Superfluid helium-4

Last updated

Superfluid helium-4 (helium II or He-II) is the superfluid form of helium-4, an isotope of the element helium. A superfluid is a state of matter in which matter behaves like a fluid with zero viscosity. The substance, which resembles other liquids such as helium I (conventional, non-superfluid liquid helium), flows without friction past any surface, which allows it to continue to circulate over obstructions and through pores in containers which hold it, subject only to its own inertia. [1]

Contents

The formation of the superfluid is a manifestation of the formation of a Bose–Einstein condensate of helium atoms. This condensation occurs in liquid helium-4 at a far higher temperature (2.17 K) than it does in helium-3 (2.5 mK) because each atom of helium-4 is a boson particle, by virtue of its zero spin. Helium-3, however, is a fermion particle, which can form bosons only by pairing with itself at much lower temperatures, in a weaker process that is similar to the electron pairing in superconductivity. [2]

History

Known as a major facet in the study of quantum hydrodynamics and macroscopic quantum phenomena, the superfluidity effect was discovered by Pyotr Kapitsa [3] and John F. Allen, and Don Misener [4] in 1937. Onnes possibly observed the superfluid phase transition on August 2, 1911, the same day that he observed superconductivity in mercury. [5] It has since been described through phenomenological and microscopic theories.

In the 1950s, Hall and Vinen performed experiments establishing the existence of quantized vortex lines in superfluid helium. [6] In the 1960s, Rayfield and Reif established the existence of quantized vortex rings. [7] Packard has observed the intersection of vortex lines with the free surface of the fluid, [8] and Avenel and Varoquaux have studied the Josephson effect in superfluid helium-4. [9] In 2006, a group at the University of Maryland visualized quantized vortices by using small tracer particles of solid hydrogen. [10]

In the early 2000s, physicists created a Fermionic condensate from pairs of ultra-cold fermionic atoms. Under certain conditions, fermion pairs form diatomic molecules and undergo Bose–Einstein condensation. At the other limit, the fermions (most notably superconducting electrons) form Cooper pairs which also exhibit superfluidity. This work with ultra-cold atomic gases has allowed scientists to study the region in between these two extremes, known as the BEC-BCS crossover.

Supersolids may also have been discovered in 2004 by physicists at Penn State University. When helium-4 is cooled below about 200 mK under high pressures, a fraction (≈1%) of the solid appears to become superfluid. [11] [12] By quench cooling or lengthening the annealing time, thus increasing or decreasing the defect density respectively, it was shown, via torsional oscillator experiment, that the supersolid fraction could be made to range from 20% to completely non-existent. This suggested that the supersolid nature of helium-4 is not intrinsic to helium-4 but a property of helium-4 and disorder. [13] [14] Some emerging theories posit that the supersolid signal observed in helium-4 was actually an observation of either a superglass state [15] or intrinsically superfluid grain boundaries in the helium-4 crystal. [16]

Applications

Recently in the field of chemistry, superfluid helium-4 has been successfully used in spectroscopic techniques as a quantum solvent. Referred to as superfluid helium droplet spectroscopy (SHeDS), it is of great interest in studies of gas molecules, as a single molecule solvated in a superfluid medium allows a molecule to have effective rotational freedom, allowing it to behave similarly to how it would in the "gas" phase. Droplets of superfluid helium also have a characteristic temperature of about 0.4 K which cools the solvated molecule(s) to its ground or nearly ground rovibronic state.

Superfluids are also used in high-precision devices such as gyroscopes, which allow the measurement of some theoretically predicted gravitational effects (for an example, see Gravity Probe B).

The Infrared Astronomical Satellite IRAS, launched in January 1983 to gather infrared data was cooled by 73 kilograms of superfluid helium, maintaining a temperature of 1.6 K (−271.55 °C). When used in conjunction with helium-3, temperatures as low as 40 mK are routinely achieved in extreme low temperature experiments. The helium-3, in liquid state at 3.2 K, can be evaporated into the superfluid helium-4, where it acts as a gas due to the latter's properties as a Bose–Einstein condensate. This evaporation pulls energy from the overall system, which can be pumped out in a way completely analogous to normal refrigeration techniques.

Superfluid-helium technology is used to extend the temperature range of cryocoolers to lower temperatures. So far the limit is 1.19 K, but there is a potential to reach 0.7 K. [17]

Properties

Superfluids, such as helium-4 below the lambda point, exhibit many unusual properties. A superfluid acts as if it were a mixture of a normal component, with all the properties of a normal fluid, and a superfluid component. The superfluid component has zero viscosity and zero entropy. Application of heat to a spot in superfluid helium results in a flow of the normal component which takes care of the heat transport at relatively high velocity (up to 20 cm/s) which leads to a very high effective thermal conductivity.

Film flow

Many ordinary liquids, like alcohol or petroleum, creep up solid walls, driven by their surface tension. Liquid helium also has this property, but, in the case of He-IV, the flow of the liquid in the layer is not restricted by its viscosity but by a critical velocity which is about 20 cm/s. This is a fairly high velocity so superfluid helium can flow relatively easily up the wall of containers, over the top, and down to the same level as the surface of the liquid inside the container, in a siphon effect. It was, however, observed, that the flow through nanoporous membrane becomes restricted if the pore diameter is less than 0.7 nm (i.e. roughly three times the classical diameter of helium atom), suggesting the unusual hydrodynamic properties of He arise at larger scale than in the classical liquid helium. [18]

Rotation

Another fundamental property becomes visible if a superfluid is placed in a rotating container. Instead of rotating uniformly with the container, the rotating state consists of quantized vortices. That is, when the container is rotated at speeds below the first critical angular velocity, the liquid remains perfectly stationary. Once the first critical angular velocity is reached, the superfluid will form a vortex. The vortex strength is quantized, that is, a superfluid can only spin at certain "allowed" values. Rotation in a normal fluid, like water, is not quantized. If the rotation speed is increased more and more quantized vortices will be formed which arrange in nice patterns similar to the Abrikosov lattice in a superconductor.

Comparison with helium-3

Although the phenomenologies of the superfluid states of helium-4 and helium-3 are very similar, the microscopic details of the transitions are very different. Helium-4 atoms are bosons, and their superfluidity can be understood in terms of the Bose–Einstein statistics that they obey. Specifically, the superfluidity of helium-4 can be regarded as a consequence of Bose–Einstein condensation in an interacting system. On the other hand, helium-3 atoms are fermions, and the superfluid transition in this system is described by a generalization of the BCS theory of superconductivity. In it, Cooper pairing takes place between atoms rather than electrons, and the attractive interaction between them is mediated by spin fluctuations rather than phonons. (See fermion condensate.) A unified description of superconductivity and superfluidity is possible in terms of gauge symmetry breaking.

Macroscopic theory

Thermodynamics

Fig. 1. Phase diagram of He. In this diagram is also given the l-line. Phase diagram of 4He 01.svg
Fig. 1. Phase diagram of He. In this diagram is also given the λ-line.
Fig. 2. Heat capacity of liquid He at saturated vapor pressure as function of the temperature. The peak at T=2.17 K marks a (second-order) phase transition. He liquid tbotevadobis damokidebuleba TemperaTurastan 350-en.svg
Fig. 2. Heat capacity of liquid He at saturated vapor pressure as function of the temperature. The peak at T=2.17 K marks a (second-order) phase transition.
Fig. 3. Temperature dependence of the relative superfluid and normal components rn/r and rs/r as functions of T. Normal and superfluid density 01.jpg
Fig. 3. Temperature dependence of the relative superfluid and normal components ρn/ρ and ρs/ρ as functions of T.

Figure 1 is the phase diagram of 4He. [19] It is a pressure-temperature (p-T) diagram indicating the solid and liquid regions separated by the melting curve (between the liquid and solid state) and the liquid and gas region, separated by the vapor-pressure line. This latter ends in the critical point where the difference between gas and liquid disappears. The diagram shows the remarkable property that 4He is liquid even at absolute zero. 4He is only solid at pressures above 25 bar.

Figure 1 also shows the λ-line. This is the line that separates two fluid regions in the phase diagram indicated by He-I and He-II. In the He-I region the helium behaves like a normal fluid; in the He-II region the helium is superfluid.

The name lambda-line comes from the specific heat – temperature plot which has the shape of the Greek letter λ. [20] [21] See figure 2, which shows a peak at 2.172 K, the so-called λ-point of 4He.

Below the lambda line the liquid can be described by the so-called two-fluid model. It behaves as if it consists of two components: a normal component, which behaves like a normal fluid, and a superfluid component with zero viscosity and zero entropy. The ratios of the respective densities ρn/ρ and ρs/ρ, with ρns) the density of the normal (superfluid) component, and ρ (the total density), depends on temperature and is represented in figure 3. [22] By lowering the temperature, the fraction of the superfluid density increases from zero at Tλ to one at zero kelvins. Below 1 K the helium is almost completely superfluid.

It is possible to create density waves of the normal component (and hence of the superfluid component since ρn + ρs = constant) which are similar to ordinary sound waves. This effect is called second sound. Due to the temperature dependence of ρn (figure 3) these waves in ρn are also temperature waves.

Fig. 4. Helium II will "creep" along surfaces in order to find its own level - after a short while, the levels in the two containers will equalize. The Rollin film also covers the interior of the larger container; if it were not sealed, the helium II would creep out and escape. Helium-II-creep.svg
Fig. 4. Helium II will "creep" along surfaces in order to find its own level – after a short while, the levels in the two containers will equalize. The Rollin film also covers the interior of the larger container; if it were not sealed, the helium II would creep out and escape.
Fig. 5. The liquid helium is in the superfluid phase. As long as it remains superfluid, it creeps up the wall of the cup as a thin film. It comes down on the outside, forming a drop which will fall into the liquid below. Another drop will form - and so on - until the cup is empty. Liquid helium Rollin film.jpg
Fig. 5. The liquid helium is in the superfluid phase. As long as it remains superfluid, it creeps up the wall of the cup as a thin film. It comes down on the outside, forming a drop which will fall into the liquid below. Another drop will form – and so on – until the cup is empty.

Superfluid hydrodynamics

The equation of motion for the superfluid component, in a somewhat simplified form, [23] is given by Newton's law

The mass M4 is the molar mass of 4He, and is the velocity of the superfluid component. The time derivative is the so-called hydrodynamic derivative, i.e. the rate of increase of the velocity when moving with the fluid. In the case of superfluid 4He in the gravitational field the force is given by [24] [25]

In this expression μ is the molar chemical potential, g the gravitational acceleration, and z the vertical coordinate. Thus we get the equation which states that the thermodynamics of a certain constant will be amplified by the force of the natural gravitational acceleration

(1)

Eq.  (1) only holds if vs is below a certain critical value, which usually is determined by the diameter of the flow channel. [26] [27]

In classical mechanics the force is often the gradient of a potential energy. Eq.  (1) shows that, in the case of the superfluid component, the force contains a term due to the gradient of the chemical potential. This is the origin of the remarkable properties of He-II such as the fountain effect.

Fig. 6. Integration path for calculating m at arbitrary p and T. Integration path in pT diagram 01.jpg
Fig. 6. Integration path for calculating μ at arbitrary p and T.
Fig. 7. Demonstration of the fountain pressure. The two vessels are connected by a superleak through which only the superfluid component can pass. Demo fountain pressure 01.jpg
Fig. 7. Demonstration of the fountain pressure. The two vessels are connected by a superleak through which only the superfluid component can pass.
Fig. 8. Demonstration of the fountain effect. A capillary tube is "closed" at one end by a superleak and is placed into a bath of superfluid helium and then heated. The helium flows up through the tube and squirts like a fountain. Helium fountain 01.jpg
Fig. 8. Demonstration of the fountain effect. A capillary tube is "closed" at one end by a superleak and is placed into a bath of superfluid helium and then heated. The helium flows up through the tube and squirts like a fountain.

Fountain pressure

In order to rewrite Eq. (1) in more familiar form we use the general formula

(2)

Here Sm is the molar entropy and Vm the molar volume. With Eq. (2) μ(p,T) can be found by a line integration in the p-T plane. First we integrate from the origin (0,0) to (p,0), so at T =0. Next we integrate from (p,0) to (p,T), so with constant pressure (see figure 6). In the first integral dT=0 and in the second dp=0. With Eq. (2) we obtain

(3)

We are interested only in cases where p is small so that Vm is practically constant. So

(4)

where Vm0 is the molar volume of the liquid at T =0 and p =0. The other term in Eq. (3) is also written as a product of Vm0 and a quantity pf which has the dimension of pressure

(5)

The pressure pf is called the fountain pressure. It can be calculated from the entropy of 4He which, in turn, can be calculated from the heat capacity. For T =Tλ the fountain pressure is equal to 0.692 bar. With a density of liquid helium of 125 kg/m3 and g = 9.8 m/s2 this corresponds with a liquid-helium column of 56 meter height. So, in many experiments, the fountain pressure has a bigger effect on the motion of the superfluid helium than gravity.

With Eqs. (4) and (5) , Eq. (3) obtains the form

(6)

Substitution of Eq. (6) in (1) gives

(7)

with ρ0 = M4/Vm0 the density of liquid 4He at zero pressure and temperature.

Eq. (7) shows that the superfluid component is accelerated by gradients in the pressure and in the gravitational field, as usual, but also by a gradient in the fountain pressure.

So far Eq. (5) has only mathematical meaning, but in special experimental arrangements pf can show up as a real pressure. Figure 7 shows two vessels both containing He-II. The left vessel is supposed to be at zero kelvins (Tl=0) and zero pressure (pl = 0). The vessels are connected by a so-called superleak. This is a tube, filled with a very fine powder, so the flow of the normal component is blocked. However, the superfluid component can flow through this superleak without any problem (below a critical velocity of about 20 cm/s). In the steady state vs=0 so Eq. (7) implies

(8)

where the index l (r) applies to the left (right) side of the superleak. In this particular case pl = 0, zl = zr, and pfl = 0 (since Tl = 0). Consequently,

This means that the pressure in the right vessel is equal to the fountain pressure at Tr.

In an experiment, arranged as in figure 8, a fountain can be created. The fountain effect is used to drive the circulation of 3He in dilution refrigerators. [28] [29]

Fig. 9. Transport of heat by a counterflow of the normal and superfluid components of He-II Counterflow heat exchange 01.jpg
Fig. 9. Transport of heat by a counterflow of the normal and superfluid components of He-II

Heat transport

Figure 9 depicts a heat-conduction experiment between two temperatures TH and TL connected by a tube filled with He-II. When heat is applied to the hot end a pressure builds up at the hot end according to Eq. (7) . This pressure drives the normal component from the hot end to the cold end according to

(9)

Here ηn is the viscosity of the normal component, [30] Z some geometrical factor, and the volume flow. The normal flow is balanced by a flow of the superfluid component from the cold to the hot end. At the end sections a normal to superfluid conversion takes place and vice versa. So heat is transported, not by heat conduction, but by convection. This kind of heat transport is very effective, so the thermal conductivity of He-II is very much better than the best materials. The situation is comparable with heat pipes where heat is transported via gas–liquid conversion. The high thermal conductivity of He-II is applied for stabilizing superconducting magnets such as in the Large Hadron Collider at CERN.

Microscopic theory

Landau two-fluid approach

L. D. Landau's phenomenological and semi-microscopic theory of superfluidity of helium-4 earned him the Nobel Prize in physics, in 1962. Assuming that sound waves are the most important excitations in helium-4 at low temperatures, he showed that helium-4 flowing past a wall would not spontaneously create excitations if the flow velocity was less than the sound velocity. In this model, the sound velocity is the "critical velocity" above which superfluidity is destroyed. (Helium-4 actually has a lower flow velocity than the sound velocity, but this model is useful to illustrate the concept.) Landau also showed that the sound wave and other excitations could equilibrate with one another and flow separately from the rest of the helium-4, which is known as the "condensate".

From the momentum and flow velocity of the excitations he could then define a "normal fluid" density, which is zero at zero temperature and increases with temperature. At the so-called Lambda temperature, where the normal fluid density equals the total density, the helium-4 is no longer superfluid.

To explain the early specific heat data on superfluid helium-4, Landau posited the existence of a type of excitation he called a "roton", but as better data became available he considered that the "roton" was the same as a high momentum version of sound.

The Landau theory does not elaborate on the microscopic structure of the superfluid component of liquid helium. [31] The first attempts to create a microscopic theory of the superfluid component itself were done by London [32] and subsequently, Tisza. [33] [34] Other microscopical models have been proposed by different authors. Their main objective is to derive the form of the inter-particle potential between helium atoms in superfluid state from first principles of quantum mechanics. To date, a number of models of this kind have been proposed, including: models with vortex rings, hard-sphere models, and Gaussian cluster theories.

Vortex ring model

Landau thought that vorticity entered superfluid helium-4 by vortex sheets, but such sheets have since been shown to be unstable. Lars Onsager and, later independently, Feynman showed that vorticity enters by quantized vortex lines. They also developed the idea of quantum vortex rings. Arie Bijl in the 1940s, [35] and Richard Feynman around 1955, [36] developed microscopic theories for the roton, which was shortly observed with inelastic neutron experiments by Palevsky. Later on, Feynman admitted that his model gives only qualitative agreement with experiment. [37] [38]

Hard-sphere models

The models are based on the simplified form of the inter-particle potential between helium-4 atoms in the superfluid phase. Namely, the potential is assumed to be of the hard-sphere type. [39] [40] [41] In these models the famous Landau (roton) spectrum of excitations is qualitatively reproduced.

Gaussian cluster approach

This is a two-scale approach which describes the superfluid component of liquid helium-4. It consists of two nested models linked via parametric space. The short-wavelength part describes the interior structure of the fluid element using a non-perturbative approach based on the logarithmic Schrödinger equation; it suggests the Gaussian-like behaviour of the element's interior density and interparticle interaction potential. The long-wavelength part is the quantum many-body theory of such elements which deals with their dynamics and interactions. [42] The approach provides a unified description of the phonon, maxon and roton excitations, and has noteworthy agreement with experiment: with one essential parameter to fit one reproduces at high accuracy the Landau roton spectrum, sound velocity and structure factor of superfluid helium-4. [43] This model utilizes the general theory of quantum Bose liquids with logarithmic nonlinearities [44] which is based on introducing a dissipative-type contribution to energy related to the quantum Everett–Hirschman entropy function. [45] [46]

See also

Related Research Articles

<span class="mw-page-title-main">Bose–Einstein condensate</span> State of matter

In condensed matter physics, a Bose–Einstein condensate (BEC) is a state of matter that is typically formed when a gas of bosons at very low densities is cooled to temperatures very close to absolute zero. Under such conditions, a large fraction of bosons occupy the lowest quantum state, at which microscopic quantum-mechanical phenomena, particularly wavefunction interference, become apparent macroscopically. More generally, condensation refers to the appearance of macroscopic occupation of one or several states: for example, in BCS theory, a superconductor is a condensate of Cooper pairs. As such, condensation can be associated with phase transition, and the macroscopic occupation of the state is the order parameter.

<span class="mw-page-title-main">Liquid helium</span> Liquid state of the element helium

Liquid helium is a physical state of helium at very low temperatures at standard atmospheric pressures. Liquid helium may show superfluidity.

Quantum turbulence is the name given to the turbulent flow – the chaotic motion of a fluid at high flow rates – of quantum fluids, such as superfluids. The idea that a form of turbulence might be possible in a superfluid via the quantized vortex lines was first suggested by Richard Feynman. The dynamics of quantum fluids are governed by quantum mechanics, rather than classical physics which govern classical (ordinary) fluids. Some examples of quantum fluids include superfluid helium, Bose–Einstein condensates (BECs), polariton condensates, and nuclear pasta theorized to exist inside neutron stars. Quantum fluids exist at temperatures below the critical temperature at which Bose-Einstein condensation takes place.

<span class="mw-page-title-main">Roton</span> Collective excitation in superfluid helium-4 (a quasiparticle)

In theoretical physics, a roton is an elementary excitation, or quasiparticle, seen in superfluid helium-4 and Bose–Einstein condensates with long-range dipolar interactions or spin-orbit coupling. The dispersion relation of elementary excitations in this superfluid shows a linear increase from the origin, but exhibits first a maximum and then a minimum in energy as the momentum increases. Excitations with momenta in the linear region are called phonons; those with momenta close to the minimum are called rotons. Excitations with momenta near the maximum are called maxons.

<span class="mw-page-title-main">Supersolid</span> State of matter

In condensed matter physics, a supersolid is a spatially ordered material with superfluid properties. In the case of helium-4, it has been conjectured since the 1960s that it might be possible to create a supersolid. Starting from 2017, a definitive proof for the existence of this state was provided by several experiments using atomic Bose–Einstein condensates. The general conditions required for supersolidity to emerge in a certain substance are a topic of ongoing research.

<span class="mw-page-title-main">Fermionic condensate</span> State of matter

A fermionic condensate is a superfluid phase formed by fermionic particles at low temperatures. It is closely related to the Bose–Einstein condensate, a superfluid phase formed by bosonic atoms under similar conditions. The earliest recognized fermionic condensate described the state of electrons in a superconductor; the physics of other examples including recent work with fermionic atoms is analogous. The first atomic fermionic condensate was created by a team led by Deborah S. Jin using potassium-40 atoms at the University of Colorado Boulder in 2003.

<span class="mw-page-title-main">Helium-4</span> Isotope of helium

Helium-4 is a stable isotope of the element helium. It is by far the more abundant of the two naturally occurring isotopes of helium, making up about 99.99986% of the helium on Earth. Its nucleus is identical to an alpha particle, and consists of two protons and two neutrons.

<span class="mw-page-title-main">Lambda point</span>

The lambda point is the temperature at which normal fluid helium makes the transition to superfluid helium II. The lowest pressure at which He-I and He-II can coexist is the vapor−He-I−He-II triple point at 2.1768 K (−270.9732 °C) and 5.0418 kPa (0.049759 atm), which is the "saturated vapor pressure" at that temperature. The highest pressure at which He-I and He-II can coexist is the bcc−He-I−He-II triple point with a helium solid at 1.762 K (−271.388 °C), 29.725 atm (3,011.9 kPa).

<span class="mw-page-title-main">Quantum vortex</span> Quantized flux circulation of some physical quantity

In physics, a quantum vortex represents a quantized flux circulation of some physical quantity. In most cases, quantum vortices are a type of topological defect exhibited in superfluids and superconductors. The existence of quantum vortices was first predicted by Lars Onsager in 1949 in connection with superfluid helium. Onsager reasoned that quantisation of vorticity is a direct consequence of the existence of a superfluid order parameter as a spatially continuous wavefunction. Onsager also pointed out that quantum vortices describe the circulation of superfluid and conjectured that their excitations are responsible for superfluid phase transitions. These ideas of Onsager were further developed by Richard Feynman in 1955 and in 1957 were applied to describe the magnetic phase diagram of type-II superconductors by Alexei Alexeyevich Abrikosov. In 1935 Fritz London published a very closely related work on magnetic flux quantization in superconductors. London's fluxoid can also be viewed as a quantum vortex.

Superfluidity is a phenomenon where a fluid, or a fraction of a fluid, loses all its viscosity and can flow without resistance. A superfluid film is the thin film it may then form as a result.

In condensed matter physics, second sound is a quantum mechanical phenomenon in which heat transfer occurs by wave-like motion, rather than by the more usual mechanism of diffusion. Its presence leads to a very high thermal conductivity. It is known as "second sound" because the wave motion of entropy and temperature is similar to the propagation of pressure waves in air (sound). The phenomenon of second sound was first described by Lev Landau in 1941.

In a standard superconductor, described by a complex field fermionic condensate wave function, vortices carry quantized magnetic fields because the condensate wave function is invariant to increments of the phase by . There a winding of the phase by creates a vortex which carries one flux quantum. See quantum vortex.

<span class="mw-page-title-main">Moses H. W. Chan</span> Chinese-American physicist

Moses Hung-Wai Chan is a Chinese-American physicist who is Evan Pugh Professor at Pennsylvania State University. He is an alumnus of Bridgewater College and Cornell University, where he earned his Ph.D. in 1974 and was a postdoctoral associate at Duke University. He has been a professor at Penn State's University Park Campus since 1979.

Macroscopic quantum phenomena are processes showing quantum behavior at the macroscopic scale, rather than at the atomic scale where quantum effects are prevalent. The best-known examples of macroscopic quantum phenomena are superfluidity and superconductivity; other examples include the quantum Hall effect, Josephson effect and topological order. Since 2000 there has been extensive experimental work on quantum gases, particularly Bose–Einstein condensates.

<span class="mw-page-title-main">Superfluidity</span> Fluid which flows without losing kinetic energy

Superfluidity is the characteristic property of a fluid with zero viscosity which therefore flows without any loss of kinetic energy. When stirred, a superfluid forms vortices that continue to rotate indefinitely. Superfluidity occurs in two isotopes of helium when they are liquefied by cooling to cryogenic temperatures. It is also a property of various other exotic states of matter theorized to exist in astrophysics, high-energy physics, and theories of quantum gravity. The theory of superfluidity was developed by Soviet theoretical physicists Lev Landau and Isaak Khalatnikov.

Bose–Einstein condensation can occur in quasiparticles, particles that are effective descriptions of collective excitations in materials. Some have integer spins and can be expected to obey Bose–Einstein statistics like traditional particles. Conditions for condensation of various quasiparticles have been predicted and observed. The topic continues to be an active field of study.

<span class="mw-page-title-main">Electron-on-helium qubit</span> Quantum bit

An electron-on-helium qubit is a quantum bit for which the orthonormal basis states |0⟩ and |1⟩ are defined by quantized motional states or alternatively the spin states of an electron trapped above the surface of liquid helium. The electron-on-helium qubit was proposed as the basic element for building quantum computers with electrons on helium by Platzman and Dykman in 1999. 

Robert Everett Ecke is an American experimental physicist who is a laboratory fellow and director emeritus of the Center for Nonlinear Studies (CNLS) at Los Alamos National Laboratory and Affiliate Professor of Physics at the University of Washington. His research has included chaotic nonlinear dynamics, pattern formation, rotating Rayleigh-Bénard convection, two-dimensional turbulence, granular materials, and stratified flows. He is a Fellow of the American Physical Society (APS) and of the American Association for the Advancement of Science (AAAS), was chair of the APS Topical Group on Statistical and Nonlinear Physics, served in numerous roles in the APS Division of Fluid Dynamics, and was the Secretary of the Physics Section of the AAAS.

Turbulent phenomena are observed universally in energetic fluid dynamics, associated with highly chaotic fluid motion involving excitations spread over a wide range of length scales. The particular features of turbulence are dependent on the fluid and geometry, and specifics of forcing and dissipation.

References

  1. "Superfluidity". Encyclopedia of Condensed Matter Physics. Elsevier. 2005. pp. 128–133.
  2. "The Nobel Prize in Physics 1996 - Advanced Information". Nobel Foundation. Retrieved February 10, 2017.
  3. Kapitza, P. (1938). "Viscosity of Liquid Helium Below the λ-Point". Nature. 141 (3558): 74. Bibcode:1938Natur.141...74K. doi: 10.1038/141074a0 . S2CID   3997900.
  4. Allen, J. F.; Misener, A. D. (1938). "Flow of Liquid Helium II". Nature. 142 (3597): 643. Bibcode:1938Natur.142..643A. doi:10.1038/142643a0. S2CID   4135906.
  5. van Delft, Dirk; Kes, Peter (September 1, 2010). "The discovery of superconductivity". Physics Today. 63 (9): 38–43. Bibcode:2010PhT....63i..38V. doi: 10.1063/1.3490499 . ISSN   0031-9228.
  6. Hall, H. E.; Vinen, W. F. (1956). "The Rotation of Liquid Helium II. II. The Theory of Mutual Friction in Uniformly Rotating Helium II". Proceedings of the Royal Society A: Mathematical, Physical and Engineering Sciences. 238 (1213): 215. Bibcode:1956RSPSA.238..215H. doi:10.1098/rspa.1956.0215. S2CID   120738827.
  7. Rayfield, G.; Reif, F. (1964). "Quantized Vortex Rings in Superfluid Helium". Physical Review. 136 (5A): A1194. Bibcode:1964PhRv..136.1194R. doi:10.1103/PhysRev.136.A1194.
  8. Packard, Richard E. (1982). "Vortex photography in liquid helium" (PDF). Physica B. 109–110: 1474–1484. Bibcode:1982PhyBC.109.1474P. CiteSeerX   10.1.1.210.8701 . doi:10.1016/0378-4363(82)90510-1. Archived from the original (PDF) on November 7, 2017. Retrieved November 7, 2017.
  9. Avenel, O.; Varoquaux, E. (1985). "Observation of Singly Quantized Dissipation Events Obeying the Josephson Frequency Relation in the Critical Flow of Superfluid ^{4}He through an Aperture". Physical Review Letters. 55 (24): 2704–2707. Bibcode:1985PhRvL..55.2704A. doi:10.1103/PhysRevLett.55.2704. PMID   10032216.[ permanent dead link ]
  10. Bewley, Gregory P.; Lathrop, Daniel P.; Sreenivasan, Katepalli R. (2006). "Superfluid helium: Visualization of quantized vortices" (PDF). Nature. 441 (7093): 588. Bibcode:2006Natur.441..588B. doi: 10.1038/441588a . PMID   16738652. S2CID   4429923.
  11. E. Kim and M. H. W. Chan (2004). "Probable Observation of a Supersolid Helium Phase". Nature. 427 (6971): 225–227. Bibcode:2004Natur.427..225K. doi:10.1038/nature02220. PMID   14724632. S2CID   3112651.
  12. Moses Chan's Research Group. "Supersolid Archived 2013-04-08 at the Wayback Machine ." Penn State University, 2004.
  13. Sophie, A; Rittner C (2006). "Observation of Classical Rotational Inertia and Nonclassical Supersolid Signals in Solid 4 He below 250 mK". Phys. Rev. Lett. 97 (16): 165301. arXiv: cond-mat/0604528 . Bibcode:2006PhRvL..97p5301R. doi:10.1103/PhysRevLett.97.165301. PMID   17155406. S2CID   45453420.
  14. Sophie, A; Rittner C (2007). "Disorder and the Supersolid State of Solid 4 He". Phys. Rev. Lett. 98 (17): 175302. arXiv: cond-mat/0702665 . Bibcode:2007PhRvL..98q5302R. doi:10.1103/PhysRevLett.98.175302. S2CID   119469548.
  15. Boninsegni, M; Prokofev (2006). "Superglass Phase of 4 He". Phys. Rev. Lett. 96 (13): 135301. arXiv: cond-mat/0603003 . Bibcode:2006PhRvL..96m5301W. doi:10.1103/PhysRevLett.96.135301. PMID   16711998. S2CID   41657202.
  16. Pollet, L; Boninsegni M (2007). "Superfuididty of Grain Boundaries in Solid 4 He". Phys. Rev. Lett. 98 (13): 135301. arXiv: cond-mat/0702159 . Bibcode:2007PhRvL..98m5301P. doi:10.1103/PhysRevLett.98.135301. PMID   17501209. S2CID   20038102.
  17. Tanaeva, I. A. (2004). "Superfluid Vortex Cooler". AIP Conference Proceedings (PDF). Vol. 710. pp. 034911–1/8. doi:10.1063/1.1774894. S2CID   109758743.
  18. Ohba, Tomonori (2016). "Limited Quantum Helium Transportation through Nano-channels by Quantum Fluctuation". Scientific Reports. 6: 28992. Bibcode:2016NatSR...628992O. doi:10.1038/srep28992. PMC   4929499 . PMID   27363671.
  19. Swenson, C. (1950). "The Liquid-Solid Transformation in Helium near Absolute Zero". Physical Review. 79 (4): 626. Bibcode:1950PhRv...79..626S. doi:10.1103/PhysRev.79.626.
  20. Keesom, W.H.; Keesom, A.P. (1935). "New measurements on the specific heat of liquid helium". Physica. 2 (1): 557. Bibcode:1935Phy.....2..557K. doi:10.1016/S0031-8914(35)90128-8.
  21. Buckingham, M.J.; Fairbank, W.M. (1961). "Chapter III The Nature of the λ-Transition in Liquid Helium". The nature of the λ-transition in liquid helium. Progress in Low Temperature Physics. Vol. 3. p. 80. doi:10.1016/S0079-6417(08)60134-1. ISBN   978-0-444-53309-8.
  22. E.L. Andronikashvili Zh. Éksp. Teor. Fiz, Vol.16 p.780 (1946), Vol.18 p. 424 (1948)
  23. S. J. Putterman, Superfluid Hydrodynamics (North-Holland Publishing Company, Amsterdam, 1974) ISBN   0-444-10681-2.
  24. L. D. Landau, "The theory of superfluidity of helium II", J. Phys. USSR, Vol. 5 (1941) p. 71.
  25. I. M. Khalatnikov, An introduction to the theory of superfluidity (W. A. Benjamin, Inc., New York, 1965) ISBN   0-7382-0300-9.
  26. Van Alphen, W. M.; Van Haasteren, G. J.; De Bruyn Ouboter, R.; Taconis, K. W. (1966). "The dependence of the critical velocity of the superfluid on channel diameter and film thickness". Physics Letters. 20 (5): 474. Bibcode:1966PhL....20..474V. doi:10.1016/0031-9163(66)90958-9.
  27. De Waele, A. Th. A. M.; Kuerten, J. G. M. (1992). "Chapter 3: Thermodynamics and Hydrodynamics of 3He–4He Mixtures". Thermodynamics and hydrodynamics of 3He–4He mixtures. Progress in Low Temperature Physics. Vol. 13. p. 167. doi:10.1016/S0079-6417(08)60052-9. ISBN   978-0-444-89109-9.
  28. Staas, F.A.; Severijns, A.P.; Van Der Waerden, H.C.M. (1975). "A dilution refrigerator with superfluid injection". Physics Letters A. 53 (4): 327. Bibcode:1975PhLA...53..327S. doi:10.1016/0375-9601(75)90087-0.
  29. Castelijns, C.; Kuerten, J.; De Waele, A.; Gijsman, H. (1985). "3He flow in dilute 3He-4He mixtures at temperatures between 10 and 150 mK". Physical Review B. 32 (5): 2870–2886. Bibcode:1985PhRvB..32.2870C. doi:10.1103/PhysRevB.32.2870. PMID   9937394.
  30. J.C.H. Zeegers Critical velocities and mutual friction in 3He-4He mixtures at low temperatures below 100 mK', thesis, Appendix A, Eindhoven University of Technology, 1991.
  31. Alonso, J. L.; Ares, F.; Brun, J. L. (October 5, 2018). "Unraveling the Landau's consistence criterion and the meaning of interpenetration in the "Two-Fluid" Model". The European Physical Journal B. 91 (10): 226. arXiv: 1806.11034 . Bibcode:2018EPJB...91..226A. doi:10.1140/epjb/e2018-90105-x. ISSN   1434-6028. S2CID   53464405.
  32. F. London (1938). "The λ-Phenomenon of Liquid Helium and the Bose-Einstein Degeneracy". Nature. 141 (3571): 643–644. Bibcode:1938Natur.141..643L. doi:10.1038/141643a0. S2CID   4143290.
  33. L. Tisza (1938). "Transport Phenomena in Helium II". Nature. 141 (3577): 913. Bibcode:1938Natur.141..913T. doi: 10.1038/141913a0 . S2CID   4116542.
  34. L. Tisza (1947). "The Theory of Liquid Helium". Phys. Rev. 72 (9): 838–854. Bibcode:1947PhRv...72..838T. doi:10.1103/PhysRev.72.838.
  35. Bijl, A; de Boer, J; Michels, A (1941). "Properties of liquid helium II". Physica. 8 (7): 655–675. Bibcode:1941Phy.....8..655B. doi:10.1016/S0031-8914(41)90422-6.
  36. Braun, L. M., ed. (2000). Selected papers of Richard Feynman with commentary. World Scientific Series in 20th century Physics. Vol. 27. World Scientific. ISBN   978-9810241315. Section IV (pages 313 to 414) deals with liquid helium.
  37. R. P. Feynman (1954). "Atomic Theory of the Two-Fluid Model of Liquid Helium" (PDF). Phys. Rev. 94 (2): 262. Bibcode:1954PhRv...94..262F. doi:10.1103/PhysRev.94.262.
  38. R. P. Feynman & M. Cohen (1956). "Energy Spectrum of the Excitations in Liquid Helium" (PDF). Phys. Rev. 102 (5): 1189–1204. Bibcode:1956PhRv..102.1189F. doi:10.1103/PhysRev.102.1189.
  39. T. D. Lee; K. Huang & C. N. Yang (1957). "Eigenvalues and Eigenfunctions of a Bose System of Hard Spheres and Its Low-Temperature Properties". Phys. Rev. 106 (6): 1135–1145. Bibcode:1957PhRv..106.1135L. doi:10.1103/PhysRev.106.1135.
  40. L. Liu; L. S. Liu & K. W. Wong (1964). "Hard-Sphere Approach to the Excitation Spectrum in Liquid Helium II". Phys. Rev. 135 (5A): A1166–A1172. Bibcode:1964PhRv..135.1166L. doi:10.1103/PhysRev.135.A1166.
  41. A. P. Ivashin & Y. M. Poluektov (2011). "Short-wave excitations in non-local Gross-Pitaevskii model". Cent. Eur. J. Phys. 9 (3): 857–864. arXiv: 1004.0442 . Bibcode:2011CEJPh...9..857I. doi:10.2478/s11534-010-0124-7. S2CID   118633189.
  42. Santos, L.; Shlyapnikov, G. V.; Lewenstein, M. (2003). "Roton-Maxon Spectrum and Stability of Trapped Dipolar Bose-Einstein Condensates". Physical Review Letters. 90 (25): 250403. arXiv: cond-mat/0301474 . Bibcode:2003PhRvL..90y0403S. doi:10.1103/PhysRevLett.90.250403. PMID   12857119. S2CID   25309672.
  43. K. G. Zloshchastiev (2012). "Volume element structure and roton-maxon-phonon excitations in superfluid helium beyond the Gross-Pitaevskii approximation". Eur. Phys. J. B. 85 (8): 273. arXiv: 1204.4652 . Bibcode:2012EPJB...85..273Z. doi:10.1140/epjb/e2012-30344-3. S2CID   118545094.
  44. A. V. Avdeenkov & K. G. Zloshchastiev (2011). "Quantum Bose liquids with logarithmic nonlinearity: Self-sustainability and emergence of spatial extent". J. Phys. B: At. Mol. Opt. Phys. 44 (19): 195303. arXiv: 1108.0847 . Bibcode:2011JPhB...44s5303A. doi:10.1088/0953-4075/44/19/195303. S2CID   119248001.
  45. Hugh Everett, III. The Many-Worlds Interpretation of Quantum Mechanics: the theory of the universal wave function. Everett's Dissertation
  46. I.I. Hirschman, Jr., A note on entropy. American Journal of Mathematics (1957) pp. 152–156

Further reading