Unconventional superconductor

Last updated

Unconventional superconductors are materials that display superconductivity which does not conform to conventional BCS theory or its extensions.

Contents

History

The superconducting properties of CeCu2Si2, a type of heavy fermion material, were reported in 1979 by Frank Steglich. [1] For a long time it was believed that CeCu2Si2 was a singlet d-wave superconductor, but since the mid-2010s, this notion has been strongly contested. [2] In the early eighties, many more unconventional, heavy fermion superconductors were discovered, including UBe13, [3] UPt3 [4] and URu2Si2. [5] In each of these materials, the anisotropic nature of the pairing was implicated by the power-law dependence of the nuclear magnetic resonance (NMR) relaxation rate and specific heat capacity on temperature. The presence of nodes in the superconducting gap of UPt3 was confirmed in 1986 from the polarization dependence of the ultrasound attenuation. [6]

The first unconventional triplet superconductor, organic material (TMTSF)2PF6, was discovered by Denis Jerome, Klaus Bechgaard and coworkers in 1980. [7] Experimental works by Paul Chaikin's and Michael Naughton's groups as well as theoretical analysis of their data by Andrei Lebed have firmly confirmed unconventional nature of superconducting pairing in (TMTSF)2X (X=PF6, ClO4, etc.) organic materials. [8]

High-temperature singlet d-wave superconductivity was discovered by J.G. Bednorz and K.A. Müller in 1986, who also discovered that the lanthanum-based cuprate perovskite material LaBaCuO4 develops superconductivity at a critical temperature (Tc) of approximately 35  K (-238 degrees Celsius). This was well above the highest critical temperature known at the time (Tc = 23 K), and thus the new family of materials was called high-temperature superconductors. Bednorz and Müller received the Nobel prize in Physics for this discovery in 1987. Since then, many other high-temperature superconductors have been synthesized.

LSCO (La2−xSrxCuO4) was discovered the same year (1986). Soon after, in January 1987, yttrium barium copper oxide (YBCO) was discovered to have a Tc of 90 K, the first material to achieve superconductivity above the boiling point of liquid nitrogen (77 K). [9] This was highly significant from the point of view of the technological applications of superconductivity because liquid nitrogen is far less expensive than liquid helium, which is required to cool conventional superconductors down to their critical temperature. In 1988 bismuth strontium calcium copper oxide (BSCCO) with Tc up to 107 K, [10] and thallium barium calcium copper oxide (TBCCO) (T=thallium) with Tc of 125 K were discovered. The current record critical temperature is about Tc = 133 K (−140 °C) at standard pressure, and somewhat higher critical temperatures can be achieved at high pressure. Nevertheless, at present it is considered unlikely that cuprate perovskite materials will achieve room-temperature superconductivity.

On the other hand, other unconventional superconductors have been discovered. These include some that do not superconduct at high temperatures, such as strontium ruthenate Sr2RuO4, but that, like high-temperature superconductors, are unconventional in other ways. (For example, the origin of the attractive force leading to the formation of Cooper pairs may be different from the one postulated in BCS theory.) In addition to this, superconductors that have unusually high values of Tc but that are not cuprate perovskites have been discovered. Some of them may be extreme examples of conventional superconductors (this is suspected of magnesium diboride, MgB2, with Tc = 39 K). Others could display more unconventional features.

In 2008 a new class that does not include copper (layered oxypnictide superconductors), for example LaOFeAs, was discovered. [11] [12] [13] An oxypnictide of samarium seemed to have a Tc of about 43 K, which was higher than predicted by BCS theory. [14] Tests at up to 45  T [15] [16] suggested the upper critical field of LaFeAsO0.89F0.11 to be around 64 T. Some other iron-based superconductors do not contain oxygen.

As of 2009, the highest-temperature superconductor (at ambient pressure) is mercury barium calcium copper oxide (HgBa2Ca2Cu3Ox), at 138 K and is held by a cuprate-perovskite material, [17] possibly 164 K under high pressure. [18]

Other unconventional superconductors not based on cuprate structure have too been found. [11] Some have unusually high values of the critical temperature, Tc, and hence they are sometimes also called high-temperature superconductors.

Graphene

In 2017, scanning tunneling microscopy and spectroscopy experiments on graphene proximitized to the electron-doped (non-chiral) d-wave superconductor Pr2−xCexCuO4 (PCCO) revealed evidence for an unconventional superconducting density of states induced in graphene. [19] Publications in March 2018 provided evidence for unconventional superconducting properties of a graphene bilayer where one layer was offset by a "magic angle" of 1.1° relative to the other. [20]

Ongoing research

After more than twenty years of intensive research, the origin of high-temperature superconductivity is still not clear, being one of the major unsolved problems of theoretical condensed matter physics. But it seems that instead of electron-phonon attraction mechanisms, as in conventional superconductivity, genuine electronic mechanisms (e.g. by antiferromagnetic correlations) are occurring. In addition, instead of s-wave pairing, d-waves are substantial.

One goal of much research is room-temperature superconductivity. [21]

Despite intensive research and many promising leads, an explanation has so far eluded scientists. One reason for this is that the materials in question are generally very complex, multi-layered crystals (for example, BSCCO), making theoretical modeling difficult.

Possible mechanisms

The most controversial topic in condensed matter physics has been the mechanism for high-Tc superconductivity (HTS). There have been two representative theories on the HTS : (See also Resonating valence bond theory )

Weak-coupling theory

Firstly, it has been suggested that the HTS emerges by antiferromagnetic spin fluctuation in a doped system. [22] According to this weak-coupling theory, the pairing wave function of the HTS should have a dx2y2 symmetry. Thus, whether the symmetry of the pairing wave function is the d symmetry or not is essential to demonstrate on the mechanism of the HTS in respect of the spin fluctuation. That is, if the HTS order parameter (pairing wave function) does not have d symmetry, then a pairing mechanism related to spin fluctuation can be ruled out. The tunnel experiment (see below) seems to detect d symmetry in some HTS.

Interlayer coupling model

Secondly, there is the interlayer coupling model, according to which a layered structure consisting of BCS-type (s symmetry) superconductor can enhance the superconductivity by itself. [23] By introducing an additional tunneling interaction between each layer, this model successfully explained the anisotropic symmetry of the order parameter in the HTS as well as the emergence of the HTS.[ citation needed ]

In order to solve this unsettled problem, there have been numerous experiments such as photoelectron spectroscopy, NMR, specific heat measurement, etc. Unfortunately, the results were ambiguous, where some reports supported the d symmetry for the HTS but others supported the s symmetry.[ citation needed ] This muddy situation possibly originated from the indirect nature of the experimental evidence, as well as experimental issues such as sample quality, impurity scattering, twinning, etc.

Superexchange

Promising experimental results from various researchers in September 2022, including Weijiong Chen, J.C. Séamus Davis and H. Eisiaki revealed that superexchange of electrons is possibly the most probable reason for high-temperature superconductivity. [24] [25]

Previous studies on the symmetry of the HTS order parameter

The symmetry of the HTS order parameter has been studied in nuclear magnetic resonance measurements and, more recently, by angle-resolved photoemission and measurements of the microwave penetration depth in a HTS crystal. NMR measurements probe the local magnetic field around an atom and hence reflect the susceptibility of the material. They have been of special interest for the HTS materials because many researchers have wondered whether spin correlations might play a role in the mechanism of the HTS.

NMR measurements of the resonance frequency on YBCO indicated that electrons in the copper oxide superconductors are paired in spin-singlet states. This indication came from the behavior of the Knight shift, the frequency shift that occurs when the internal field is different from the applied field: In a normal metal, the magnetic moments of the conduction electrons in the neighborhood of the ion being probed align with the applied field and create a larger internal field. As these metals go superconducting, electrons with oppositely directed spins couple to form singlet states. In the anisotropic HTS, perhaps NMR measurements have found that the relaxation rate for copper depends on the direction of the applied static magnetic field, with the rate being higher when the static field is parallel to one of the axes in the copper oxide plane. While this observation by some group supported the d symmetry of the HTS, other groups could not observe it.

Also, by measuring the penetration depth, the symmetry of the HTS order parameter can be studied. The microwave penetration depth is determined by the superfluid density responsible for screening the external field. In the s wave BCS theory, because pairs can be thermally excited across the gap Δ, the change in superfluid density per unit change in temperature goes as exponential behavior, exp(-Δ/kBT). In that case, the penetration depth also varies exponentially with temperature T. If there are nodes in the energy gap as in the d symmetry HTS, electron pair can more easily be broken, the superfluid density should have a stronger temperature dependence, and the penetration depth is expected to increase as a power of T at low temperatures. If the symmetry is specially dx2-y2 then the penetration depth should vary linearly with T at low temperatures. This technique is increasingly being used to study superconductors and is limited in application largely by the quality of available single crystals.

Photoemission spectroscopy also could provide information on the HTS symmetry. By scattering photons off electrons in the crystal, one can sample the energy spectra of the electrons. Because the technique is sensitive to the angle of the emitted electrons one can determine the spectrum for different wave vectors on the Fermi surface. However, within the resolution of the angle-resolved photoemission spectroscopy (ARPES), researchers could not tell whether the gap goes to zero or just gets very small. Also, ARPES are sensitive only to the magnitude and not to the sign of the gap, so it could not tell if the gap goes negative at some point. This means that ARPES cannot determine whether the HTS order parameter has the d symmetry or not.

Junction experiment supporting the d-wave symmetry

There was a clever experimental design to overcome the muddy situation. An experiment based on pair tunneling and flux quantization in a three-grain ring of YBa2Cu3O7 (YBCO) was designed to test the symmetry of the order parameter in YBCO. [26] Such a ring consists of three YBCO crystals with specific orientations consistent with the d-wave pairing symmetry to give rise to a spontaneously generated half-integer quantum vortex at the tricrystal meeting point. Furthermore, the possibility that junction interfaces can be in the clean limit (no defects) or with maximum zig-zag disorder was taken into account in this tricrystal experiment. [26] A proposal of studying vortices with half magnetic flux quanta in heavy-fermion superconductors in three polycrystalline configurations was reported in 1987 by V. B. Geshkenbein, A. Larkin and A. Barone in 1987. [27]

In the first tricrystal pairing symmetry experiment, [26] the spontaneous magnetization of half flux quantum was clearly observed in YBCO, which convincingly supported the d-wave symmetry of the order parameter in YBCO. Because YBCO is orthorhombic, it might inherently have an admixture of s-wave symmetry. So, by tuning their technique further, it was found that there was an admixture of s-wave symmetry in YBCO within about 3%. [28] Also, it was demonstrated by Tsuei, Kirtley et al. that there was a pure dx2-y2 order parameter symmetry in the tetragonal Tl2Ba2CuO6. [29]

Related Research Articles

<span class="mw-page-title-main">BCS theory</span> Microscopic theory of superconductivity

In physics, theBardeen–Cooper–Schrieffer (BCS) theory is the first microscopic theory of superconductivity since Heike Kamerlingh Onnes's 1911 discovery. The theory describes superconductivity as a microscopic effect caused by a condensation of Cooper pairs. The theory is also used in nuclear physics to describe the pairing interaction between nucleons in an atomic nucleus.

<span class="mw-page-title-main">Superconductivity</span> Electrical conductivity with exactly zero resistance

Superconductivity is a set of physical properties observed in certain materials where electrical resistance vanishes and magnetic fields are expelled from the material. Any material exhibiting these properties is a superconductor. Unlike an ordinary metallic conductor, whose resistance decreases gradually as its temperature is lowered, even down to near absolute zero, a superconductor has a characteristic critical temperature below which the resistance drops abruptly to zero. An electric current through a loop of superconducting wire can persist indefinitely with no power source.

<span class="mw-page-title-main">High-temperature superconductivity</span> Superconductive behavior at temperatures much higher than absolute zero

High-temperature superconductors are defined as materials with critical temperature above 77 K, the boiling point of liquid nitrogen. They are only "high-temperature" relative to previously known superconductors, which function at even colder temperatures, close to absolute zero. The "high temperatures" are still far below ambient, and therefore require cooling. The first break through of high-temperature superconductor was discovered in 1986 by IBM researchers Georg Bednorz and K. Alex Müller. Although the critical temperature is around 35.1 K, this new type of superconductor was readily modified by Ching-Wu Chu to make the first high-temperature superconductor with critical temperature 93 K. Bednorz and Müller were awarded the Nobel Prize in Physics in 1987 "for their important break-through in the discovery of superconductivity in ceramic materials". Most high-Tc materials are type-II superconductors.

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.

<span class="mw-page-title-main">Bismuth strontium calcium copper oxide</span> Family of high-temperature superconductors

Bismuth strontium calcium copper oxide (BSCCO, pronounced bisko), is a type of cuprate superconductor having the generalized chemical formula Bi2Sr2Can−1CunO2n+4+x, with n = 2 being the most commonly studied compound (though n = 1 and n = 3 have also received significant attention). Discovered as a general class in 1988, BSCCO was the first high-temperature superconductor which did not contain a rare-earth element.

<span class="mw-page-title-main">Pseudogap</span> State at which a Fermi surface has a partial energy gap in condensed matter physics

In condensed matter physics, a pseudogap describes a state where the Fermi surface of a material possesses a partial energy gap, for example, a band structure state where the Fermi surface is gapped only at certain points.

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.

A Josephson junction 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.

A charge density wave (CDW) is an ordered quantum fluid of electrons in a linear chain compound or layered crystal. The electrons within a CDW form a standing wave pattern and sometimes collectively carry an electric current. The electrons in such a CDW, like those in a superconductor, can flow through a linear chain compound en masse, in a highly correlated fashion. Unlike a superconductor, however, the electric CDW current often flows in a jerky fashion, much like water dripping from a faucet due to its electrostatic properties. In a CDW, the combined effects of pinning and electrostatic interactions likely play critical roles in the CDW current's jerky behavior, as discussed in sections 4 & 5 below.

In solid-state physics, heavy fermion materials are a specific type of intermetallic compound, containing elements with 4f or 5f electrons in unfilled electron bands. Electrons are one type of fermion, and when they are found in such materials, they are sometimes referred to as heavy electrons. Heavy fermion materials have a low-temperature specific heat whose linear term is up to 1000 times larger than the value expected from the free electron model. The properties of the heavy fermion compounds often derive from the partly filled f-orbitals of rare-earth or actinide ions, which behave like localized magnetic moments. The name "heavy fermion" comes from the fact that the fermion behaves as if it has an effective mass greater than its rest mass. In the case of electrons, below a characteristic temperature (typically 10 K), the conduction electrons in these metallic compounds behave as if they had an effective mass up to 1000 times the free particle mass. This large effective mass is also reflected in a large contribution to the resistivity from electron-electron scattering via the Kadowaki–Woods ratio. Heavy fermion behavior has been found in a broad variety of states including metallic, superconducting, insulating and magnetic states. Characteristic examples are CeCu6, CeAl3, CeCu2Si2, YbAl3, UBe13 and UPt3.

Ferromagnetic superconductors are materials that display intrinsic coexistence of ferromagnetism and superconductivity. They include UGe2, URhGe, and UCoGe. Evidence of ferromagnetic superconductivity was also reported for ZrZn2 in 2001, but later reports question these findings. These materials exhibit superconductivity in proximity to a magnetic quantum critical point.

Superconductors can be classified in accordance with several criteria that depend on physical properties, current understanding, and the expense of cooling them or their material.

<span class="mw-page-title-main">122 iron arsenide</span>

The 122 iron arsenide unconventional superconductors are part of a new class of iron-based superconductors. They form in the tetragonal I4/mmm, ThCr2Si2 type, crystal structure. The shorthand name "122" comes from their stoichiometry; the 122s have the chemical formula AEFe2Pn2, where AE stands for alkaline earth metal (Ca, Ba Sr or Eu) and Pn is pnictide (As, P, etc.). These materials become superconducting under pressure and also upon doping. The maximum superconducting transition temperature found to date is 38 K in the Ba0.6K0.4Fe2As2. The microscopic description of superconductivity in the 122s is yet unclear.

Heavy fermion superconductors are a type of unconventional superconductor.

The Fulde–Ferrell–Larkin–Ovchinnikov (FFLO) phase can arise in a superconductor in large magnetic field. Among its characteristics are Cooper pairs with nonzero total momentum and a spatially non-uniform order parameter, leading to normal conducting areas in the superconductor.

<span class="mw-page-title-main">Distrontium ruthenate</span> Chemical compound

Distrontium ruthenate, also known as strontium ruthenate, is an oxide of strontium and ruthenium with the chemical formula Sr2RuO4. It was the first reported perovskite superconductor that did not contain copper. Strontium ruthenate is structurally very similar to the high-temperature cuprate superconductors, and in particular, is almost identical to the lanthanum doped superconductor (La, Sr)2CuO4. However, the transition temperature for the superconducting phase transition is 0.93 K (about 1.5 K for the best sample), which is much lower than the corresponding value for cuprates.

CeCoIn5 ("Cerium-Cobalt-Indium 5") is a heavy-fermion superconductor with a layered crystal structure, with somewhat two-dimensional electronic transport properties. The critical temperature of 2.3 K is the highest among all of the Ce-based heavy-fermion superconductors.

<span class="mw-page-title-main">Rare-earth barium copper oxide</span> Chemical compounds known for exhibiting high temperature superconductivity

Rare-earth barium copper oxide (ReBCO) is a family of chemical compounds known for exhibiting high-temperature superconductivity (HTS). ReBCO superconductors have the potential to sustain stronger magnetic fields than other superconductor materials. Due to their high superconducting critical temperature and critical magnetic field, this class of materials are proposed for future use in technical applications where conventional low-temperature superconductors do not suffice. This includes magnetic confinement fusion reactors such as the ARC reactor, allowing a more compact and potentially more economical construction, and superconducting magnets to use in future particle accelerators to come after the Large Hadron Collider at CERN, which utilizes low-temperature superconductors.

Dale J. Van Harlingen is an American condensed matter physicist.

In physics, the Matthias rules refers to a historical set of empirical guidelines on how to find superconductors. These rules were authored Bernd T. Matthias who discovered hundreds of superconductors using these principles in the 1950s and 1960s. Deviations from these rules have been found since the end of the 1970s with the discovery of unconventional superconductors.

References

  1. Steglich, F.; Aarts, J.; Bredl, C.D.; Lieke, W.; Meschede, D.; Franz, W.; Schäfer, H. (1979). "Superconductivity in the Presence of Strong Pauli Paramagnetism: CeCu2Si2". Physical Review Letters. 43 (25): 1892–1896. Bibcode:1979PhRvL..43.1892S. doi:10.1103/PhysRevLett.43.1892. hdl: 1887/81461 . S2CID   123497750.
  2. Kittaka, Shunichiro; Aoki, Yuya; Shimura, Yasuyuki; Sakakibara, Toshiro; Seiro, Silvia; Geibel, Christoph; Steglich, Frank; Ikeda, Hiroaki; Machida, Kazushige (February 12, 2014). "Multiband Superconductivity with Unexpected Deficiency of Nodal Quasiparticles in CeCu2Si2". Physical Review Letters. 112 (6): 067002. arXiv: 1307.3499 . Bibcode:2014PhRvL.112f7002K. doi:10.1103/PhysRevLett.112.067002. PMID   24580704. S2CID   13367098.
  3. Ott, H. R.; Rudigier, H.; Fisk, Z.; Smith, J. (1983). "UBe_{13}: An Unconventional Actinide Superconductor". Physical Review Letters. 50 (20): 1595–1598. Bibcode:1983PhRvL..50.1595O. doi:10.1103/PhysRevLett.50.1595.
  4. Stewart, G. R.; Fisk, Z.; Willis, J. O.; Smith, J. L. (1984). "Possibility of Coexistence of Bulk Superconductivity and Spin Fluctuations in UPt3". Physical Review Letters. 52 (8): 679–682. Bibcode:1984PhRvL..52..679S. doi:10.1103/PhysRevLett.52.679. S2CID   73591098.
  5. Palstra, T. T. M.; Menovsky, A. A.; Berg, J. van den; Dirkmaat, A. J.; Kes, P. H.; Nieuwenhuys, G. J.; Mydosh, J. A. (1985). "Superconducting and Magnetic Transitions in the Heavy-Fermion System URu_{2}Si_{2}". Physical Review Letters. 55 (24): 2727–2730. Bibcode:1985PhRvL..55.2727P. doi:10.1103/PhysRevLett.55.2727. PMID   10032222.
  6. Shivaram, B. S.; Jeong, Y. H.; Rosenbaum, T.F.; Hinks, D. (1986). "Anisotropy of Transverse Sound in the Heavy-Fermion Superconductor UPt3" (PDF). Physical Review Letters. 56 (10): 1078–1081. Bibcode:1986PhRvL..56.1078S. doi:10.1103/PhysRevLett.56.1078. PMID   10032562.
  7. Jérome, D.; Mazaud, A.; Ribault, M.; Bechgaard, K. (1980). "Superconductivity in a synthetic organic conductor (TMTSF)2PF 6" (PDF). Journal de Physique Lettres. 41 (4): 95. doi:10.1051/jphyslet:0198000410409500.
  8. Bechgaard, Klaus; Carneiro, Claus S.; Olsen, Malte; Rasmussen, Finn; Jacobsen, Claus (1981). "Zero-Pressure Organic Superconductor: Di-(Tetramethyltetraselenafulvalenium)-Perchlorate [(TMTSF)2ClO4]" (PDF). Physical Review Letters. 46 (13): 852. Bibcode:1981PhRvL..46..852B. doi:10.1103/PhysRevLett.46.852.
  9. K. M. Wu; et al. (1987). "Superconductivity at 93 K in a new mixed-phase Yb-Ba-Cu-O compound system at ambient pressure". Phys. Rev. Lett. 58 (9): 908–910. Bibcode:1987PhRvL..58..908W. doi: 10.1103/PhysRevLett.58.908 . PMID   10035069.
  10. H. Maeda; Y. Tanaka; M. Fukutumi & T. Asano (1988). "A New High-Tc Oxide Superconductor without a Rare Earth Element". Jpn. J. Appl. Phys. 27 (2): L209–L210. Bibcode:1988JaJAP..27L.209M. doi: 10.1143/JJAP.27.L209 .
  11. 1 2 Hiroki Takahashi; Kazumi Igawa; Kazunobu Arii; Yoichi Kamihara; Masahiro Hirano; Hideo Hosono (2008). "Superconductivity at 43K in an iron-based layered compound LaO1−xFxFeAs". Nature. 453 (7193): 376–378. Bibcode:2008Natur.453..376T. doi:10.1038/nature06972. PMID   18432191. S2CID   498756.
  12. Choi, Charles Q. (June 1, 2008). "A New Iron Age: New class of superconductor may help pin down mysterious physics". Scientific American. Retrieved October 29, 2009.
  13. National Institute of Standards and Technology (June 1, 2008). "New High-Temperature Superconductors Are Iron-based With Unusual Magnetic Properties". ScienceDaily.
  14. Chen, X. H.; Wu, T.; Wu, G.; Liu, R. H.; Chen, H.; Fang, D. F. (2008). "Superconductivity at 43 K in SmFeAsO1−xFX". Nature. 453 (7196): 761–762. arXiv: 0803.3603 . Bibcode:2008Natur.453..761C. doi:10.1038/nature07045. PMID   18500328. S2CID   205213713.
  15. High-temp superconductors pave way for 'supermagnets' [ permanent dead link ]
  16. Hunte, F.; Jaroszynski, J.; Gurevich, A.; Larbalestier, D. C.; Jin, R.; Sefat, A. S.; McGuire, M. A.; Sales, B. C.; et al. (2008). "Very High Field Two-band Superconductivity in LaFeAsO0.89F0.11 at very high magnetic fields". Nature. 453 (7197): 903–5. arXiv: 0804.0485 . Bibcode:2008Natur.453..903H. doi:10.1038/nature07058. PMID   18509332. S2CID   115211939.
  17. P. Dai; B. C. Chakoumakos; G. F. Sun; K. W. Wong; Y. Xin; D. F. Lu (1995). "Synthesis and neutron powder diffraction study of the superconductor HgBa2Ca2Cu3O8+δ by Tl substitution". Physica C. 243 (3–4): 201–206. Bibcode:1995PhyC..243..201D. doi:10.1016/0921-4534(94)02461-8.
  18. L. Gao; Y. Y. Xue; F. Chen; Q. Xiong; R. L. Meng; D. Ramirez; C. W. Chu; J. H. Eggert & H. K. Mao (1994). "Superconductivity up to 164 K in HgBa2Cam-1CumO2m+2+δ (m=1, 2, and 3) under quasihydrostatic pressures". Phys. Rev. B. 50 (6): 4260–4263. Bibcode:1994PhRvB..50.4260G. doi:10.1103/PhysRevB.50.4260. PMID   9976724.
  19. Di Bernardo, A.; Millo, O.; Barbone, M.; Alpern, H.; Kalcheim, Y.; Sassi, U.; Ott, A. K.; Fazio, D. De; Yoon, D. (January 19, 2017). "p-wave triggered superconductivity in single-layer graphene on an electron-doped oxide superconductor". Nature Communications. 8: 14024. arXiv: 1702.01572 . Bibcode:2017NatCo...814024D. doi:10.1038/ncomms14024. ISSN   2041-1723. PMC   5253682 . PMID   28102222.
  20. Gibney, Elizabeth (March 5, 2018). "Surprise graphene discovery could unlock secrets of superconductivity". News. Nature. 555 (7695): 151–2. Bibcode:2018Natur.555..151G. doi: 10.1038/d41586-018-02773-w . PMID   29517044. Physicists now report that arranging two layers of atom-thick graphene so that the pattern of their carbon atoms is offset by an angle of 1.1º makes the material a superconductor.
  21. A. Mourachkine (2004). Room-Temperature Superconductivity. Cambridge International Science Publishing. arXiv: cond-mat/0606187 . Bibcode:2006cond.mat..6187M. ISBN   1-904602-27-4.
  22. P. Monthoux; Balatsky, A.; Pines, D.; et al. (1992). "Weak-coupling theory of high-temperature superconductivity in the antiferromagnetically correlated copper oxides". Phys. Rev. B. 46 (22): 14803–14817. Bibcode:1992PhRvB..4614803M. doi:10.1103/PhysRevB.46.14803. PMID   10003579.
  23. S. Chakravarty; Sudbo, A.; Anderson, P. W.; Strong, S.; et al. (1993). "Interlayer Tunneling and Gap Anisotropy in High-Temperature Superconductors". Science. 261 (5119): 337–40. Bibcode:1993Sci...261..337C. doi:10.1126/science.261.5119.337. PMID   17836845. S2CID   41404478.
  24. O’Mahony, Shane M.; Ren, Wangping; Chen, Weijiong; Chong, Yi Xue; Liu, Xiaolong; Eisaki, H.; Uchida, S.; Hamidian, M. H.; Davis, J. C. Séamus (September 13, 2022). "On the electron pairing mechanism of copper-oxide high temperature superconductivity". Proceedings of the National Academy of Sciences. 119 (37): e2207449119. arXiv: 2108.03655 . Bibcode:2022PNAS..11907449O. doi:10.1073/pnas.2207449119. ISSN   0027-8424. PMC   9477408 . PMID   36067325.
  25. Wood, Charlie. "The High-Temperature Superconductivity Mystery Is Finally Solved". Wired. ISSN   1059-1028 . Retrieved December 26, 2022.
  26. 1 2 3 C. C. Tsuei; Kirtley, J. R.; Chi, C. C.; Yu-Jahnes, Lock See; Gupta, A.; Shaw, T.; Sun, J. Z.; Ketchen, M. B.; et al. (1994). "Pairing Symmetry and Flux Quantization in a Tricrystal Ring of Superconductin YBa2Cu3O7- delta". Phys. Rev. Lett. 73 (4): 593–596. Bibcode:1994PhRvL..73..593T. doi:10.1103/PHYSREVLETT.73.593. PMID   10057486.
  27. V. B. Geshkenbein; Larkin, A.; Barone, A.; et al. (1987). "Vortices with half magnetic flux quanta in heavy-fermion superconductors". Phys. Rev. B. 36 (1): 235–238. Bibcode:1987PhRvB..36..235G. doi:10.1103/PhysRevB.36.235. PMID   9942041.
  28. J. R. Kirtley; Tsuei, C. C.; Ariando, A.; Verwijs, C. J. M.; Harkema, S.; Hilgenkamp, H.; et al. (2006). "Angle-resolved phase-sensitive determination of the in-plane gap symmetry in YBa2Cu3O7-delta". Nat. Phys. 2 (3): 190. Bibcode:2006NatPh...2..190K. doi:10.1038/nphys215. S2CID   118447968.
  29. C. C. Tsuei; Kirtley, J. R.; Ren, Z. F.; Wang, J. H.; Raffy, H.; Li, Z. Z.; et al. (1997). "Pure dx2 – y2 order-parameter symmetry in the tetragonal superconductor TI2Ba2CuO6+delta". Nature. 387 (6632): 481. Bibcode:1997Natur.387..481T. doi:10.1038/387481a0. S2CID   4314494.