Quantum vortex

Last updated
Vortices in a 200-nm-thick YBCO film imaged by scanning SQUID microscopy YBCO vortices.jpg
Vortices in a 200-nm-thick YBCO film imaged by scanning SQUID microscopy

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. [2] 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 [3] and in 1957 were applied to describe the magnetic phase diagram of type-II superconductors by Alexei Alexeyevich Abrikosov. [4] 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.

Contents

Quantum vortices are observed experimentally in type-II superconductors (the Abrikosov vortex), liquid helium, and atomic gases [5] (see Bose–Einstein condensate), as well as in photon fields (optical vortex) and exciton-polariton superfluids.

In a superfluid, a quantum vortex "carries" quantized orbital angular momentum, thus allowing the superfluid to rotate; in a superconductor, the vortex carries quantized magnetic flux.

The term "quantum vortex" is also used in the study of few body problems. [6] [7] Under the de Broglie–Bohm theory, it is possible to derive a "velocity field" from the wave function. In this context, quantum vortices are zeros on the wave function, around which this velocity field has a solenoidal shape, similar to that of irrotational vortex on potential flows of traditional fluid dynamics.

Vortex-quantisation in a superfluid

In a superfluid, a quantum vortex is a hole with the superfluid circulating around the vortex axis; the inside of the vortex may contain excited particles, air, vacuum, etc. The thickness of the vortex depends on a variety of factors; in liquid helium, the thickness is of the order of a few Angstroms.

A superfluid has the special property of having phase, given by the wavefunction, and the velocity of the superfluid is proportional to the gradient of the phase (in the parabolic mass approximation). The circulation around any closed loop in the superfluid is zero if the region enclosed is simply connected. The superfluid is deemed irrotational; however, if the enclosed region actually contains a smaller region with an absence of superfluid, for example a rod through the superfluid or a vortex, then the circulation is:

where is Planck's constant divided by , m is the mass of the superfluid particle, and is the total phase difference around the vortex. Because the wave-function must return to its same value after an integer number of turns around the vortex (similar to what is described in the Bohr model), then , where n is an integer. Thus, the circulation is quantized:

London's flux quantization in a superconductor

A principal property of superconductors is that they expel magnetic fields; this is called the Meissner effect. If the magnetic field becomes sufficiently strong it will, in some cases, “quench” the superconductive state by inducing a phase transition. In other cases, however, it will be energetically favorable for the superconductor to form a lattice of quantum vortices, which carry quantized magnetic flux through the superconductor. A superconductor that is capable of supporting vortex lattices is called a type-II superconductor, vortex-quantization in superconductors is general.

Over some enclosed area S, the magnetic flux is

where is the vector potential of the magnetic induction

Substituting a result of London's equation: , we find (with ):

where ns, m, and es are, respectively, number density, mass, and charge of the Cooper pairs.

If the region, S, is large enough so that along , then

The flow of current can cause vortices in a superconductor to move, causing the electric field due to the phenomenon of electromagnetic induction. This leads to energy dissipation and causes the material to display a small amount of electrical resistance while in the superconducting state. [8]

Constrained vortices in ferromagnets and antiferromagnets

The vortex states in ferromagnetic or antiferromagnetic material are also important, mainly for information technology [9] They are exceptional, since in contrast to superfluids or superconducting material one has a more subtle mathematics: instead of the usual equation of the type where is the vorticity at the spatial and temporal coordinates, and where is the Dirac function, one has:

where now at any point and at any time there is the constraint . Here is constant, the constant magnitude of the non-constant magnetization vector . As a consequence the vector in eqn. (*) has been modified to a more complex entity . This leads, among other points, to the following fact:

In ferromagnetic or antiferromagnetic material a vortex can be moved to generate bits for information storage and recognition, corresponding, e.g., to changes of the quantum number n. [9] But although the magnetization has the usual azimuthal direction, and although one has vorticity quantization as in superfluids, as long as the circular integration lines surround the central axis at far enough perpendicular distance, this apparent vortex magnetization will change with the distance from an azimuthal direction to an upward or downward one, as soon as the vortex center is approached.

Thus, for each directional element there are now not two, but four bits to be stored by a change of vorticity: The first two bits concern the sense of rotation, clockwise or counterclockwise; the remaining bits three and four concern the polarization of the central singular line, which may be polarized up- or downwards. The change of rotation and/or polarization involves subtle topology. [10]

Statistical mechanics of vortex lines

As first discussed by Onsager and Feynman, if the temperature in a superfluid or a superconductor is raised, the vortex loops undergo a second-order phase transition. This happens when the configurational entropy overcomes the Boltzmann factor which suppresses the thermal or heat generation of vortex lines. The lines form a condensate. Since the center of the lines, the vortex cores, are normal liquid or normal conductors, respectively, the condensation transforms the superfluid or superconductor into the normal state. The ensembles of vortex lines and their phase transitions can be described efficiently by a gauge theory.

Statistical mechanics of point vortices

In 1949 Onsager analysed a toy model consisting of a neutral system of point vortices confined to a finite area. [2] He was able to show that, due to the properties of two-dimensional point vortices the bounded area (and consequently, bounded phase space), allows the system to exhibit negative temperatures. Onsager provided the first prediction that some isolated systems can exhibit negative Boltzmann temperature. Onsager's prediction was confirmed experimentally for a system of quantum vortices in a Bose-Einstein condensate in 2019. [11] [12]

Pair-interactions of quantum vortices

In a nonlinear quantum fluid, the dynamics and configurations of the vortex cores can be studied in terms of effective vortex-vortex pair interactions. The effective intervortex potential is predicted to affect quantum phase transitions and giving rise to different few-vortex molecules and many-body vortex patterns. [13] [14] Preliminary experiments in the specific system of exciton-polaritons fluids showed an effective attractive-repulsive intervortex dynamics between two cowinding vortices, whose attractive component can be modulated by the nonlinearity amount in the fluid. [15]

Spontaneous vortices

Quantum vortices can form via the Kibble–Zurek mechanism. As a condensate forms by quench cooling, separate protocondensates form with independent phases. As these phase domains merge quantum vortices can be trapped in the emerging condensate order parameter. Spontaneous quantum vortices were observed in atomic Bose-Einstein condensates in 2008. [16]

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">Lorentz force</span> Force acting on charged particles in electric and magnetic fields

In physics, specifically in electromagnetism, the Lorentz force is the combination of electric and magnetic force on a point charge due to electromagnetic fields. A particle of charge q moving with a velocity v in an electric field E and a magnetic field B experiences a force of

The magnetic flux, represented by the symbol Φ, threading some contour or loop is defined as the magnetic field B multiplied by the loop area S, i.e. Φ = BS. Both B and S can be arbitrary, meaning that the flux Φ can be as well but increments of flux can be quantized. The wave function can be multivalued as it happens in the Aharanov-Bohm effect or quantized as in superconductors. The unit of quantization is therefore called magnetic flux quantum.

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.

In physics, Ginzburg–Landau theory, often called Landau–Ginzburg theory, named after Vitaly Ginzburg and Lev Landau, is a mathematical physical theory used to describe superconductivity. In its initial form, it was postulated as a phenomenological model which could describe type-I superconductors without examining their microscopic properties. One GL-type superconductor is the famous YBCO, and generally all cuprates.

<span class="mw-page-title-main">Josephson effect</span> Quantum physical phenomenon

In physics, the Josephson effect is a phenomenon that occurs when two superconductors are placed in proximity, with some barrier or restriction between them. The effect is named after the British physicist Brian Josephson, who predicted in 1962 the mathematical relationships for the current and voltage across the weak link. It is an example of a macroscopic quantum phenomenon, where the effects of quantum mechanics are observable at ordinary, rather than atomic, scale. The Josephson effect has many practical applications because it exhibits a precise relationship between different physical measures, such as voltage and frequency, facilitating highly accurate measurements.

<span class="mw-page-title-main">Magnetic vector potential</span> Integral of the magnetic field

In classical electromagnetism, magnetic vector potential is the vector quantity defined so that its curl is equal to the magnetic field: . Together with the electric potential φ, the magnetic vector potential can be used to specify the electric field E as well. Therefore, many equations of electromagnetism can be written either in terms of the fields E and B, or equivalently in terms of the potentials φ and A. In more advanced theories such as quantum mechanics, most equations use potentials rather than fields.

The Little–Parks effect was discovered in 1962 by William A. Little and Ronald D. Parks in experiments with empty and thin-walled superconducting cylinders subjected to a parallel magnetic field. It was one of the first experiments to indicate the importance of Cooper-pairing principle in BCS theory.

The Berezinskii–Kosterlitz–Thouless (BKT) transition is a phase transition of the two-dimensional (2-D) XY model in statistical physics. It is a transition from bound vortex-antivortex pairs at low temperatures to unpaired vortices and anti-vortices at some critical temperature. The transition is named for condensed matter physicists Vadim Berezinskii, John M. Kosterlitz and David J. Thouless. BKT transitions can be found in several 2-D systems in condensed matter physics that are approximated by the XY model, including Josephson junction arrays and thin disordered superconducting granular films. More recently, the term has been applied by the 2-D superconductor insulator transition community to the pinning of Cooper pairs in the insulating regime, due to similarities with the original vortex BKT transition.

<span class="mw-page-title-main">Abrikosov vortex</span>

In superconductivity, a fluxon is a vortex of supercurrent in a type-II superconductor, used by Alexei Abrikosov to explain magnetic behavior of type-II superconductors. Abrikosov vortices occur generically in the Ginzburg–Landau theory of superconductivity.

<span class="mw-page-title-main">Type-II superconductor</span> Superconductor characterized by the formation of magnetic vortices in an applied magnetic field

In superconductivity, a type-II superconductor is a superconductor that exhibits an intermediate phase of mixed ordinary and superconducting properties at intermediate temperature and fields above the superconducting phases. It also features the formation of magnetic field vortices with an applied external magnetic field. This occurs above a certain critical field strength Hc1. The vortex density increases with increasing field strength. At a higher critical field Hc2, superconductivity is destroyed. Type-II superconductors do not exhibit a complete Meissner effect.

The superconductor–insulator transition is an example of a quantum phase transition, whereupon tuning some parameter in the Hamiltonian, a dramatic change in the behavior of the electrons occurs. The nature of how this transition occurs is disputed, and many studies seek to understand how the order parameter, , changes. Here is the amplitude of the order parameter, and is the phase. Most theories involve either the destruction of the amplitude of the order parameter - by a reduction in the density of states at the Fermi surface, or by destruction of the phase coherence; which results from the proliferation of vortices.

In quantum mechanics, the Pauli equation or Schrödinger–Pauli equation is the formulation of the Schrödinger equation for spin-½ particles, which takes into account the interaction of the particle's spin with an external electromagnetic field. It is the non-relativistic limit of the Dirac equation and can be used where particles are moving at speeds much less than the speed of light, so that relativistic effects can be neglected. It was formulated by Wolfgang Pauli in 1927. In its linearized form it is known as Lévy-Leblond equation.

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.

An LC circuit can be quantized using the same methods as for the quantum harmonic oscillator. An LC circuit is a variety of resonant circuit, and consists of an inductor, represented by the letter L, and a capacitor, represented by the letter C. When connected together, an electric current can alternate between them at the circuit's resonant frequency:

The Aharonov–Casher effect is a quantum mechanical phenomenon predicted in 1984 by Yakir Aharonov and Aharon Casher, in which a traveling magnetic dipole is affected by an electric field. It is dual to the Aharonov–Bohm effect, in which the quantum phase of a charged particle depends upon which side of a magnetic flux tube it comes through. In the Aharonov–Casher effect, the particle has a magnetic moment and the tubes are charged instead. It was observed in a gravitational neutron interferometer in 1989 and later by fluxon interference of magnetic vortices in Josephson junctions. It has also been seen with electrons and atoms.

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.

Lagrangian field theory is a formalism in classical field theory. It is the field-theoretic analogue of Lagrangian mechanics. Lagrangian mechanics is used to analyze the motion of a system of discrete particles each with a finite number of degrees of freedom. Lagrangian field theory applies to continua and fields, which have an infinite number of degrees of freedom.

In quantum mechanics, the Byers–Yang theorem states that all physical properties of a doubly connected system enclosing a magnetic flux through the opening are periodic in the flux with period . The theorem was first stated and proven by Nina Byers and Chen-Ning Yang (1961), and further developed by Felix Bloch (1970).

The Peierls substitution method, named after the original work by Rudolf Peierls is a widely employed approximation for describing tightly-bound electrons in the presence of a slowly varying magnetic vector potential.

References

  1. Wells, Frederick S.; Pan, Alexey V.; Wang, X. Renshaw; Fedoseev, Sergey A.; Hilgenkamp, Hans (2015). "Analysis of low-field isotropic vortex glass containing vortex groups in YBa2Cu3O7−x thin films visualized by scanning SQUID microscopy". Scientific Reports. 5: 8677. arXiv: 1807.06746 . Bibcode:2015NatSR...5E8677W. doi:10.1038/srep08677. PMC   4345321 . PMID   25728772.
  2. 1 2 Onsager, L. (1949). "Statistical Hydrodynamics". Il Nuovo Cimento. 6(Suppl 2) (2): 279–287. Bibcode:1949NCim....6S.279O. doi:10.1007/BF02780991. ISSN   1827-6121. S2CID   186224016.
  3. Feynman, R. P. (1955). "Application of quantum mechanics to liquid helium". Progress in Low Temperature Physics. 1: 17–53. doi:10.1016/S0079-6417(08)60077-3. ISBN   978-0-444-53307-4.
  4. Abrikosov, A. A. (1957) "On the Magnetic properties of superconductors of the second group", Sov. Phys. JETP 5:1174–1182 and Zh. Eksp. Teor. Fiz. 32:1442–1452.
  5. Matthews, M. R.; Anderson, B. P.; Haljan, P. C.; Hall, D. S; Wieman, C. E.; Cornell, E. A. (1999). "Vortices in a Bose-Einstein Condensate". Physical Review Letters. 83 (13): 2498–2501. arXiv: cond-mat/9908209 . Bibcode:1999PhRvL..83.2498M. doi:10.1103/PhysRevLett.83.2498. S2CID   535347.
  6. Macek, J. H.; Sternberg, J. B.; Ovchinnikov, S. Y.; Briggs, J. S. (2010-01-20). "Theory of Deep Minima in $(e,2e)$ Measurements of Triply Differential Cross Sections". Physical Review Letters. 104 (3): 033201. Bibcode:2010PhRvL.104c3201M. doi:10.1103/PhysRevLett.104.033201. PMID   20366640.
  7. Navarrete, F; Picca, R Della; Fiol, J; Barrachina, R O (2013). "Vortices in ionization collisions by positron impact". Journal of Physics B: Atomic, Molecular and Optical Physics. 46 (11): 115203. arXiv: 1302.4357 . Bibcode:2013JPhB...46k5203N. doi:10.1088/0953-4075/46/11/115203. hdl:11336/11099. S2CID   119277044.
  8. "First vortex 'chains' observed in engineered superconductor". Physorg.com. June 20, 2017. Retrieved 2011-03-23.
  9. 1 2 Magnetic vortices in nanodisks reveal information. Phys.org (March 3, 2015).
  10. Pylipovskyi, O.V. et al. (January 2015) "Polarity Switching in Magnets with Surface Anisotropy. arxiv.org
  11. Gauthier, G.; Reeves, M. T.; Yu, X.; Bradley, A. S.; Baker, M. A.; Bell, T. A.; Rubinsztein-Dunlop, H.; Davis, M. J.; Neely, T. W. (2019). "Giant vortex clusters in a two-dimensional quantum fluid". Science. 364 (6447): 1264–1267. arXiv: 1801.06951 . Bibcode:2019Sci...364.1264G. doi:10.1126/science.aat5718. PMID   31249054. S2CID   195750381.
  12. Johnstone, S. P.; Groszek, A. J.; Starkey, P. T.; Billinton, C. J.; Simula, T. P.; Helmerson, K. (2019). "Evolution of large-scale flow from turbulence in a two-dimensional superfluid". Science. 365 (6447): 1267–1271. arXiv: 1801.06952 . Bibcode:2019Sci...364.1267J. doi:10.1126/science.aat5793. PMID   31249055. S2CID   4948239.
  13. Zhao, H. J.; Misko, V. R.; Tempere, J.; Nori, F. (2017). "Pattern formation in vortex matter with pinning and frustrated intervortex interactions". Phys. Rev. B. 95 (10): 104519. arXiv: 1704.00225 . Bibcode:2017PhRvB..95j4519Z. doi:10.1103/PhysRevB.95.104519. S2CID   52245546.
  14. Wei, C.A.; Xu, X.B.; Xu, X.N.; Wang, Z.H.; Gu, M. (2018). "Equilibrium vortex structures of type-II/1 superconducting films with washboard pinning landscapes". Physica C: Superconductivity and Its Applications. 548: 55–60. Bibcode:2018PhyC..548...55W. doi: 10.1016/j.physc.2018.02.005 .
  15. Dominici, L; Carretero-González, R; Gianfrate, A; et al. (2018). "Interactions and scattering of quantum vortices in a polariton fluid". Nature Communications. 9 (1): 1467. arXiv: 1706.00143 . Bibcode:2018NatCo...9.1467D. doi: 10.1038/s41467-018-03736-5 . PMC   5899148 . PMID   29654228.
  16. Weiler, C. N.; Neely, T. W.; Scherer, D. R.; Bradley, A. S.; Davis, M. J.; Anderson, B. P. (2009). "Spontaneous vortices in the formation of Bose-Einstein condensates". Nature. 455 (7215): 948–951. arXiv: 0807.3323 . Bibcode:2008Natur.455..948W. doi:10.1038/nature07334. S2CID   459795.