and phonon softening in monolayer ...
    Sep 12, 2017 · The small applied gate voltage and slightly inhomogeneous superconductivity may be possible reasons and will require further investigations. To ...
  83. [83]
    [PDF] Gate-induced Superconductivity in atomically thin MoS2 crystals.
    Upon sweeping the gate voltage VG for positive values, electrons are accumulated, resulting in an increase of surface conductivity (Fig. 1c). At room ...
  84. [84]
    Hard Superconducting Gap and Diffusion-Induced Superconductors ...
    Nov 26, 2019 · We show a hard superconducting gap in a Ge–Si nanowire Josephson transistor up to in-plane magnetic fields of 250 mT, an important step toward creating and ...
  85. [85]
    Interface-enhanced superconductivity in monolayer 1T′-MoTe 2 on ...
    Aug 25, 2023 · This study shows that interfacial engineering is an efficient way to tune monolayer 1T′-MoTe 2 film into superconducting states.
  86. [86]
    Two-Dimensional van der Waals Superconductor Heterostructures
    Mar 12, 2024 · For example, the Josephson junction, a specific structure in which a nonsuperconducting barrier is inserted between two proximity-coupled SCs, ...Introduction · Fabrication Strategies · Device Performance · Applications
  87. [87]
    MIT Discovers Three New 2D Superconductors That Challenge ...
    Feb 14, 2025 · Two of MIT's three newly discovered superconductive materials are composed of graphene and transition metal dichalcogenides (TMDs) sheets; ...
  88. [88]
    New superconducting material discovered in transition-metal ...
    Jan 19, 2024 · New superconducting material discovered in transition-metal dichalcogenides materials · Thin film reveals origins of pre-superconducting phase.
  89. [89]
    A basic science breakthrough: Evidence of a new type ... - YaleNews
    Nov 13, 2024 · Yale physicist Eduardo H. da Silva Neto led an experiment that supports the existence of a new type of superconductor.<|control11|><|separator|>
  90. [90]
    In a first, researchers stabilize a promising new class of high ...
    Feb 4, 2025 · Researchers have made a significant step in the study of a new class of high-temperature superconductors: creating superconductors that work at room pressure.
  91. [91]
    Scientists turn common semiconductor into a superconductor
    Researchers have made germanium superconducting for the first time, a feat that could transform computing and quantum technologies.
  92. [92]
    Superconducting germanium made with industry-compatible methods
    Superconducting germanium made with industry-compatible methods. 31 October 2025. Gloved hands hold a large silicon disc. A semiconductor wafer similar to ...
  93. [93]
    HTSC-2025: A Benchmark Dataset of Ambient-Pressure High ... - arXiv
    Jun 4, 2025 · HTSC-2025 is a benchmark dataset of ambient-pressure high-temperature superconductors for AI-driven critical temperature prediction, including ...Missing: hydrides X2YH6 MXH3
  94. [94]
    New tool makes generative AI models more likely to ... - MIT News
    Sep 22, 2025 · A new tool called SCIGEN allows researchers to implement design rules that AI models must follow when generating new materials.
  95. [95]
    LK-99 isn't a superconductor — how science sleuths solved ... - Nature
    Aug 16, 2023 · Researchers seem to have solved the puzzle of LK-99. Scientific detective work has unearthed evidence that the material is not a superconductor.
  96. [96]
    Superconductive magnet design - Questions and Answers ​in MRI
    The conductor used in nearly all modern superconducting MR scanners is niobium-titanium (NbTi) that becomes superconductive below 9.4°K. Each wire is ...
  97. [97]
    Conductors for commercial MRI magnets beyond NbTi
    Advantages of superconducting MRI systems include better performance, the highest temporal and spatial homogeneity of the magnetic field, high signal-to-noise ...
  98. [98]
    [PDF] Superconducting Magnets and RF Cavities for the LHC
    Jun 19, 2000 · The heart of the 8.3 T main dipole magnet is a coil of 15.3 mm wide “Rutherford” conductor wound on a circular mandrel of diameter 56 mm, in ...<|separator|>
  99. [99]
    [PDF] High-field Magnet Development toward the High Luminosity LHC
    In addition, Nb3Sn 11 T dipoles will replace, in few location, the main 8.3 T dipoles to allow the insertion of additional collimators needed for HL-LHC ...
  100. [100]
    [PDF] SUPERCONDUCTING MAGLEV (SCMAGLEV) - JR Central
    This leading edge Japanese technology is the next generation of super fast train travel. Propulsion System. By passing current through the Propulsion Coils ...
  101. [101]
    Superconducting Maglev - Technology
    The current SCMAGLEV train, known as the L0 (L-zero) holds the Guinness record for the world's fastest maglev train – clocking in at 374.68 mph (603 km/h)!.
  102. [102]
    Superconducting Magnetic Energy Storage (SMES) Market
    The global Superconducting Magnetic Energy Storage Systems market size was exhibited at USD 75.3 million in 2023 and is projected to hit around USD 167.72 ...
  103. [103]
    Review on high-temperature superconducting trapped field magnets
    Nov 25, 2024 · Flux pumping involves sweeping a smaller magnetic field across the superconductors in a wave, inducing a consistent current direction ...
  104. [104]
    Review of high temperature superconducting flux pumps
    High temperature superconducting (HTS) magnets conduct DC currents ranging from hundreds to tens of thousands of amperes. To achieve such DC output ...
  105. [105]
    [PDF] SQUID-based multichannel system for Magnetoencephalography
    Oct 17, 2013 · Here we present a multichannel system based on superconducting quantum interference devices (SQUIDs) for magnetoencephalography (MEG).
  106. [106]
    HTS rf-SQUID applications to geophysical exploration methods
    SQUID systems offer an alternative to commonly used vector magnetometers (induction coils, fluxgate) in geophysical exploration methods like magnetotellurics ( ...
  107. [107]
  108. [108]
  109. [109]
    Charge insensitive qubit design derived from the Cooper pair box
    Feb 28, 2007 · Here, we introduce a new type of superconducting qubit called the "transmon." Unlike the charge qubit, the transmon is designed to operate in a ...
  110. [110]
    [0906.0831] Fluxonium: single Cooper pair circuit free of charge offsets
    Jun 4, 2009 · Access Paper: View a PDF of the paper titled Fluxonium: single Cooper pair circuit free of charge offsets, by Vladimir E. Manucharyan and 7 ...
  111. [111]
  112. [112]
    Operation of Longest Superconducting Cable Worldwide Started
    In Essen, the 10,000 volt superconducting cable replaces a conventional 100,000 volt line. Prior to the AmpaCity project, technical feasibility and economic ...
  113. [113]
    Superconducting fault current limiter protects national electricity grids
    Apr 23, 2021 · Today, engineers achieve the same result of limiting (switching off) the fault current by various passive and non-ideal means. In 2014, the ...
  114. [114]
    Resistive superconducting fault current limiter - SuperGrid Institute
    Sep 9, 2025 · The RSFCL relieves congestion on your AC network by limiting short circuit currents in your substation without upgrading your existing switchyard.
  115. [115]
    A superconductive breakthrough for wind energy
    Aug 29, 2018 · The world's first superconducting wind turbine will be installed this year off the coast of Denmark - a landmark achievement by an EU-funded project.
  116. [116]
    Magnets - ITER
    For the most technically challenging raw material—the niobium-tin (Nb3Sn) superconducting strands used in ITER's toroidal field and central solenoid magnet ...
  117. [117]
    The Nobel Prize in Physics 1913 - NobelPrize.org
    The Nobel Prize in Physics 1913 was awarded to Heike Kamerlingh Onnes for his investigations on the properties of matter at low temperatures.
  118. [118]
    Heike Kamerlingh Onnes – Facts - NobelPrize.org
    The same year he received the Nobel Prize for the production of liquid helium, he also discovered the phenomenon of superconductivity—when current passes ...
  119. [119]
    The Nobel Prize in Physics 1972 - NobelPrize.org
    ... Cooper and John Robert Schrieffer "for their jointly developed theory of superconductivity, usually called the BCS-theory". To cite this section. MLA style ...
  120. [120]
    Press release: The 1972 Nobel Prize in Physics - NobelPrize.org
    The award is given for their jointly developed theory of superconductivity, usually called the BCS-theory.
  121. [121]
    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.
  122. [122]
    Press release: The 1973 Nobel Prize in Physics - NobelPrize.org
    Half of the prize is divided equally between Esaki and Giaever for their experimental discoveries regarding tunneling phenomena in semiconductors and ...
  123. [123]
    The Nobel Prize in Physics 1987 - NobelPrize.org
    The Nobel Prize in Physics 1987 was awarded jointly to J. Georg Bednorz and K. Alexander Müller for their important break-through in the discovery of ...
  124. [124]
    [PDF] Nobelpris i fysik 2003 (Abrikososv, Ginzburg, Leggett)
    Oct 7, 2003 · Abrikosov's theory of superconductors in a magnetic field created a new field of physics – the study of type-II superconductors. After the ...
  125. [125]
    [PDF] Nb3Sn Wires and Cables for High-Field Accelerator Magnets
    These developments on wires and cables were crucial for the first reproducible accelerator-quality Nb3Sn dipole and quadrupole magnets and their scale-up in the ...Missing: 1958 | Show results with:1958
  126. [126]
    Nb3Sn Wires and Cables for High-Field Accelerator Magnets
    Aug 31, 2019 · The first laboratory attempt to produce Nb3Sn wires was in 1961 at Bell Laboratories (Kunzler et al. 1961) by filling Nb tubes with crushed ...
  127. [127]
    First-Hand:Discovery of Superconductivity at 93 K in YBCO
    In February of 1987, scientists at the University of Houston and The University of Alabama in Huntsville separately announced the discovery of superconductivity ...
  128. [128]
    Scientists Create New Type of Semiconductor that Holds ... - NYU
    “Germanium is already a workhorse material for advanced semiconductor technologies, so by showing it can also become superconducting under ...