Cottrell atmosphere

Last updated
A carbon atom below a dislocation in iron, forming a Cottrell atmosphere CottrellAtmosphere.png
A carbon atom below a dislocation in iron, forming a Cottrell atmosphere

In materials science, the concept of the Cottrell atmosphere was introduced by A. H. Cottrell and B. A. Bilby in 1949 [1] to explain how dislocations are pinned in some metals by boron, carbon, or nitrogen interstitials.

Contents

Cottrell atmospheres occur in body-centered cubic (BCC) and face-centered cubic (FCC) materials, such as iron or nickel, with small impurity atoms, such as boron, [2] carbon, [3] or nitrogen.[ citation needed ] As these interstitial atoms distort the lattice slightly, there will be an associated residual stress field surrounding the interstitial. This stress field can be relaxed by the interstitial atom diffusing towards a dislocation[ citation needed ], which contains a small gap at its core (as it is a more open structure), see Figure 1. Once the atom has diffused into the dislocation core the atom will stay. Typically only one interstitial atom is required per lattice plane of the dislocation.[ citation needed ] The collection of solute atoms around the dislocation core due to this process is the Cottrell atmosphere.

Influence on Mechanical Behavior

A dislocation moving with a Cottrell Atmosphere around it. At high stresses (top), the dislocation can "break free" of the atmosphere, while at low stresses (bottom), the dislocation must drag the solutes with it, and motion is much slower. CottrellAtmosphereDrag.gif
A dislocation moving with a Cottrell Atmosphere around it. At high stresses (top), the dislocation can "break free" of the atmosphere, while at low stresses (bottom), the dislocation must drag the solutes with it, and motion is much slower.

The collection of solute atoms at the dislocation relieves the stresses associated with the dislocation, which lowers the energy of the dislocation's presence. Thus, moving the dislocation out of this Cottrell atmosphere constitutes an increase in energy, so it is not favorable for the dislocation to move forward in the crystal. As a result, the dislocation is effectively pinned by the Cottrell atmosphere.

Once a dislocation has become pinned, a large force is required to unpin the dislocation prior the yielding, thus at room temperature, the dislocation will not get unpinned. [4] This produces an observed upper yield point in a stress–strain graph. Beyond the upper yield point, the pinned dislocation will act as Frank–Read source to generate new dislocations that are not pinned. These dislocations are free to move in the crystal, which results in a subsequent lower yield point, and the material will deform in a more plastic manner.

Leaving the sample to age, by holding it at room temperature for a few hours, enables the carbon atoms to rediffuse back to dislocation cores, resulting in a return of the upper yield point.

Cottrell atmospheres lead to formation of Lüders bands and large forces for deep drawing and forming large sheets, making them a hindrance to manufacture. Some steels are designed to remove the Cottrell atmosphere effect by removing all the interstitial atoms. Steels such as interstitial free steel are decarburized and small quantities of titanium are added to remove nitrogen.

The Cottrell atmosphere also has important consequences for material behavior at high homologous temperatures, i.e. when the material is experiencing creep conditions. Moving a dislocation with an associated Cottrell atmosphere introduces viscous drag, an effective frictional force that makes moving the dislocation more difficult [5] (and thus slowing plastic deformation). This drag force can be expressed according to the equation:

,

where is the diffusivity of the solute atom in the host material, is the atomic volume, is the velocity of the dislocation, is the diffusion flux density, and is the solute concentration. [5] The existence of the Cottrell atmosphere and the effects of viscous drag have been proven to be important in high temperature deformation at intermediate stresses, as well as contributing to the power-law breakdown regime. [6]

Similar phenomena

While the Cottrell atmosphere is a general effect, there are additional related mechanisms that occur under more specialized circumstances.

Suzuki effect

The Suzuki effect is characterized by the segregation of solutes to stacking fault defects. When dislocations in an FCC system split into two partial dislocations, a hexagonal close-packed (HCP) stacking fault is formed between the two partials. H. Suzuki predicted that the concentration of solute atoms at this boundary would differ from the bulk. Moving through this field of solute atoms would therefore produce a similar drag on dislocations as the Cottrell atmosphere. [7] Suzuki later observed such segregation in 1961. [8]  The Suzuki effect is often associated with adsorption of substitutional solute atoms to the stacking fault, but it has also been found to occur with interstitial atoms diffusing out of the stacking fault. [9]

Once two partial dislocations have split, they cannot cross-slip around obstacles anymore. Just as the Cottrell atmosphere provided a force against dislocation motion, the Suzuki effect in the stacking fault will lead to increased stresses for recombination of partials, leading to increased difficulty in bypassing obstacles (such as precipitates or particles), and therefore resulting in a stronger material.

Snoek effect

Under an applied stress, interstitial solute atoms, such as carbon and nitrogen can migrate within the α-Fe lattice, a BCC metal. These short-range migrations of carbon and nitrogen solute atoms result in an internal friction or an elastic effect, called the Snoek effect. The Snoek effect was discovered by J. L. Snoek in 1941. At room temperature, the solubility of carbon and nitrogen in solid solutions is exceedingly small. [10] By raising, the temperature beyond 400oC and cooling at a moderate rate, it is easy to keep a few hundredths of a percent of either element within the solution, while the remainder is supersaturated. [10] This revelation led to observed special magnetic phenomena in iron, mainly the presence of magnetism and time decrease of permeability due to small amount of carbon and nitrogen remaining in the iron. [10] Moreover, the additional presence of magnetism leads to an elastic-after effect. [11]

By preparing samples containing a larger amount of carbon or nitrogen in solid solution, magnetic and elastic phenomena are greatly enhanced. The solubility of nitrogen is much larger than the solubility of carbon in solid solution. [10] The study of the Snoek effect on annealed irons provides a reliable mechanism for calculating the solubility of carbon and nitrogen in α-iron. [12] A sample in a mixture of hydrogen and ammonia (or carbon monoxide) is mixed and heated until a stationary state was reached, where the mass of carbon and nitrogen taken up during the process can be found by estimating the changes in the weight of the sample. [10]

Carbon and nitrogen atoms occupy octahedral interstices at the midpoints of the cube edges and at the centers of the cube faces. [13] If a stress is applied a long the z, or [001] direction, the octahedral interstices along the x- and y-axes will contract, while the octahedral interstices along the z-direction expand. [13] Eventually, the interstitial atoms move to sites along the z-axis. [13] When the interstitial atoms move, this leads to a reduction in strain energy. In BCC metals, interstitial sites of an unstrained lattice are equally favorable. The interstitial solutes create elastic dipoles. [14] However, once a strain is applied on the lattice, such as that formed by a dislocation, 1/3 of the sites become more favorable than the other 2/3. Solute atoms will therefore move to occupy the favorable sites, forming a short ranged order of solutes immediately within the vicinity of the dislocation. [15] The motion of the interstitial solutes to these other sites constitutes a change in the elastic dipoles, so there is a relaxation time associated with this change which can be connected to the diffusivity and migration enthalpy of the solute atoms. [14] In the new, relaxed solute configuration, more energy is therefore required to break a dislocation from this order.

However, a stress applied in the [111] direction will not lead to any changes in the locations of the interstitial atoms as the three directions of the cube will be equally stressed, and on average, equally occupied by carbon atoms. [13] When a stress is applied along a cube edge and at an amount below the yield stress, the interstitial atom will lead to strain lagging before stress, showing the presence of internal friction. [13] A torsional pendulum is typically used as a means of studying this lagging effect. The angle of lag is taken to be δ and tan δ is considered a measure of internal friction. [13] The internal friction is expressed according to the equation:

Where the logarithmic decrement is the ratio of consecutive magnitudes of one cycle of the pendulum. [13] When the magnitude of one cycle decreases to of its original value in time , then the internal fraction behaves according to the equation:

Where is the vibrational frequency of the pendulum. [13]

The interstitials that occupy the normal sites in an unstressed lattice will promote internal friction. [13] Substituted solute atoms and interstitials in strain fields of a dislocation or at grain boundaries have their internal friction changed. [13] Therefore, the Snoek effect can measure carbon and nitrogen concentration in BCC alpha-Fe and other solutes present in ternary alloys. [16]

Materials

Materials in which dislocations described by Cottrell atmosphere include metals and semiconductor materials such silicon crystals.

Related Research Articles

Crystallographic defect Disruption of the periodicity of a crystal lattice

Crystallographic defects are interruptions of regular patterns in crystalline solids. They are common because positions of atoms or molecules at repeating fixed distances determined by the unit cell parameters in crystals, which exhibit a periodic crystal structure, are usually imperfect.

Neutron radiation Ionizing radiation that presents as free neutrons

Neutron radiation is a form of ionizing radiation that presents as free neutrons. Typical phenomena are nuclear fission or nuclear fusion causing the release of free neutrons, which then react with nuclei of other atoms to form new isotopes—which, in turn, may trigger further neutron radiation. Free neutrons are unstable, decaying into a proton, an electron, plus an electron antineutrino with a mean lifetime of 887 seconds.

Creep (deformation) Tendency of a solid material to move slowly or deform permanently under mechanical stress

In materials science, creep is the tendency of a solid material to move slowly or deform permanently under the influence of persistent mechanical stresses. It can occur as a result of long-term exposure to high levels of stress that are still below the yield strength of the material. Creep is more severe in materials that are subjected to heat for long periods and generally increases as they near their melting point.

Dislocation Linear crystallographic defect or irregularity

In materials science, a dislocation or Taylor's dislocation is a linear crystallographic defect or irregularity within a crystal structure that contains an abrupt change in the arrangement of atoms. The movement of dislocations allow atoms to slide over each other at low stress levels and is known as glide or slip. The crystalline order is restored on either side of a glide dislocation but the atoms on one side have moved by one position. The crystalline order is not fully restored with a partial dislocation. A dislocation defines the boundary between slipped and unslipped regions of material and as a result, must either form a complete loop, intersect other dislocations or defects, or extend to the edges of the crystal. A dislocation can be characterised by the distance and direction of movement it causes to atoms which is defined by the Burgers vector. Plastic deformation of a material occurs by the creation and movement of many dislocations. The number and arrangement of dislocations influences many of the properties of materials.

Work hardening Strengthening a material through plastic deformation

In materials science, work hardening, also known as strain hardening, is the strengthening of a metal or polymer by plastic deformation. Work hardening may be desirable, undesirable, or inconsequential, depending on the context.

Precipitation hardening, also called age hardening or particle hardening, is a heat treatment technique used to increase the yield strength of malleable materials, including most structural alloys of aluminium, magnesium, nickel, titanium, and some steels and stainless steels. In superalloys, it is known to cause yield strength anomaly providing excellent high-temperature strength.

Hardening is a metallurgical metalworking process used to increase the hardness of a metal. The hardness of a metal is directly proportional to the uniaxial yield stress at the location of the imposed strain. A harder metal will have a higher resistance to plastic deformation than a less hard metal.

The Portevin–Le Chatelier (PLC) effect describes a serrated stress–strain curve or jerky flow, which some materials exhibit as they undergo plastic deformation, specifically inhomogeneous deformation. This effect has been long associated with dynamic strain aging or the competition between diffusing solutes pinning dislocations and dislocations breaking free of this stoppage.

Yield (engineering) Phenomenon of deformation due to structural stress

In materials science and engineering, the yield point is the point on a stress-strain curve that indicates the limit of elastic behavior and the beginning of plastic behavior. Below the yield point, a material will deform elastically and will return to its original shape when the applied stress is removed. Once the yield point is passed, some fraction of the deformation will be permanent and non-reversible and is known as plastic deformation.

In metallurgy and materials science, annealing is a heat treatment that alters the physical and sometimes chemical properties of a material to increase its ductility and reduce its hardness, making it more workable. It involves heating a material above its recrystallization temperature, maintaining a suitable temperature for an appropriate amount of time and then cooling.

Slip (materials science)

In materials science, slip is the large displacement of one part of a crystal relative to another part along crystallographic planes and directions. Slip occurs by the passage of dislocations on close packed planes which are planes containing the greatest number of atoms per area and in close-packed directions. Close-packed planes are known as slip or glide planes. A slip system describes the set of symmetrically identical slip planes and associated family of slip directions for which dislocation motion can easily occur and lead to plastic deformation. The magnitude and direction of slip are represented by the Burgers vector.

Nanocrystalline material

A nanocrystalline (NC) material is a polycrystalline material with a crystallite size of only a few nanometers. These materials fill the gap between amorphous materials without any long range order and conventional coarse-grained materials. Definitions vary, but nanocrystalline material is commonly defined as a crystallite (grain) size below 100 nm. Grain sizes from 100–500 nm are typically considered "ultrafine" grains.

Solid solution strengthening is a type of alloying that can be used to improve the strength of a pure metal. The technique works by adding atoms of one element to the crystalline lattice of another element, forming a solid solution. The local nonuniformity in the lattice due to the alloying element makes plastic deformation more difficult by impeding dislocation motion through stress fields. In contrast, alloying beyond the solubility limit can form a second phase, leading to strengthening via other mechanisms.

Radiation damage is the effect of ionizing radiation on physical objects including non-living structural materials. It can be either detrimental or beneficial for materials.

A deformation mechanism, in geotechnical engineering, is a process occurring at a microscopic scale that is responsible for changes in a material's internal structure, shape and volume. The process involves planar discontinuity and/or displacement of atoms from their original position within a crystal lattice structure. These small changes are preserved in various microstructures of materials such as rocks, metals and plastics, and can be studied in depth using optical or digital microscopy.

Methods have been devised to modify the yield strength, ductility, and toughness of both crystalline and amorphous materials. These strengthening mechanisms give engineers the ability to tailor the mechanical properties of materials to suit a variety of different applications. For example, the favorable properties of steel result from interstitial incorporation of carbon into the iron lattice. Brass, a binary alloy of copper and zinc, has superior mechanical properties compared to its constituent metals due to solution strengthening. Work hardening has also been used for centuries by blacksmiths to introduce dislocations into materials, increasing their yield strengths.

In materials science, segregation is the enrichment of atoms, ions, or molecules at a microscopic region in a materials system. While the terms segregation and adsorption are essentially synonymous, in practice, segregation is often used to describe the partitioning of molecular constituents to defects from solid solutions, whereas adsorption is generally used to describe such partitioning from liquids and gases to surfaces. The molecular-level segregation discussed in this article is distinct from other types of materials phenomena that are often called segregation, such as particle segregation in granular materials, and phase separation or precipitation, wherein molecules are segregated in to macroscopic regions of different compositions. Segregation has many practical consequences, ranging from the formation of soap bubbles, to microstructural engineering in materials science, to the stabilization of colloidal suspensions.

Low hydrogen annealing, commonly known as "baking" is a heat treatment in metallurgy for the reduction or elimination of hydrogen in a material to prevent hydrogen embrittlement. Hydrogen embrittlement is the hydrogen-induced cracking of metals, particularly steel which results in degraded mechanical properties such as plasticity, ductility and fracture toughness at low temperature. Low hydrogen annealing is called a de-embrittlement process. Low hydrogen annealing is an effective method compared to alternatives such as electroplating the material with zinc to provide a barrier for hydrogen ingress which results in coating defects.

Dislocation creep is a deformation mechanism in crystalline materials. Dislocation creep involves the movement of dislocations through the crystal lattice of the material, in contrast to diffusion creep, in which diffusion is the dominant creep mechanism. It causes plastic deformation of the individual crystals, and thus the material itself.

Dynamic strain aging (DSA) for materials science is an instability in plastic flow of materials, associated with interaction between moving dislocations and diffusing solutes. Although sometimes dynamic strain aging is used interchangeably with the Portevin–Le Chatelier effect, dynamic strain aging refers specifically to the microscopic mechanism that induces the Portevin–Le Chatelier effect. This strengthening mechanism is related to solid-solution strengthening and has been observed in a variety of fcc and bcc substitutional and interstitial alloys, metalloids like silicon, and ordered intermetallics within specific ranges of temperature and strain rate.

References

  1. Cottrell, A. H.; Bilby, B. A. (1949), "Dislocation Theory of Yielding and Strain Ageing of Iron", Proceedings of the Physical Society, 62 (1): 49–62, Bibcode:1949PPSA...62...49C, doi:10.1088/0370-1298/62/1/308
  2. Blavette, D.; Cadel, E.; Fraczkiewicz, A.; Menand, A. (1999). "Three-Dimensional Atomic-Scale Imaging of Impurity Segregation to Line Defects". Science. 286 (5448): 2317–2319. doi:10.1126/science.286.5448.2317. PMID   10600736.
  3. Waseda, Osamu; Veiga, Roberto GA; Morthomas, Julien; Chantrenne, Patrice; Becquart, Charlotte S.; Ribeiro, Fabienne; Jelea, Andrei; Goldenstein, Helio; Perez, Michel (March 2017). "Formation of carbon Cottrell atmospheres and their effect on the stress field around an edge dislocation". Scripta Materialia. 129: 16–19. doi:10.1016/j.scriptamat.2016.09.032. ISSN   1359-6462.
  4. Veiga, R.G.A.; Goldenstein, H.; Perez, M.; Becquart, C.S. (1 November 2015). "Monte Carlo and molecular dynamics simulations of screw dislocation locking by Cottrell atmospheres in low carbon Fe–C alloys". Scripta Materialia. 108: 19–22. doi:10.1016/j.scriptamat.2015.06.012. ISSN   1359-6462.
  5. 1 2 Takeuchi, S.; Argon, A. S. (1976-10-01). "Steady-state creep of alloys due to viscous motion of dislocations". Acta Metallurgica. 24 (10): 883–889. doi:10.1016/0001-6160(76)90036-5. ISSN   0001-6160.
  6. Mohamed, Farghalli A. (1979-04-01). "Creep behavior of solid solution alloys". Materials Science and Engineering. 38 (1): 73–80. doi:10.1016/0025-5416(79)90034-X. ISSN   0025-5416.
  7. Suzuki, Hideji (1952-01-01). "Chemical Interaction of Solute Atoms with Dislocations". Science Reports of the Research Institutes, Tohoku University. Ser. A, Physics, Chemistry and Metallurgy (in Japanese). 4: 455–463.
  8. Suzuki, Hideji (1962-02-15). "Segregation of Solute Atoms to Stacking Faults". Journal of the Physical Society of Japan. 17 (2): 322–325. Bibcode:1962JPSJ...17..322S. doi:10.1143/JPSJ.17.322. ISSN   0031-9015.
  9. Hickel, T.; Sandlöbes, S.; Marceau, R. K. W.; Dick, A.; Bleskov, I.; Neugebauer, J.; Raabe, D. (2014-08-15). "Impact of nanodiffusion on the stacking fault energy in high-strength steels". Acta Materialia. 75: 147–155. doi:10.1016/j.actamat.2014.04.062. ISSN   1359-6454.
  10. 1 2 3 4 5 Snoek, J. L. (1941-07-01). "Effect of small quantities of carbon and nitrogen on the elastic and plastic properties of iron". Physica. 8 (7): 711–733. doi:10.1016/S0031-8914(41)90517-7. ISSN   0031-8914.
  11. Koiwa, M. (1971-09-01). "Theory of the snoek effect in ternary b.c.c. alloys". The Philosophical Magazine. 24 (189): 539–554. doi:10.1080/14786437108217028. ISSN   0031-8086.
  12. Gavriljuk, V. G.; Shyvaniuk, V. N.; Teus, S. M. (2021-12-15). "Mobility of dislocations in the iron-based C-, N-, H-solid solutions measured using internal friction: Effect of electron structure". Journal of Alloys and Compounds. 886: 161260. doi:10.1016/j.jallcom.2021.161260. ISSN   0925-8388.
  13. 1 2 3 4 5 6 7 8 9 10 Marc Meyers, Krishan Chawla (2009). Mechanical Behavior of Materials. Cambridge, UK: Cambridge. pp. 569–570. ISBN   978-0-511-45557-5.
  14. 1 2 Weller, M. (2006-12-20). "The Snoek relaxation in bcc metals—From steel wire to meteorites". Materials Science and Engineering: A. Proceedings of the 14th International Conference on Internal Friction and Mechanical Spectroscopy. 442 (1): 21–30. doi:10.1016/j.msea.2006.02.232. ISSN   0921-5093.
  15. Hosford, William F. (2005). Mechanical behavior of materials. Cambridge: Cambridge University Press. ISBN   0-521-84670-6. OCLC   56482243.
  16. Koiwa, M. (1971-09-01). "Theory of the snoek effect in ternary b.c.c. alloys". The Philosophical Magazine: A Journal of Theoretical Experimental and Applied Physics. 24 (189): 539–554. doi:10.1080/14786437108217028. ISSN   0031-8086.