Fact-checked by Grok 2 weeks ago

Quantum fluid

A quantum fluid is a state of matter in which quantum mechanical phenomena, such as Bose-Einstein condensation, dominate the macroscopic behavior of the system, leading to unique properties like frictionless flow and quantized excitations. These fluids typically arise in bosonic systems at ultralow temperatures, where a significant fraction of particles occupy the same , enabling collective quantum coherence over large scales. The most prominent examples include (^4He) below its lambda transition temperature of 2.17 K and Bose-Einstein condensates (BECs) of dilute atomic gases cooled to nanokelvin temperatures. Unlike classical fluids, quantum fluids exhibit irrotational flow except at discrete singularities and display critical velocities beyond which dissipates. Superfluidity, a hallmark of quantum fluids, manifests as zero viscosity, allowing the fluid to flow without energy dissipation through narrow channels or along surfaces, as observed in the Rollin film effect in ^4He. This property stems from the macroscopic wave function describing the fluid, governed by the Gross-Pitaevskii equation for BECs or the two-fluid model for helium superfluids, where a superfluid component coexists with a normal viscous component at finite temperatures. Quantized vortices, with circulation multiples of h/m where h is Planck's constant and m is the mass of the bosonic constituent (helium-4 atom or helium-3 Cooper pair), form the elementary excitations, enabling phenomena like vortex tangles in quantum turbulence. In fermionic systems like superfluid ^3He, pairing into bosonic Cooper pairs at around 2.7 mK produces analogous behavior, though with p-wave symmetry. The study of quantum fluids originated with the discovery of superfluidity in ^4He in 1938 by Pyotr Kapitza, John F. Allen, and Donald Misener, building on earlier predictions of Bose-Einstein condensation by and in 1924-1925. Experimental realization of BECs in 1995 using alkali atoms like rubidium-87 marked a milestone, providing tunable systems for probing in controlled environments. Quantum fluids also extend to exotic realms, such as the superfluid cores of neutron stars and high-temperature superconductors where pairs behave fluid-like. These systems continue to inform fundamental physics, from understanding to applications in precision sensing and quantum simulation. As of 2025, research continues to explore quantum fluids of light and novel phases for quantum technologies.

Fundamentals

Definition

A quantum fluid is a system in which quantum mechanical effects, including wave-particle duality and adherence to Bose or Fermi , manifest on a within fluid-like states of . These systems arise at sufficiently low temperatures where quantum coherence dominates, leading to collective behaviors not observed in classical regimes. In contrast to classical fluids, which follow Boltzmann and exhibit due to particle collisions, quantum fluids display anomalous properties such as zero (superfluidity) or perfect , emerging when the thermal de Broglie wavelength becomes comparable to the average interparticle spacing. represents a hallmark property of many such fluids, enabling frictionless flow. Prominent examples include liquid helium-4 below the λ-point at 2.17 , where it transitions to a superfluid state; electron fluids in superconductors exhibiting zero electrical resistance; and ultracold atomic gases forming Bose-Einstein condensates. The role of quantum statistics is central: bosonic quantum fluids, composed of integer-spin particles, can undergo Bose-Einstein condensation into a single , fostering macroscopic coherence; fermionic quantum fluids, made of half-integer-spin particles, achieve analogous ordered states through mechanisms.

Quantum Mechanical Basis

Quantum fluids exhibit macroscopic quantum behavior when the wave-like nature of their constituent particles becomes prominent on scales comparable to the interparticle spacing. This arises from wave-particle duality, where the thermal de Broglie , \lambda_{dB} = h / \sqrt{2\pi m k_B T}, of the particles reaches or exceeds the average interparticle distance d. In such conditions, the wave functions of neighboring particles overlap significantly, allowing quantum interference effects to influence the collective dynamics of the fluid rather than being confined to individual particles. The quantum statistics obeyed by the particles play a crucial role in enabling these effects. For bosonic particles, such as atoms, Bose-Einstein statistics permit multiple particles to occupy the same , leading to Bose-Einstein where a macroscopic fraction of particles resides in the at low temperatures. This underpins in bosonic quantum fluids. In contrast, fermionic systems, like electrons in superconductors or atoms, follow Fermi-Dirac statistics and the , preventing single-particle ground-state occupation. However, attractive interactions can induce pairing of fermions into composite bosons, known as Cooper pairs, which then undergo , resulting in macroscopic quantum coherence. This statistical behavior facilitates macroscopic , where the entire fluid can be described by a single, coherent quantum with a well-defined . The locking across the system allows the fluid to exhibit phenomena on macroscopic scales, as the collective behaves like a single quantum entity rather than a collection of independent particles. first proposed this macroscopic description for superfluids, emphasizing how quantum extends to the bulk scale. In systems like , weak interparticle interactions are essential for quantum effects to dominate over . The van der Waals forces between helium atoms are sufficiently feeble, combined with the light atomic mass, to sustain large zero-point motion that prevents solidification even at and allows quantum delocalization to prevail. This weak coupling minimizes scattering and preserves the coherence necessary for quantum fluid properties.

Historical Development

Theoretical Foundations

The theoretical foundations of quantum fluids emerged in the 1920s with the development of quantum statistics, which provided the framework for understanding collective quantum behaviors in many-body systems at low temperatures. In 1924, derived the Planck distribution for photons using a novel counting method for , laying the groundwork for statistics applicable to bosons. extended this approach in 1925 to a monatomic , predicting Bose-Einstein condensation (BEC), a where a macroscopic number of bosons occupy the lowest below a critical , potentially manifesting in quantum fluids like dilute gases or liquids. This work highlighted how quantum degeneracy could lead to coherent, in bosonic systems. Complementing Bose-Einstein statistics, Fermi-Dirac statistics were formulated in 1926 to describe fermions, particles obeying the , such as electrons and nuclei. developed the statistical distribution for identical fermions, ensuring no two occupy the same , which applies to fermionic quantum fluids where degeneracy pressure arises at low temperatures. Independently, arrived at the same formulation, emphasizing its role in for systems like liquid , where fermionic pairing could enable analogous to bosonic condensation. These advancements in quantum statistics established the particle-specific behaviors essential for predicting quantum fluid states. By , theorists began applying these statistical foundations to real liquids, particularly , to explain anomalous low-temperature properties. proposed in 1938 a two-fluid model for superfluid -II, positing that below the lambda transition temperature, the liquid separates into an inviscid superfluid component (arising from Bose-Einstein degeneracy) and a viscous normal fluid component (behaving like a of excitations). This model qualitatively predicted phenomena like zero and persistent flow, framing superfluid as a quantum fluid where quantum dominates macroscopic hydrodynamics. Extending these ideas to electronic systems, the Bardeen-Cooper-Schrieffer ( of described as a quantum fluid state in metals, where electrons form Cooper pairs (bosonic pairs of fermions) mediated by interactions, leading to a with zero electrical resistance below a critical . This microscopic theory unified with quantum fluid concepts, demonstrating how attractive interactions could overcome fermionic repulsion to produce macroscopic quantum coherence.

Key Discoveries

The discovery of marked the first experimental observation of a quantum fluid , occurring in 1911 when and his team at measured the electrical resistance of mercury cooled to 4.2 K using , finding it abruptly dropped to zero. This transition, unexpected at the time, laid the groundwork for understanding macroscopic quantum effects in solids. In 1937, was independently discovered in liquid helium-4 by in and by John F. Allen and Don Misener in , who observed a at approximately 2.17 K—the —below which the fluid exhibited zero and flowed without resistance through narrow channels. 's work, conducted at the Mond Laboratory, highlighted the anomalous thermal properties, earning him the 1978 for achievements in low-temperature physics. The superfluid phase of , a fermionic quantum fluid, was identified in 1972 by , David M. Lee, and Robert C. Richardson at through pulsed experiments on pure liquid helium-3, revealing phase transitions at around 2.6 mK into superfluid states with paired fermions. This breakthrough, which demonstrated in a system of fermions requiring Cooper pairing analogous to , earned the trio the 1996 . In 1995, the first Bose-Einstein condensate (BEC) was achieved as a quantum fluid by Eric Cornell and at , using evaporative cooling to condense approximately 2,000 rubidium-87 atoms into their at 170 nK, confirming the theoretical prediction of macroscopic occupation of a single . This milestone, shared with Wolfgang Ketterle's independent work, was recognized with the 2001 . During the 2000s, experimental progress in ultracold atomic gases led to the realization of unitary Fermi gases exhibiting , with key evidence of pairing and the BCS-BEC crossover observed in 2003–2004 by teams including Deborah Jin's group using atoms tuned to infinite scattering length via Feshbach resonance, enabling studies of strongly interacting fermionic matter.

Types

Superfluid Helium

Superfluid helium represents the archetypal quantum fluid, where liquid isotopes exhibit at cryogenic temperatures. Liquid helium-4 (^4He), composed of bosonic atoms with total spin zero, undergoes a second-order known as the lambda transition at 2.17 K under saturated , marking the onset of superfluidity in the helium II phase. This transition was first indicated by anomalies in the specific heat observed in the late 1920s, with the precise temperature value refined through subsequent thermodynamic measurements. Below this temperature, ^4He demonstrates zero shear viscosity, enabling frictionless flow through narrow capillaries, as independently demonstrated by experiments measuring flow rates far exceeding classical limits. Additionally, superfluid ^4He exhibits extraordinarily high thermal conductivity, on the order of 10^5 times that of typical liquids, due to the ballistic propagation of excitations in the absence of viscous scattering. In contrast, liquid (^3He), an isotopic fermionic system with half-integer , does not Bose-condense directly but achieves through the of atoms into composite bosonic entities, analogous to Cooper pairs in . This occurs at much lower , around 1 mK at zero , with the increasing to a maximum of about 2.5 mK at intermediate , requiring advanced cooling techniques to access. The superfluid phases of ^3He include the A-phase, characterized by anisotropic p-wave with equal-spin and broken , stable in , and the B-phase, featuring isotropic with higher under zero field and fully gapped excitations. These phases were discovered in through specific heat anomalies and NMR measurements in Pomeranchuk-cooled samples, revealing distinct that lead to different superfluid properties, such as anisotropic mass transport in the A-phase versus isotropic behavior in the B-phase. The isotopic differences stem from quantum statistics: ^4He atoms, being indistinguishable bosons, readily form a coherent macroscopic wavefunction via Bose-Einstein condensation at the , facilitating without pairing. Conversely, ^3He's fermionic nature enforces the , suppressing direct condensation and necessitating attractive interactions for pairing at millikelvin scales, resulting in a transition temperature orders of magnitude lower than for ^4He. Experimental investigations of ^4He often involve setups exploiting its unique behaviors, such as the fountain effect—where heat applied to one side of a capillary-connected causes a jet of to fountain upward due to a gradient—and Rollin films, ultra-thin (∼100 nm) superfluid layers that creep along surfaces without resistance, enabling transfer between vessels. For ^3He, studies rely on dilution refrigerators, which achieve millikelvin temperatures by leveraging the and entropy-driven dilution of ^3He in superfluid ^4He, providing continuous cooling essential for probing its delicate superfluid phases.

Superconductors

Superconductors exhibit quantum fluid behavior through the formation of Cooper pairs, where electrons pair up via phonon-mediated attraction to create bosonic quasiparticles that condense into a macroscopic quantum state, enabling zero electrical resistance and perfect diamagnetism below a critical temperature. This pairing mechanism, described by Bardeen-Cooper-Schrieffer (BCS) theory, transforms the electron gas into a coherent quantum fluid analogous to superfluids, with the pairs occupying the same ground state across the material. Superconductivity was first observed in mercury in 1911, marking the initial experimental evidence of this quantum phenomenon. Superconductors are classified into Type I and Type II based on their response to magnetic fields. Type I superconductors, such as pure metals like lead and tin, completely expel magnetic fields via the up to a critical field strength, maintaining perfect in their quantum fluid state. In contrast, Type II superconductors, including alloys like niobium-titanium, allow partial penetration of magnetic flux in the form of quantized vortices once the lower critical field is exceeded, enabling higher field tolerance while preserving up to an upper critical field; this vortex lattice arises from the Abrikosov theory and is crucial for practical applications. The critical temperature T_c, below which the quantum fluid phase emerges, varies widely: conventional low-temperature superconductors like achieve T_c \approx 9.2 K, requiring cooling. High-temperature , discovered in the late and advanced in the , reach T_c up to 133 K in compounds like \mathrm{HgBa_2Ca_2Cu_3O_{8+\delta}}, allowing operation with and revolutionizing potential applications despite ongoing debates on their pairing mechanism beyond BCS. A key manifestation of the superconducting quantum fluid is the , predicted in 1962, where a supercurrent tunnels coherently through a thin insulating barrier between two superconductors without applied voltage, driven by the phase difference of the macroscopic wavefunction. This effect enables devices like superconducting quantum interference devices (SQUIDs), which exploit interference of supercurrents for ultrasensitive magnetic field detection down to femtotesla levels.

Bose-Einstein Condensates

Bose-Einstein condensates (BECs) in ultracold dilute gases represent a class of quantum fluids where bosonic atoms, such as ^{87}Rb and ^{23}Na, are cooled to temperatures on the order of nanokelvins, enabling the macroscopic occupation of the system's . This , theoretically predicted by Einstein in 1924–1925 based on Bose's for photons, occurs when a significant fraction of atoms collapse into a single , forming a coherent that behaves as a single giant atom. The formation process begins with to reduce the atomic velocity distribution, followed by evaporative cooling in a magnetic trap, where hotter atoms are selectively removed, allowing the remaining gas to thermalize at progressively lower temperatures until degeneracy sets in. In these condensates, up to 10^6 or more atoms can occupy the , resulting in a macroscopic quantum wavefunction that exhibits phase across the entire , analogous to in but realized in a dilute gaseous medium. This manifests in phenomena such as patterns when two condensates are released and overlapped, confirming the wave-like nature of the . The dilute nature of these gases, with densities around 10^{13}–10^{15} cm^{-3}, minimizes interactions compared to liquid systems, allowing precise control over the quantum behavior. A key feature of gaseous BECs is their tunable interactions, achieved through magnetic Feshbach resonances, which adjust the atomic scattering length by coupling open-channel collisions to bound molecular states near a magnetic-field-tuned crossing. By varying the scattering length from positive (repulsive) to negative (attractive) values, researchers can drive phase transitions, such as from a superfluid to a state in optical lattices, where atoms localize due to strong repulsion. This tunability has enabled studies of quantum many-body physics, including the exploration of unitary gases and novel quantum phases. Extensions of the bosonic BEC paradigm include fermionic condensates formed via pairing of fermionic atoms, such as ^6Li, into bosonic molecules that then condense at ultralow temperatures near a Feshbach resonance, bridging the BCS superfluid regime to BEC behavior. Additionally, spinor condensates incorporate internal , allowing multiple hyperfine components (e.g., F=1 in ^{87}Rb) to coexist and evolve coherently, revealing spin textures, domain formation, and magnetic phase transitions driven by spin-exchange interactions. These systems provide a versatile platform for simulating complex quantum Hamiltonians with internal structure.

Properties

Superfluidity

Superfluidity manifests as the remarkable ability of quantum fluids to exhibit zero viscosity, enabling without energy when velocities remain below a critical threshold. This property was first demonstrated through experiments involving the of II through narrow capillaries, where no was required to maintain the flow, contrasting sharply with classical viscous fluids and indicating an absence of internal in the superfluid component. In flow setups, the superfluid displays anomalous behavior: the exceeds predictions from classical hydrodynamics, as the superfluid portion moves without viscous drag, while any arises solely from interactions with the container walls or excitations above the critical velocity. The theoretical framework for this frictionless flow is provided by the two-fluid model, proposed by , which describes the quantum fluid as a superposition of two interpenetrating components: a normal fluid that carries all and behaves viscously, and a superfluid component with zero and zero that flows ideally without dissipation. In this model, the normal fluid fraction increases with temperature, accounting for the gradual loss of as the system approaches the transition temperature, while the superfluid fraction dominates at lower temperatures, enabling the observed dissipationless motion. This separation explains why persists even as thermal excitations are present, with the superfluid component decoupled from dissipative processes. The regime of zero viscosity is limited by a critical , beyond which the superflow destabilizes and energy dissipation occurs through the of quantized vortices. A rough for this critical in channel flows is v_c \approx \frac{h}{m d}, where h is Planck's constant, m is the of the constituent particles, and d is the characteristic dimension of the flow channel (such as its diameter); more precise models involve logarithmic or power-law dependencies on d. For , this threshold is observed in channel flows and influences practical limits in low-temperature experiments. A key thermal consequence of superfluidity is the effectively infinite thermal conductivity arising from the two-fluid dynamics, where heat is transported via counterflow of the normal and superfluid components without generating temperature gradients or dissipative losses in the bulk. This high conductivity enables exceptional efficiency but, at elevated heat fluxes, can lead to , where a stable vapor layer forms at the heated surface, temporarily insulating it and altering the heat transfer regime. Such behavior underscores the unique interplay between frictionless flow and thermal transport in quantum fluids.

Quantized Phenomena

In quantum fluids, quantized vortices represent a fundamental discrete phenomenon arising from the wave-like nature of the superfluid order parameter. The circulation of the superfluid velocity \mathbf{v}_s around any closed path enclosing a vortex core is quantized, given by \oint \mathbf{v}_s \cdot d\mathbf{l} = n \kappa, where n is an integer, \kappa = h/m is the quantum of circulation, h is Planck's constant, and m is the mass of the bosonic constituent particles (e.g., ^4He atoms in superfluid ). This quantization stems from the single-valuedness of the macroscopic wavefunction, ensuring the phase changes by $2\pi n around the loop. The concept was first proposed by in 1949 and elaborated by in 1955, who described vortices as stable topological defects in the superfluid. Feynman's seminal posits that quantized vortices manifest as nodes where the wavefunction vanishes, analogous to zeros in a complex , with stability arising from energy minimization in the presence of or . In rotating superfluids, such as ^4He below the , these vortices arrange into lattices to accommodate the imposed , effectively simulating rigid-body while preserving the irrotational nature of the superfluid elsewhere; the vortex density scales linearly with speed, as n_v = 2\Omega / \kappa, where \Omega is the . Experimental visualizations, including neutron scattering and , confirm these lattice structures with inter-vortex spacings on the order of micrometers under typical rates. Another quantized effect occurs during phase slippage in confined geometries, such as narrow channels or orifices, where sustained superflow exceeds a critical velocity, leading to discrete dissipation events. Here, the superfluid phase slips by integer multiples of $2\pi, corresponding to the nucleation and passage of vortices across the channel, quantizing the flow rate in units related to \kappa; each slip event releases a fixed energy quantum, observable as reproducible voltage pulses in flow measurements. This phenomenon, studied in submicron apertures in ^4He, highlights the discrete nature of momentum transfer in one-dimensional-like superfluid transport. Quantized vortices are also observed in Bose-Einstein condensates (BECs) of dilute atomic gases, where they form singly or in lattices under rotation, providing a highly controllable platform for studying quantum turbulence and vortex dynamics. Analogous quantized structures appear in other quantum fluids, notably type-II superconductors, where Abrikosov vortices form in the mixed state under applied magnetic fields. Each vortex carries a quantized magnetic flux \Phi_0 = h/(2e), with e the elementary charge, threading the core as a normal region surrounded by supercurrents; these arrange into triangular lattices, mirroring superfluid vortex arrays but coupled to electromagnetic fields. This flux quantization, predicted by Abrikosov in 1957, underpins the partial penetration of magnetic fields in materials like niobium alloys.

Theoretical Derivation

De Broglie Wavelength Criterion

The de Broglie wavelength \lambda of a particle is given by the relation \lambda = h / p, where h is Planck's constant and p is the particle's . For particles in a fluid at , the relevant momentum scale is the thermal momentum p \approx \sqrt{2 m k_B T}, where m is the particle mass, k_B is Boltzmann's constant, and T is the temperature; this yields an estimate for the wavelength associated with thermal motion. In the quantum regime of a fluid, quantum effects dominate when the de Broglie wavelength exceeds the average interparticle spacing d = n^{-1/3}, where n is the number density; this condition implies significant overlap of the particles' wavefunctions, marking the onset of quantum degeneracy. To derive the precise criterion, the thermal de Broglie wavelength is defined as \lambda_{th} = h / \sqrt{2 \pi m k_B T}, which arises from the quantum partition function for non-interacting particles and accounts for the spread in thermal momenta. Quantum degeneracy occurs when \lambda_{th} n^{1/3} \gtrsim 1, as this ensures the wave packets overlap sufficiently for collective quantum behavior to emerge. For dense fluids such as liquid helium, with interparticle spacings on the order of angstroms, the condition \lambda_{th} n^{1/3} > 1 is satisfied at temperatures ranging from millikelvin to a few kelvin, enabling quantum fluid phenomena.

Temperature Thresholds

The temperature thresholds for quantum fluid transitions mark the points at which thermal energy becomes comparable to quantum mechanical energy scales, leading to macroscopic quantum coherence in bosonic or fermionic systems. These thresholds are derived using statistical mechanics, where the occupation numbers of quantum states determine the onset of condensation or pairing. For bosons, the critical temperature arises when the chemical potential reaches the ground-state energy, allowing macroscopic occupation of the lowest energy state; for fermions, degeneracy sets in at the Fermi temperature, with superfluidity emerging at lower temperatures through pairing mechanisms. For an ideal Bose gas, the Bose-Einstein condensation temperature T_c is given by k_B T_c = \frac{h^2}{2\pi m} \left( \frac{n}{\zeta(3/2)} \right)^{2/3}, where h is Planck's constant, m is the particle mass, n is the , k_B is Boltzmann's constant, and \zeta(3/2) \approx 2.612 is the value. This formula indicates that T_c scales with density and inversely with mass, setting the scale for quantum degeneracy in dilute gases. In interacting systems, such as ultracold atomic gases, corrections shift T_c by a few percent due to mean-field effects and beyond. For fermionic systems, the Fermi temperature T_F defines the degeneracy scale, beyond which Pauli exclusion dominates, given by T_F = \frac{\hbar^2}{2 m k_B} (3\pi^2 n)^{2/3}, with \hbar = h / 2\pi. Superfluidity in fermions, as in paired states, occurs well below T_F via mechanisms like BCS pairing, where the transition temperature T_c is exponentially suppressed relative to T_F in weakly attractive systems. This pairing enables superfluid transitions in fermionic quantum fluids despite the absence of a direct condensate. In liquid , a strongly interacting bosonic system, the superfluid transition occurs at the T_\lambda \approx 2.17 K under saturated , significantly lower than the ideal gas prediction of about 3.13 K due to repulsive interactions that harden the spectrum and reduce for condensation. For , a fermionic liquid, the superfluid transition temperature is pressure-dependent, reaching a maximum of ≈2.5 mK near 3 bar and ≈0.93 mK at saturated , arising from p-wave of fermions near the , with T_c / T_F \sim 10^{-3} reflecting the strength of the attractive interaction in this system. The derivation of these thresholds begins with the grand-canonical partition function for non-interacting particles, \mathcal{Z} = \prod_k \frac{1}{1 - [z](/page/Z) e^{-\beta \epsilon_k}} for bosons (where [z](/page/Z) = e^{\beta \mu} is the and \beta = 1 / k_B T), leading to the total particle number N = \sum_k \frac{1}{[z](/page/Z)^{-1} e^{\beta \epsilon_k} - 1}. For a in three dimensions, the excited-state integrates to N_e = g_{3/2}([z](/page/Z)) (V / \lambda^3), where \lambda = h / \sqrt{2\pi m k_B T} is the and g_{3/2}([z](/page/Z)) is the ; occurs when [z](/page/Z) \to 1 and N_e < N, yielding T_c such that N = \zeta(3/2) V / \lambda_c^3. For fermions, the analogous Fermi-Dirac integral at T = 0 fills states up to \epsilon_F, with T_F = \epsilon_F / k_B, and finite-temperature numbers f_k = 1 / (e^{\beta (\epsilon_k - \mu)} + 1) broaden the . Interactions introduce corrections via or methods, shifting thresholds by altering effective masses and scattering lengths, as seen in where strong correlations require beyond-mean-field treatments.

Applications

Low-Temperature Experiments

Low-temperature experiments on quantum fluids require sophisticated cryogenic infrastructure to access the millikelvin regime, where quantum effects dominate. The is the cornerstone of such setups, leveraging the between and isotopes in a superfluid mixture to achieve continuous cooling below 300 mK, with base temperatures as low as 2-10 mK depending on the design and cooling power. These systems typically feature a mixing chamber where 3He dissolves into superfluid 4He, releasing that is extracted via circulating 3He, enabling stable operation for extended periods in studies of superfluid helium and Bose-Einstein condensates. For even lower temperatures in helium-3 experiments, Pomeranchuk cooling complements dilution refrigeration by adiabatically compressing solid , exploiting its anomalous property where the solid phase has higher entropy than the liquid below 0.32 K, thus reducing temperature upon solidification. This method, first applied in the of superfluid , reaches sub-millikelvin temperatures and is particularly useful for probing fermionic superfluid phases under controlled . Flow visualization techniques in superfluids utilize waves—temperature waves propagating without mass flow—to quantify counterflow dynamics between and superfluid components, revealing dissipation and thresholds. Experiments often employ optical methods, such as tracing metastable molecules or particles, to image these waves and map velocity profiles in channels, providing insights into the two-fluid model's validity at high relative velocities. Nuclear magnetic resonance (NMR) and ultrasound serve as key probes for vortex dynamics and phase transitions in quantum fluids. In superfluid helium, NMR detects spatial variations in the order parameter around quantized vortices through frequency shifts and linewidth broadening, enabling mapping of vortex textures in 3He phases. Ultrasound, meanwhile, measures sound attenuation and speed to track vortex pinning and reconnection in helium, as well as to identify symmetry-breaking transitions in superconductors by monitoring acoustic anomalies near critical points. Prominent facilities hosting these experiments include dilution refrigerators at NIST, where compact closed-cycle systems achieve 70 mK for neutron scattering studies of quantum fluids, and at CERN, where large-scale units cool multi-ton masses to below 250 mK for precision particle physics investigations involving superfluid targets.

Emerging Technologies

Superconducting qubits, which rely on the quantum fluid properties of superconductors, form the basis of many current architectures. These qubits are typically implemented using Josephson junctions, nonlinear superconducting elements that enable tunable coupling and control of quantum states. In 2019, Google's , a 53-qubit device built with qubits incorporating Josephson junctions, demonstrated by performing a random circuit sampling task in 200 seconds—a estimated to take classical supercomputers up to years. This milestone highlighted the potential of superconducting quantum fluids for scalable processing, with ongoing improvements in times and gate fidelities exceeding 99.9% in recent iterations. Helium cryogenics plays a critical role in maintaining the low temperatures required for superconducting technologies, with below 2.17 K used in advanced applications such as particle accelerators. In , at 4.2 K cools the niobium-titanium coils of MRI magnets, enabling stable magnetic fields up to 3 T for high-resolution scans without electrical resistance. For , the (LHC) at employs at 1.9 K to cool its 8.3 T magnets, distributing heat efficiently through the fluid's zero-viscosity flow and high thermal conductivity, which sustains the collider's 27 km ring during high-energy proton collisions. Ultracold atomic gases, including Bose-Einstein condensates (BECs), serve as versatile quantum simulators to model complex quantum fluid behaviors inaccessible to classical computation. These systems replicate the to investigate mechanisms, such as pairing in fermionic lattices, providing insights into materials where transition temperatures exceed 100 K. Similarly, dipolar ultracold gases simulate the extreme conditions inside stars, capturing superfluid vortex dynamics and crust-core transitions through controlled interactions in optical traps. In precision timekeeping, BECs enhance atomic clocks by suppressing thermal noise and enabling coherent matter-wave ; recent developments include continuous-wave BECs that maintain coherence for hours, improving clock stability to 10^{-18} fractional frequency uncertainty over daily timescales. Post-2020 advances have expanded quantum fluid applications into hybrid systems and exotic many-body regimes. Hybrid magnet-superconductor platforms integrate quantum fluids with topological states, enabling robust encoding via Majorana fermions for fault-tolerant , with prototypes demonstrating protected edge modes at millikelvin temperatures. Unitary Fermi gases, tuned to infinite scattering length, have advanced understanding of strongly interacting quantum fluids through neural-network simulations, revealing universal superfluid properties like Tan contact correlations that bridge few-body and thermodynamic limits. In , research has explored trapped electrons on quantum fluids and solids as a promising route for building s, offering new approaches to hardware. These developments underscore quantum fluids' role in probing non-equilibrium dynamics, such as many-body in lattice gases, paving the way for next-generation quantum technologies.

References

  1. [1]
    [PDF] QUANTUM FLUIDS & SUPERFLUIDS - UBC Physics
    Superfluidity is seen, eg., in the absence of viscous resistance to the flow of a fluid. The superfluid flows freely, ad infinitum, through holes hardly larger ...
  2. [2]
    Bose–Einstein condensation - PMC - NIH
    Understanding the quantum fluid nature of the condensate should help us understand some of the properties of other quantum fluids (like liquid helium). Recent ...
  3. [3]
    [PDF] Super-Fluids and Quantum-Fluids – A Brief Review - HAL
    Feb 23, 2018 · Abstract. This particular paper gives a brief account of quantum fluids and Super-fluids, their intrinsic connection, the insights they.
  4. [4]
    Introduction to quantum turbulence - PMC - NIH
    Mar 25, 2014 · The term quantum turbulence denotes the turbulent motion of quantum fluids, systems such as superfluid helium and atomic Bose–Einstein ...Missing: key | Show results with:key
  5. [5]
    A century of Bose-Einstein condensation | Communications Physics
    Jul 1, 2025 · Bose-Einstein Condensation is a phenomenon at the heart of many of the past century's most intriguing and fundamental manifestations, such as superfluidity and ...
  6. [6]
    Quantum Fluid - an overview | ScienceDirect Topics
    Quantum fluids are defined as fluids characterized by a velocity field and probability density, where the flows are generated by probability rather than ...Missing: facts | Show results with:facts
  7. [7]
    Quantum fluids of light | Rev. Mod. Phys.
    Feb 21, 2013 · This article reviews recent theoretical and experimental advances in the fundamental understanding and active control of quantum fluids of light in nonlinear ...
  8. [8]
    Superfluid helium-4 hydrodynamics with discrete topological defects
    Oct 8, 2018 · In the helium-4 case, however, below T = 2.17 K (the so-called lambda point), the global U ( 1 ) symmetry of the microscopic quantum system is ...
  9. [9]
  10. [10]
  11. [11]
    Dynamical indistinguishability and statistics in quantum fluids
    This is readily formulated as a condition that the de Broglie wavelength λ dB should reach or exceed inter-particle separation x . The second criterion is ...
  12. [12]
    [PDF] David M. Lee - Nobel Lecture
    Fritz London's great contribution¹³ was to note that superfluidity could be ... de Broglie wavelength is on the order of the mean interparticle spacing.
  13. [13]
    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.
  14. [14]
    When quantum mechanics became huge | Nature Physics
    Sep 5, 2017 · The ground-breaking realization that macroscopic wavefunctions underlie these effects was introduced by Fritz London (pictured) in 1946.
  15. [15]
    None
    ### Summary of Content on Liquid Helium from Feynman (1957)
  16. [16]
    [PDF] Bose-Einstein Condensation in Alkali Gases - Nobel Prize
    In 1924 the physicist S. N. Bose from Dacca University in India sent Einstein a paper, in which the Planck distribution law for photons was derived by ...Missing: Satyendra Nath
  17. [17]
    This Month in Physics History | American Physical Society
    The story of superconductivity begins in 1911, when Heike Kamerlingh Onnes first discovered the phenomenon.
  18. [18]
    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] ...
  19. [19]
    The Discovery of Superfluidity | Journal of Low Temperature Physics
    Jan 11, 2007 · Superfluidity is a remarkable manifestation of quantum mechanics at the macroscopic level. This article describes the history of its discovery.Missing: Kapitsa | Show results with:Kapitsa
  20. [20]
    Press release: The 1996 Nobel Prize in Physics - NobelPrize.org
    Osheroff and Robert C. Richardson discovered at the beginning of the 1970s, in the low-temperature laboratory at Cornell University, that the helium isotope ...
  21. [21]
    Nobel Prize in Physics to Lee, Osheroff and Richardson for ...
    Richardson have been awarded the 1996 Nobel Prize in. Physics for their discovery of super- fluidity in helium-3. Their experi- ments, done in the low ...
  22. [22]
    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.
  23. [23]
    January 1938: Discovery of Superfluidity - American Physical Society
    Jan 1, 2006 · When helium-4 is chilled to below about 2.2 K, it starts to behave in some very weird ways. The fluid passes through narrow tubes with ...
  24. [24]
    The Nobel Prize in Physics 1996 - Advanced information
    Superfluidity in helium-3 only appears at very low temperatures, below about 2 mK, and has found practical applications only for specialists in the extreme low ...
  25. [25]
    Superconductivity | CERN
    In 1911, while studying the properties of matter at very low temperature, the Dutch physicist Heike Kamerlingh Onnes and his team discovered that the ...<|separator|>
  26. [26]
    Possible new effects in superconductive tunnelling - ScienceDirect
    1 July 1962, Pages 251-253. Physics Letters. Possible new effects in superconductive tunnelling☆. Author links open overlay panel B.D. Josephson. Show more. Add ...Missing: SQUIDs | Show results with:SQUIDs
  27. [27]
    BEC Introduction (BEC@MIT) - RLE at MIT
    In our experiment we take a gas of many atoms (about 5 to 10 million) and by lowering the temperature we make the atoms wavelengths very long. This makes our ...
  28. [28]
    Feshbach resonances in ultracold gases | Rev. Mod. Phys.
    Apr 29, 2010 · Feshbach resonances are the essential tool to control the interaction between atoms in ultracold quantum gases. They have found numerous ...Abstract · Article Text · Control of Atomic Quantum... · Ultracold Feshbach Molecules
  29. [29]
    Condensation of Pairs of Fermionic Atoms near a Feshbach ...
    We have observed Bose-Einstein condensation of pairs of fermionic atoms in an ultracold 6 L i gas at magnetic fields above a Feshbach resonance.
  30. [30]
    Observation of Spinor Dynamics in Optically Trapped Bose-Einstein ...
    Apr 7, 2004 · A spinor BEC can be described by a multicomponent order parameter which is invariant under gauge transformation and rotation in spin spaceMissing: paper | Show results with:paper
  31. [31]
    Theory of the Superfluidity of Helium II | Phys. Rev.
    Phys. Rev. 60, 356 (1941) ... Theory of the Superfluidity of Helium II. L. Landau. Institute of Physical Problems, Academy of Sciences USSR, Moscow, USSR. PDF ...
  32. [32]
    Calculation of the recovery heat flux from film boiling in superfluid ...
    A method for calculating the recovery heat flux in saturated He II heat transfer is presented. The approach combines the molecular kinetic equations for film ...
  33. [33]
    Violation of the London law and Onsager–Feynman quantization in ...
    Jun 17, 2007 · Onsager and Feynman pointed out that the superfluid velocity (v) in these vortices in single-component systems is quantized, and the circulation ...
  34. [34]
    QUANTIZED VORTICES AND TURBULENCE IN HELIUM II
    Although Feynman's (1955) conception of vortices uses Onsager's. (1949) basic circulation hypothesis, the idea of discrete vortex lines leads to a different ...
  35. [35]
    Quantized vortices in superfluid | Rev. Mod. Phys.
    Jul 1, 1987 · In this review, the concept of broken symmetry is applied to investigate the quantized vortex lines in superfluid 3 H e . In the superfluid A ...
  36. [36]
    Filming vortex lines reconnecting in a turbulent superfluid
    Jul 1, 2010 · (a) Circulation of superfluid around the vortex line is quantized so that the contour integral of its velocity around any loop enclosing one ...
  37. [37]
    Quantized phase slippage in superfluid | Phys. Rev. Lett.
    Apr 27, 1992 · We have studied dissipative phase-slip events in the oscillatory superflow of 4 H e through two different orifices of submicron size.Missing: narrow | Show results with:narrow
  38. [38]
    [PDF] Quantized Phase Slippage in Superfluid *He - Berkeley Physics
    Apr 27, 1992 · In conclusion, we have shown that quantized phase-slip events characterize the two orifices under study. The critical velocity v for the events ...Missing: narrow | Show results with:narrow
  39. [39]
    De Broglie Wavelength - Richard Fitzpatrick
    Quantum effects become important if the mean inter-particle distance, $ r_d$ , becomes comparable, or less than, the de Broglie wavelength.
  40. [40]
    Strongly correlated quantum fluids: ultracold quantum gases ...
    In such a state, fermions of opposite spin join to make Cooper pairs, but their average pair size ξc is greater than the interparticle spacing n−1/3, so that ...<|separator|>
  41. [41]
    Theory of ultracold atomic Fermi gases | Rev. Mod. Phys.
    Oct 2, 2008 · For a noninteracting gas, the distribution function follows the ballistic law f ( r , p , t ) = f 0 ( r − p t ∕ m , p ) , where f 0 is the ...<|separator|>
  42. [42]
    Anderson's considerations on the flow of superfluid helium: Some ...
    Aug 17, 2015 · Helium-4 undergoes an ordering transition toward a superfluid state at T ∼ 2.17 K under its saturated vapor pressure (SVP), which is now ...
  43. [43]
    Nobel Lecture: Superfluid : the early days as seen by a theorist
    Dec 2, 2004 · The element helium comes in two (stable) forms, He 4 and He 3 ; at low temperatures and pressures both form liquids rather than solids. The ...
  44. [44]
    Bose-Einstein condensation in a canonical ensemble with fixed total ...
    The theory of Bose-Einstein condensation (BEC), at the simplest level of an ideal gas, relies on the balance equation for a total number N of particles ...
  45. [45]
    Theoretical Models for the Cooling Power and Base Temperature of ...
    Aug 7, 2025 · 3He/4He dilution refrigerators are widely used for applications requiring continuous cooling at temperatures below approximately 300 mK.
  46. [46]
    [PDF] Robert C. Richardson - Nobel Lecture
    The entropies of the liquid and solid phases are equal at 0.32 K. At lower temperatures Pomeranchuk's suggestion for cooling will work. (Entropy per mole).
  47. [47]
    Second-sound studies of coflow and counterflow of superfluid 4He ...
    Jun 2, 2015 · We report a comprehensive study of turbulent superfluid 4He flow through a channel of square cross section. We study for the first time two ...
  48. [48]
    Visualization of two-fluid flows of superfluid helium-4 - PMC - NIH
    Mar 25, 2014 · Cryogenic flow visualization techniques have been proved in recent years to be a very powerful experimental method to study superfluid turbulence.
  49. [49]
    [PDF] Vortex-bound solitons in topological superfluid 3He - arXiv
    Nov 30, 2022 · Vortex-bound solitons in 3He form from spatial variation of order parameters, terminating in a vortex with trapped superfluid currents. These ...
  50. [50]
    Ultrasonically generated quantized vorticity in HeII - ScienceDirect
    Mar 20, 1978 · Ultrasound at one megahertz and electron bubble currents are used to probe superfluid helium. Onset of electron bubble trapping on sonically ...
  51. [51]
    Ultrasound response to time-reversal symmetry breaking below the ...
    Sep 23, 2025 · In this work, we develop a framework to model the ultrasound response of multi-component superconductors undergoing a time-reversal symmetry ...
  52. [52]
    Performance of a Miniature, Closed-Cycle Dilution Refrigerator at ...
    Dec 1, 2022 · Here, we show that a closed-cycle DR from Chase Research Cryogenics can successfully cool to about 70 mK at tilt angles up to and including 30°.
  53. [53]
    ICE Dil Fridge | NIST - National Institute of Standards and Technology
    Nov 22, 2017 · Low temperature control is achieved using the dilution fridge. For temperatures above 1.5 Kelvin, exchange gas is added to the dilution fridge ...
  54. [54]
    Dilution Refrigeration of Multi-Ton Cold Masses
    May 15, 2007 · Dilution refrigeration is the only means to provide continuous cooling at temperatures below 250 mK. Future experiments featuring multi-ton ...
  55. [55]
    Quantum supremacy using a programmable superconducting ...
    Oct 23, 2019 · Quantum supremacy using a programmable superconducting processor. Nature 574, 505–510 (2019). https://doi.org/10.1038/s41586-019-1666-5.A Suitable Computational... · Building A High-Fidelity... · Fidelity Estimation In The...
  56. [56]
    The liquid gold of MRI - Siemens Healthineers
    Mar 4, 2021 · Liquid helium is the only medium cold enough to rid metals of their electrical resistance, enabling them to generate powerful and stable magnetic fields.
  57. [57]
    Cryogenics: Low temperatures, high performance - CERN
    Superfluid helium has remarkable properties, including very high thermal conductivity; it is an efficient heat conductor. These qualities make helium an ...
  58. [58]
    Quantum simulations with ultracold atoms in optical lattices - Science
    Sep 8, 2017 · Ultracold atoms in optical lattices represent an ideal platform for simulations of quantum many-body problems.
  59. [59]
    Investigating the crust of neutron stars with neural-network quantum ...
    Mar 19, 2025 · In this work, we introduce a variational Monte Carlo method based on a neural Pfaffian-Jastrow quantum state, which allows us to model the transition from the ...
  60. [60]
    Continuous Bose–Einstein condensation - Nature
    Jun 8, 2022 · Bose–Einstein condensates (BECs) are macroscopic coherent matter waves that have revolutionized quantum science and atomic physics.
  61. [61]
    Magnet-superconductor hybrid quantum systems: a materials ...
    Mar 3, 2025 · Magnet–superconductor hybrid (MSH) systems have recently emerged as one of the most significant developments in condensed matter physics.
  62. [62]
    Neural-network quantum states for ultra-cold Fermi gases - Nature
    May 8, 2024 · The unitary Fermi gas (UFG) is a strongly interacting system that exhibits stable superfluid behavior. Studying the BCS-BEC crossover near the ...
  63. [63]
    Manybody interferometry of quantum fluids | Science Advances
    Jul 19, 2024 · We present an approach that we term “manybody Ramsey interferometry” that combines adiabatic state preparation and Ramsey spectroscopy.