Interstitial defect

From Wikipedia, the free encyclopedia
Interstitial atoms (blue) occupy some of the spaces within a lattice of larger atoms (red)

An interstitial defect is a type of point crystallographic defect where an atom of the same or of a different type, occupies a normally unoccupied site in the crystal structure. When the atom is of the same type as those already present they are known as a self-interstitial defect. Alternatively, small atoms in some crystals may occupy interstitial sites, such as hydrogen in palladium. Interstitials can be produced by bombarding a crystal with elementary particles having energy above the displacement threshold for that crystal, but they may also exist in small concentrations in thermodynamic equilibrium. The presence of interstitial defects can modify the physical and chemical properties of a material.

Self-interstitials[]

Self-interstitial defects are interstitial defects which contain only atoms which are the same as those already present in the lattice.

Structure of self-interstitial in some common metals. The left-hand side of each crystal type shows the perfect crystal and the right-hand side the one with a defect.

The structure of interstitial defects has been experimentally determined in some metals and semiconductors.

Contrary to what one might intuitively expect, most self-interstitials in metals with a known structure have a 'split' structure, in which two atoms share the same lattice site.[1][2] Typically the center of mass of the two atoms is at the lattice site, and they are displaced symmetrically from it along one of the principal lattice directions. For instance, in several common face-centered cubic (fcc) metals such as copper, nickel and platinum, the ground state structure of the self-interstitial is the split [100] interstitial structure, where two atoms are displaced in a positive and negative [100] direction from the lattice site. In body-centered cubic (bcc) iron the ground state interstitial structure is similarly a [110] split interstitial.

These split interstitials are often called dumbbell interstitials, because plotting the two atoms forming the interstitial with two large spheres and a thick line joining them makes the structure resemble a dumbbell weight-lifting device.

In other bcc metals than iron, the ground state structure is believed based on recent density-functional theory calculations to be the [111] crowdion interstitial,[3] which can be understood as a long chain (typically some 10–20) of atoms along the [111] lattice direction, compressed compared to the perfect lattice such that the chain contains one extra atom.

Structure of dumbbell self-interstitial in silicon. Note that the structure of the interstitial in silicon may depend on charge state and doping level of the material.

In semiconductors the situation is more complex, since defects may be charged and different charge states may have different structures. For instance, in silicon, the interstitial may either have a split [110] structure or a tetrahedral truly interstitial one.[4]

Carbon, notably in graphite and diamond, has a number of interesting self-interstitials - recently discovered using Local-density approximation-calculations is the "spiro-interestitial" in graphite, named after spiropentane, as the interstitial carbon atom is situated between two basal planes and bonded in a geometry similar to spiropentane.[5]

Impurity interstitials[]

Small impurity interstitial atoms are usually on true off-lattice sites between the lattice atoms. Such sites can be characterized by the symmetry of the interstitial atom position with respect to its nearest lattice atoms. For instance, an impurity atom I with 4 nearest lattice atom A neighbours (at equal distances) in an fcc lattice is in a tetrahedral symmetry position, and thus can be called a tetrahedral interstitial.

Large impurity interstitials can also be in split interstitial configurations together with a lattice atom, similar to those of the self-interstitial atom.

Octahedral (red) and tetrahedral (blue) interstitial symmetry polyhedra in a face-centered cubic lattice. The actual interstitial atom would ideally be in the middle of one of the polyhedra.

Effects of interstitials[]

Interstitials modify the physical and chemical properties of materials.

  • Interstitial carbon atoms have a crucial role for the properties and processing of steels, in particular carbon steels.
  • Impurity interstitials can be used e.g. for storage of hydrogen in metals.
  • The crystal lattice can expand with the concentration of impurity interstitials
  • The amorphization of semiconductors such as silicon during ion irradiation is often explained by the build up of a high concentration of interstitials leading eventually to the collapse of the lattice as it becomes unstable.[6][7]
  • Creation of large amounts of interstitials in a solid can lead to a significant energy buildup, which on release can even lead to severe accidents in certain old types of nuclear reactors (Wigner effect). The high-energy states can be released by annealing.
  • At least in fcc lattice, interstitials have a large diaelastic softening effect on the material.[8]
  • It has been proposed that interstitials are related to the onset of melting and the glass transition.[9][10][11]

References[]

  1. ^ Ehrhart, P. (1991) Properties and interactions of atomic defects in metals and alloys, H. Ullmaier (ed.), Landolt-Börnstein, New Series III vol. 25 ch. 2, pp. 88 ff. Springer, Berlin.
  2. ^ Schilling, W. (1978). "Self-interstitial atoms in metals". Journal of Nuclear Materials. 69–70: 465. Bibcode:1978JNuM...69..465S. doi:10.1016/0022-3115(78)90261-1.
  3. ^ Derlet, P. M.; D. Nguyen-Manh; S. L. Dudarev (2007). "Multiscale modeling of crowdion and vacancy defects in body-centered-cubic transition metals". Phys. Rev. B. 76 (5): 054107. Bibcode:2007PhRvB..76e4107D. doi:10.1103/physrevb.76.054107.
  4. ^ Watkins, G. D. (1991) "Native defects and their interactions with impurities in silicon", p. 139 in Defects and Diffusion in Silicon Processing, T. Diaz de la Rubia, S. Coffa, P. A. Stolk and C. S. Rafferty (eds.), MRS Symposium Proceedings vol. 469. Materials Research Society, Pittsburg.
  5. ^ Heggie, M.; Eggen, B.R.; Ewels, C.P.; et al. (1998). "LDF calculations of point defects in graphites and fullerenes". Electrochem Soc Proc. 98 (?): 60.
  6. ^ Seidman, D. N.; Averback, R. S.; Okamoto, P. R.; Baily, A. C. (1987). "Amorphization Processes in Electron- and/or Ion-Irradiated Silicon" (PDF). Phys. Rev. Lett. 58 (9): 900–903. Bibcode:1987PhRvL..58..900S. doi:10.1103/PhysRevLett.58.900. PMID 10035067.
  7. ^ Cerofilini, G. F.; Meda, L.; Volpones, C. (1988). "A model for damage release in ion-implanted silicon". J. Appl. Phys. 63 (10): 4911. Bibcode:1988JAP....63.4911C. doi:10.1063/1.340432.
  8. ^ Rehn, L. E.; Holder, J.; Granato, A. V.; Coltman, R. R.; Young, J. F. W. (1974). "Effects of thermal-neutron irradiation on the elastic constants of copper". Phys. Rev. B. 10 (2): 349. Bibcode:1974PhRvB..10..349R. doi:10.1103/PhysRevB.10.349.
  9. ^ Granato, A. V. (1992). "Interstitialcy Model for Condensed Matter States of Face-Centered-Cubic Metals". Phys. Rev. Lett. 68 (7): 974–977. Bibcode:1992PhRvL..68..974G. doi:10.1103/PhysRevLett.68.974. PMID 10046046.
  10. ^ Forsblom, M.; Grimvall, G. (2005). "Homogeneous melting of superheated crystals: Molecular dynamics simulations". Phys. Rev. B. 72 (5): 054107. Bibcode:2005PhRvB..72e4107F. doi:10.1103/PhysRevB.72.054107.
  11. ^ Nordlund, K.; Ashkenazy, Y.; Averback, R. S.; Granato, A. V. (2005). "Strings and interstitials in liquids, glasses and crystals" (PDF). Europhys. Lett. 71 (4): 625. Bibcode:2005EL.....71..625N. doi:10.1209/epl/i2005-10132-1.
Retrieved from ""