Second sound

Last updated

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). [1] The phenomenon of second sound was first described by Lev Landau in 1941. [2]

Contents

Description

Normal sound waves are fluctuations in the displacement and density of molecules in a substance; [3] [4] second sound waves are fluctuations in the density of quasiparticle thermal excitations (rotons and phonons [5] ). Second sound can be observed in any system in which most phonon-phonon collisions conserve momentum, like superfluids [6] and in some dielectric crystals [1] [7] [8] when Umklapp scattering is small.

Contrary to molecules in a gas, quasiparticles are not necessarily conserved. Also gas molecules in a box conserve momentum (except at the boundaries of box), while quasiparticles can sometimes not conserve momentum in the presence of impurities or Umklapp scattering. Umklapp phonon-phonon scattering exchanges momentum with the crystal lattice, so phonon momentum is not conserved, but Umklapp processes can be reduced at low temperatures. [9]

Normal sound in gases is a consequence of the collision rate τ between molecules being large compared to the frequency of the sound wave ω 1/τ. For second sound, the Umklapp rate τu has to be small compared to the oscillation frequency ω 1/τu for energy and momentum conservation. However analogous to gasses, the relaxation time τN describing the collisions has to be large with respect to the frequency ω 1/τN, leaving a window: [9]

for sound-like behaviour or second sound. The second sound thus behaves as oscillations of the local number of quasiparticles (or of the local energy carried by these particles). Contrary to the normal sound where energy is related to pressure and temperature, in a crystal the local energy density is purely a function of the temperature. In this sense, the second sound can also be considered as oscillations of the local temperature. Second sound is a wave-like phenomena which makes it very different from usual heat diffusion. [9]

In helium II

Second sound is observed in liquid helium at temperatures below the lambda point, 2.1768  K, where 4He becomes a superfluid known as helium II. Helium II has the highest thermal conductivity of any known material (several hundred times higher than copper). [10] Second sound can be observed either as pulses or in a resonant cavity. [11]

The speed of second sound is close to zero near the lambda point, increasing to approximately 20 m/s around 1.8 K, [12] about ten times slower than normal sound waves. [13] At temperatures below 1 K, the speed of second sound in helium II increases as the temperature decreases. [14]

Second sound is also observed in superfluid helium-3 below its lambda point 2.5 mK. [15]

As per the two-fluid, the speed of second sound is given by

where is the temperature, is the entropy, is the specific heat, is the superfluid density and is the normal fluid density. As , , where is the ordinary (or first) sound speed.

In other media

Second sound has been observed in solid 4He and 3He, [16] [17] and in some dielectric solids such as Bi in the temperature range of 1.2 to 4.0 K with a velocity of 780 ± 50 m/s, [18] or solid sodium fluoride (NaF) around 10 to 20 K. [19] In 2021 this effect was observed in a BKT superfluid [20] as well as in a germanium semiconductor [21] [22]

In graphite

In 2019 it was reported that ordinary graphite exhibits second sound at 120 K. This feature was both predicted theoretically and observed experimentally, and was by far the highest temperature at which second sound has been observed. [23] However, this second sound is observed only at the microscale, because the wave dies out exponentially with characteristic length 1-10 microns. Therefore, presumably graphite in the right temperature regime has extraordinarily high thermal conductivity but only for the purpose of transferring heat pulses distances of order 10 microns, and for pulses of duration on the order of 10 nanoseconds. For more "normal" heat-transfer, graphite's observed thermal conductivity is less than that of, e.g., copper. The theoretical models, however, predict longer absorption lengths would be seen in isotopically pure graphite, and perhaps over a wider temperature range, e.g. even at room temperature. (As of March 2019, that experiment has not yet been tried.)

Applications

Measuring the speed of second sound in 3He-4He mixtures can be used as a thermometer in the range 0.01-0.7 K. [24]

Oscillating superleak transducers (OST) [25] use second sound to locate defects in superconducting accelerator cavities. [26] [27]

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.

Superfluid helium-4 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 looks like a normal liquid, 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.

<span class="mw-page-title-main">Fermi liquid theory</span> Theoretical model in physics

Fermi liquid theory is a theoretical model of interacting fermions that describes the normal state of the conduction electrons in most metals at sufficiently low temperatures. The theory describes the behavior of many-body systems of particles in which the interactions between particles may be strong. The phenomenological theory of Fermi liquids was introduced by the Soviet physicist Lev Davidovich Landau in 1956, and later developed by Alexei Abrikosov and Isaak Khalatnikov using diagrammatic perturbation theory. The theory explains why some of the properties of an interacting fermion system are very similar to those of the ideal Fermi gas, and why other properties differ.

<span class="mw-page-title-main">Polariton</span> Quasiparticles arising from EM wave coupling

In physics, polaritons are quasiparticles resulting from strong coupling of electromagnetic waves with an electric or magnetic dipole-carrying excitation. They are an expression of the common quantum phenomenon known as level repulsion, also known as the avoided crossing principle. Polaritons describe the crossing of the dispersion of light with any interacting resonance. To this extent polaritons can also be thought of as the new normal modes of a given material or structure arising from the strong coupling of the bare modes, which are the photon and the dipolar oscillation. The polariton is a bosonic quasiparticle, and should not be confused with the polaron, which is an electron plus an attached phonon cloud.

<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">Homes's law</span>

In superconductivity, Homes's law is an empirical relation that states that a superconductor's critical temperature (Tc) is proportional to the strength of the superconducting state for temperatures well below Tc close to zero temperature (also referred to as the fully formed superfluid density, ) multiplied by the electrical resistivity measured just above the critical temperature. In cuprate high-temperature superconductors the relation follows the form

<span class="mw-page-title-main">Magnon</span> Spin 1 quasiparticle; quantum of a spin wave

A magnon is a quasiparticle, a collective excitation of the spin structure of an electron in a crystal lattice. In the equivalent wave picture of quantum mechanics, a magnon can be viewed as a quantized spin wave. Magnons carry a fixed amount of energy and lattice momentum, and are spin-1, indicating they obey boson behavior.

<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">Rollin film</span> Liquid film of superfluid helium

A Rollin film, named after Bernard V. Rollin, is a 30 nm-thick liquid film of helium in the helium II state. It exhibits a "creeping" effect in response to surfaces extending past the film's level. Helium II can escape from any non-closed container via creeping toward and eventually evaporating from capillaries of 10 to 100 nm or greater.

A two-dimensional electron gas (2DEG) is a scientific model in solid-state physics. It is an electron gas that is free to move in two dimensions, but tightly confined in the third. This tight confinement leads to quantized energy levels for motion in the third direction, which can then be ignored for most problems. Thus the electrons appear to be a 2D sheet embedded in a 3D world. The analogous construct of holes is called a two-dimensional hole gas (2DHG), and such systems have many useful and interesting properties.

Relativistic heat conduction refers to the modelling of heat conduction in a way compatible with special relativity. In special relativity, the usual heat equation for non-relativistic heat conduction must be modified, as it leads to faster-than-light signal propagation. Relativistic heat conduction, therefore, encompasses a set of models for heat propagation in continuous media that are consistent with relativistic causality, namely the principle that an effect must be within the light-cone associated to its cause. Any reasonable relativistic model for heat conduction must also be stable, in the sense that differences in temperature propagate both slower than light and are damped over time.

A composite fermion is the topological bound state of an electron and an even number of quantized vortices, sometimes visually pictured as the bound state of an electron and, attached, an even number of magnetic flux quanta. Composite fermions were originally envisioned in the context of the fractional quantum Hall effect, but subsequently took on a life of their own, exhibiting many other consequences and phenomena.

As the devices continue to shrink further into the sub-100 nm range following the trend predicted by Moore’s law, the topic of thermal properties and transport in such nanoscale devices becomes increasingly important. Display of great potential by nanostructures for thermoelectric applications also motivates the studies of thermal transport in such devices. These fields, however, generate two contradictory demands: high thermal conductivity to deal with heating issues in sub-100 nm devices and low thermal conductivity for thermoelectric applications. These issues can be addressed with phonon engineering, once nanoscale thermal behaviors have been studied and understood.

Phonons can scatter through several mechanisms as they travel through the material. These scattering mechanisms are: Umklapp phonon-phonon scattering, phonon-impurity scattering, phonon-electron scattering, and phonon-boundary scattering. Each scattering mechanism can be characterised by a relaxation rate 1/ which is the inverse of the corresponding relaxation time.

The Kibble–Zurek mechanism (KZM) describes the non-equilibrium dynamics and the formation of topological defects in a system which is driven through a continuous phase transition at finite rate. It is named after Tom W. B. Kibble, who pioneered the study of domain structure formation through cosmological phase transitions in the early universe, and Wojciech H. Zurek, who related the number of defects it creates to the critical exponents of the transition and to its rate—to how quickly the critical point is traversed.

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. 

References

  1. 1 2 Srinivasan, R (June 1999). "Second Sound: The Role of Elastic Waves" (PDF). Resonance. 4: 15–19. doi:10.1007/bf02834631. S2CID   124849291.
  2. Landau, L. (1941). Theory of the superfluidity of helium II. Physical Review, 60(4), 356.
  3. Feynman, Richard (4 October 2011). Feynman Lectures on Physics. Basic Books. ISBN   978-0465024933.
  4. Feynman. "Sound. The wave equation". feynmanlectures.caltech.edu. Caltech. Retrieved 20 July 2021.
  5. Smith, Henrik; Jensen, H. Hojgaard (1989). "Section 4.3: Second Sound". Transport Phenomena. Oxford University Press. ISBN   0-19-851985-0.
  6. Srinivasan, R (March 1999). "Second Sound: Waves of Entropy and Temperature" (PDF). Resonance. 3: 16–24. doi:10.1007/BF02838720. S2CID   123957486.
  7. Prohofsky, E.; Krumhansl, J. (1964). "Second-Sound Propagation in Dielectric Solids". Physical Review. 133 (5A): A1403. Bibcode:1964PhRv..133.1403P. doi:10.1103/PhysRev.133.A1403.
  8. Chester, M. (1963). "Second Sound in Solids". Physical Review. 131 (5): 2013–2015. Bibcode:1963PhRv..131.2013C. doi:10.1103/PhysRev.131.2013.
  9. 1 2 3 Ashcroft, Neil W.; Mermin, N. David (1976). Solid State Physics. Holt, Rinehart and Winston. ISBN   978-0-03-083993-1.
  10. Lebrun, Phillipe (July 17, 1997). Superfluid helium as a technical coolant (PDF) (LHC-Project-Report-125). CERN. p. 4.
  11. Van Der Boog, A. G. M.; Husson, L. P. J.; Disatnik, Y.; Kramers, H. C. (1981). "Experimental results on the velocity of second sound and the viscosity in dilute 3He-4He mixtures". Physica B+C. 104 (3): 303–315. Bibcode:1981PhyBC.104..303V. doi:10.1016/0378-4363(81)90176-5.
  12. Wang, R. T.; Wagner, W. T.; Donnelly, R. J. (1987). "Precision second-sound velocity measurements in helium II". Journal of Low Temperature Physics. 68 (5–6): 409–417. Bibcode:1987JLTP...68..409W. doi:10.1007/BF00682305. S2CID   120789592.
  13. Lane, C.; Fairbank, H.; Fairbank, W. (1947). "Second Sound in Liquid Helium II". Physical Review. 71 (9): 600–605. Bibcode:1947PhRv...71..600L. doi:10.1103/PhysRev.71.600.
  14. De Klerk, D.; Hudson, R.; Pellam, J. (1954). "Second Sound Propagation below 1K". Physical Review. 93 (1): 28–37. Bibcode:1954PhRv...93...28D. doi:10.1103/PhysRev.93.28.
  15. Lu, S.; Kojima, H. (1985). "Observation of Second Sound in Superfluid ^{3}He-B". Physical Review Letters. 55 (16): 1677–1680. Bibcode:1985PhRvL..55.1677L. doi:10.1103/PhysRevLett.55.1677. PMID   10031890.
  16. Ackerman, C.; Bertman, B.; Fairbank, H.; Guyer, R. (1966). "Second Sound in Solid Helium". Physical Review Letters. 16 (18): 789–791. Bibcode:1966PhRvL..16..789A. doi:10.1103/PhysRevLett.16.789.
  17. Ackerman, C.; Overton, W. (1969). "Second Sound in Solid Helium-3". Physical Review Letters. 22 (15): 764–766. Bibcode:1969PhRvL..22..764A. doi:10.1103/PhysRevLett.22.764.
  18. Narayanamurti, V.; Dynes, R. (1972). "Observation of Second Sound in Bismuth". Physical Review Letters. 28 (22): 1461–1465. Bibcode:1972PhRvL..28.1461N. doi:10.1103/PhysRevLett.28.1461.
  19. Jackson, H.; Walker, C.; McNelly, T. (1970). "Second Sound in NaF". Physical Review Letters. 25 (1): 26–28. Bibcode:1970PhRvL..25...26J. doi:10.1103/PhysRevLett.25.26.
  20. Christodoulou P, Gałka M, Dogra N, et al. (10 June 2021). "Observation of first and second sound in a BKT superfluid". Nature. 594 (7862): 191–194. arXiv: 2008.06044 . Bibcode:2021Natur.594..191C. doi:10.1038/s41586-021-03537-9. PMID   34108696. S2CID   235394222.
  21. Beardo, Albert; López-Suárez, Miquel; Pérez, Luis Alberto; Sendra, Lluc; Alonso, Maria Isabel; Melis, Claudio; Bafaluy, Javier; Camacho, Juan; Colombo, Luciano; Rurali, Riccardo; Alvarez, Francesc Xavier; Reparaz, Sebastian (2021-06-01). "Observation of second sound in a rapidly varying temperature field in Ge". Science Advances. 7 (27): eabg4677. arXiv: 2007.05487 . Bibcode:2021SciA....7.4677B. doi:10.1126/sciadv.abg4677. ISSN   2375-2548. PMC   8245038 . PMID   34193427.
  22. "'Second sound' appears in germanium". Physics World. 2021-07-18. Retrieved 2021-07-20.
  23. Huberman, S.; Duncan, R.A. (2019). "Observation of second sound in graphite at temperatures above 100 K". Science. 364 (6438): 375–379. arXiv: 1901.09160 . Bibcode:2019Sci...364..375H. doi:10.1126/science.aav3548. PMID   30872535. S2CID   78091609.
  24. Pitre, L. (2003). "The Comparison between a Second-Sound Thermometer and a Melting-Curve Thermometer from 0.8 K Down to 20 mK". AIP Conference Proceedings. Vol. 684. pp. 101–106. doi:10.1063/1.1627108.
  25. Sherlock, R. A. (1970). "Oscillating Superleak Second Sound Transducers". Review of Scientific Instruments. 41 (11): 1603–1609. Bibcode:1970RScI...41.1603S. doi: 10.1063/1.1684354 .
  26. Hesla, Leah (21 April 2011). "The sound of accelerator cavities". ILC Newsline. Retrieved 26 October 2012.
  27. Quadt, A.; Schröder, B.; Uhrmacher, M.; Weingarten, J.; Willenberg, B.; Vennekate, H. (2012). "Response of an oscillating superleak transducer to a pointlike heat source". Physical Review Special Topics: Accelerators and Beams. 15 (3): 031001. arXiv: 1111.5520 . Bibcode:2012PhRvS..15c1001Q. doi:10.1103/PhysRevSTAB.15.031001. S2CID   118996515.

Bibliography