Fact-checked by Grok 2 weeks ago

Superfluidity

Superfluidity is a quantum mechanical phenomenon in which certain fluids, at temperatures near absolute zero, exhibit zero viscosity and the ability to flow without friction through narrow channels or around obstacles, defying classical fluid dynamics. This state of matter was first observed in liquid helium-4 (^4He) below the lambda transition temperature of 2.17 K (the so-called lambda point), where the fluid, known as helium II, displays remarkable properties such as anomalously high thermal conductivity and the capacity to climb container walls via a thin superfluid film known as the Rollin film. The discovery occurred in 1938, when physicists Pyotr Kapitza, John F. Allen, and Donald Misener independently measured the drastic drop in viscosity of liquid helium below this temperature, coining the term "superfluidity" to describe the frictionless flow. Liquid helium-4 had been liquefied in 1908 by Heike Kamerlingh Onnes, but its peculiar behaviors were only systematically investigated in the 1930s. The theoretical framework for superfluidity emerged from , with proposing in 1938 that it arises from Bose-Einstein condensation (BEC) of helium atoms, which are bosons, allowing macroscopic occupation of the and coherent flow without dissipation. This two-fluid model, developed by in 1941, describes helium II as a mixture of a viscous "normal" fluid component and an inviscid "superfluid" component, with their proportions varying with temperature; the superfluid fraction approaches 100% as temperature nears 0 K. Experimental hallmarks include the fountain effect (self-siphoning up narrow tubes) and quantized vortices, where circulation is restricted to discrete multiples of h/(2m), with h as Planck's constant and m the helium atom mass. Superfluidity also manifests in liquid helium-3 (^3He), a fermionic system, but requires much lower temperatures around 2.5 due to the need for Cooper pairing of atoms to form bosonic pairs before ; this was discovered in 1972 by , David Lee, and Robert Richardson, earning them the 1996 . Unlike ^4He, ^3He superfluids exhibit multiple phases (A and B) with anisotropic pairing, leading to complex topological defects and exotic excitations. Beyond helium, superfluidity has been realized in ultracold gases since 2005, where fermionic atoms form superfluids via tunable interactions, providing clean analogs for studying and matter. These properties make superfluidity a cornerstone of low-temperature physics, bridging condensed phenomena like —where electrons pair similarly—and enabling applications such as cryogenic cooling for particle detectors and MRI magnets, as well as probing fundamental quantum behaviors in macroscopic systems.

Definition and Fundamental Properties

Core Definition

Superfluidity is a of in which a flows with zero , enabling persistent motion without due to the establishment of long-range quantum across a many-body system at low temperatures. This quantum phenomenon manifests as a macroscopic occupation of a single by the system's particles, distinguishing it fundamentally from classical where arises from dissipative particle interactions. In bosonic systems, superfluidity typically emerges through Bose-Einstein condensation, where bosons follow Bose-Einstein statistics and a macroscopic number condense into the lowest energy state, fostering coherent collective behavior. For fermionic systems, which obey Fermi-Dirac statistics and cannot all occupy the same state due to the , superfluidity requires the formation of Cooper pairs—bound pairs of fermions that effectively behave as composite bosons and undergo condensation. Unlike classical fluids, where flow eventually dissipates due to and circulation decays, superfluid flow is dissipationless and irreversible once initiated, with persistent currents that can circulate indefinitely. Moreover, in superfluids, the circulation of around any closed path is quantized in units of h / m (where h is Planck's constant and m is the mass of the constituent particles), leading to the formation of quantized vortices rather than continuous . Superfluidity arises via a second-order at a critical , below which the superfluid component begins to form while a normal fluid component may coexist, as captured in phenomenological models like the two-fluid description. For liquid , this transition occurs at the , marking the onset of superfluid behavior.

Characteristic Phenomena

One of the defining features of superfluidity is the absence of in the superfluid component, allowing persistent flow without energy dissipation. This frictionless motion is vividly demonstrated by the , a thin layer of superfluid helium that creeps along surfaces, such as the walls of a , at speeds up to several centimeters per second, driven solely by gravitational or capillary forces without measurable viscous drag. Experiments measuring the of these films out of cylindrical vessels confirm that the transfer rate remains finite and temperature-dependent below the , highlighting the superfluid's ability to maintain coherent motion over macroscopic distances. Another hallmark is the formation of quantized vortices, where the circulation of the superfluid velocity around a closed loop is restricted to discrete multiples of the quantum of circulation, \kappa = h / m, with h being Planck's constant and m the mass of the constituent particles. These vortices possess a singular core of atomic size, approximately 1–10 Å in radius for helium, where the superfluid density vanishes, surrounded by irrotational flow that decays inversely with distance from the core. The dynamics of these vortices, including their nucleation, reconnection, and mutual interactions, govern the transition from laminar to turbulent superflow, as observed in rotating superfluids where vortex lines align into lattices. Superfluid flow remains dissipationless only below a critical , beyond which energy dissipation sets in due to the of quantized vortices. This threshold, typically on the order of 10–50 cm/s in experiments depending on and , marks the onset of vortex formation near obstacles or walls, leading to a of vortex shedding and eventual quantum . Measurements in cylindrical channels reveal that the critical for equilibrium vortex entry scales with the container size and rotation rate, providing a direct probe of the superfluid's stability limits. The superfluid fraction, \rho_s / \rho, represents the proportion of the total density \rho that participates in the frictionless flow, while the remainder behaves as a viscous normal fluid. This fraction, which approaches unity at absolute zero and decreases with temperature, is quantified through torsional oscillation experiments where stacked disks immersed in the superfluid experience drag primarily from the normal component. Pioneering measurements using such oscillators demonstrated that the effective moment of inertia reflects only the normal fluid's contribution, yielding \rho_s / \rho \approx 1 - (T / T_\lambda)^7 near the lambda transition, consistent with theoretical expectations. Thermomechanical effects arise from the coupling between gradients and differences in the two-fluid model, driving superfluid counterflow without transport by the superfluid component. The Rollin film creep exemplifies this, as the thin superfluid layer migrates to equalize across surfaces, facilitating heat transfer via film flow. The fountain effect, where superfluid emerges from a heated as a against , results from a induced by the , with the of the fountain scaling as \Delta P / \rho = (S / \rho_s) \Delta T, where S is the per mass; this phenomenon underscores the irreversible nature of normal fluid production.

Theoretical Frameworks

Macroscopic Descriptions

The two-fluid model offers a phenomenological framework for describing the macroscopic behavior of superfluids, particularly below the lambda transition temperature, by representing the system as an intimate mixture of two interpenetrating, non-interacting fluid components in the ideal case. The normal fluid component exhibits classical properties, including , thermal conductivity, and the capacity to carry all the system's , while the superfluid component flows without or and transports no . This separation allows the model to account for the coexistence of dissipative and non-dissipative flows observed in superfluids. The model was rigorously formulated by , building on earlier ideas, and provides the foundation for understanding bulk superfluid dynamics without reference to microscopic quantum details. In the two-fluid model, the total mass density \rho is given by \rho = \rho_n + \rho_s, where \rho_n and \rho_s denote the partial densities of and superfluid components, respectively; these densities depend on temperature, with \rho_s approaching \rho as temperature nears . Each component possesses its own independent velocity field, \mathbf{v}_n and \mathbf{v}_s, enabling counterflow where the superfluid moves oppositely to fluid without net . The hydrodynamic description includes separate continuity equations for the components: \frac{\partial \rho_s}{\partial t} + \nabla \cdot (\rho_s \mathbf{v}_s) = 0, \frac{\partial \rho_n}{\partial t} + \nabla \cdot (\rho_n \mathbf{v}_n) = 0. The superfluid motion, being irrotational (\nabla \times \mathbf{v}_s = 0), follows the inviscid Euler equation in the form \frac{\partial \mathbf{v}_s}{\partial t} + \nabla \left( \frac{v_s^2}{2} + \mu \right) = 0, where \mu is the specific chemical potential (chemical potential per unit mass). The normal fluid, in contrast, obeys the full Navier-Stokes equations, incorporating pressure gradients, viscous stresses, and entropy advection, as it behaves like an ordinary fluid. These equations close with thermodynamic relations, such as the entropy density s = s_n \rho_n / \rho confined to the normal component. A key application of the two-fluid model lies in the propagation of , which reveals two distinct modes due to the decoupled components. First sound corresponds to ordinary or , where \mathbf{v}_n and \mathbf{v}_s oscillate in , compressing both fluids simultaneously and yielding a speed on the order of 200–300 m/s in helium II, akin to classical . Second , predicted within this framework, manifests as temperature or where the components oscillate out of with no net mass current; here, the normal fluid's entropy variation drives counterflow, propagating at speeds around 20 m/s in helium II at low temperatures. The existence of , experimentally confirmed shortly after the model's proposal, provides direct evidence for the two-fluid nature of superfluids. The ideal two-fluid model assumes no momentum exchange between components, but this approximation fails at high relative velocities (exceeding a critical value, typically ~10–100 cm/s in helium II channels) or near boundaries, where interactions lead to dissipation via mutual friction. This friction originates from the scattering of normal fluid excitations (like rotons) by quantized vortices in the superfluid, coupling the velocity fields and introducing dissipative terms in the momentum equations, such as a force density \mathbf{f} = \alpha' \hat{\mathbf{s}} \times (\mathbf{v}_n - \mathbf{v}_s) \times \hat{\mathbf{s}} - \alpha \hat{\mathbf{s}} \times (\mathbf{v}_n - \mathbf{v}_s), where \alpha and \alpha' are temperature-dependent coefficients, and \hat{\mathbf{s}} relates to vortex orientation. These terms, incorporated into extended two-fluid hydrodynamics, describe phenomena like vortex motion and turbulent decay but mark the regime where the simple model requires modifications for accuracy.

Microscopic Explanations

Superfluidity arises from quantum mechanical effects in many-body systems of bosons or fermions at low temperatures, where interactions lead to a macroscopic characterized by long-range . In bosonic systems, such as liquid helium-4, the phenomenon is linked to Bose-Einstein condensation (BEC), in which a significant fraction of particles occupies the , enabling coherent flow without . This idea was first proposed by , who connected the λ-transition in helium to the degeneracy predicted by Bose-Einstein statistics. For weakly interacting bosons, the dynamics of the condensate are described by the Gross-Pitaevskii equation, a that captures the mean-field effects of interparticle interactions. Independently derived by Eugene Gross and Lev Pitaevskii, the time-dependent form is i \hbar \frac{\partial \psi}{\partial t} = -\frac{\hbar^2}{2m} \nabla^2 \psi + V \psi + g |\psi|^2 \psi, where \psi is the condensate wavefunction, V is the external potential, m is the particle mass, g is the interaction strength proportional to the s-wave scattering length, and \hbar is the reduced Planck's constant. This equation governs the evolution of the order parameter \psi, whose phase determines the superfluid velocity, and its stationary solutions yield the ground-state density |\psi|^2. The validity of this description holds in the dilute limit, where the gas parameter n a^3 \ll 1 (with n the density and a the scattering length), ensuring perturbative treatment of interactions. In fermionic systems, superfluidity emerges through the formation of Cooper pairs, bound states of opposite-spin fermions mediated by attractive interactions, as explained by . This microscopic theory, developed for but applicable to neutral fermionic superfluids like , predicts pairing in momentum space near the , leading to a gapped spectrum and zero-resistance flow. The superconducting gap \Delta(k) satisfies the self-consistent gap equation \Delta(\mathbf{k}) = -\sum_{\mathbf{k}'} V(\mathbf{k}, \mathbf{k}') \langle c_{-\mathbf{k}' \downarrow} c_{\mathbf{k}' \uparrow} \rangle, where V is the pairing potential, c^\dagger and c are fermionic creation and annihilation operators, and the expectation value reflects the anomalous average in the paired state. In the weak-coupling limit, the gap scales exponentially with the attraction strength, \Delta \sim \hbar \omega_D \exp(-1/N(0)V), with \omega_D the Debye frequency and N(0) the density of states at the Fermi level. A unifying microscopic signature of superfluidity in both bosonic and fermionic systems is off-diagonal long-range order (ODLRO), introduced by Chen-Ning Yang, which quantifies the persistence of over large distances. ODLRO is present when the one-body density matrix exhibits a non-zero eigenvalue in the limit of infinite separation, specifically \lim_{|\mathbf{r}| \to \infty} \langle \psi^\dagger(\mathbf{0}) \psi(\mathbf{r}) \rangle \neq 0, where \psi annihilates a particle at position \mathbf{r}. This long-range correlation implies a macroscopic eigenvalue of the , corresponding to the fraction in bosons or the pair in fermions, distinguishing superfluids from normal fluids. The nature of interactions profoundly influences the superfluid state, transitioning from dilute weakly interacting regimes—where mean-field approximations like the Gross-Pitaevskii equation suffice—to strongly interacting cases, such as , where correlations beyond mean-field are essential. In both limits, superfluidity involves spontaneous breaking of U(1) symmetry, leading to the emergence of Goldstone modes: gapless excitations with linear dispersion \omega = c k at low momentum k, where c is the sound determined by the superfluid . These modes, predicted by Goldstone's , represent the restoring force against phase twists and underpin the rigidity of the ordered state.

Historical Development

Early Discoveries

The of helium marked a pivotal advancement in low-temperature physics, achieved by in at the University of Leiden, where he produced the first samples of at its atmospheric of 4.2 K. This breakthrough enabled systematic studies of matter near , revealing helium's unique reluctance to solidify under normal pressures and its exceptionally low even above the superfluid transition. Initial observations noted deviations from classical behavior, such as minimal , but these were not fully anomalous until further cooling experiments. Significant peculiarities emerged in the mid-1920s through measurements of 's thermodynamic properties. In , Hendrik Keesom, a successor to Onnes at , conducted specific heat experiments with collaborators, identifying a sharp peak at approximately 2.17 —later termed the λ-point—indicating a from normal I to the more exotic helium II, without but with a λ-shaped anomaly in the curve. Concurrently, studies revealed anomalous contraction below this temperature, defying expectations for most liquids and suggesting underlying quantum effects, as the volume decreased sharply while density increased. These findings, plotted as the characteristic λ-curve, highlighted helium's non-classical response to cooling. Early experimental probes in the early further illuminated these anomalies, particularly in confined geometries. Researchers observed that helium II could flow through narrow capillaries and porous media with negligible resistance, enabling persistent supercurrents that circulated indefinitely without dissipation, as demonstrated in setups using fine glass tubes and annuli. Such flows, exceeding classical hydrodynamic limits, implied a approaching zero and foreshadowed the superfluid state. The culmination came in late 1937, when Pyotr Kapitza at the Institute for Physical Problems in (having been detained in the USSR since 1934) measured the of helium II below 2.17 K and found it vanishingly small, reporting results in early 1938. Independently, John F. Allen and Don Misener at the confirmed this zero- behavior through flow experiments in fine channels, establishing superfluidity as a distinct . Theoretical insight followed swiftly, with proposing in March 1938 that superfluidity stemmed from Bose-Einstein condensation, wherein a macroscopic fraction of atoms occupy the ground , enabling coherent, frictionless flow. This quantum hypothesis unified the observed phenomena under wave mechanics, contrasting with classical hydrodynamics. Kapitza's pioneering work on superfluidity and related low-temperature phenomena earned him the 1978 .

Key Theoretical Advances

In 1938, proposed the two-fluid model for superfluid , describing it as a mixture of a superfluid component with zero and a normal fluid component carrying all the , which provided a phenomenological framework to explain observed transport properties like persistent flow and reduced thermal conductivity. In 1941, independently developed a similar two-fluid hydrodynamics, emphasizing the role of and introducing the concept of a superfluid tied to a macroscopic wavefunction phase, laying the groundwork for subsequent dissipative extensions. During the 1940s and 1950s, Landau advanced a phenomenological theory of superfluidity that incorporated elementary excitations to account for thermodynamic and hydrodynamic behaviors, predicting the existence of first and second sound waves in helium II. A key element was the introduction of the roton spectrum for excitations, characterized by an energy-momentum relation E(p) \approx \Delta + \frac{(p - p_0)^2}{2\mu}, where \Delta is the energy gap, p_0 the momentum at the minimum, and \mu an effective mass, which explained the specific heat anomaly and scattering processes at low temperatures. This spectrum, refined in later works, highlighted the gapped nature of rotons contrasting with gapless phonons, influencing calculations of superfluid density and viscosity. In 1957, , , and Robert Schrieffer formulated the , which explained through the formation of Cooper pairs of electrons mediated by phonons, establishing a microscopic mechanism for pairing in fermionic systems that directly inspired analogous descriptions of superfluidity. This pairing mechanism was crucial for understanding the superfluidity of liquid helium-3, experimentally discovered in 1972 by , David M. Lee, and Robert C. Richardson through specific heat measurements at millikelvin temperatures, for which they received the 1996 . Theoretical models extended BCS concepts to ^3He, where p-wave pairing of ^3He atoms leads to anisotropic order parameters and multiple superfluid phases (A and B), accounting for the observed transition below approximately 2.5 mK and finite-temperature properties. In the 1960s, David Feenberg pioneered microscopic calculations using correlated basis functions to describe the of liquid helium-4, incorporating strong interatomic correlations beyond simple Bose-Einstein condensation to compute binding energies, momentum distributions, and excitation spectra with improved accuracy over earlier variational methods. These approaches, building on hypernetted-chain approximations, provided quantitative insights into the depletion of the condensate fraction (around 7-10%) and the , bridging phenomenological models with quantum many-body theory. Recent advances up to 2025 have leveraged simulations, particularly path-integral Monte Carlo methods, to confirm ground-state properties of , such as the equation of state, pair correlation functions, and superfluid fraction, achieving convergence with and resolving discrepancies in earlier approximations for both bulk and confined systems. These simulations have validated the minimum parameters and fraction to within 1% of experimental values, enhancing understanding of quantum coherence at zero temperature.

Superfluidity in Liquid Helium

Helium-4 Superfluidity

Liquid (^4He) exhibits superfluidity in its He II phase, which emerges below the lambda transition temperature T_\lambda = 2.17 K at saturated . This marks the onset of zero and other quantum behaviors in the bosonic fluid. The of ^4He shows the He II region extending from low pressures up to the solidification curve at approximately 25 atm, where the liquid remains stable down to without solidifying under its own . Above T_\lambda, the normal He I phase dominates, characterized by finite . The superfluid component is quantified by the superfluid density \rho_s(T), which represents the fraction of the total density that participates in frictionless flow. Near the lambda transition, \rho_s(T) approaches zero with a power-law dependence \rho_s(T) \propto (T_\lambda - T)^{2/3}, reflecting critical behavior associated with the three-dimensional XY universality class. At lower temperatures, \rho_s approaches the total density \rho, indicating nearly complete superfluidity. This temperature dependence has been precisely measured using techniques like torsion oscillators and second sound attenuation. Key experimental hallmarks of He II include the Kapitza resistance, a thermal boundary resistance at solid-liquid arising from mismatch between the solid and the superfluid. Discovered in 1941, this resistance leads to a jump across the interface during flow, with magnitude scaling as T^{-3} at low temperatures due to acoustic mismatch. Another hallmark is the persistence of supercurrents in toroidal containers, where circulation can be maintained indefinitely without energy dissipation, demonstrating the irrotational nature of superfluid flow except at singularities. In rotating containers, He II forms arrays of quantized vortices rather than classical solid-body , stabilizing the superfluid against . These vortices, predicted by Onsager and Feynman, carry quantized circulation \kappa = h / m = 9.97 \times 10^{-4} cm²/s, where h is Planck's constant and m the ^4He , with vortex density proportional to the rotation rate. The vortex core size is set by the healing length \xi = \hbar / \sqrt{2 m \mu}, where \mu is the , typically on the order of 1 nm near T_\lambda and smaller at lower temperatures, marking the distance over which the superfluid order parameter recovers from the core . The ease of superfluidity in ^4He stems from its bosonic nature, with atomic spin zero allowing Bose-Einstein condensation at accessible temperatures around 2 K. In contrast, the fermionic ^3He isotope, with half-integer spin, cannot condense directly and requires Cooper pairing mechanisms, leading to superfluid transitions only at millikelvin temperatures under specific conditions. This isotopic distinction underscores the role of quantum statistics in .

Helium-3 Superfluidity

Liquid , consisting of fermionic atoms, undergoes a transition to superfluidity at ultralow temperatures below approximately 2.5 , depending on . Unlike bosonic , this superfluidity arises from the formation of Cooper pairs in a p-wave pairing state, leading to two distinct superfluid s: the A phase and the B phase. The reveals these phases below the critical temperature T_c, with the A phase stable at higher temperatures and lower pressures, transitioning to the B phase at lower temperatures; the polycritical point occurs around 21 bar and 2.2 , separating regions where either phase is favored. The superfluid phases feature anisotropic order parameters due to p-wave pairing with triplet spin and odd-parity pairs. In the B phase, the order parameter corresponds to the isotropic Balian-Werthamer state, a uniform spin-triplet ^3P_0 that gaps the entire . The A phase, in contrast, exhibits a more anisotropic chiral structure akin to the Anderson-Brinkman-Morel state, with nodes in the gap function along specific directions. The superfluid energy gap \Delta(T) in both phases follows a BCS-like weak-coupling form, \Delta(T) \approx \Delta_0 \tanh\left(1.74 \sqrt{T_c / T - 1}\right) near T_c, though strong-coupling effects modify it at lower temperatures. Experimental probes have confirmed the symmetry of these order parameters. Nuclear magnetic resonance (NMR) shifts in the A phase reveal a transverse shift due to the anisotropic susceptibility, distinguishing it from the isotropic B phase response. Ultrasound attenuation measurements show sharp drops below T_c, reflecting pair-breaking and collective mode excitations, with distinct behaviors in the A and B phases due to their gap structures. In comparison to helium-4 superfluidity, helium-3's critical temperature exhibits strong pressure dependence, rising from ~1 mK at low pressure to a maximum of ~2.5 mK near 30 bar before decreasing toward the solid phase. Additionally, the spin-triplet nature enables half-quantum vortices in the A phase, where circulation is half the standard quantum unit due to combined orbital and spin contributions. Achieving these millikelvin temperatures relies on the Pomeranchuk effect, an adiabatic compression of liquid-solid mixtures that cools via the higher entropy of the solid phase below ~0.3 K.

Superfluidity in Ultracold Gases

Bose-Einstein Condensate Superfluids

Superfluidity in Bose-Einstein condensates (BECs) of dilute atomic gases represents a highly controllable realization of quantum coherence in laboratory settings, distinct from the cryogenic liquids like . These systems are formed by cooling bosonic atoms, such as rubidium-87, to temperatures near using and evaporative cooling techniques in magnetic or optical traps. The first experimental achievement of a BEC occurred in 1995, when a team led by Eric A. Cornell and Carl E. Wieman at produced a condensate of approximately 2,000 rubidium-87 atoms at a temperature of 170 nanokelvin. Shortly thereafter, Wolfgang at created a sodium-23 BEC, enabling comparative studies. This breakthrough earned Cornell, Wieman, and Ketterle the 2001 for demonstrating Bose-Einstein condensation in dilute gases of alkali atoms. In trapped BECs, superfluid properties emerge from the macroscopic occupation of the , described mean-field theoretically by the Gross-Pitaevskii equation, which captures the balance between kinetic energy, trapping potential, and interatomic interactions. Key length scales include the healing length \xi = \frac{\hbar}{\sqrt{2m\mu}}, where \mu is the , m the , and \hbar the reduced Planck's constant; this characterizes the distance over which the superfluid order parameter recovers from perturbations, such as defects or boundaries. The , often comparable to \xi in uniform systems but extended by the trap size in inhomogeneous ones, quantifies phase correlations essential for superflow. For large particle numbers, the Thomas-Fermi approximation simplifies the density profile, neglecting the kinetic term to yield n(\mathbf{r}) = \frac{1}{g} \left[ \mu - V(\mathbf{r}) \right] for \mu > V(\mathbf{r}) and zero otherwise, where g = \frac{4\pi \hbar^2 a_s}{m} is the interaction strength, a_s the s-wave scattering length, and V(\mathbf{r}) the trapping potential; this parabolic profile has been verified in experiments with and sodium BECs. Quantized vortices and solitons in BECs provide direct probes of superfluidity, analogous to those in helium but tunable in dilute gases. Vortices, with singly quantized circulation \kappa = \frac{h}{m} (h Planck's constant), were first observed in 1999 through stirring a rubidium BEC with a laser beam, forming stable lattices under rotation that mimic Abrikosov arrays in superconductors. In optical lattices—periodic potentials created by interfering lasers—dynamical instabilities arise when the superflow velocity exceeds the local sound speed, leading to the spontaneous formation of gray solitons (density dips with phase jumps) or vortices via modulational growth of perturbations; these instabilities were experimentally confirmed in 2004 using time-of-flight imaging of expanding sodium BECs. Additionally, drag effects manifest as frictional dissipation on moving vortices or soliton trains in lattices, where lattice phonons or band structure induce anomalous velocity-dependent forces, contrasting frictionless flow in uniform superfluids. Interactions in dilute BECs can be precisely tuned using magnetic Feshbach resonances, where an external field adjusts the scattering length a_s across positive (repulsive) and negative (attractive) values near molecular bound-state thresholds. The first observation of such tuning in a sodium BEC in demonstrated enhanced atom loss rates near resonance, confirming control over collision dynamics; subsequent experiments with and enabled studies from weakly interacting to unitarity-limited regimes. This tunability facilitates exploration of superfluid phase transitions, such as from BEC to molecular . Unlike , where strong, short-range interactions in a dense liquid (n a_s^3 \approx 10^{-2}) limit dimensionality to , dilute BECs feature weak interactions (n a_s^3 \ll 1) that allow realization of quasi- and 1D superfluids by tightening traps perpendicular to the desired dimension. In , phase fluctuations are suppressed below a Berezinskii-Kosterlitz-Thouless transition, enabling true long-range order, as observed in BECs in 2006. In 1D, the transition to a Tonks-Girardeau gas of impenetrable bosons mimics fermionic behavior while retaining superfluid correlations, achieved via Feshbach tuning in elongated traps. These low-dimensional systems highlight the versatility of gaseous BECs for probing quantum many-body effects inaccessible in helium. Recent advances include the realization of supersolid phases in dipolar BECs of atoms like and , where anisotropic dipole-dipole interactions enable simultaneous superfluidity and density-wave order. These states, first reported in 2017, have been further characterized through observation of quantized vortices confirming superfluidity as of 2024. Additionally, Floquet engineering using periodic driving has allowed precise control of Feshbach resonances in bosonic systems, opening new regimes for dynamic superfluid manipulation as demonstrated in 2025 experiments.

Fermionic Superfluids

Fermionic superfluidity in ultracold gases represents a tunable realization of paired systems, distinct from bosonic counterparts due to the requiring antisymmetric wavefunctions and s-wave pairing. The first experimental evidence emerged in 2003–2004 through studies of spin-polarized ⁶Li atoms, where magnetic Feshbach resonances enabled precise control of interatomic interactions to access the superfluid phase. In these experiments, hydrodynamic expansion and quantized vortices confirmed the onset of superfluid flow in a strongly interacting regime, marking the initial observation of fermionic superfluidity below a critical of approximately 0.2 times the Fermi temperature T_F. A hallmark of these systems is the BCS-BEC crossover, where the s-wave scattering length a_s is tuned across resonance via external magnetic fields, interpolating between the Bardeen-Cooper-Schrieffer (BCS) limit of weakly overlapping Cooper pairs (a_s < 0) and the Bose-Einstein condensate (BEC) limit of tightly bound bosonic molecules (a_s > 0). In the BEC regime, the two-body of fermion pairs is given by E_b = \frac{\hbar^2}{m a_s^2}, where m is the mass, allowing pairs to behave as composite bosons that condense. At the unitary limit (a_s → ∞), interactions reach a universal strength, and the system exhibits scale-invariant properties with a pairing gap Δ on the order of the E_F. Radio-frequency (RF) spectroscopy measurements have directly probed this gap, revealing Δ ≈ 0.5 E_F in trapped ⁶Li gases at unitarity, consistent with theoretical predictions from extended to strong coupling. Collective excitations in these fermionic superfluids, such as and radial modes, have been observed through time-of-flight after sudden perturbations, showing frequencies matching hydrodynamic predictions with minimal in the superfluid phase. propagation, manifesting as first waves, has been detected via density correlations, with speeds approaching the Bogoliubov value c = √(Δ E_F / m) at unitarity, and rates analyzed to probe and thermal effects in inhomogeneous s. These tunable systems serve as quantum simulators for condensed matter phenomena, including the pairing mechanisms in high-temperature superconductors and the equation of state in interiors, offering insights unattainable in solid-state or astrophysical settings. Recent progress as of 2024 includes studies of stability and sensitivity in interacting fermionic superfluids under optical , revealing phase transitions and enhanced times. Hybrid-pair superfluidity in driven Fermi gases has also been explored theoretically and experimentally, enabling new paired phases beyond BCS-BEC crossover. These developments further solidify ultracold fermionic gases as platforms for simulating complex quantum matter.

Astrophysical Superfluids

In Neutron Stars

Neutron stars host superfluid matter in their interiors, particularly in where reach approximately $10^{17} to $10^{18} kg/m³, corresponding to several times the saturation n_s \approx 0.16 fm⁻³. At these extreme conditions, neutrons form Cooper pairs primarily through the spin-triplet ^3P_2 channel, enabling superfluidity with pairing gaps \Delta \sim 0.1 - 1 MeV, which vary with and are suppressed near the crust-core . This superfluid phase arises from the attractive interaction in the at high , contrasting with lower- regions where ^1S_0 dominates. Protons in may also exhibit superfluidity via ^1S_0 , though with smaller gaps, influencing the overall . One key observable manifestation of neutron superfluidity is pulsar glitches, sudden increases in rotation rate attributed to the unpinning and outward motion of quantized superfluid vortices in the inner crust. In this model, vortices are pinned to lattice ions in the crust, building angular momentum lag until a critical threshold triggers catastrophic release, transferring angular momentum to the crust and causing spin-up. The Vela pulsar exemplifies this, exhibiting large glitches with fractional size \Delta \Omega / \Omega \approx 10^{-6} every few years, consistent with superfluid reservoir dynamics where the entrained superfluid mass fraction exceeds 10% of the star's total. Post-glitch recovery involves gradual re-coupling via mutual friction, providing constraints on the superfluid's viscosity and pinning strength. Superfluidity profoundly impacts cooling by altering emission rates, the primary cooling mechanism in the early phases. The pairing gap suppresses standard processes like modified Urca below a critical T_c \sim \Delta / k_B, reducing emissivity exponentially, but introduces enhanced channels such as pair breaking and formation (PBF), where pairs are emitted during creation or disruption, peaking at T \approx 0.4 T_c. For gaps \Delta \sim 0.1 - 1 MeV, PBF can dominate cooling in the core, leading to steeper drops observable in young pulsars like . This luminosity, integrated over the core , sets the cooling timescale to \sim 10^3 - 10^5 years. The presence of superfluidity modifies the nuclear equation of state (EOS) by contributing an additional energy term from the pairing interaction, stiffening the at high densities and affecting mass-radius relations. Microscopic calculations incorporating ^3P_2 neutron pairing yield EOS models with maximum masses M_{\max} \sim 1.8 - 2.2 M_\odot and radii R \sim 11 - 13 km for 1.4 M_\odot stars, testable via pulsar timing measurements of binary systems like PSR J0737-3039. These effects are subtle, on the order of a few percent in , but crucial for reconciling observations with theoretical models, as superfluid gaps influence and sound speed. Glitches and cooling curves further probe the EOS by constraining the superfluid fraction and coefficients.

Cosmological Contexts

Superfluid phase transitions in the early are invoked in models of electroweak , where the Higgs acquires a , forming a analogous to the order parameter in a Bose superfluid. This process breaks the SU(2) × U(1) electroweak symmetry, much like the transition to a superfluid state in , where collective excitations emerge below a critical temperature. The analogy draws from the A-B phase transition in superfluid ^3He, which serves as a model for such cosmological events, highlighting bubble nucleation and dynamics that could produce stochastic gravitational wave backgrounds observable by detectors like . In superfluid dark matter paradigms, the dark matter component is hypothesized to condense into a superfluid state at low temperatures and densities prevalent in cosmic structures, with the superfluid vacuum mediating long-range interactions that modify Newtonian on galactic scales. These models reconcile the successes of the \LambdaCDM framework on large cosmological scales with MOND-like in galaxies, where phonons in the superfluid excite a scalar countering for accelerations below a_0 \approx 10^{-10} m/s^2. The effective description treats the dark matter as a non-relativistic condensate, with self-interactions parameterized by a length that suppresses small-scale . Recent developments, including 2025 analyses, extend this to include finite-temperature effects and vortex dynamics in halos, predicting observable signatures in rotation curves and lensing. Cosmic strings arise as stable topological defects during superfluid-like phase transitions in the early , formed via the Kibble-Zurek mechanism when the system is quenched through the critical point faster than the correlation length can adjust. In these scenarios, the strings correspond to quantized vortices in the superfluid order parameter, with linear energy density (tension) \mu \sim \eta^2, where \eta denotes the symmetry-breaking scale, typically around $10^{16} GeV for grand unified theories. simulations in superfluid ^4He confirm the topological origin of string tension and proliferation during rapid cooling, mirroring cosmic defect formation and offering tests of scaling laws in an expanding . Links to cosmic emerge in effective theories where superfluid phonons replicate the Goldstone modes of a broken , sourcing primordial density perturbations akin to those from an . In protoinflationary phases dominated by a relativistic superfluid, phonon fluctuations provide a spectrum of curvature perturbations with nearly scale-invariant power, transitioning smoothly to post-inflationary reheating via vortex reconnections. Recent models up to 2025 incorporate Bose-Einstein condensate halos with repulsive self-interactions, forming solitonic cores that exhibit superfluid flow and influence large-scale structure via acoustic oscillations.

Analogies in High-Energy Physics

Superfluid Vacuum Models

Superfluid vacuum models conceptualize the as a superfluid medium, analogous to a (BEC) composed of gravitons or fluctuations in the metric, in which low-energy excitations appear as phonons that represent perturbations to the metric. This framework posits that geometry and gravitational interactions emerge from the collective dynamics of this underlying superfluid state, drawing parallels between condensed matter phenomena and fundamental effects. The superfluid nature ensures dissipationless flow at low temperatures, with the vacuum's acting as the from which all physical phenomena arise. These models, which remain speculative proposals in research, trace their origins to ideas in the 1960s by John Wheeler and collaborators, who explored quantum foam-like structures in as a way to unify and , laying groundwork for viewing the as a dynamic, fluctuating medium. The concepts were revived and formalized in the 2000s by researchers including George Chapline, Pawel Mazur, and Emil Mottola, who developed the idea of gravitational vacuum stars, treating collapsed matter as stabilized BEC states in a superfluid to avoid singularities. Their work extended Bose-Einstein condensation principles to gravitational systems, proposing that the 's superfluid phase could address issues like the by linking to properties. In superfluid vacuum theories, the emerge naturally from the hydrodynamic equations governing the superfluid, where long-wavelength perturbations in the density and yield relativistic curvature in the low-momentum limit. For instance, the effective arises from the of the wavefunction, with gravitational effects encoded in the superfluid's velocity field, recovering while allowing deviations at high energies. masses are interpreted as resulting from interactions of excitations or impurities with this superfluid background, akin to gap generation in condensed matter systems, where the vacuum's internal structure imparts effective to quasiparticles. These models offer testable predictions through analog systems, such as interpretations of phenomena using vortex dynamics in superfluids. Goldstone modes associated with broken symmetries in the provide a brief link to microscopic explanations of superfluidity, manifesting as massless excitations that align with spectra. Superfluidity offers a powerful analog framework for investigating phenomena, where collective excitations in a flowing superfluid, such as or Bogoliubov quasiparticles, propagate along an effective determined by the background and local speed. This , rooted in the acoustic of fluids, allows acoustic horizons to emerge in regions where the superfluid flow exceeds the local , mimicking the event horizons of black holes in . Across such horizons, pairs of excitations analogous to particle-antiparticle creations can be separated, leading to Hawking-like with a temperature inversely proportional to the surface of the analog horizon. Pioneering theoretical proposals in the late laid the groundwork for experimental realizations, with William Unruh suggesting in 1981 that sonic analogs in fluids could simulate evaporation. During the 1990s and 2000s, initial laboratory demonstrations focused on classical fluids like and , creating stationary and dynamical horizons to observe mode mixing and amplified scattering indicative of Hawking radiation. A landmark advancement came with Bose-Einstein condensates (BECs) in the , where experiments realized tunable black holes by engineering supersonic flows in ultracold atomic gases, enabling precise control over horizon parameters. These setups revealed the Bogoliubov dispersion relation, \omega(k) = \sqrt{(c k)^2 + (\hbar k^2 / 2m)^2}, where low-momentum excitations follow a relativistic linear dispersion (\omega \approx c k) mimicking light cones, while higher momenta exhibit nonlinear deviations that parallel effects like modified relations. In the context of holographic duality, superfluids serve as boundary realizations of the AdS/CFT correspondence, modeling strongly coupled quantum theories dual to gravitational dynamics in . Holographic superfluid models, introduced around 2008, describe phase transitions to superfluid states as in the boundary theory, with applications to the quark-gluon produced in heavy-ion collisions, where transport properties like and are computed via black hole horizons in the bulk. These models predict universal scaling behaviors near criticality, validated against simulations for the quark-gluon plasma's near-perfect fluidity. Superfluid vortex tangles, arising in quantum turbulence, provide an analog for structures resembling foam—a fluctuating, foamy at the in theories. In these systems, interconnected quantized vortices form dense, self-similar tangles that persist over long timescales, with their statistical properties, such as vortex line density L \propto t^{-3/2} during decay, mirroring the entangled quantum fluctuations expected in a foamy vacuum. This analogy suggests that the macroscopic quantum coherence in superfluids captures non-local entanglement akin to that in or foams.

References

  1. [1]
    Superfluidity: what is it and why does it matter? - Illinois News Bureau
    Dec 20, 2018 · The most obvious definition of superfluidity is the ability of a liquid to flow through narrow channels without apparent friction.
  2. [2]
    [PDF] THE DISCOVERY OF SUPERFLUIDITY
    Superfluidity is the friction-less flow of superfluid liquid helium, discovered in 1938, and named, after hints in 1911.
  3. [3]
    A theoretical description of the new phases of liquid | Rev. Mod. Phys.
    Apr 1, 1975 · This paper reviews the theory of anisotropic superfluid phases and its application to the new A and B phases of liquid $^{3}\mathrm{He}$.
  4. [4]
    [PDF] Nobel Lecture: Superfluid He: the early days as seen by a theorist*
    Dec 2, 2004 · Leggett: Superfluid 3He: the early days as seen by a theorist. Rev ... and published 共Leggett, 1972兲 in Physical Review Letters later ...
  5. [5]
    Rollin Film Rates in Liquid Helium | Phys. Rev.
    Quantitative measurements have been made on the flow of the Rollin film out of a cylindrical vessel containing liquid helium. The range of temperature ...
  6. [6]
    Quantum Theory of Superfluid Vortices. I. Liquid Helium II | Phys. Rev.
    The theory is applied to two distinct configurations in He II: a single vortex and a rotating vortex lattice. The specific heat associated with the vortex waves ...
  7. [7]
    Chaotic quantized vortices and hydrodynamic processes in ...
    Jan 1, 1995 · Superfluid turbulence in He II flows appears as a stochastic tangle of quantized vortex lines. Interest in this system extends beyond the field of superfluid ...
  8. [8]
    Measurement of Equilibrium Critical Velocities for Vortex Formation ...
    Sep 15, 1975 · We have made measurements of the critical velocity for vortex formation in right cylinders of circular, elliptical, and rectangular cross ...
  9. [9]
    Measuring the Superfluid Fraction of an Ultracold Atomic Gas
    In the Andronikashvili experiment [10] a torsional oscillator is used to measure the moment of inertia of the fluid coupled to the oscillator, I . A frequency ...
  10. [10]
    Forces Associated with Heat Flow in Helium II - Nature
    ... fountain effect', was associated in some way with the surface of the capillary or of the powder particles. Recent experiments by Daunt and Mendelssohn2, who ...
  11. [11]
  12. [12]
  13. [13]
  14. [14]
    [PDF] The liquefaction of helium - KNAW
    Kamerlingh Onnes, The liquefaction of helium., in: KNAW, Proceedings, 11, 1908-1909, Amsterdam, 1909, pp. 168-185. This PDF was made on 24 September 2010 ...
  15. [15]
    [PDF] The Specific Heats of Solid Substances at the Temperatures ...
    17 March 1926. 767.6. 4.210 .4121 3). ,767.6. 4.210 .4128. Between liquid helium and liquid hydrogen temperatures interpolation was made with the help of a ...Missing: paper | Show results with:paper
  16. [16]
    Superfluid Helium - an overview | ScienceDirect Topics
    Phase diagram of 4He at low temperatures. 4He remains liquid at zero temperature if the pressure is below 2.5 MPa (approximately 25 atm). The liquid has a phase ...
  17. [17]
    52.11: Superfluids - Physics LibreTexts
    Aug 7, 2024 · Helium II can flow through capillaries much less than 1 ⁢ μ ⁢ m in diameter, and in such experiments behaves as though it has zero viscosity.
  18. [18]
    Superfluid density as a function of temperature and pressure near ...
    As the lambda transition temperature is approached from below, the superfluid density, ρs, decreases as ρs=ρ0(P)t−ζ, where t=1−T/Tλ(P).
  19. [19]
    [PDF] The calculated thermodynamic properties of superfluid helium-4
    Oct 15, 2009 · Comprehensive tables of the primary thermodynamic properties of superfluid helium-4, such ~s the specific heat and entropy. are presented as ...
  20. [20]
    Press release: The 1996 Nobel Prize in Physics - NobelPrize.org
    But it was not until the end of the 1930s that Pjotr Kapitsa (Nobel Prize in Physics 1978) discovered experimentally the phenomenon of superfluidity in helium-4 ...
  21. [21]
    A-B Transition in Superfluid $$^3$$ He and Cosmological Phase ...
    Jun 8, 2024 · The transition from a normal Fermi liquid to the superfluid takes place between 1- − 2.5 mK, at which temperatures superfluid 3 He is ...
  22. [22]
    Superconductors take an odd turn - Physics World
    Nov 11, 2004 · States with p-wave symmetry are also formed when liquid helium-3 becomes a superfluid. ... spin-triplet” or “odd-parity” state. Although ...
  23. [23]
    Chiral Superfluid Helium-3 in the Quasi-Two-Dimensional Limit
    The data clearly deviate from the weak-coupling gap Δ 0 , BCS ( T ) obtained within the p -wave BCS theory. This is a manifestation of strong-coupling ...
  24. [24]
  25. [25]
    Superfluid 3He - Theory and Recent Experiments - Europhysics News
    The basic reason for the complexity of superfluid 3He is that there exist three p-wave states as well as three spin triplet states. This means that the.
  26. [26]
    Superfluid T c T_c of Helium-3 and its Pressure Dependence
    Superfluid T c T_c of liquid helium-3 and its pressure dependence are calculated by using a relation obtained from our macro-orbital microscopic theory.
  27. [27]
    Physics - Half-Quantum Vortices in Superfluid Helium
    Dec 14, 2016 · Researchers working with superfluid helium in a porous medium have generated vortices with half the quantum unit of fluid flow.
  28. [28]
    [PDF] Robert C. Richardson - Nobel Lecture
    A central part of the story of the discovery of superfluid ³He is the cooling technique used for the experiments, the Pomeranchuk Effect. Although it is not an ...Missing: millikelvin | Show results with:millikelvin
  29. [29]
    Observation of Bose-Einstein Condensation in a Dilute Atomic Vapor
    A Bose-Einstein condensate was produced in a vapor of rubidium-87 atoms that was confined by magnetic fields and evaporatively cooled.
  30. [30]
    Theory of Bose-Einstein condensation in trapped gases
    Apr 1, 1999 · The phenomenon of Bose-Einstein condensation of dilute gases in traps is reviewed from a theoretical perspective.
  31. [31]
    Vortices in a Bose-Einstein Condensate | Phys. Rev. Lett.
    Sep 27, 1999 · We have created vortices in two-component Bose-Einstein condensates. The vortex state was created through a coherent process involving the spatial and temporal ...
  32. [32]
    Observation of Dynamical Instability for a Bose-Einstein Condensate ...
    Sep 29, 2004 · We have experimentally studied the unstable dynamics of a harmonically trapped Bose-Einstein condensate loaded into a 1D moving optical lattice.Abstract · Article Text
  33. [33]
    Evidence for Superfluidity in a Resonantly Interacting Fermi Gas
    Apr 13, 2004 · These observations provide the first evidence for superfluid hydrodynamics in a resonantly interacting Fermi gas. Figure 1; Figure 2; Figure 3.
  34. [34]
    The Nature of Superfluidity in Ultracold Fermi Gases Near Feshbach ...
    Sep 12, 2003 · We study the superfluid state of atomic Fermi gases using a BCS-BEC crossover theory. Our approach emphasizes non-condensed fermion pairs.
  35. [35]
    [1709.10340] Superfluidity and Superconductivity in Neutron Stars
    This review focuses on applications of the ideas of superfluidity and superconductivity in neutron stars in a broader context.
  36. [36]
    [1406.6109] Pairing and superfluidity of nucleons in neutron stars
    Jun 23, 2014 · We survey the current status of understanding of pairing and superfluidity of neutrons and protons in neutron stars from a theoretical perspective.Missing: 1P1 | Show results with:1P1
  37. [37]
    Vortex Pinning in Neutron Stars, Slipstick Dynamics, and the Origin ...
    Dec 19, 2022 · We study pinning and unpinning of superfluid vortices in the inner crust of a neutron star using three-dimensional dynamical simulations.
  38. [38]
    Pinning down the superfluid and measuring masses using pulsar ...
    Oct 2, 2015 · However, glitching pulsars such as Vela have been shown to require a superfluid reservoir that greatly exceeds that available in the crust. We ...Pulsar Glitches · Results · Pulsar Mass From Glitches
  39. [39]
    Post-glitch Recovery and the Neutron Star Structure: The Vela Pulsar
    Jun 2, 2025 · The post-glitch recovery phase, characterized by a gradual recovery, reveals insights into the superfluid dynamics within the star. The post- ...
  40. [40]
    Cooling neutron stars and superfluidity in their interiors - arXiv
    Jun 28, 1999 · We study the heat capacity and neutrino emission reactions (direct and modified Urca processes, nucleon-nucleon bremsstrahlung, Cooper pairing of nucleons)
  41. [41]
    [1412.7759] Tests of the nuclear equation of state and superfluid ...
    Dec 24, 2014 · Here we present two new Chandra ACIS-S Graded observations of this neutron star and measurements of the neutron star mass M and radius R ...
  42. [42]
    A-B transition in superfluid $^3$He and cosmological phase ... - arXiv
    Jan 15, 2024 · A-B transition in superfluid ^3He and cosmological phase transitions. Authors:Mark Hindmarsh, J.A. Sauls, Kuang Zhang, S.Autti, Richard P. Haley ...
  43. [43]
    Theory of dark matter superfluidity | Phys. Rev. D
    Nov 9, 2015 · We propose a novel theory of dark matter (DM) superfluidity that matches the successes of the Λ cold dark matter ( Λ ⁢ C D M ) model on ...
  44. [44]
    [2505.23900] Superfluid Dark Matter - arXiv
    May 29, 2025 · The superfluid dark matter model offers an elegant solution to reconcile discrepancies between the predictions of the cold dark matter paradigm ...
  45. [45]
    [1304.4832] Fluid phonons, protoinflationary dynamics and large ...
    Apr 17, 2013 · Abstract:We explore what can be said on the effective temperature and sound speed of a statistical ensemble of fluid phonons present at the ...Missing: superfluid | Show results with:superfluid
  46. [46]
    A review of basic results on the Bose–Einstein condensate dark ...
    We review basic results on the Bose–Einstein condensate dark matter (BECDM) model. Self-gravitating BECs experience a collisionless process of gravitational ...
  47. [47]
  48. [48]
    Superfluid vacuum theory and deformed dispersion relations - arXiv
    Nov 24, 2020 · We present an analytical example of the dispersion relation and argue that it should have a Landau "roton" form which ensures the suppression of dissipative ...Missing: historical proposal Wheeler Chapline Mottola
  49. [49]
    [PDF] John A. Wheeler - National Academy of Sciences
    Robert Geroch (a PhD student of Wheeler's in the mid 1960s) has described Wheeler's ... an (approximate) solution of the vacuum Einstein equations: vacuum ...Missing: superfluid | Show results with:superfluid
  50. [50]
    [gr-qc/0407075] Gravitational Vacuum Condensate Stars - arXiv
    Jul 20, 2004 · Title:Gravitational Vacuum Condensate Stars. Authors:Pawel O. Mazur, Emil Mottola (University of South Carolina, Los Alamos National Laboratory).Missing: superfluid revival 2000s<|separator|>
  51. [51]
    None
    Nothing is retrieved...<|separator|>
  52. [52]
    Analogue Gravity - PMC - PubMed Central
    In this review article we will focus on “analogue gravity”, the development of analogies (typically but not always based on condensed matter physics)
  53. [53]
    Analogue gravity and the Hawking effect: historical perspective and ...
    Dec 4, 2023 · Zel'dovich's analogy suggested that rotating black holes could radiate, and the thermodynamic analogy suggested the black hole would have a ...
  54. [54]
    Realization of a Sonic Black Hole Analog in a Bose-Einstein ...
    Dec 7, 2010 · The solid line is the Bogoliubov dispersion relation. The green curve indicates the trapped excitations with negative energy. (f) The real ...
  55. [55]
    Holographic Superfluids and the Dynamics of Symmetry Breaking
    Jan 2, 2013 · We explore the far-from-equilibrium response of a holographic superfluid using the AdS/CFT correspondence. We establish the dynamical phase diagram.
  56. [56]
    The next generation of analogue gravity experiments - Journals
    Jul 20, 2020 · This article introduces the theme issue for the Scientific Discussion Meeting on The next generation of analogue gravity experiments held at the Royal Society ...