Unconventional superconductor

Unconventional superconductors are materials that display superconductivity which does not conform to either the conventional BCS theory or the Nikolay Bogolyubov's theory or its extensions.

The first unconventional singlet d-wave superconductor, CeCu2Si2, a type of heavy fermion metal, was discovered in 1978 by Frank Steglich.[1] In the early eighties, many more unconventional, heavy fermion superconductors were discovered, including UBe13,[2] UPt3 [3] and URu2Si2.[4] In each of these materials, the anisotropic nature of the pairing is 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.[5]

The first unconventional triplet superconductor, organic material (TMTSF)2PF6, was discovered by Denis Jerome and Klaus Bechgaard in 1979.[6] Recent 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.[7]

High-temperature singlet d-wave superconductivity was discovered by J.G. Bednorz and K.A. Müller in 1986, who discovered that the lanthanum-based cuprate perovskite material LaBaCuO4 develops superconductivity at a critical temperature (Tc) of approximately 35 K (-238 degrees Celsius). This is well above the highest critical temperature known at the time (Tc = 23 K) and thus the new family of materials were 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. As early as 1987, superconductivity above 77 K, the boiling point of nitrogen, was achieved. This is 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. 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, in recent years other unconventional superconductors have been discovered. These include some that do not superconduct at high temperatures, such as the strontium ruthenate oxide compounds, but that, like the 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 display more unconventional features.

In 2008 a new class (layered oxypnictide superconductors), for example LaOFeAs, were discovered that do not include copper.[8][9][10] An oxypnictide of samarium seems to have a Tc of about 43 K which is higher than predicted by BCS theory.[11] Tests at up to 45 teslas[12][13] suggest the upper critical field of LaFeAsO0.89F0.11 may be around 64 teslas. Some other iron-based superconductors do not contain oxygen.

History and progress

After more than twenty years of intensive research the origin of high-temperature superconductivity is still not clear, but it seems that instead of electron-phonon attraction mechanisms, as in conventional superconductivity, one is dealing with genuine electronic mechanisms (e.g. by antiferromagnetic correlations), and instead of s-wave pairing, d-waves are substantial.

One goal of all this research is room-temperature superconductivity.[20]

Examples

Examples of high-Tc cuprate superconductors include La1.85Ba0.15CuO4, and YBCO (yttrium-barium-copper-oxide), which is famous as the first material to achieve superconductivity above the boiling point of liquid nitrogen.

Process

Perovskites are made by mixing oxides in stoichiometric quantities and then heating in a furnace at high temperatures in an oxygen-rich atmosphere.

Ongoing research

The question of how superconductivity arises in high-temperature superconductors is one of the major unsolved problems of theoretical condensed matter physics as of 2011. The mechanism that causes the electrons in these crystals to form pairs is not known.

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 mechanism

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. Firstly, it has been suggested that the HTS emerges by antiferromagnetic spin fluctuation in a doped system.[21] According to this 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 HTS order parameter (pairing wave function) does not have d symmetry, then a pairing mechanism related to spin fluctuation can be ruled out. Secondly, there was the interlayer coupling model, according to which a layered structure consisting of BCS-type (s symmetry) superconductor can enhance the superconductivity by itself.[22] 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. Thus, 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. 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.

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 symmetry

There was a clever experimental design to overcome the muddy situation. An experiment based on flux quantization of a three-grain ring of YBa2Cu3O7 (YBCO) was proposed to test the symmetry of the order parameter in the HTS. The symmetry of the order parameter could best be probed at the junction interface as the Cooper pairs tunnel across a Josephson junction or weak link.[23] It was expected that only for a junction of d symmetry superconductors there could occur a half-integer flux, that is, a spontaneous magnetization. However, even if the junction experiment is the strongest method to determine the symmetry of the HTS order parameter, there have been ambiguous results of the junction experiments. J. R. Kirtley and C. C. Tsuei thought that the ambiguous results came from the defect inside the HTS, so that they designed the experiment where both of clean limit (no defect) and dirty limit (maximum of defects) were simultaneously considered.[24] In the experiment, the spontaneous magnetization was clearly observed in YBCO, which absolutely supported the d symmetry of the order parameter in YBCO. Because YBCO is orthorhombic, it might inherently have an admixture of s symmetry. So, by tuning their technique further, they found that there was an admixture of s symmetry in YBCO within about 3%.[25] Also, they found that there was a pure dx2-y2 order parameter symmetry in the tetragonal Tl2Ba2CuO6.[26]

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. Bibcode:1979PhRvL..43.1892S. doi:10.1103/PhysRevLett.43.1892.
  2. Ott, H. R.; Rudigier, H.; Fisk, Z.; Smith, J. (1983). "UBe_{13}: An Unconventional Actinide Superconductor". Physical Review Letters 50 (20): 1595. Bibcode:1983PhRvL..50.1595O. doi:10.1103/PhysRevLett.50.1595.
  3. 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. Bibcode:1984PhRvL..52..679S. doi:10.1103/PhysRevLett.52.679.
  4. 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.
  5. Shivaram, B. S.; Jeong, Y. H.; Rosenbaum, T.F.; Hinks, D. (1986). "Anisotropy of Transverse Sound in the Heavy-Fermion Superconductor UPt3". Physical Review Letters 56 (10): 1009. Bibcode:1986PhRvL..56.1078S. doi:10.1103/PhysRevLett.56.1078. PMID 10032562.
  6. Jérome, D.; Mazaud, A.; Ribault, M.; Bechgaard, K. (1980). "Superconductivity in a synthetic organic conductor (TMTSF)2PF 6". Journal de Physique Lettres 41 (4): 95. doi:10.1051/jphyslet:0198000410409500.
  7. Bechgaard, Klaus; Carneiro, Claus S.; Olsen, Malte; Rasmussen, Finn; Jacobsen, Claus (1981). "Zero-Pressure Organic Superconductor: Di-(Tetramethyltetraselenafulvalenium)-Perchlorate [(TMTSF)2ClO4]". Physical Review Letters 46 (13): 852. Bibcode:1981PhRvL..46..852B. doi:10.1103/PhysRevLett.46.852.
  8. Hiroki Takahashi, Kazumi Igawa, Kazunobu Arii, Yoichi Kamihara, Masahiro Hirano, Hideo Hosono; Igawa; Arii; Kamihara; Hirano; 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.
  9. "Iron Exposed as High-Temperature Superconductor: Scientific American". Sciam.com. 2008-04-23. Retrieved 2009-10-29.
  10. New High-Temperature Superconductors Are Iron-based With Unusual Magnetic Properties
  11. Samarium oxypnictide
  12. High-temp superconductors pave way for 'supermagnets'
  13. Hunte, F.; Jaroszynski, J.; Gurevich, A.; Larbalestier, D. C.; Jin, R.; Sefat, A. S.; McGuire, M. A.; Sales, B. C. et al. (2008). "Two-band superconductivity in LaFeAsO0.89F0.11 at very high magnetic fields". Nature 453 (7197): 903–5. Bibcode:2008Natur.453..903H. doi:10.1038/nature07058. PMID 18509332.
  14. J. G. Bednorz and K. A. Müller (1986). "Possible high Tc superconductivity in the Ba−La−Cu−O system". Z. Physik, B 64 (1): 189–193. Bibcode:1986ZPhyB..64..189B. doi:10.1007/BF01303701.
  15. 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.
  16. H. Maeda, Y. Tanaka, M. Fukutumi, and T. Asano (1988). "A New High-Tc Oxide Superconductor without a Rare Earth Element". Jpn. J. Appl. Phys. 27: L209–L210. Bibcode:1988JaJAP..27L.209M. doi:10.1143/JJAP.27.L209.
  17. P. Dai, B. C. Chakoumakos, G. F. Sun, K. W. Wong, Y. Xin and 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, and 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.
  19. Hiroki Takahashi, Kazumi Igawa, Kazunobu Arii, Yoichi Kamihara, Masahiro Hirano, Hideo Hosono; Igawa; Arii; Kamihara; Hirano; Hosono (2008). "Superconductivity at 43 K in an iron-based layered compound LaO1-xFxFeAs". Nature 453 (7193): 376–378. Bibcode:2008Natur.453..376T. doi:10.1038/nature06972. PMID 18432191.
  20. A. Mourachkine (2004). Room-Temperature Superconductivity. Cambridge International Science Publishing. arXiv:cond-mat/0606187. ISBN 1-904602-27-4.
  21. 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. Bibcode:1992PhRvB..4614803M. doi:10.1103/PhysRevB.46.14803.
  22. 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.
  23. V. B. Geshkenbein; Larkin, A.; Barone, A. et al. (1987). "Vortices with half magnetic flux quanta in heavy-fermion superconductors". Phys. Rev. B 36: 235. Bibcode:1987PhRvB..36..235G. doi:10.1103/PhysRevB.36.235.
  24. J. R. Kirtley; Tsuei, C. C.; Sun, J. Z.; Chi, C. C.; Yu-Jahnes, Lock See; Gupta, A.; Rupp, M.; Ketchen, M. B. et al. (1995). "Symmetry of the order parameter in the high-Tc superconductor YBa2Cu3O7- delta". Nature 373 (6511): 225. Bibcode:1995Natur.373..225K. doi:10.1038/373225a0.
  25. 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.
  26. 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.