You are on page 1of 2

NEWS & VIEWS

SUPERCONDUCTIVITY

Death of a Fermi surface


Every metal, semimetal and doped semiconductor has a Fermi surface that determines its physical properties. A new state of matter within the pseudogap state of a high-temperature superconductor destroys the Fermi surface, the process of which provides information about the new state.

KYLE MCELROY
is in the Materials Sciences Division, Berkeley National Laboratory, Berkeley, California 94720, USA. e-mail: kpmcelroy@lbl.gov
T*

aterials with strong electronic correlations are breeding grounds for new states of matter the copper oxide superconductors1 being a perfect example. Although the parent (nonsuperconducting) oxide compound is expected to be a metal, Coulomb repulsion between electrons leads to localization and insulating behaviour. Such insulators, in which correlation effects dominate, are known as Mott insulators. Upon chemical doping, which adds charge carriers, many strange states emerge from this Mott insulator. Among these are not only superconductivity with incredibly high critical temperatures of up to 150 K, but, for low doping concentration, a most mysterious state of matter known as the (much-debated) pseudogap phase. On page 447 of this issue2, Kanigel et al. bolster one view of the pseudogap state, as a precursor to superconductivity, by carefully showing that it has a simple phenomenology that, when taken in the low-temperature limit, looks very similar to the superconducting state. By varying the hole concentration and temperature, researchers have found a wide variety of unexpected phenomena in the copper oxide superconductors (Fig. 1). The states close to the Mott insulator region that is, for low hole density are particularly intriguing. In this regime, called the pseudogap state, indications of a most unusual state of matter have been found. This regime differs from those in conventional conductors in the behaviour of electronic states right near the Fermi surface. In most materials, like in the non-interacting electron gas, electrons simply fill up the lowest available electronic states in momentum space. The top of this filled volume creates a surface of states called the Fermi surface that governs the low-energy (room temperature) physics. The situation is different in the pseudogap state: below the temperature T* an energy gap opens up on part of the Fermi surface3, removing states from this low-energy region. Although the development of such a gap often indicates a breaking of some symmetry by a new

Mott insulator

Pseudogap Metal Superconductor

1/2 filling

Hole doping (p)

electronic phase, in this case the gap consumes only part of the Fermi surface (hence the pseudo in its name) and leaves behind small, disconnected Fermi arcs4 (as shown in Fig. 2) where low-energy, metallic excitations seem to survive. Despite previous work, the identity of this phase remains unclear. On the one hand, a completely new electronic state may exist inside the pseudogap phase, eating up many of the states and therefore leaving few available to pair up and superconduct. On the other hand, the similarity between this partial gap and that in the superconducting state has led many to the conclusion that the partial gap results from preformed Cooper pairs (above Tc). According to this theory, at low enough temperatures these pairs will condense into a collective superconducting state, but above this temperature in the pseudogap state fluctuations in the phase of the wavefunctions prevent the condensation5. Hence, the electronic states should look similar to the superconducting ones except that they lack the hallmark of zero resistance. Recently, transport studies6 of the very-low-energy states near the lowest hole-doping levels in YBa2Cu3O6.33 that still superconduct have supported this conclusion. Across the critical doping range bridging the pseudogap and the superconductivity (marked by the red ellipse in Fig. 1), the thermal conductivity behaves in a continuous manner, suggesting little difference between the states in the pseudogap and those in the superconductor.

Figure 1 Phase diagram of a typical hole-doped copper oxide superconductor. As the number of holes is increased from the half-lled Mott insulator (one electron per state) an assortment of electronic states appear. At high doping these include both superconducting and metallic phases. At low doping other strange states of matter can be found, including the pseudogap state that persists up to T *. The red ellipse indicates previous work connecting the pseudogap to the superconducting state.

nature physics | VOL 2 | JULY 2006 | www.nature.com/naturephysics

441
2006 Nature Publishing Group

NEWS & VIEWS


Figure 2 The square Brillouin zone of a copper oxide superconductor. The normal state Fermi surface above T*, is characterized by straight (dashed) and curved sections (green), making pockets centred on the corners of the zone. In the pseudogap phase, the low-energy states occupy ungapped regions that exist over four small, disconnected metallic arcs (green) known as Fermi arcs. Within this state the gap slowly engulfs the states (black arrows) as the temperature is lowered, leaving only the four blue points ungapped. The straight sections of the Fermi surface (shaded ellipses), which share a connecting wave vector q* (orange arrows), behave differently their gap abruptly opens all at once, as if another electronic state is gapping them.
Zone face

Fermi arc

q*

In their latest work, Kanigel et al. study the electronic states throughout the pseudogap phase to characterize the shape of the gap. To do this, the authors use angle-resolved photoemission spectroscopy to carefully map out the electronic states of the pseudogap throughout the Brillouin zone of Bi2Sr2CaCu2O8+x. By following where these states make their closest approach to the Fermi surface, they are able to trace how the gap in this state forms below T* and quickly consumes the states starting near the flat faces of the zone (Fig. 2), leaving only the Fermi arcs that become shorter and shorter as the temperature is lowered. An extrapolation to zero temperature shows that eventually the pseudogap engulfs the whole Fermi surface, and the disconnected Fermi arcs shrink down to just four points along the zone diagonals. Interestingly,

the shape of this gap is the same as that of the superconducting state found at higher dopings and lower temperatures. So throughout the pseudogap phase, the electronic states on the Fermi arcs look very much like they are preparing, and indeed prepairing, themselves for the superconducting state. Along the curved parts of the Fermi surface, the pseudogap structure seems to be a precursor to superconductivity. But near the zone face (Fig. 2) the straight parts of the Fermi surface reveal something different. Along these sections the arcs do not continue to grow as the temperature increases. Instead, the missing electronic states begin to fill in along these sections until the pseudogap transition is reached, where the gap abruptly closes. This qualitatively different behaviour seems to indicate that as the Fermi arcs are getting ready for superconductivity, these straight segments are doing something entirely different. Significantly, it is exactly these straight segments that interact most strongly with other degrees of freedom, and may lead to other states of matter. Additionally, their parallel nature allows them to interact more strongly with a particular wave vector that connects them, q* (orange arrows in Fig. 2), thereby opening a gap by breaking another symmetry, such as translation. So perhaps, while the electrons along the Fermi arcs are preparing for superconductivity, yet another state of matter for instance, one with charge ordering7 lives along these straight segments inside the pseudogap.
REFERENCES
1. 2. 3. 4. 5. 6. 7. Bednorz, J. G. & Mller, K. A. Z. Phys. B 64, 189193 (1986). Kanigel, A. et al. Nature Phys. 2, 447451 (2006). Timusk, T. & Statt, B. Rep. Prog. Phys. 62, 61122 (1999). Norman, M. R. et al. Nature 392, 157160 (1998). Nagaosa, N. & Lee, P. A. Phys. Rev. B 45, 966970 (1992). Sutherland, M. et al. Phys. Rev. Lett. 94, 147004 (2005). Hanaguri, T. et al. Nature 430, 10011005 (2004).

442
2006 Nature Publishing Group

nature physics | VOL 2 | JULY 2006 | www.nature.com/naturephysics

You might also like