Difference between revisions of "Crystallite" - New World Encyclopedia

From New World Encyclopedia
(added credit and category tags, deleted foreign language links)
Line 32: Line 32:
 
Because of the dangers of grain boundaries in certain materials such as [[superalloy]] turbine blades, great technological leaps were made to minimize as much as possible the effect of grain boundaries in the blades. The result was [[directional solidification]] processing in which grain boundaries were eliminated by producing columnar grain structures aligned parallel to the axis of the blade, since this is usually the direction of maximum tensile stress felt by a blade during its rotation in an airplane. The resulting turbine blades consisted of a single grain, improving reliability.  
 
Because of the dangers of grain boundaries in certain materials such as [[superalloy]] turbine blades, great technological leaps were made to minimize as much as possible the effect of grain boundaries in the blades. The result was [[directional solidification]] processing in which grain boundaries were eliminated by producing columnar grain structures aligned parallel to the axis of the blade, since this is usually the direction of maximum tensile stress felt by a blade during its rotation in an airplane. The resulting turbine blades consisted of a single grain, improving reliability.  
  
Generally, polycrystals cannot be [[superheating|superheated]]; they will melt promptly once they are brought to a high enough temperature.  This is because grain boundaries are amorphous, and serve as [[Nucleation|nucleation points]] for the liquid [[phase (matter)|phase]].  By contrast, if no solid nucleus is present as a liquid cools, it tends to become [[supercooling|supercooled]].  Since this is undesirable for mechanical materials, [[alloy]] designers often take steps against it.  See [[grain refinement]].
+
Generally, polycrystals cannot be [[superheating|superheated]]; they will melt promptly once they are brought to a high enough temperature.  This is because grain boundaries are amorphous, and serve as [[Nucleation|nucleation points]] for the liquid [[phase (matter)|phase]].  By contrast, if no solid nucleus is present as a liquid cools, it tends to become [[supercooling|supercooled]].  Since this is undesirable for mechanical materials, [[alloy]] designers often take steps against it.
  
 
== See also ==
 
== See also ==
Line 38: Line 38:
 
* [[Crystal]]
 
* [[Crystal]]
 
* [[Crystallography]]
 
* [[Crystallography]]
 +
* [[Gemstone]]
 +
* [[Grain refinement]]
  
== Footnotes ==
+
== Notes ==
{{reflist}}
+
<references/>
  
 
== References ==
 
== References ==
*Allen, Samuel and Thomas, Edwin. The Structure of Materials. New York: John Wiley & Sons, Inc. 1999.  
+
 
*Jiles, David. Introduction to Magnetism and Magnetic Materials. London: Chapman & Hall/CRC, 1998.
+
*Allen, Samuel, and Thomas, Edwin. 1999. The Structure of Materials. New York: John Wiley & Sons, Inc.  
 +
*Jiles, David. 1998. Introduction to Magnetism and Magnetic Materials. London: Chapman & Hall/CRC.
  
 
[[Category:Physical sciences]]
 
[[Category:Physical sciences]]

Revision as of 22:17, 11 February 2009

Galvanized surface with visible crystallites of zinc. Crystallites in the steel under the coating are microscopic.

A crystallite is a domain of solid-state matter that has the same structure as a single crystal. Metallurgists often refer to crystallites as "grains".

Solid objects that are large enough to see and handle are rarely composed of a single crystal, except for a few cases (gems, silicon single crystals for the electronics industry, certain types of fiber, and single crystals of a nickel-based superalloy for turbojet engines). Most materials are polycrystalline; they are made of a large number of single crystals — crystallites — held together by thin layers of amorphous solid. The crystallite size can vary from a few nanometers to several millimeters.

If the individual crystallites are oriented randomly (that is, if they lack texture), a large enough volume of polycrystalline material will be approximately isotropic. This property helps the simplifying assumptions of continuum mechanics to apply to real-world solids. However, most manufactured materials have some alignment to their crystallites, which must be taken into account for accurate predictions of their behavior and characteristics.

Material fractures can be intergranular fracture or a transgranular fracture. There is an ambiguity with powder grains: a powder grain can be made of several crystallites. Thus, the (powder) "grain size" found by laser granulometry can be different from the "grain size" (or, rather, crystallite size) found by X-ray diffraction (e.g. Scherrer method), by optical microscopy under polarised light, or by scanning electron microscopy (backscattered electrons).

Coarse grained rocks are formed very slowly, while fine grained rocks are formed quickly, on geological time scales. If a rock forms very quickly, such as the solidification of lava ejected from a volcano, there may be no crystals at all. This is how obsidian forms.

Grain boundaries



Grain boundaries are interfaces where crystals of different orientations meet. A grain boundary is a single-phase interface, with crystals on each side of the boundary being identical except in orientation. The term "crystallite boundary" is sometimes, though rarely, used. Grain boundary areas contain those atoms that have been perturbed from their original lattice sites, dislocations, and impurities that have migrated to the lower energy grain boundary.

Treating a grain boundary geometrically as an interface of a single crystal cut into two parts, one of which is rotated, we see that there are five variables required to define a grain boundary. The first two numbers come from the unit vector that specifies a rotation axis. The third number designates the angle of rotation of the grain. The final two numbers specify the plane of the grain boundary (or a unit vector that is normal to this plane).

Grain boundaries disrupt the motion of dislocations through a material. Dislocation propagation is impeded because of the stress field of the grain boundary defect region and the lack of slip planes and slip directions and overall alignment across the boundaries. Reducing grain size is therefore a common way to improve strength, often without any sacrifice in toughness because the smaller grains create more obstacles per unit area of slip plane. This crystallite size-strength relationship is given by the Hall-Petch relationship. The high interfacial energy and relatively weak bonding in grain boundaries makes them preferred sites for the onset of corrosion and for the precipitation of new phases from the solid.

Grain boundary migration plays an important role in many of the mechanisms of creep. Grain boundary migration occurs when a shear stress acts on the grain boundary plane and causes the grains to slide. This means that fine-grained materials actually have a poor resistance to creep relative to coarser grains, especially at high temperatures, because smaller grains contain more atoms in grain boundary sites. Grain boundaries also cause deformation in that they are sources and sinks of point defects. Voids in a material tend to gather in a grain boundary, and if this happens to a critical extent, the material could fracture.

During grain boundary migration, the rate determining step depends on the angle between two adjacent grains. In a small angle dislocation boundary, the migration rate depends on vacancy diffusion between dislocations. In a high angle dislocation boundary, this depends on the atom transport by single atom jumps from the shrinking to the growing grains [1].

Grain boundaries are generally only a few nanometers wide. In common materials, crystallites are large enough that grain boundaries account for a small fraction of the material. However, very small grain sizes are achievable. In nanocrystalline solids, grain boundaries become a significant volume fraction of the material, with profound effects on such properties as diffusion and plasticity. In the limit of small crystallites, as the volume fraction of grain boundaries approaches 100%, the material ceases to have any crystalline character, and thus becomes an amorphous solid.

Grain boundaries are also present in magnetic domains in magnetic materials. A computer hard disk, for example, is made of a hard ferromagnetic material that contains regions of atoms whose magnetic moments can be realigned by an inductive head. The magnetization varies from region to region, and the misalignment between these regions forms boundaries that are key to data storage. The inductive head measures the orientation of the magnetic moments of these domain regions and reads out either a “1” or “0”. These bits are the data being read. Grain size is important in this technology because it limits the number of bits that can fit on one hard disk. The smaller the grain sizes, the more data that can be stored.

Because of the dangers of grain boundaries in certain materials such as superalloy turbine blades, great technological leaps were made to minimize as much as possible the effect of grain boundaries in the blades. The result was directional solidification processing in which grain boundaries were eliminated by producing columnar grain structures aligned parallel to the axis of the blade, since this is usually the direction of maximum tensile stress felt by a blade during its rotation in an airplane. The resulting turbine blades consisted of a single grain, improving reliability.

Generally, polycrystals cannot be superheated; they will melt promptly once they are brought to a high enough temperature. This is because grain boundaries are amorphous, and serve as nucleation points for the liquid phase. By contrast, if no solid nucleus is present as a liquid cools, it tends to become supercooled. Since this is undesirable for mechanical materials, alloy designers often take steps against it.

See also

Notes

  1. R.D. Doherty et al., Current issues in recrystallization: a review, Materials Science and Engineering A238 (1997), p. 222 (Access to the article for subscribers only)

References
ISBN links support NWE through referral fees

  • Allen, Samuel, and Thomas, Edwin. 1999. The Structure of Materials. New York: John Wiley & Sons, Inc.
  • Jiles, David. 1998. Introduction to Magnetism and Magnetic Materials. London: Chapman & Hall/CRC.

Credits

New World Encyclopedia writers and editors rewrote and completed the Wikipedia article in accordance with New World Encyclopedia standards. This article abides by terms of the Creative Commons CC-by-sa 3.0 License (CC-by-sa), which may be used and disseminated with proper attribution. Credit is due under the terms of this license that can reference both the New World Encyclopedia contributors and the selfless volunteer contributors of the Wikimedia Foundation. To cite this article click here for a list of acceptable citing formats.The history of earlier contributions by wikipedians is accessible to researchers here:

The history of this article since it was imported to New World Encyclopedia:

Note: Some restrictions may apply to use of individual images which are separately licensed.