Saturday 13 November 2010

YouTube - small blue thing - Suzanne Vega

Myanmar democracy leader Aung San Suu Kyi released

A smiling Suu Kyi, wearing a traditional jacket and a flower in her hair, appeared at the gate of her compound as the crowd chanted, cheered and sang the national anthem.

Speaking briefly in Burmese, she thanked the well-wishers, who quickly swelled to as many as 5,000, and said they would see each other again Sunday at the headquarters of her political party.

The 65-year-old Nobel Peace Prize laureate, whose latest period of detention spanned 7 1/2 years, has come to symbolize the struggle for democracy in the Southeast Asian nation ruled by the military since 1962.

The release from house arrest of one of the world's most prominent political prisoners came a week after an election that was swept by the military's proxy political party and decried by Western nations as a sham designed to perpetuate authoritarian control.

Supporters had been waiting most of the day near her residence and the headquarters of her party. Suu Kyi has been jailed or under house arrest for more than 15 of the last 21 years.

As her release was under way, riot police stationed in the area left the scene and a barbed-wire barricade near her residence was removed, allowing the waiting supporters to surge forward.

Her release was immediately welcomed by several activist groups around the world, and British Prime Minister David Cameron said it was long overdue.

"Aung San Suu Kyi is an inspiration for all of us who believe in freedom of speech, democracy and human rights," he said in a statement.

Critics allege the Nov. 7 elections were manipulated to give the pro-military party a sweeping victory. Results have been released piecemeal and already have given the junta-backed Union Solidarity and Development Party a majority in both houses of Parliament.

The last elections in 1990 were won overwhelmingly by Suu Kyi's National League for Democracy party, but the military refused to hand over power and instead clamped down on opponents.

Suu Kyi's release gives the junta some ammunition against critics of the election and the government's human rights record, which includes the continued detention of some 2,200 political prisoners and brutal military campaigns against ethnic minorities.

A 200ft Bungee Jump in Harlech - 01766 780 992

Bungee jumpers are not required to provide any equipment. Harnesses, ropes, etc. are supplied by the club. Sensible outdoor clothes are recommended.

Records

Highest bungee jump down a vertical wall, Verscaze Dam, Lugano, Switzerland, 700 ft.
Greatest altitude, from hot air balloon at 12,610 ft.
Longest unstretched jumpline, from a helicopter, 816 ft.
Highest jump from a Bridge, 1053 ft from Royal Gorge Bridge, Colorado, USA.
Oldest bungee jumper, 100 year old man.

Competitions

The first Bungee Jumping World Championships were staged in Rhode Island, USA, in 1995.

Books Photo by David Newall

Bungee Jumping for Fun and Profit (Frase)

Pros

Low cost
No equipment needed by the jumper
Excellent opportunity to raise money for charity
Low risk of injury

Cons

Only two permanent BERSA bungee sites in UK
Strong winds prevent play
Each buzz only lasts for a few seconds

Photo � David Newall

 

High-temperature superconductivity - Wikipedia, the free encyclopedia

From Wikipedia, the free encyclopedia

Jump to: navigation, search
Unsolved problems in physics What causes superconductivity at temperatures above 50 kelvin? Question mark2.svg

High-temperature superconductors (abbreviated high-Tc or HTS) are materials that have a superconducting transition temperature (Tc) above 30 K (−243.2 °C). From 1960 to 1980, 30 K was thought to be the highest theoretically possible Tc. The first high-Tc superconductor[1] was discovered in 1986 by IBM Researchers Karl Müller and Johannes Bednorz, for which they were awarded the Nobel Prize in Physics in 1987. Until Fe-based superconductors were discovered in 2008,[2][3] the term high-temperature superconductor was used interchangeably with cuprate superconductor for compounds such as bismuth strontium calcium copper oxide (BSCCO) and yttrium barium copper oxide (YBCO).

"High-temperature" has three common definitions in the context of superconductivity:

  1. Above the temperature of 30 K that had historically been taken as the upper limit allowed by BCS theory. This is also above the 1973 record of 23 K that had lasted until copper-oxide materials were discovered in 1986.
  2. Having a transition temperature that is a larger fraction of the Fermi temperature than for conventional superconductors such as elemental mercury or lead. This definition encompasses a wider variety of unconventional superconductors and is used in the context of theoretical models.
  3. Greater than the boiling point of liquid nitrogen (77 K or −196 °C). This is significant for technological applications of superconductivity because liquid nitrogen is a relatively inexpensive and easily handled coolant.

Technological applications benefit from both the higher critical temperature being above the boiling point of liquid nitrogen and also the higher critical magnetic field (and critical current density) at which superconductivity is destroyed. In magnet applications the high critical magnetic field may be more valuable than the high Tc itself. Some cuprates have an upper critical field around 100 teslas. However, cuprate materials are brittle ceramics which are expensive to manufacture and not easily turned into wires or other useful shapes.

Two decades of intense experimental and theoretical research, with over 100,000 published papers on the subject,[4] have discovered many common features in the properties of high-temperature superconductors,[5] but as of 2009[update], there is no widely accepted theory to explain their properties. Cuprate superconductors (and other unconventional superconductors) differ in many important ways from conventional superconductors, such as elemental mercury or lead, which are adequately explained by the BCS theory. There also has been much debate as to high-temperature superconductivity coexisting with magnetic ordering in YBCO,[6] iron-based superconductors, several ruthenocuprates and other exotic superconductors, and the search continues for other families of materials. HTS are Type-II superconductors, which allow magnetic fields to penetrate their interior in quantized units of flux, meaning that much higher magnetic fields are required to suppress superconductivity. The layered structure also gives a directional dependence to the magnetic field response.

Contents

[edit] History and progress

  • April 1986 - The term high-temperature superconductor was first used to designate the new family of cuprate-perovskite ceramic materials discovered by Johannes Georg Bednorz and Karl Alexander Müller,[1] for which they won the Nobel Prize in Physics the following year. Their discovery of the first high-temperature superconductor, LaBaCuO, with a transition temperature of 30 K, generated great excitement.
  • LSCO (La2-xSrxCuO4) discovered the same year.
  • January 1987 - YBCO was discovered to have a Tc of 90 K.[7]
  • 1988 - BSCCO discovered with Tc up to 108 K,[8] and TBCCO (T=thallium) discovered to have Tc of 127 K.[9]
  • As of 2009[update], the highest-temperature superconductor (at ambient pressure) is mercury barium calcium copper oxide (HgBa2Ca2Cu3Ox), at 135 K and is held by a cuprate-perovskite material,[10] possibly 164 K under high pressure.[11]
  • Recently, iron-based superconductors with critical temperatures as high as 56 K have been discovered.[12][13] These are often also referred to as high-temperature superconductors.

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.[14]

[edit] 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.

Transition temperatures of well-known superconductors
(Boiling point of liquid nitrogen for comparison) Transition temperature
(in kelvins) Material Class 133 HgBa2Ca2Cu3Ox Copper-oxide superconductors 110 Bi2Sr2Ca2Cu3O10(BSCCO) 90 YBa2Cu3O7 (YBCO) 77 Boiling point of liquid nitrogen 55 SmFeAs(O,F) Iron-based superconductors 41 CeFeAs(O,F) 26 LaFeAs(O,F) 20 Boiling point of liquid hydrogen 18 Nb3Sn Metallic low-temperature superconductors 10 NbTi 4.2 Hg (mercury)

[edit] Cuprates

Simplified doping dependent phase diagram of cuprate superconductors for both electron (n) and hole (p) doping. The phases shown are the antiferromagnetic (AF) phase close to zero doping, the superconducting phase around optimal doping, and the pseudogap phase. Doping ranges possible for some common compounds are also shown. After.[15]

Cuprate superconductors are generally considered to be quasi-two-dimensional materials with their superconducting properties determined by electrons moving within weakly coupled copper-oxide (CuO2) layers. Neighbouring layers containing ions such as La, Ba, Sr, or other atoms act to stabilize the structure and dope electrons or holes onto the copper-oxide layers. The undoped 'parent' or 'mother' compounds are Mott insulators with long-range antiferromagnetic order at low enough temperature. Single band models are generally considered to be sufficient to describe the electronic properties.

The cuprate superconductors adopt a perovskite structure. The copper-oxide planes are checkerboard lattices with squares of O2− ions with a Cu2+ ion at the centre of each square. The unit cell is rotated by 45° from these squares. Chemical formulae of superconducting materials generally contain fractional numbers to describe the doping required for superconductivity. There are several families of cuprate superconductors and they can be categorized by the elements they contain and the number of adjacent copper-oxide layers in each superconducting block. For example, YBCO and BSCCO can alternatively be referred to as Y123 and Bi2201/Bi2212/Bi2223 depending on the number of layers in each superconducting block (n). The superconducting transition temperature has been found to peak at an optimal doping value (p=0.16) and an optimal number of layers in each superconducting block, typically n = 3.

A small sample of the high-temperature superconductor BSCCO-2223.

Possible mechanisms for superconductivity in the cuprates are still the subject of considerable debate and further research. Certain aspects common to all materials have been identified.[5] Similarities between the antiferromagnetic low-temperature state of the undoped materials and the superconducting state that emerges upon doping, primarily the dx2-y2 orbital state of the Cu2+ ions, suggest that electron-electron interactions are more significant than electron-phonon interactions in cuprates – making the superconductivity unconventional. Recent work on the Fermi surface has shown that nesting occurs at four points in the antiferromagnetic Brillouin zone where spin waves exist and that the superconducting energy gap is larger at these points. The weak isotope effects observed for most cuprates contrast with conventional superconductors that are well described by BCS theory.

Similarities and differences in the properties of hole-doped and electron doped cuprates:

  • Presence of a pseudogap phase up to at least optimal doping.
  • Different trends in the Uemura plot relating transition temperature to the superfluid density. The inverse square of the London penetration depth appears to be proportional to the critical temperature for a large number of underdoped cuprate superconductors, but the constant of proportionality is different for hole- and electron-doped cuprates. The linear trend implies that the physics of these materials is strongly two-dimensional.
  • Universal hourglass-shaped feature in the spin excitations of cuprates measured using inelastic neutron diffraction.
  • Nernst effect evident in both the superconducting and pseudogap phases.

[edit] Iron-based superconductors

Simplified doping dependent phase diagrams of iron-based superconductors for both Ln-1111 and Ba-122 materials. The phases shown are the antiferromagnetic/spin density wave (AF/SDW) phase close to zero doping and the superconducting phase around optimal doping. The Ln-1111 phase diagrams for La[16] and Sm[17][18] were determined using muon spin spectroscopy, the phase diagram for Ce[19] was determined using neutron diffraction. The Ba-122 phase diagram is based on.[20]

Iron-based superconductors contain layers of iron and a pnictogen, such as arsenic, phosphorus, or chalcogens. This is currently the family with the second highest critical temperature, behind the cuprates. Interest in their superconducting properties began in 2006 with the discovery of superconductivity in LaFePO at 4 K[21] and gained much greater attention in 2008 after the analogous material LaFeAs(O,F)[12] was found to superconduct at up to 43 K under pressure.[13]

Since the original discoveries several families of iron-based superconductors have emerged:

  • LnFeAs(O,F) or LnFeAsO1-x with Tc up to 56 K, referred to as 1111 materials.[3] A fluoride variant of these materials was subsequently found with similar Tc values.[22]
  • (Ba,K)Fe2As2 and related materials with pairs of iron-arsenide layers, referred to as 122 compounds. Tc values range up to 38 K.[23][24] These materials also superconduct when iron is replaced with cobalt
  • LiFeAs and NaFeAs with Tc up to around 20 K. These materials superconduct close to stoichiometric composition and are referred to as 111 compounds.[25][26][27]
  • FeSe with small off-stoichiometry or tellurium doping.[28]

Most undoped iron-based superconductors show a tetragonal-orthorhombic structural phase transition followed at lower temperature by magnetic ordering, similar to the cuprate superconductors.[19] However, they are poor metals rather than Mott insulators and have five bands at the Fermi surface rather than one. The phase diagram emerging as the iron-arsenide layers are doped is remarkably similar, with the superconducting phase close to or overlapping the magnetic phase. Strong evidence that the Tc value varies with the As-Fe-As bond angles has already emerged and shows that the optimal Tc value is obtained with undistorted FeAs4 tetrahedra.[29] The symmetry of the pairing wavefunction is still widely debated, but an extended s-wave scenario is currently favoured.

[edit] Other materials sometimes referred to as high-temperature superconductors

Magnesium diboride is occasionally referred to as a high-temperature superconductor because its Tc value of 39 K is above that historically expected for BCS superconductors. However, it is more generally regarded as the highest Tc conventional superconductor, the increased Tc resulting from two separate bands being present at the Fermi energy.

Fulleride superconductors[30] where alkali-metal atoms are intercalated into C60 molecules show superconductivity at temperatures of up to 38 K for Cs3C60.[31]

Some organic superconductors and heavy fermion compounds are considered to be high-temperature superconductors because of their high Tc values relative to their Fermi energy, despite the Tc values being lower than for many conventional superconductors. This description may relate better to common aspects of the superconducting mechanism than the superconducting properties.

Theoretical work by Neil Ashcroft predicted that liquid metallic hydrogen at extremely high pressure should become superconducting at approximately room-temperature because of its extremely high speed of sound and expected strong coupling between the conduction electrons and the lattice vibrations.[32] This prediction is yet to be experimentally verified.

All known high-Tc superconductors are Type-II superconductors. In contrast to Type-I superconductors, which expel all magnetic fields due to the Meissner Effect, Type-II superconductors allow magnetic fields to penetrate their interior in quantized units of flux, creating "holes" or "tubes" of normal metallic regions in the superconducting bulk. Consequently, high-Tc superconductors can sustain much higher magnetic fields.

[edit] 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 2010[update]. The mechanism that causes the electrons in these crystals to form pairs is not known.[5] 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 modelling difficult. Improving the quality and variety of samples also gives rise to considerable research, both with the aim of improved characterisation of the physical properties of existing compounds, and synthesizing new materials, often with the hope of increasing Tc. Technological research focusses on making HTS materials in sufficient quantities to make their use economically viable and optimizing their properties in relation to applications.

[edit] Possible mechanism

There have been two representative theories for HTS. Firstly, it has been suggested that the HTS emerges from antiferromagnetic spin fluctuations in a doped system.[33] According to this theory, the pairing wave function of the cuprate HTS should have a dx2-y2 symmetry. Thus, determining whether the pairing wave function has d-wave symmetry is essential to test the spin fluctuation mechanism. That is, if the HTS order parameter (pairing wave function) does not have d-wave symmetry, then a pairing mechanism related to spin fluctuations can be ruled out. (Similar arguments can be made for iron-based superconductors but the different material properties allow a different pairing symmetry.) Secondly, there was the interlayer coupling model, according to which a layered structure consisting of BCS-type (s-wave symmetry) superconductors can enhance the superconductivity by itself.[34] By introducing an additional tunnelling interaction between each layer, this model successfully explained the anisotropic symmetry of the order parameter as well as the emergence of the HTS. Thus, in order to solve this unsettled problem, there have been numerous experiments such as photoemission spectroscopy, NMR, specific heat measurements, etc. But, unfortunately, the results were ambiguous, some reports supported the d symmetry for the HTS whereas 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.

[edit] Junction experiment supporting the d symmetry

The Meissner effect or a magnet levitating above a superconductor (cooled by liquid nitrogen).

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.[35] It was expected that a half-integer flux, that is, a spontaneous magnetization could only occur for a junction of d symmetry superconductors. But, even if the junction experiment is the strongest method to determine the symmetry of the HTS order parameter, the results have been ambiguous. J. R. Kirtley and C. C. Tsuei thought that the ambiguous results came from the defects inside the HTS, so that they designed an experiment where both clean limit (no defects) and dirty limit (maximal defects) were considered simultaneously.[36] In the experiment, the spontaneous magnetization was clearly observed in YBCO, which supported the d symmetry of the order parameter in YBCO. But, since 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%.[37] Also, they found that there was a pure dx2-y2 order parameter symmetry in the tetragonal Tl2Ba2CuO6.[38]

[edit] Qualitative explanation of the spin-fluctuation mechanism

While, despite all these years, the mechanism of high-Tc superconductivity is still highly controversial, this being due to mostly the lack of exact theoretical computations on such strongly interacting electron systems, most rigorous theoretical calculations, including phenomenological and diagrammatic approaches, converge on magnetic fluctuations as the pairing mechanism for these systems. The qualitative explanation is as follows. (Note that, in the following argument, one can replace “electron” with “hole” and vice versa depending on the actual material.)

In a normal conductor, a hole is created whenever an electron is moved. This causes a resistivity because charge neutrality must be conserved and as electrons move under an electric field, they drag holes behind them through defects and thermal oscillations in the system. In contrast, in a superconductor, one gets an unlimited supply of electrons without creating holes behind. This is through the creation of so-called Cooper pairs in a superconductor. Cooper pairs are pairs of electrons. In a normal conductor, creation of an electron leads to creation of a hole, which conserves the number of particles. But in a superconductor, it's possible to create a Cooper pair without creating holes and therefore not to conserve the number of particles, hence leading to the unlimited supply of electrons.

In a conventional superconductor, Cooper pairs are created as follows. When an electron moves through the system, it creates a depression in the atomic lattice through lattice vibrations known as phonons. If the depression of the lattice is strong enough, another electron can fall into the depression created by the first electron—the so-called water-bed effect—and a Cooper pair is formed. When this effect becomes strong enough, Cooper pairs win over the creation of holes behind the electrons, and the normal conductor turns into a superconductor through an unlimited supply of electrons by the creation of Cooper pairs.

In a high-Tc superconductor, the mechanism is extremely similar to a conventional superconductor. Except, in this case, phonons virtually play no role and their role is replaced by spin-density waves. As all conventional superconductors are strong phonon systems, all high-Tc superconductors are strong spin-density wave systems, within close vicinity of a magnetic transition to, for example, an antiferromagnet. When an electron moves in a high-Tc superconductor, its spin creates a spin-density wave around it. This spin-density wave in turn causes a nearby electron to fall into the spin depression created by the first electron (water-bed effect again). Hence, again, a Cooper pair is formed. Eventually, when the system temperature is lowered, more spin density waves and Cooper pairs are created and superconductivity begins when an unlimited supply of Cooper pairs, denoted as a phase transition, happens. Note that in high-Tc systems, as these systems are magnetic systems due to the Coulomb interaction, there is a strong Coulomb repulsion between electrons. This Coulomb repulsion prevents pairing of the Cooper pairs on the same lattice site. The pairing of the electrons occur at near-neighbor lattice sites as a result. This is the so-called d-wave pairing, where the pairing state has a node (zero) at the origin.

[edit] See also

[edit] References

  1. ^ a b J.G. Bednorz and K.A. Mueller (1986). "Possible high TC superconductivity in the Ba-La-Cu-O system". Z. Phys. B64 (2): 189–193. doi:10.1007/BF01303701. 
  2. ^ Iron Exposed as High-Temperature Superconductor: Scientific American
  3. ^ a b Ren, Zhi-An; Che, Guang-Can; Dong, Xiao-Li; Yang, Jie; Lu, Wei; Yi, Wei; Shen, Xiao-Li; Li, Zheng-Cai et al. (2008). ""Superconductivity and phase diagram in iron-based arsenic-oxides ReFeAsO1−δ (Re = rare-earth metal) without fluorine doping". EPL (Europhysics Letters) 83: 17002. doi:10.1209/0295-5075/83/17002. 
  4. ^ Mark Buchanan (2001). "Mind the pseudogap". Nature 409 (6816): 8. doi:10.1038/35051238. PMID 11343081. 
  5. ^ a b c Anthony Leggett (2006). "What DO we know about high Tc?". Nature Physics 2: 134. doi:10.1038/nphys254. 
  6. ^ S. Sanna, G. Allodi, G. Concas, A. D. Hillier, and R. De Renzi (2004). "Nanoscopic Coexistence of Magnetism and Superconductivity in YBa2Cu3O6+x Detected by Muon Spin Rotation". Phys. Rev. Lett. 93: 207001. doi:10.1103/PhysRevLett.93.207001. 
  7. ^ 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. doi:10.1103/PhysRevLett.58.908. PMID 10035069. 
  8. ^ 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. doi:10.1143/JJAP.27.L209. 
  9. ^ Sheng, Z. Z.; Hermann A. M. (1988). "Bulk superconductivity at 120 K in the Tl–Ca/Ba–Cu–O system". Nature 332: 138–139. doi:10.1038/332138a0. 
  10. ^ Chu, C. W.; Gao, L.; Chen, F.; Huang, Z. J.; Meng, R. L.; Xue, Y. Y. (1993). "Superconductivity above 150 K in HgBa2Ca2Cu3O8+δ at high pressures". Nature 365: 323. doi:10.1038/365323a0. 
  11. ^ 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. doi:10.1103/PhysRevB.50.4260. 
  12. ^ a b Yoichi Kamihara, Takumi Watanabe, Masahiro Hirano, and Hideo Hosono (2008). "Iron-Based Layered Superconductor La[O1-xFxFeAs (x = 0.05−0.12) with Tc = 26 K"]. J. Am. Chem. Soc. 130 (11): 3296–3297. doi:10.1021/ja800073m. PMID 18293989. http://pubs.acs.org/doi/abs/10.1021/ja800073m. 
  13. ^ a b Hiroki Takahashi, Kazumi Igawa, Kazunobu Arii, Yoichi Kamihara, Masahiro Hirano, Hideo Hosono (2008). "Superconductivity at 43 K in an iron-based layered compound LaO1-xFxFeAs". Nature 453 (7193): 376–378. doi:10.1038/nature06972. PMID 18432191. 
  14. ^ A. Mourachkine (2004). Room-Temperature Superconductivity. Cambridge International Science Publishing (Cambridge, UK) (also http://xxx.lanl.gov/abs/cond-mat/0606187). ISBN 1904602274. 
  15. ^ Christine Hartinger. "DFG FG 538 - Doping Dependence of Phase transitions and Ordering Phenomena in Cuprate Superconductors". Wmi.badw-muenchen.de. http://www.wmi.badw-muenchen.de/FG538/projects/P4_crystal_growth/index.htm. Retrieved 2009-10-29. 
  16. ^ H. Luetkens, H.-H. Klauss, M. Kraken, F. J. Litterst, T. Dellmann, R. Klingeler, C. Hess, R. Khasanov, A. Amato, C. Baines, J. Hamann-Borrero, N. Leps, A. Kondrat, G. Behr, J. Werner, B. Buechner (2009). "Electronic phase diagram of the LaO1-xFxFeAs superconductor". doi:10.1038/nmat2397. http://www.nature.com/nmat/journal/vaop/ncurrent/abs/nmat2397.html. 
  17. ^ A. J. Drew, Ch. Niedermayer, P. J. Baker, F. L. Pratt, S. J. Blundell, T. Lancaster, R. H. Liu, G. Wu, X. H. Chen, I. Watanabe, V. K. Malik, A. Dubroka, M. Rössle, K. W. Kim, C. Baines and C. Bernhard (2009). "Coexistence of static magnetism and superconductivity in SmFeAsO1-xFx as revealed by muon spin rotation". Nature Materials 8 (4): 310–4. doi:10.1038/nmat2396. PMID 19234446. http://www.nature.com/nmat/journal/vaop/ncurrent/abs/nmat2396.html. 
  18. ^ S. Sanna, R. De Renzi, G. Lamura, C. Ferdeghini, A. Palenzona, M. Putti, M. Tropeano, and T. Shiroka (2009). "Competition between magnetism and superconductivity at the phase boundary of doped SmFeAsO pnictides". http://arxiv.org/abs/0902.2156. 
  19. ^ a b Jun Zhao, Q. Huang, Clarina de la Cruz, Shiliang Li, J. W. Lynn, Y. Chen, M. A. Green, G. F. Chen, G. Li, Z. Li, J. L. Luo, N. L. Wang & Pengcheng Dai (2008). "Structural and magnetic phase diagram of CeFeAsO1-xFx and its relation to high-temperature superconductivity". Nature Materials 7 (12): 953–959. doi:10.1038/nmat2315. PMID 18953342. http://www.nature.com/nmat/journal/v7/n12/abs/nmat2315.html. 
  20. ^ Jiun-Haw Chu, James G. Analytis, Chris Kucharczyk, Ian R. Fisher (2008). "Determination of the phase diagram of the electron doped superconductor Ba(Fe1-x}Cox)2As2". http://arxiv.org/abs/0811.2463. 
  21. ^ Yoichi Kamihara, Hidenori Hiramatsu, Masahiro Hirano, Ryuto Kawamura, Hiroshi Yanagi, Toshio Kamiya, and Hideo Hosono (2006). "Iron-Based Layered Superconductor: LaOFeP". J. Am. Chem. Soc. 128 (31): 10012–10013. doi:10.1021/ja063355c. PMID 16881620. http://pubs.acs.org/doi/abs/10.1021/ja063355c. 
  22. ^ G. Wu, Y. L. Xie, H. Chen, M. Zhong, R. H. Liu, B. C. Shi, Q. J. Li, X. F. Wang, T. Wu, Y. J. Yan, J. J. Ying and X. H. Chen (2008). "Superconductivity at 56 K in Samarium-doped SrFeAsF". http://arxiv.org/abs/0811.0761v2. 
  23. ^ Marianne Rotter, Marcus Tegel, and Dirk Johrendt (2008). "Superconductivity at 38 K in the Iron Arsenide (Ba1-xKx)Fe2As2". Physical Review Letters 101 (10): 107006. doi:10.1103/PhysRevLett.101.107006. PMID 18851249. http://link.aps.org/doi/10.1103/PhysRevLett.101.107006. 
  24. ^ Kalyan Sasmal, Bing Lv, Bernd Lorenz, Arnold M. Guloy, Feng Chen, Yu-Yi Xue, and Ching-Wu Chu (2008). "Superconducting Fe-Based Compounds (A1-xSrx)Fe2As2 with A=K and Cs with Transition Temperatures up to 37 K". Physical Review Letters 101 (10): 107007. doi:10.1103/PhysRevLett.101.107007. PMID 18851250. http://link.aps.org/doi/10.1103/PhysRevLett.101.107007. 
  25. ^ Michael J. Pitcher, Dinah R. Parker, Paul Adamson, Sebastian J. C. Herkelrath, Andrew T. Boothroyd, Richard M. Ibberson, Michela Brunelli and Simon J. Clarke (2008). "Structure and superconductivity of LiFeAs". Chem. Commun. 2008 (45): 5918–5920. doi:10.1039/b813153h. PMID 19030538. http://xlink.rsc.org/?doi=b813153h. 
  26. ^ Joshua H. Tapp, Zhongjia Tang, Bing Lv, Kalyan Sasmal, Bernd Lorenz, Paul C. W. Chu, and Arnold M. Guloy (2008). "LiFeAs: An intrinsic FeAs-based superconductor with Tc=18 K". Physical Review B 78: 060505. doi:10.1103/PhysRevB.78.060505. 
  27. ^ Dinah R. Parker, Michael J. Pitcher, Peter J. Baker, Isabel Franke, Tom Lancaster, Stephen J. Blundell and Simon J. Clarke (2009). "Structure, antiferromagnetism and superconductivity of the layered iron arsenide NaFeAs". Chem. Commun. 2009 (16): 2189–2191. doi:10.1039/b818911k. PMID 19360189. 
  28. ^ Fong-Chi Hsu, Jiu-Yong Luo, Kuo-Wei Yeh, Ta-Kun Chen, Tzu-Wen Huang, Phillip M. Wu, Yong-Chi Lee, Yi-Lin Huang, Yan-Yi Chu, Der-Chung Yan, and Maw-Kuen Wu (2008). "Superconductivity in the PbO-type structure α-FeSe". PNAS 105 (38): 14262–14264. doi:10.1073/pnas.0807325105. PMID 18776050. PMC 2531064. http://www.pnas.org/content/105/38/14262. 
  29. ^ C.-H. Lee, A. Iyo, H. Eisaki, H. Kito, M. T. Fernandez-Diaz, T. Ito, K. Kihou, H. Matsuhata, M. Braden, and K. Yamada (2008). "Effect of Structural Parameters on Superconductivity in Fluorine-Free LnFeAsO1-y (Ln = La, Nd)". J. Phys. Soc. Jpn. 77: 083704. doi:10.1143/JPSJ.77.083704. 
  30. ^ A. F. Hebard, M. J. Rosseinsky, R. C. Haddon, D. W. Murphy, S. H. Glarum, T. T. M. Palstra, A. P. Ramirez, and A. R. Kortan (1991). "Superconductivity at 18 K in potassium-doped C60". Nature 350: 600. doi:10.1038/350600a0. http://www.nature.com/nature/journal/v350/n6319/abs/350600a0.html. 
  31. ^ A. Y. Ganin, Y. Takabayashi, Y. Z. Khimyak, S. Margadonna, A. Tamai, M. J. Rosseinsky, and K. Prassides (2008). "Bulk superconductivity at 38 K in a molecular system". Nature Materials 7 (5): 367. doi:10.1038/nmat2179. PMID 18425134. http://www.nature.com/nmat/journal/v7/n5/abs/nmat2179.html. 
  32. ^ N. W. Ashcroft (1968). "Metallic Hydrogen: A High-Temperature Superconductor?". Physical Review Letters 21: 1748–1749. doi:10.1103/PhysRevLett.21.1748. http://prola.aps.org/abstract/PRL/v21/i26/p1748_1. 
  33. ^ P. Monthoux, A. V. Balatsky, and D. Pines (1992). "Weak-coupling theory of high-temperature superconductivity in the antiferromagnetically correlated copper oxides". Phys. Rev. B 46: 14803–14817. doi:10.1103/PhysRevB.46.14803. http://prola.aps.org/abstract/PRB/v46/i22/p14803_1. 
  34. ^ S. Chakravarthy, A. Sudbø, P. W. Anderson, and S. Strong (1993). "Interlayer Tunneling and Gap Anisotropy in High-Temperature Superconductors". Science 261 (5119): 337–340. doi:10.1126/science.261.5119.337. PMID 17836845. http://www.sciencemag.org/cgi/content/abstract/261/5119/337. 
  35. ^ V. B. Geshkenbein, A. I. Larkin, and A. Barone (1987). "Vortices with half magnetic flux quanta in ‘‘heavy-fermion’’ superconductors". Phys. Rev. B 36: 235–238. doi:10.1103/PhysRevB.36.235. http://prola.aps.org/abstract/PRB/v36/i1/p235_1. 
  36. ^ J. R. Kirtley, C. C. Tsuei, J. Z. Sun, C. C. Chi, Lock See Yu-Jahnes, A. Gupta, M. Rupp & M. B. Ketchen (1995). "Symmetry of the order parameter in the high-Tc superconductor YBa2Cu3O7-δ". Nature 373: 225–228. doi:10.1038/373225a0. http://www.nature.com/nature/journal/v373/n6511/abs/373225a0.html. 
  37. ^ J. R. Kirtley, C. C. Tsuei, Ariando, C. J. M. Verwijs, S. Harkema, and H. Hilgenkamp (2006). "Angle-resolved phase-sensitive determination of the in-plane gap symmetry in YBa2Cu3O7-δ". Nature Physics 2: 190–194. doi:10.1038/nphys215. http://www.nature.com/nphys/journal/v2/n3/abs/nphys215.html. 
  38. ^ C. C. Tsuei J. R. Kirtley, Z. F. Ren, J. H. Wang, H. Raffy, and Z. Z. Li (1997). "Pure dx2-y2 order-parameter symmetry in the tetragonal superconductor Tl2Ba2CuO6+δ". Nature 387: 481. doi:10.1038/387481a0. http://www.nature.com/nature/journal/v387/n6632/abs/387481a0.html. 

[edit] External links

Flickr - projectbrainsaver

www.flickr.com
projectbrainsaver's A Point of View photoset projectbrainsaver's A Point of View photoset