superconductivity

superconductivity
superconduction /sooh'peuhr keuhn duk"sheuhn/, n.superconductive /sooh'peuhr keuhn duk"tiv/, superconducting, adj.superconductor /sooh'peuhr keuhn duk"teuhr/, n.
/sooh'peuhr kon'deuhk tiv"i tee/, n. Physics.
the phenomenon of almost perfect conductivity shown by certain substances at temperatures approaching absolute zero. The recent discovery of materials that are superconductive at temperatures hundreds of degrees above absolute zero raises the possibility of revolutionary developments in the production and transmission of electrical energy.
[1915-20; SUPER- + CONDUCTIVITY]

* * *

Almost total lack of electrical resistance in certain materials when they are cooled to a temperature near absolute zero.

Superconducting materials allow low power dissipation, high-speed operation, and high sensitivity. They also have the ability to prevent external magnetic fields from penetrating their interiors and are perfect diamagnets (see diamagnetism). Since it was first discovered in mercury by Heike Kamerlingh Onnes in 1911, similar behaviour has been found in some 25 other chemical elements and in thousands of alloys and compounds. Superconductors have applications in medical imaging, magnetic energy-storage systems, motors, generators, transformers, computer components, and sensitive magnetic-field measuring devices.

* * *

Introduction

      complete disappearance of electrical resistance in various solids when they are cooled below a characteristic temperature. This temperature, called the transition temperature, varies for different materials but generally is below 20 K (−253 °C).

      The use of superconductors in magnets is limited by the fact that strong magnetic fields (magnetic field) above a certain critical value, depending upon the material, cause a superconductor to revert to its normal, or nonsuperconducting, state, even though the material is kept well below the transition temperature.

      Suggested uses for superconducting materials include medical magnetic-imaging devices, magnetic energy-storage systems, motors, generators, transformers, computer parts, and very sensitive devices for measuring magnetic fields, voltages, or currents. The main advantages of devices made from superconductors are low power dissipation, high-speed operation, and high sensitivity.

Discovery
      Superconductivity was discovered in 1911 by the Dutch physicist Heike Kamerlingh Onnes (Kamerlingh Onnes, Heike); he was awarded the Nobel Prize for Physics in 1913 for his low-temperature research. Kamerlingh Onnes found that the electrical resistivity of a mercury wire disappears suddenly when it is cooled below a temperature of about 4 K (−269 °C); absolute zero is 0 K, the temperature at which all matter loses its disorder. He soon discovered that a superconducting material can be returned to the normal (i.e., nonsuperconducting) state either by passing a sufficiently large current through it or by applying a sufficiently strong magnetic field to it.

      For many years it was believed that, except for the fact that they had no electrical resistance (i.e., that they had infinite electrical conductivity), superconductors had the same properties as normal materials. This belief was shattered in 1933 by the discovery that a superconductor is highly diamagnetic (diamagnetism); that is, it is strongly repelled by and tends to expel a magnetic field. This phenomenon, which is very strong in superconductors, is called the Meissner effect for one of the two men who discovered it. Its discovery made it possible to formulate, in 1934, a theory of the electromagnetic properties of superconductors that predicted the existence of an electromagnetic penetration depth, which was first confirmed experimentally in 1939. In 1950 it was clearly shown for the first time that a theory of superconductivity must take into account the fact that free electrons in a crystal are influenced by the vibrations of atoms that define the crystal structure, called the lattice vibrations. In 1953, in an analysis of the thermal conductivity of superconductors, it was recognized that the distribution of energies of the free electrons in a superconductor is not uniform but has a separation called the energy gap.

      The theories referred to thus far served to show some of the interrelationships between observed phenomena but did not explain them as consequences of the fundamental laws of physics. For almost 50 years after Kamerlingh Onnes's discovery, theorists were unable to develop a fundamental theory of superconductivity. Finally, in 1957 such a theory was presented by the physicists John Bardeen (Bardeen, John), Leon N. Cooper (Cooper, Leon N.), and John Robert Schrieffer (Schrieffer, John Robert) of the United States; it won for them the Nobel Prize for Physics in 1972. It is now called the BCS theory in their honour, and most later theoretical work is based on it. The BCS theory also provided a foundation for an earlier model that had been introduced by the Russian physicists Lev Davidovich Landau and Vitaly Lazarevich Ginzburg (1950). This model has been useful in understanding electromagnetic properties, including the fact that any internal magnetic flux in superconductors exists only in discrete amounts (instead of in a continuous spectrum of values), an effect called the quantization of magnetic flux. This flux quantization, which had been predicted from quantum mechanical principles, was first observed experimentally in 1961.

      In 1962 the British physicist Brian D. Josephson (Josephson, Brian D.) predicted that two superconducting objects placed in electric contact would display certain remarkable electromagnetic properties. These properties (Josephson effect) have since been observed in a wide variety of experiments, demonstrating quantum mechanical effects on a macroscopic scale.

      The theory of superconductivity has been tested in a wide range of experiments, involving, for example, ultrasonic absorption studies, nuclear-spin phenomena, low-frequency infrared absorption, and electron-tunneling experiments. The results of these measurements have brought understanding to many of the detailed properties of various superconductors.

Thermal properties of superconductors
      Superconductivity is a startling departure from the properties of normal (i.e., nonsuperconducting) conductors of electricity. In materials that are electric conductors, some of the electrons are not bound to individual atoms but are free to move through the material; their motion constitutes an electric current. In normal conductors these so-called conduction electrons are scattered by impurities, dislocations, grain boundaries, and lattice vibrations (phonons). In a superconductor, however, there is an ordering among the conduction electrons that prevents this scattering. Consequently, electric current can flow with no resistance at all. The ordering of the electrons, called Cooper pairing, involves the momenta of the electrons rather than their positions. The energy per electron that is associated with this ordering is extremely small, typically about one thousandth of the amount by which the energy per electron changes when a chemical reaction takes place. One reason that superconductivity remained unexplained for so long is the smallness of the energy changes that accompany the transition between normal and superconducting states. In fact, many incorrect theories of superconductivity were advanced before the BCS theory was proposed. For additional details on electric conduction in metals and the effects of temperature and other influences, see the article electricity.

      Hundreds of materials are known to become superconducting at low temperatures. Twenty-seven of the chemical elements, all of them metals, are superconductors in their usual crystallographic forms at low temperatures and low (atmospheric) pressure. Among these are commonly known metals such as aluminum, tin, lead, and mercury and less common ones such as rhenium, lanthanum, and protactinium. In addition, 11 chemical elements that are metals, semimetals, or semiconductors are superconductors at low temperatures and high pressures. Among these are uranium, cerium, silicon, and selenium. Bismuth and five other elements, though not superconducting in their usual crystallographic form, can be made superconducting by preparing them in a highly disordered form, which is stable at extremely low temperatures. Superconductivity is not exhibited by any of the magnetic elements chromium, manganese, iron, cobalt, or nickel.

      Most of the known superconductors are alloys or compounds. It is possible for a compound to be superconducting even if the chemical elements constituting it are not; examples are disilver fluoride (Ag2F) and a compound of carbon and potassium (C8K). Some semiconducting compounds, such as tin telluride (SnTe), become superconducting if they are properly doped with impurities.

      Since 1986 some compounds containing copper and oxygen (called cuprates) have been found to have extraordinarily high transition temperatures, denoted Tc. This is the temperature below which a substance is superconducting. The properties of these high-Tc compounds are different in some respects from those of the types of superconductors known prior to 1986, which will be referred to as classic superconductors in this discussion. For the most part, the high-Tc superconductors are treated explicitly toward the end of this section. In the discussion that immediately follows, the properties possessed by both kinds of superconductors will be described, with attention paid to specific differences for the high-Tc materials. A further classification problem is presented by the superconducting compounds of carbon (sometimes doped with other atoms) in which the carbon atoms are on the surface of a cluster with a spherical or spheroidal crystallographic structure. These compounds, discovered in the 1980s, are called fullerenes (fullerene) (if only carbon is present) or fullerides (if doped). They have superconducting transition temperatures higher than those of the classic superconductors. It is not yet known whether these compounds are fundamentally similar to the cuprate high-temperature superconductors.

Transition temperatures
      The vast majority of the known superconductors have transition temperatures that lie between 1 K and 10 K. Of the chemical elements, tungsten has the lowest transition temperature, 0.015 K, and niobium the highest, 9.2 K. The transition temperature is usually very sensitive to the presence of magnetic impurities. A few parts per million of manganese in zinc, for example, lowers the transition temperature considerably.

Specific heat and thermal conductivity
      The thermal properties of a superconductor can be compared with those of the same material at the same temperature in the normal state. (The material can be forced into the normal state at low temperature by a large enough magnetic field.)

 When a small amount of heat is put into a system, some of the energy is used to increase the lattice vibrations (an amount that is the same for a system in the normal and in the superconducting state), and the remainder is used to increase the energy of the conduction electrons. The electronic specific heat (Ce) of the electrons is defined as the ratio of that portion of the heat used by the electrons to the rise in temperature of the system. Figure 1—> shows how the specific heat of the electrons in a superconductor varies with the absolute temperature (T ) in the normal and in the superconducting state. It is evident from the figure that the electronic specific heat in the superconducting state (designated Ces) is smaller than in the normal state (designated Cen) at low enough temperatures but that Ces becomes larger than Cen as the transition temperature Tc is approached, at which point it drops abruptly to Cen for the classic superconductors, although the curve has a cusp shape near Tc for the high-Tc superconductors. Precise measurements have indicated that, at temperatures considerably below the transition temperature, the logarithm of the electronic specific heat is inversely proportional to the temperature. This temperature dependence, together with the principles of statistical mechanics, strongly suggests that there is a gap in the distribution of energy levels available to the electrons in a superconductor, so that a minimum energy is required for the excitation of each electron from a state below the gap to a state above the gap. Some of the high-Tc superconductors provide an additional contribution to the specific heat, which is proportional to the temperature. This behaviour indicates that there are electronic states lying at low energy; additional evidence of such states is obtained from optical properties and tunneling measurements.

      The heat flow per unit area of a sample equals the product of the thermal conductivity (K) and the temperature gradient △T: JQ = -KT, the minus sign indicating that heat always flows from a warmer to a colder region of a substance.

      The thermal conductivity in the normal state (Kn) approaches the thermal conductivity in the superconducting state (Ks) as the temperature (T ) approaches the transition temperature (Tc) for all materials, whether they are pure or impure. This suggests that the energy gap (Δ) for each electron approaches zero as the temperature (T ) approaches the transition temperature (Tc). This would also account for the fact that the electronic specific heat in the superconducting state (Ces) is higher than in the normal state (Cen) near the transition temperature: as the temperature is raised toward the transition temperature (Tc), the energy gap in the superconducting state decreases, the number of thermally excited electrons increases, and this requires the absorption of heat.

Energy gaps
      As stated above, the thermal properties of superconductors indicate that there is a gap in the distribution of energy levels available to the electrons, and so a finite amount of energy, designated as delta (Δ), must be supplied to an electron to excite it. This energy is maximum (designated Δ0) at absolute zero and changes little with increase of temperature until the transition temperature is approached, where Δ decreases to zero, its value in the normal state. The BCS theory predicts an energy gap with just this type of temperature dependence.

      According to the BCS theory, there is a type of electron pairing (electrons of opposite spin acting in unison) in the superconductor that is important in interpreting many superconducting phenomena. The electron pairs, called Cooper pairs, are broken up as the superconductor is heated. Each time a pair is broken, an amount of energy that is at least as much as the energy gap (Δ) must be supplied to each of the two electrons in the pair, so an energy at least twice as great (2Δ) must be supplied to the superconductor. The value of twice the energy gap at 0 K (which is 2Δ0) might be assumed to be higher when the transition temperature of the superconductor is higher. In fact, the BCS theory predicts a relation of this type—namely, that the energy supplied to the superconductor at absolute zero would be 2Δ0 = 3.53 kTc, where k is Boltzmann's constant (1.38 × 10−23 joule per kelvin). In the high-Tc cuprate compounds, values of 2Δ0 range from approximately three to eight multiplied by kTc.

      The energy gap (Δ) can be measured most precisely in a tunneling experiment (a process in quantum mechanics that allows an electron to escape from a metal without acquiring the energy required along the way according to the laws of classical physics). In this experiment, a thin insulating junction is prepared between a superconductor and another metal, assumed here to be in the normal state. In this situation, electrons can quantum mechanically tunnel from the normal metal to the superconductor if they have sufficient energy. This energy can be supplied by applying a negative voltage (V) to the normal metal, with respect to the voltage of the superconductor.

      Tunneling will occur if eV—the product of the electron charge, e (−1.60 × 10−19 coulomb), and the voltage—is at least as large as the energy gap Δ. The current flowing between the two sides of the junction is small up to a voltage equal to V = Δ/e, but then it rises sharply. This provides an experimental determination of the energy gap (Δ). In describing this experiment it is assumed here that the tunneling electrons must get their energy from the applied voltage rather than from thermal excitation.

Magnetic and electromagnetic properties of superconductors

Critical field
      One of the ways in which a superconductor can be forced into the normal state is by applying a magnetic field. The weakest magnetic field that will cause this transition is called the critical field (Hc) if the sample is in the form of a long, thin cylinder or ellipsoid and the field is oriented parallel to the long axis of the sample. (In other configurations the sample goes from the superconducting state into an intermediate state, in which some regions are normal and others are superconducting, and finally into the normal state.) The critical field increases with decreasing temperature. For the superconducting elements, its values (H0) at absolute zero range from 1.1 oersted for tungsten to 830 oersteds for tantalum.

      These remarks about the critical field apply to ordinary (so-called type I) superconductors. In the following section the behaviour of other (type II) superconductors is examined.

      As was stated above, a type I superconductor in the form of a long, thin cylinder or ellipsoid remains superconducting at a fixed temperature as an axially oriented magnetic field is applied, provided the applied field does not exceed a critical value (Hc). Under these conditions, superconductors exclude the magnetic field from their interior, as could be predicted from the laws of electromagnetism and the fact that the superconductor has no electric resistance. A more astonishing effect occurs if the magnetic field is applied in the same way to the same type of sample at a temperature above the transition temperature and is then held at a fixed value while the sample is cooled. It is found that the sample expels the magnetic flux as it becomes superconducting. This is called the Meissner effect. Complete expulsion of the magnetic flux (a complete Meissner effect) occurs in this way for certain superconductors, called type I superconductors, but only for samples that have the described geometry. For samples of other shapes, including hollow structures, some of the magnetic flux can be trapped, producing an incomplete or partial Meissner effect.

 Type II superconductors have a different magnetic behaviour. Examples of materials of this type are niobium and vanadium (the only type II superconductors among the chemical elements) and some alloys and compounds, including the high-Tc compounds. As a sample of this type, in the form of a long, thin cylinder or ellipsoid, is exposed to a decreasing magnetic field that is axially oriented with the sample, the increase of magnetization, instead of occurring suddenly at the critical field (Hc), sets in gradually. Beginning at the upper critical field (Hc2), it is completed at a lower critical field (Hc1; see Figure 2—>). If the sample is of some other shape, is hollow, or is inhomogeneous or strained, some magnetic flux remains trapped, and some magnetization of the sample remains after the applied field is completely removed. Known values of the upper critical field extend up to 6 × 105 oersteds, the value for the compound of lead, molybdenum, and sulfur with formula PbMo6S8.

      The expulsion of magnetic flux by type I superconductors in fields below the critical field (Hc) or by type II superconductors in fields below Hc1 is never quite as complete as has been stated in this simplified presentation, because the field always penetrates into a sample for a small distance, known as the electromagnetic penetration depth. Values of the penetration depth for the superconducting elements at low temperature lie in the range from about 390 to 1,300 angstroms. As the temperature approaches the critical temperature, the penetration depth becomes extremely large.

High-frequency electromagnetic properties
      The foregoing descriptions have pertained to the behaviour of superconductors in the absence of electromagnetic fields or in the presence of steady or slowly varying fields; the properties of superconductors in the presence of high-frequency electromagnetic fields, however, have also been studied.

      The energy gap in a superconductor has a direct effect on the absorption of electromagnetic radiation. At low temperatures, at which a negligible fraction of the electrons are thermally excited to states above the gap, the superconductor can absorb energy only in a quantized amount that is at least twice the gap energy (at absolute zero, 2Δ0). In the absorption process, a photon (a quantum of electromagnetic energy) is absorbed, and a Cooper pair is broken; both electrons in the pair become excited. The photon's energy (E) is related to its frequency (ν) by the Planck relation, E = hν, in which h is Planck's constant (6.63 × 10−34 joule second). Hence the superconductor can absorb electromagnetic energy only for frequencies at least as large as 2Δ0/h.

Magnetic-flux quantization
      The laws of quantum mechanics dictate that electrons have wave properties and that the properties of an electron can be summed up in what is called a wave function. If several wave functions are in phase (i.e., act in unison), they are said to be coherent. The theory of superconductivity indicates that there is a single, coherent, quantum mechanical wave function that determines the behaviour of all the superconducting electrons. As a consequence, a direct relationship can be shown to exist between the velocity of these electrons and the magnetic flux (Φ) enclosed within any closed path inside the superconductor. Indeed, inasmuch as the magnetic flux arises because of the motion of the electrons, the magnetic flux can be shown to be quantized; i.e., the intensity of this trapped flux can change only by units of Planck's constant divided by twice the electron charge.

      When a magnetic field enters a type II superconductor (in an applied field between the lower and upper critical fields, Hc1 and Hc2), it does so in the form of quantized fluxoids, each carrying one quantum of flux. These fluxoids tend to arrange themselves in regular patterns that have been detected by electron microscopy and by neutron diffraction. If a large enough current is passed through the superconductor, the fluxoids move. This motion leads to energy dissipation that can heat the superconductor and drive it into the normal state. The maximum current per unit area that a superconductor can carry without being forced into the normal state is called the critical current density (Jc). In making wire for superconducting high-field magnets, manufacturers try to fix the positions of the fluxoids by making the wire inhomogeneous in composition.

Josephson currents
      If two superconductors are separated by an insulating film that forms a low-resistance junction between them, it is found that Cooper pairs can tunnel from one side of the junction to the other. (This process occurs in addition to the single-particle tunneling already described.) Thus, a flow of electrons, called the Josephson current, is generated and is intimately related to the phases of the coherent quantum mechanical wave function for all the superconducting electrons on the two sides of the junction. It was predicted that several novel phenomena should be observable, and experiments have demonstrated them. These are collectively called the Josephson effect or effects.

 The first of these phenomena is the passage of current through the junction in the absence of a voltage across the junction. The maximum current that can flow at zero voltage depends on the magnetic flux (Φ) passing through the junction as a result of the magnetic field generated by currents in the junction and elsewhere. The dependence of the maximum zero-voltage current on the magnetic field applied to a junction between two superconductors is shown in Figure 3—>.

      A second type of Josephson effect is an oscillating current resulting from a relation between the voltage across the junction and the frequency (ν) of the currents associated with Cooper pairs passing through the junction. The frequency (ν) of this Josephson current is given by ν = 2eV/h, where e is the charge of the electron. Thus, the frequency increases by 4.84 × 1014 hertz (cycles per second) for each additional volt applied to the junction. This effect can be demonstrated in various ways. The voltage can be established with a source of direct-current (DC) power, for instance, and the oscillating current can be detected by the electromagnetic radiation of frequency (ν) that it generates. Another method is to expose the junction to radiation of another frequency (ν′) generated externally. It is found that a graph of the DC current versus voltage has current steps at values of the voltage corresponding to Josephson frequencies that are integral multiples (n) of the external frequency (ν = nν′); that is, V = nhν′/2e. The observation of current steps of this type has made it possible to measure h/e with far greater precision than by any other method and has therefore contributed to a knowledge of the fundamental constants of nature.

      The Josephson effect has been used in the invention of novel devices for extremely high-sensitivity measurements of currents, voltages, and magnetic fields.

Higher-temperature superconductivity

Discovery and composition of high-temperature superconductors
      Ever since Kamerlingh Onnes discovered that mercury becomes superconducting at temperatures less than 4 K, scientists have been searching for superconducting materials with higher transition temperatures. Until 1986 a compound of niobium and germanium (Nb3Ge) had the highest known transition temperature, 23 K, less than a 20-degree increase in 75 years. Most researchers expected that the next increase in transition temperature would be found in a similar metallic alloy and that the rise would be only one or two degrees. In 1986, however, the Swiss physicist Karl Alex Müller (Müller, Karl Alex) and his West German associate, Johannes Georg Bednorz (Bednorz, J. Georg), discovered, after a three-year search among metal oxides, a material that had an unprecedentedly high transition temperature of about 30 K. They were awarded the Nobel Prize for Physics in 1987, and their discovery immediately stimulated groups of investigators in China, Japan, and the United States to produce superconducting oxides with even higher transition temperatures.

 These high-temperature superconductors are ceramics. They contain lanthanum, yttrium, or another of the rare-earth elements or bismuth or thallium; usually barium or strontium (both alkaline-earth elements); copper; and oxygen. Other atomic species can sometimes be introduced by chemical substitution while retaining the high-Tc properties. The superconducting transition temperatures of some of the high-Tc materials are listed in the Table—>. The value 134 K is the highest known Tc value. The compounds given in the table are members of the major families of high-Tc superconductors. Within each family, only the subscripts (i.e., stoichiometry) vary from one compound to another. The compounds listed have the highest observed superconducting transition temperature in their respective families. Samples in the families containing bismuth or thallium always exhibit a great deal of atomic disorder, with atoms in the “wrong” crystallographic sites and with impurity phases. It is possible that such disorder is required to make these compounds thermodynamically stable.

Structures and properties
      The compounds have crystal structures containing planes of Cu and O atoms, and some also have chains of Cu and O atoms. The roles played by these planes and chains have come under intense investigation. Varying the oxygen content or the heat treatment of the materials dramatically changes their transition temperatures, critical magnetic fields, and other properties. Single crystals of the high-temperature superconductors are very anisotropic—i.e., their properties associated with a direction, such as the critical fields or the critical current density, are highly dependent on the angle between that direction and the rows of atoms in the crystal.

      If the number of superconducting electrons per unit volume is locally disturbed by an applied force (typically electric or magnetic), this disturbance propagates for a certain distance in the material; the distance is called the superconducting coherence length (or Ginzburg-Landau coherence length), ξ. If a material has a superconducting region and a normal region, many of the superconducting properties disappear gradually—over a distance ξ—upon traveling from the former to the latter region. In the pure (i.e., undoped) classic superconductors ξ is on the order of a few thousand angstroms, but in the high-Tc superconductors it is on the order of 1 to 10 angstroms. The small size of ξ affects the thermodynamic and electromagnetic properties of the high-Tc superconductors. For example, it is responsible for the cusp shape of the specific heat curve near Tc that was mentioned above. It is also responsible for the ability of the high-Tc superconductors to remain superconducting in extraordinarily large fields—on the order of 1,000,000 gauss (100 teslas)—at low temperatures.

      The high-Tc superconductors are type II superconductors. They exhibit zero resistance, strong diamagnetism, the Meissner effect, magnetic flux quantization, the Josephson effects, an electromagnetic penetration depth, an energy gap for the superconducting electrons, and the characteristic temperature dependencies of the specific heat and the thermal conductivity that are described above. Therefore, it is clear that the conduction electrons in these materials form the Cooper pairs used to explain superconductivity in the BCS theory. Thus, the central conclusions of the BCS theory are demonstrated. Indeed, that theory guided Bednorz and Müller in their search for high-temperature superconductors. It is not known, however, why the transition temperatures of these oxides are so high. It was generally believed that the members of a Cooper pair are bound together because of interactions between the electrons and the lattice vibrations (phonons), but it is unlikely that these interactions are strong enough to explain transition temperatures as high as 90 K. Most experts believe that interactions among the electrons generate high-temperature superconductivity. The details of this interaction are difficult to treat theoretically because the motions of the electrons are strongly correlated with each other and because magnetic phenomena play an important part in determining the microscopic properties of these materials. These strong correlations and magnetic properties may be responsible for unusual temperature dependencies of the electric resistivity ρ and Hall coefficient RH in the normal state (i.e., above Tc). (For a discussion of the Hall effect, see magnetism: Magnetic forces (magnetism).) It is observed that at temperatures above Tc the electric resistivity, although higher for superconductors than for typical metals in the normal state, is roughly proportional to the temperature T, an unusually weak temperature dependence. Measurements of RH show it to be significantly temperature-dependent in the normal state (sometimes proportional to 1/T) rather than being roughly independent of T, which is the case for ordinary materials.

Applications
      Films of the new materials can carry currents in the superconducting state that are large enough to be of importance in making many devices. Possible applications of the high-temperature superconductors in thin-film or bulk form include the construction of computer parts (logic devices, memory elements, switches, and interconnects), oscillators, amplifiers, particle accelerators, highly sensitive devices for measuring magnetic fields, voltages or currents, magnets for medical magnetic-imaging devices, magnetic energy-storage systems, levitated passenger trains for high-speed travel, motors, generators, transformers, and transmission lines. The principal advantages of these superconducting devices would be their low power dissipation, high operating speed, and extreme sensitivity.

      Equipment made with the high-temperature superconductors would also be more economical to operate because such materials can be cooled with inexpensive liquid nitrogen (boiling point, 77 K) rather than with costly liquid helium (boiling point, 4.2 K). The ceramics have problems, however, which must be overcome before useful devices can be made from them. These problems include brittleness, instabilities of the materials in some chemical environments, and a tendency for impurities to segregate at surfaces and grain boundaries, where they interfere with the flow of high currents in the superconducting state.

Donald M. Ginsberg

Additional Reading
General discussions of low-temperature physics include the nontechnical summaries F.E. Simon et al., Low Temperature Physics: Four Lectures (1952, reissued 1961); K. Mendelssohn, The Quest for Absolute Zero: The Meaning of Low Temperature Physics, 2nd ed. (1977); and Tom Schachtman, Absolute Zero and the Conquest of Cold (1999). Works at a more-advanced level include the classic presentation of the theory, Fritz London, Superfluids: Macroscopic Theory of Superconductivity, 2nd rev. ed., vol. 1 (1961), and Macroscopic Theory of Superfluid Helium, vol. 2 (1964); and the more recent works David R. Tilley and John Tilley, Superfluidity and Superconductivity, 3rd ed. (1990, reissued 1994); A.S. Alexandrov and Nevill Mott, High Temperature Superconductors and Other Superfluids (1995); and T. Tsuneto, Superconductivity and Superfluidity, trans. from Japanese by Mikio Nakahara (1998).Works dealing more specifically with superconductivity include two introductory reviews with descriptive bibliographies from the American Journal of Physics, both by D.M. Ginsberg, “Resource Letter Scy-1 on Superconductivity,” 32(2):85–89 (February 1964), and “Resource Letter Scy-2 on Superconductivity,” 38(8):949–955 (August 1970). Paul H.E. Meijer, “Kamerlingh Onnes and the Discovery of Superconductivity,” American Journal of Physics, 57(1):17–34 (December 1994), is an account of the first observation of superconductivity. A popular nontechnical account of high-temperature superconductivity is given in Robert M. Hazen, The Breakthrough: The Race for the Superconductor (1988).Donald M. Ginsberg

* * *


Universalium. 2010.

Игры ⚽ Нужно решить контрольную?

Look at other dictionaries:

  • Superconductivity — is a phenomenon occurring in certain materials generally at very low temperatures, characterized by exactly zero electrical resistance and the exclusion of the interior magnetic field (the Meissner effect).The electrical resistivity of a metallic …   Wikipedia

  • Superconductivity — Superconductivity. См. Сверхпроводимость. (Источник: «Металлы и сплавы. Справочник.» Под редакцией Ю.П. Солнцева; НПО Профессионал , НПО Мир и семья ; Санкт Петербург, 2003 г.) …   Словарь металлургических терминов

  • superconductivity —  Superconductivity  Сверхпроводимость   Свойство некоторых материалов обладать нулевым электрическим сопротивлением ниже определённой температуры. Существует множество чистых элементов, сплавов и керамик, переходящих в сверхпроводящее состояние.… …   Толковый англо-русский словарь по нанотехнологии. - М.

  • superconductivity — ► NOUN Physics ▪ the property of zero electrical resistance in some substances at very low temperatures. DERIVATIVES superconducting adjective superconductive adjective superconductor noun …   English terms dictionary

  • superconductivity — [so͞o΄pərkən duk′shənso͞o΄pərkän΄dək tiv′ə tē] n. Physics the phenomenon, exhibited by various metals, alloys, and compounds, of conducting electrical current without resistance when cooled to low temperatures: sometimes called superconduction… …   English World dictionary

  • superconductivity — noun The property of a material whereby it has no resistance to the flow of an electric current. Currently, superconductivity can only be achieved at extremely low temperatures …   Wiktionary

  • superconductivity — [[t]su͟ːpə(r)kɒndʌktɪ̱vɪti[/t]] N UNCOUNT Superconductivity is the ability of certain metals to allow electricity to pass through them without any resistance at very low temperatures. [TECHNICAL] …   English dictionary

  • superconductivity — superlaidumas statusas T sritis fizika atitikmenys: angl. superconductivity vok. Supraleitfähigkeit, f; Supraleitung, f rus. сверхпроводимость, f pranc. supraconduction, f …   Fizikos terminų žodynas

  • superconductivity — noun Date: 1913 a complete disappearance of electrical resistance in a substance especially at very low temperatures • superconductor noun …   New Collegiate Dictionary

  • superconductivity — Смотри Сверхпроводимость …   Энциклопедический словарь по металлургии

Share the article and excerpts

Direct link
Do a right-click on the link above
and select “Copy Link”