SQUID

Last updated

Sensing element of a SQUID, 2008 SQUID by Zureks.jpg
Sensing element of a SQUID, 2008

A SQUID (superconducting quantum interference device) is a very sensitive magnetometer used to measure extremely weak magnetic fields, based on superconducting loops containing Josephson junctions.

Contents

SQUIDs are sensitive enough to measure fields as low as 5×10−14 T with a few days of averaged measurements. [1] Their noise levels are as low as 3 fHz 12. [2] For comparison, a typical refrigerator magnet produces 0.01 tesla (10−2 T), and some processes in animals produce very small magnetic fields between 10−9 T and 10−6 T. SERF atomic magnetometers, invented in the early 2000s are potentially more sensitive and do not require cryogenic refrigeration but are orders of magnitude larger in size (~1 cm3) and must be operated in a near-zero magnetic field.

History and design

There are two main types of SQUID: direct current (DC) and radio frequency (RF). RF SQUIDs can work with only one Josephson junction (superconducting tunnel junction), which might make them cheaper to produce, but are less sensitive.

DC SQUID

Diagram of a DC SQUID. The current
I
{\displaystyle I}
enters and splits into the two paths, each with currents
I
a
{\displaystyle I_{a}}
and
I
b
{\displaystyle I_{b}}
. The thin barriers on each path are Josephson junctions, which together separate the two superconducting regions.
Ph
{\displaystyle \Phi }
represents the magnetic flux threading the DC SQUID loop. DC SQUID.svg
Diagram of a DC SQUID. The current enters and splits into the two paths, each with currents and . The thin barriers on each path are Josephson junctions, which together separate the two superconducting regions. represents the magnetic flux threading the DC SQUID loop.
Electrical schematic of a SQUID where
I
b
{\displaystyle I_{b}}
is the bias current,
I
0
{\displaystyle I_{0}}
is the critical current of the SQUID,
Ph
{\displaystyle \Phi }
is the flux threading the SQUID and
V
{\displaystyle V}
is the voltage response to that flux. The X-symbols represent Josephson junctions. SQUID IV.svg
Electrical schematic of a SQUID where is the bias current, is the critical current of the SQUID, is the flux threading the SQUID and is the voltage response to that flux. The X-symbols represent Josephson junctions.
Left: Plot of current vs. voltage for a SQUID. Upper and lower curves correspond to
n
[?]
Ph
0
{\displaystyle n\cdot \Phi _{0}}
and
n
+
1
2
[?]
Ph
0
{\displaystyle n+{\frac {1}{2}}\cdot \Phi _{0}}
respectively. Right: Periodic voltage response due to flux through a SQUID. The periodicity is equal to one flux quantum,
Ph
0
{\displaystyle \Phi _{0}}
. IV curve.svg
Left: Plot of current vs. voltage for a SQUID. Upper and lower curves correspond to and respectively. Right: Periodic voltage response due to flux through a SQUID. The periodicity is equal to one flux quantum, .

The DC SQUID was invented in 1964 by Robert Jaklevic, John J. Lambe, James Mercereau, and Arnold Silver of Ford Research Labs [3] after Brian Josephson postulated the Josephson effect in 1962, and the first Josephson junction was made by John Rowell and Philip Anderson at Bell Labs in 1963. [4] It has two Josephson junctions in parallel in a superconducting loop. It is based on the DC Josephson effect. In the absence of any external magnetic field, the input current splits into the two branches equally. If a small external magnetic field is applied to the superconducting loop, a screening current, , begins to circulate the loop that generates the magnetic field canceling the applied external flux, and creates an additional Josephson phase which is proportional to this external magnetic flux. [5] The induced current is in the same direction as in one of the branches of the superconducting loop, and is opposite to in the other branch; the total current becomes in one branch and in the other. As soon as the current in either branch exceeds the critical current, , of the Josephson junction, a voltage appears across the junction.

Now suppose the external flux is further increased until it exceeds , half the magnetic flux quantum. Since the flux enclosed by the superconducting loop must be an integer number of flux quanta, instead of screening the flux the SQUID now energetically prefers to increase it to . The current now flows in the opposite direction, opposing the difference between the admitted flux and the external field of just over . The current decreases as the external field is increased, is zero when the flux is exactly , and again reverses direction as the external field is further increased. Thus, the current changes direction periodically, every time the flux increases by additional half-integer multiple of , with a change at maximum amperage every half-plus-integer multiple of and at zero amps every integer multiple.

If the input current is more than , then the SQUID always operates in the resistive mode. The voltage, in this case, is thus a function of the applied magnetic field and the period equal to . Since the current-voltage characteristic of the DC SQUID is hysteretic, a shunt resistance, is connected across the junction to eliminate the hysteresis (in the case of copper oxide based high-temperature superconductors the junction's own intrinsic resistance is usually sufficient). The screening current is the applied flux divided by the self-inductance of the ring. Thus can be estimated as the function of (flux to voltage converter) [6] [7] as follows:

, where is the self inductance of the superconducting ring

The discussion in this section assumed perfect flux quantization in the loop. However, this is only true for big loops with a large self-inductance. According to the relations, given above, this implies also small current and voltage variations. In practice the self-inductance of the loop is not so large. The general case can be evaluated by introducing a parameter

where is the critical current of the SQUID. Usually is of order one. [8]

RF SQUID

A prototype SQUID Squid prototype2.jpg
A prototype SQUID

The RF SQUID was invented in 1965 by Robert Jaklevic, John J. Lambe, Arnold Silver, and James Edward Zimmerman at Ford. [7] It is based on the AC Josephson effect and uses only one Josephson junction. It is less sensitive compared to DC SQUID but is cheaper and easier to manufacture in smaller quantities. Most fundamental measurements in biomagnetism, even of extremely small signals, have been made using RF SQUIDS. [9] [10] The RF SQUID is inductively coupled to a resonant tank circuit. [11] Depending on the external magnetic field, as the SQUID operates in the resistive mode, the effective inductance of the tank circuit changes, thus changing the resonant frequency of the tank circuit. These frequency measurements can be easily taken, and thus the losses which appear as the voltage across the load resistor in the circuit are a periodic function of the applied magnetic flux with a period of . For a precise mathematical description refer to the original paper by Erné et al. [6] [12]

Materials used

The traditional superconducting materials for SQUIDs are pure niobium or a lead alloy with 10% gold or indium, as pure lead is unstable when its temperature is repeatedly changed. To maintain superconductivity, the entire device needs to operate within a few degrees of absolute zero, cooled with liquid helium. [13]

High-temperature SQUID sensors were developed in the late 1980s. [14] They are made of high-temperature superconductors, particularly YBCO, and are cooled by liquid nitrogen which is cheaper and more easily handled than liquid helium. They are less sensitive than conventional low temperature SQUIDs but good enough for many applications. [15]

In 2006, A proof of concept was shown for CNT-SQUID sensors built with an aluminium loop and a single walled carbon nanotube Josephson junction. [16] The sensors are a few 100 nm in size and operate at 1K or below. Such sensors allow to count spins. [17]

In 2022 a SQUID was constructed on magic angle twisted bilayer graphene (MATBG) [18] [19]

Uses

The inner workings of an early SQUID, circa 1990 Squid prototype.jpg
The inner workings of an early SQUID, circa 1990

The extreme sensitivity of SQUIDs makes them ideal for studies in biology. Magnetoencephalography (MEG), for example, uses measurements from an array of SQUIDs to make inferences about neural activity inside brains. Because SQUIDs can operate at acquisition rates much higher than the highest temporal frequency of interest in the signals emitted by the brain (kHz), MEG achieves good temporal resolution. Another area where SQUIDs are used is magnetogastrography, which is concerned with recording the weak magnetic fields of the stomach. A novel application of SQUIDs is the magnetic marker monitoring method, which is used to trace the path of orally applied drugs. In the clinical environment SQUIDs are used in cardiology for magnetic field imaging (MFI), which detects the magnetic field of the heart for diagnosis and risk stratification.

Probably the most common commercial use of SQUIDs is in magnetic property measurement systems (MPMS). These are turn-key systems, made by several manufacturers, that measure the magnetic properties of a material sample. This is typically done over a temperature range from that of 300 mK to roughly 400 K. [20] With the decreasing size of SQUID sensors since the last decade, such sensor can equip the tip of an AFM probe. Such device allows simultaneous measurement of roughness of the surface of a sample and the local magnetic flux. [21]

For example, SQUIDs are being used as detectors to perform magnetic resonance imaging (MRI). While high-field MRI uses precession fields of one to several teslas, SQUID-detected MRI uses measurement fields that lie in the microtesla range. In a conventional MRI system, the signal scales as the square of the measurement frequency (and hence precession field): one power of frequency comes from the thermal polarization of the spins at ambient temperature, while the second power of field comes from the fact that the induced voltage in the pickup coil is proportional to the frequency of the precessing magnetization. In the case of untuned SQUID detection of prepolarized spins, however, the NMR signal strength is independent of precession field, allowing MRI signal detection in extremely weak fields, on the order of Earth's magnetic field. SQUID-detected MRI has advantages over high-field MRI systems, such as the low cost required to build such a system, and its compactness. The principle has been demonstrated by imaging human extremities, and its future application may include tumor screening. [22]

Another application is the scanning SQUID microscope, which uses a SQUID immersed in liquid helium as the probe. The use of SQUIDs in oil prospecting, mineral exploration, [23] earthquake prediction and geothermal energy surveying is becoming more widespread as superconductor technology develops; they are also used as precision movement sensors in a variety of scientific applications, such as the detection of gravitational waves. [24] A SQUID is the sensor in each of the four gyroscopes employed on Gravity Probe B in order to test the limits of the theory of general relativity. [1]

A modified RF SQUID was used to observe the dynamical Casimir effect for the first time. [25] [26]

SQUIDs constructed from super-cooled niobium wire loops are used as the basis for D-Wave Systems 2000Q quantum computer. [27]

Transition-edge sensors

One of the largest uses of SQUIDs is to read out superconducting Transition-edge sensors. Hundreds of thousands of multiplexed SQUIDs coupled to transition-edge sensors are presently being deployed to study the Cosmic microwave background, for X-ray astronomy, to search for dark matter made up of Weakly interacting massive particles, and for spectroscopy at Synchrotron light sources.

Cold dark matter

Advanced SQUIDS called near quantum-limited SQUID amplifiers form the basis of the Axion Dark Matter Experiment (ADMX) at the University of Washington. Axions are a prime candidate for cold dark matter. [28]

Proposed uses

A potential military application exists for use in anti-submarine warfare as a magnetic anomaly detector (MAD) fitted to maritime patrol aircraft. [29]

SQUIDs are used in superparamagnetic relaxometry (SPMR), a technology that utilizes the high magnetic field sensitivity of SQUID sensors and the superparamagnetic properties of magnetite nanoparticles. [30] [31] These nanoparticles are paramagnetic; they have no magnetic moment until exposed to an external field where they become ferromagnetic. After removal of the magnetizing field, the nanoparticles decay from a ferromagnetic state to a paramagnetic state, with a time constant that depends upon the particle size and whether they are bound to an external surface. Measurement of the decaying magnetic field by SQUID sensors is used to detect and localize the nanoparticles. Applications for SPMR may include cancer detection. [32]

See also

Notes

  1. 1 2 Ran, Shannon K’doah (2004). Gravity Probe B: Exploring Einstein's Universe with Gyroscopes (PDF). NASA. p. 26. Archived (PDF) from the original on 16 May 2008.
  2. D. Drung; C. Assmann; J. Beyer; A. Kirste; M. Peters; F. Ruede & Th. Schurig (2007). "Highly sensitive and easy-to-use SQUID sensors" (PDF). IEEE Transactions on Applied Superconductivity. 17 (2): 699–704. Bibcode:2007ITAS...17..699D. doi:10.1109/TASC.2007.897403. S2CID   19682964. Archived from the original (PDF) on 19 July 2011.
  3. R. C. Jaklevic; J. Lambe; A. H. Silver & J. E. Mercereau (1964). "Quantum Interference Effects in Josephson Tunneling". Physical Review Letters. 12 (7): 159–160. Bibcode:1964PhRvL..12..159J. doi:10.1103/PhysRevLett.12.159.
  4. Anderson, P.; Rowell, J. (1963). "Probable Observation of the Josephson Superconducting Tunneling Effect". Physical Review Letters. 10 (6): 230–232. Bibcode:1963PhRvL..10..230A. doi:10.1103/PhysRevLett.10.230.
  5. "The Feynman Lectures on Physics Vol. III Ch. 21: The Schrödinger Equation in a Classical Context: A Seminar on Superconductivity, Section 21–9: The Josephson junction". feynmanlectures.caltech.edu. Retrieved 8 January 2020.
  6. 1 2 E. du Trémolet de Lacheisserie, D. Gignoux, and M. Schlenker (editors) (2005). Magnetism: Materials and Applications. Vol. 2. Springer.{{cite book}}: |author= has generic name (help)CS1 maint: multiple names: authors list (link)
  7. 1 2 J. Clarke and A. I. Braginski (Eds.) (2004). The SQUID handbook. Vol. 1. Wiley-Vch.
  8. A.TH.A.M. de Waele & R. de Bruyn Ouboter (1969). "Quantum-interference phenomena in point contacts between two superconductors". Physica. 41 (2): 225–254. Bibcode:1969Phy....41..225D. doi:10.1016/0031-8914(69)90116-5.
  9. Romani, G. L.; Williamson, S. J.; Kaufman, L. (1982). "Biomagnetic instrumentation". Review of Scientific Instruments. 53 (12): 1815–1845. Bibcode:1982RScI...53.1815R. doi:10.1063/1.1136907. PMID   6760371.
  10. Sternickel, K.; Braginski, A. I. (2006). "Biomagnetism using SQUIDs: Status and perspectives". Superconductor Science and Technology. 19 (3): S160. Bibcode:2006SuScT..19S.160S. doi:10.1088/0953-2048/19/3/024. S2CID   122140082.
  11. Nisenoff, M.; Wolf, S. (1 September 1975). "Observation of a $cos\ensuremath{\varphi}$ term in the current-phase relation for "Dayem"-type weak link contained in an rf-biased superconducting quantum interference device". Physical Review B. 12 (5): 1712–1714. doi:10.1103/PhysRevB.12.1712.
  12. S.N. Erné; H.-D. Hahlbohm; H. Lübbig (1976). "Theory of the RF biased Superconducting Quantum Interference Device for the non-hysteretic regime". J. Appl. Phys. 47 (12): 5440–5442. Bibcode:1976JAP....47.5440E. doi: 10.1063/1.322574 .
  13. Clarke, John (August 1994). "SQUIDs". Scientific American. 271 (2): 46–53. Bibcode:1994SciAm.271b..46C. doi:10.1038/scientificamerican0894-46. JSTOR   24942801 . Retrieved 18 August 2022.
  14. M.S. Colclough, C.E. Gough et al, Radiofrequency SQUID operation usinga ceramic high temperature superconductor, Nature 328, 47 (1987)
  15. LP Lee et al., Monolithic 77K DC SQUID magnetometer, Applied Physics Letters 59, 3051 (1991)
  16. Cleuziou, J.-P.; Wernsdorfer, W. (2006). "Carbon nanotube superconducting quantum interference device". Nature Nanotechnology. 1 (October): 53–59. Bibcode:2006NatNa...1...53C. doi:10.1038/nnano.2006.54. PMID   18654142. S2CID   1942814.
  17. Aprili, Marco (2006). "The nanoSQUID makes its debut". Nature Nanotechnology. 1 (October): 15–16. Bibcode:2006NatNa...1...15A. doi:10.1038/nnano.2006.78. PMID   18654132. S2CID   205441987.
  18. Portolés, Elías; Iwakiri, Shuichi; Zheng, Giulia; Rickhaus, Peter; Taniguchi, Takashi; Watanabe, Kenji; Ihn, Thomas; Ensslin, Klaus; de Vries, Folkert K. (24 October 2022). "A tunable monolithic SQUID in twisted bilayer graphene". Nature Nanotechnology. 17 (11): 1159–1164. arXiv: 2201.13276 . Bibcode:2022NatNa..17.1159P. doi:10.1038/s41565-022-01222-0. ISSN   1748-3395. PMID   36280761. S2CID   246430218.
  19. "A new quantum component made from graphene". ethz.ch. 3 November 2022. Retrieved 15 November 2022.
  20. Kleiner, R.; Koelle, D.; Ludwig, F.; Clarke, J. (2004). "Superconducting quantum interference devices: State of the art and applications". Proceedings of the IEEE. 92 (10): 1534–1548. doi:10.1109/JPROC.2004.833655. S2CID   20573644.
  21. "Microscopie à microsquid - Institut NÉEL". neel.cnrs.fr.
  22. Clarke, J.; Lee, A.T.; Mück, M.; Richards, P.L. "Chapter 8.3". Nuclear Magnetic and Quadrupole Resonance and Magnetic Resonance Imaging. pp. 56–81. in Clarke & Braginski 2006
  23. P. Schmidt; D. Clark; K. Leslie; M. Bick; D. Tilbrook & C. Foley (2004). "GETMAG—A SQUID magnetic tensor gradiometer for mineral and oil exploration". Exploration Geophysics. 35 (4): 297–305. doi:10.1071/eg04297. S2CID   14994533.
  24. Paik, Ho J. "Chapter 15.2". "Superconducting Transducer for Gravitational-Wave Detectors" in [volume 2 of] "The SQUID Handbook: Applications of SQUIDs and SQUID Systems". pp. 548–554. in Clarke & Braginski 2006
  25. "First Observation of the Dynamical Casimir Effect". Technology Review.
  26. Wilson, C. M. (2011). "Observation of the Dynamical Casimir Effect in a Superconducting Circuit". Nature. 479 (7373): 376–379. arXiv: 1105.4714 . Bibcode:2011Natur.479..376W. doi:10.1038/nature10561. PMID   22094697. S2CID   219735.
  27. "Not Magic Quantum" (PDF). Lanl.gov. July 2016. Archived (PDF) from the original on 29 July 2016. Retrieved 26 October 2021.
  28. A Squid-Based Microwave Cavity Search For Axions By ADMX; SJ Sztalos, G Carlos, C Hagman, D Kinion, K van Bibber, M Hotz, L Rosenberg, G Rybka, J Hoskins, J Hwang, P Sikivie, DB Tanner, R Bradley, J Clarke; Phys.Rev.Lett. 104:041301; 2010
  29. Ouellette, Jennifer. "SQUID Sensors Penetrate New Markets" (PDF). The Industrial Physicist. p. 22. Archived from the original (PDF) on 18 May 2008.
  30. Flynn, E R; Bryant, H C (2005). "A biomagnetic system for in vivo cancer imaging". Physics in Medicine and Biology. 50 (6): 1273–1293. Bibcode:2005PMB....50.1273F. doi:10.1088/0031-9155/50/6/016. PMC   2041897 . PMID   15798322.
  31. De Haro, Leyma P.; Karaulanov, Todor; Vreeland, Erika C.; Anderson, Bill; Hathaway, Helen J.; Huber, Dale L.; Matlashov, Andrei N.; Nettles, Christopher P.; Price, Andrew D. (1 October 2015). "Magnetic relaxometry as applied to sensitive cancer detection and localization". Biomedical Engineering / Biomedizinische Technik. 60 (5): 445–455. doi: 10.1515/bmt-2015-0053 . ISSN   1862-278X. OSTI   1227725. PMID   26035107. S2CID   13867059.
  32. Hathaway, Helen J.; Butler, Kimberly S.; Adolphi, Natalie L.; Lovato, Debbie M.; Belfon, Robert; Fegan, Danielle; Monson, Todd C.; Trujillo, Jason E.; Tessier, Trace E. (1 January 2011). "Detection of breast cancer cells using targeted magnetic nanoparticles and ultra-sensitive magnetic field sensors". Breast Cancer Research. 13 (5): R108. doi: 10.1186/bcr3050 . ISSN   1465-542X. PMC   3262221 . PMID   22035507.

Related Research Articles

<span class="mw-page-title-main">Aharonov–Bohm effect</span> Electromagnetic quantum-mechanical effect in regions of zero magnetic and electric field

The Aharonov–Bohm effect, sometimes called the Ehrenberg–Siday–Aharonov–Bohm effect, is a quantum-mechanical phenomenon in which an electrically charged particle is affected by an electromagnetic potential, despite being confined to a region in which both the magnetic field and electric field are zero. The underlying mechanism is the coupling of the electromagnetic potential with the complex phase of a charged particle's wave function, and the Aharonov–Bohm effect is accordingly illustrated by interference experiments.

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 Φ can be as well. However, if one deals with the superconducting loop or a hole in a bulk superconductor, the magnetic flux threading such a hole/loop is quantized. The (superconducting) magnetic flux quantumΦ0 = h/(2e)2.067833848...×10−15 Wb is a combination of fundamental physical constants: the Planck constant h and the electron charge e. Its value is, therefore, the same for any superconductor. The phenomenon of flux quantization was discovered experimentally by B. S. Deaver and W. M. Fairbank and, independently, by R. Doll and M. Näbauer, in 1961. The quantization of magnetic flux is closely related to the Little–Parks effect, but was predicted earlier by Fritz London in 1948 using a phenomenological model.

<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.

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.

Superconducting quantum computing is a branch of solid state quantum computing that implements superconducting electronic circuits using superconducting qubits as artificial atoms, or quantum dots. For superconducting qubits, the two logic states are the ground state and the excited state, denoted respectively. Research in superconducting quantum computing is conducted by companies such as Google, IBM, IMEC, BBN Technologies, Rigetti, and Intel. Many recently developed QPUs use superconducting architecture.

In physics, a fluxon is a quantum of electromagnetic flux. The term may have any of several related meanings.

In superconductivity, a semifluxon is a half integer vortex of supercurrent carrying the magnetic flux equal to the half of the magnetic flux quantum Φ0. Semifluxons exist in the 0-π long Josephson junctions at the boundary between 0 and π regions. This 0-π boundary creates a π discontinuity of the Josephson phase. The junction reacts to this discontinuity by creating a semifluxon. Vortex's supercurrent circulates around 0-π boundary. In addition to semifluxon, there exist also an antisemifluxon. It carries the flux −Φ0/2 and its supercurrent circulates in the opposite direction.

<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">Flux qubit</span> Superconducting qubit implementation

In quantum computing, more specifically in superconducting quantum computing, flux qubits are micrometer sized loops of superconducting metal that is interrupted by a number of Josephson junctions. These devices function as quantum bits. The flux qubit was first proposed by Terry P. Orlando et al. at MIT in 1999 and fabricated shortly thereafter. During fabrication, the Josephson junction parameters are engineered so that a persistent current will flow continuously when an external magnetic flux is applied. Only an integer number of flux quanta are allowed to penetrate the superconducting ring, resulting in clockwise or counter-clockwise mesoscopic supercurrents in the loop to compensate a non-integer external flux bias. When the applied flux through the loop area is close to a half integer number of flux quanta, the two lowest energy eigenstates of the loop will be a quantum superposition of the clockwise and counter-clockwise currents. The two lowest energy eigenstates differ only by the relative quantum phase between the composing current-direction states. Higher energy eigenstates correspond to much larger (macroscopic) persistent currents, that induce an additional flux quantum to the qubit loop, thus are well separated energetically from the lowest two eigenstates. This separation, known as the "qubit non linearity" criteria, allows operations with the two lowest eigenstates only, effectively creating a two level system. Usually, the two lowest eigenstates will serve as the computational basis for the logical qubit.

<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.

A Josephson junction (JJ) is a quantum mechanical device which is made of two superconducting electrodes separated by a barrier. A π Josephson junction is a Josephson junction in which the Josephson phase φ equals π in the ground state, i.e. when no external current or magnetic field is applied.

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">Scanning SQUID microscopy</span> Method of imaging magnetic fields at microscopic scales

In condensed matter physics, scanning SQUID microscopy is a technique where a superconducting quantum interference device (SQUID) is used to image surface magnetic field strength with micrometre-scale resolution. A tiny SQUID is mounted onto a tip which is then rastered near the surface of the sample to be measured. As the SQUID is the most sensitive detector of magnetic fields available and can be constructed at submicrometre widths via lithography, the scanning SQUID microscope allows magnetic fields to be measured with unparalleled resolution and sensitivity. The first scanning SQUID microscope was built in 1992 by Black et al. Since then the technique has been used to confirm unconventional superconductivity in several high-temperature superconductors including YBCO and BSCCO compounds.

In quantum computing, and more specifically in superconducting quantum computing, the phase qubit is a superconducting device based on the superconductor–insulator–superconductor (SIS) Josephson junction, designed to operate as a quantum bit, or qubit.

Flux pumping is a method for magnetising superconductors to fields in excess of 15 teslas. The method can be applied to any type II superconductor and exploits a fundamental property of superconductors, namely their ability to support and maintain currents on the length scale of the superconductor. Conventional magnetic materials are magnetised on a molecular scale which means that superconductors can maintain a flux density orders of magnitude bigger than conventional materials. Flux pumping is especially significant when one bears in mind that all other methods of magnetising superconductors require application of a magnetic flux density at least as high as the final required field. This is not true of flux pumping.

The superconducting tunnel junction (STJ) — also known as a superconductor–insulator–superconductor tunnel junction (SIS) — is an electronic device consisting of two superconductors separated by a very thin layer of insulating material. Current passes through the junction via the process of quantum tunneling. The STJ is a type of Josephson junction, though not all the properties of the STJ are described by the Josephson effect.

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:

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 and topological order. Since 2000 there has been extensive experimental work on quantum gases, particularly Bose–Einstein condensates.

A φ Josephson junction is a particular type of the Josephson junction, which has a non-zero Josephson phase φ across it in the ground state. A π Josephson junction, which has the minimum energy corresponding to the phase of π, is a specific example of it.

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).

References