Grain growth

Last updated

In materials science, grain growth is the increase in size of grains (crystallites) in a material at high temperature. This occurs when recovery and recrystallisation are complete and further reduction in the internal energy can only be achieved by reducing the total area of grain boundary. The term is commonly used in metallurgy but is also used in reference to ceramics and minerals. The behaviors of grain growth is analogous to the coarsening behaviors of grains, which implied that both of grain growth and coarsening may be dominated by the same physical mechanism.

Contents

Importance of grain growth

The practical performances of polycrystalline materials are strongly affected by the formed microstructure inside, which is mostly dominated by grain growth behaviors. For example, most materials exhibit the Hall–Petch effect at room-temperature and so display a higher yield stress when the grain size is reduced (assuming abnormal grain growth has not taken place). At high temperatures the opposite is true since the open, disordered nature of grain boundaries means that vacancies can diffuse more rapidly down boundaries leading to more rapid Coble creep. Since boundaries are regions of high energy they make excellent sites for the nucleation of precipitates and other second-phases e.g. Mg–Si–Cu phases in some aluminium alloys or martensite platlets[ check spelling ] in steel. Depending on the second phase in question this may have positive or negative effects.

Rules of grain growth

Grain growth has long been studied primarily by the examination of sectioned, polished and etched samples under the optical microscope. Although such methods enabled the collection of a great deal of empirical evidence, particularly with regard to factors such as temperature or composition, the lack of crystallographic information limited the development of an understanding of the fundamental physics. Nevertheless, the following became well-established features of grain growth:

  1. Grain growth occurs by the movement of grain boundaries and also by coalescence (i.e. like water droplets) [1]
  2. Grain growth competition between Ordered coalescence and the movement of grain boundaries [2]
  3. Boundary movement may be discontinuous and the direction of motion may change suddenly during abnormal grain growth.
  4. One grain may grow into another grain whilst being consumed from the other side
  5. The rate of consumption often increases when the grain is nearly consumed
  6. A curved boundary typically migrates towards its centre of curvature

Classical driving force

The boundary between one grain and its neighbour (grain boundary) is a defect in the crystal structure and so it is associated with a certain amount of energy. As a result, there is a thermodynamic driving force for the total area of boundary to be reduced. If the grain size increases, accompanied by a reduction in the actual number of grains per volume, then the total area of grain boundary will be reduced.

In the classic theory, the local velocity of a grain boundary at any point is proportional to the local curvature of the grain boundary, i.e.:

,

where is the velocity of grain boundary, is grain boundary mobility (generally depends on orientation of two grains), is the grain boundary energy and is the sum of the two principal surface curvatures. For example, shrinkage velocity of a spherical grain embedded inside another grain is

,

where is radius of the sphere. This driving pressure is very similar in nature to the Laplace pressure that occurs in foams.

In comparison to phase transformations the energy available to drive grain growth is very low and so it tends to occur at much slower rates and is easily slowed by the presence of second phase particles or solute atoms in the structure.

Recently, in contrast to the classic linear relation between grain boundary velocity and curvature, grain boundary velocity and curvature are observed to be not correlated in Ni polycrystals, [3] which conflicting results has been revealed and be theoretically interpreted by a general model of grain boundary (GB) migration in the previous literature. [4] [5] According to the general GB migration model, the classical linear relation can only be used in a specical case.

A general theory of grain growth

Development of theoretical models describing grain growth is an active field of research. Many models have been proposed for grain growth, but no theory has yet been put forth that has been independently validated to apply across the full range of conditions and many questions remain open. [6] By no means is the following a comprehensive review. One recent theory of grain growth posits that normal grain growth only occurs in the polycrystalline systems with grain boundaries which have undergone roughening transitions, and abnormal and/or stagnant grain growth can only occur in the polycrystalline systems with non-zero GB (grain boundary) step free energy of grains. [7] Other models explaining grain coarsening assert that disconnections are responsible for the motion of grain boundaries, and provide limited experimental evidence suggesting that they govern grain boundary migration and grain growth behavior. [8] Other models have indicated that triple junctions play an important role in determining the grain growth behavior in many systems. [9]

Ideal grain growth

Computer Simulation of Grain Growth in 3D using phase field model. Click to see the animation. Grgr3d small.gif
Computer Simulation of Grain Growth in 3D using phase field model. Click to see the animation.

Ideal grain growth is a special case of normal grain growth where boundary motion is driven only by local curvature of the grain boundary. It results in the reduction of the total amount of grain boundary surface area i.e. total energy of the system. Additional contributions to the driving force by e.g. elastic strains or temperature gradients are neglected. If it holds that the rate of growth is proportional to the driving force and that the driving force is proportional to the total amount of grain boundary energy, then it can be shown that the time t required to reach a given grain size is approximated by the equation

where d0 is the initial grain size, d is the final grain size and k is a temperature dependent constant given by an exponential law:

where k0 is a constant, T is the absolute temperature and Q is the activation energy for boundary mobility. Theoretically, the activation energy for boundary mobility should equal that for self-diffusion but this is often found not to be the case.

In general these equations are found to hold for ultra-high purity materials but rapidly fail when even tiny concentrations of solute are introduced.

Self-similarity

Click to see the animation. Geometry of a single growing grain is changing during grain growth. This is extracted from a large scale phase-field simulation. Here surfaces are "grain boundaries", edges are "triple junctions" and corners are vertexes or higher-order junctions. For more information please see. Evolution of a single growing grain.gif
Click to see the animation. Geometry of a single growing grain is changing during grain growth. This is extracted from a large scale phase-field simulation. Here surfaces are "grain boundaries", edges are "triple junctions" and corners are vertexes or higher-order junctions. For more information please see.

An old-standing topic in grain growth is the evolution of the grains size distribution. Inspired by the work of Lifshitz and Slyozov on Ostwald ripening, Hillert has suggested that in a normal grain growth process the size distribution function must converge to a self-similar solution, i.e. it becomes invariant when the grain size is scaled with a characteristic length of the system that is proportional to the average grain size .

Several simulation studies, however, have shown that the size distribution deviates from the Hillert's self-similar solution. [11] Hence a search for a new possible self-similar solution was initiated that indeed led to a new class of self-similar distribution functions. [12] [13] [14] Large-scale phase field simulations have shown that there is indeed a self-similar behavior possible within the new distribution functions. It was shown that the origin of the deviation from Hillert's distribution is indeed the geometry of grains specially when they are shrinking. [15]

Normal vs abnormal

Distinction between continuous (normal) grain growth, where all grains grow at roughly the same rate, and discontinuous (abnormal) grain growth, where one grain grows at a much greater rate than its neighbours. ContinuousDiscontinuousAnnealing.png
Distinction between continuous (normal) grain growth, where all grains grow at roughly the same rate, and discontinuous (abnormal) grain growth, where one grain grows at a much greater rate than its neighbours.

In common with recovery and recrystallisation, growth phenomena can be separated into continuous and discontinuous mechanisms. In the former the microstructure evolves from state A to B (in this case the grains get larger) in a uniform manner. In the latter, the changes occur heterogeneously and specific transformed and untransformed regions may be identified. Abnormal or discontinuous grain growth is characterised by a subset of grains growing at a high rate and at the expense of their neighbours and tends to result in a microstructure dominated by a few very large grains. In order for this to occur the subset of grains must possess some advantage over their competitors such as a high grain boundary energy, locally high grain boundary mobility, favourable texture or lower local second-phase particle density. [16]

Factors hindering growth

If there are additional factors preventing boundary movement, such as Zener pinning by particles, then the grain size may be restricted to a much lower value than might otherwise be expected. This is an important industrial mechanism in preventing the softening of materials at high temperature.

Inhibition

Certain materials especially refractories which are processed at high temperatures end up with excessively large grain size and poor mechanical properties at room temperature. To mitigate this problem in a common sintering procedure, a variety of dopants are often used to inhibit grain growth.

Related Research Articles

<span class="mw-page-title-main">Sintering</span> Process of forming and bonding material by heat or pressure

Sintering or frittage is the process of compacting and forming a solid mass of material by pressure or heat without melting it to the point of liquefaction. Sintering happens as part of a manufacturing process used with metals, ceramics, plastics, and other materials. The nanoparticles in the sintered material diffuse across the boundaries of the particles, fusing the particles together and creating a solid piece.

In materials science, superplasticity is a state in which solid crystalline material is deformed well beyond its usual breaking point, usually over about 400% during tensile deformation. Such a state is usually achieved at high homologous temperature. Examples of superplastic materials are some fine-grained metals and ceramics. Other non-crystalline materials (amorphous) such as silica glass and polymers also deform similarly, but are not called superplastic, because they are not crystalline; rather, their deformation is often described as Newtonian fluid. Superplastically deformed material gets thinner in a very uniform manner, rather than forming a "neck" that leads to fracture. Also, the formation of microvoids, which is another cause of early fracture, is inhibited. Superplasticity must not be confused with superelasticity.

<span class="mw-page-title-main">Creep (deformation)</span> 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 undergo slow deformation while subject to 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 increase as they near their melting point.

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, stainless steels, and duplex stainless steel. In superalloys, it is known to cause yield strength anomaly providing excellent high-temperature strength.

<span class="mw-page-title-main">Grain boundary</span> Interface between crystallites in a polycrystalline material

In materials science, a grain boundary is the interface between two grains, or crystallites, in a polycrystalline material. Grain boundaries are two-dimensional defects in the crystal structure, and tend to decrease the electrical and thermal conductivity of the material. Most grain boundaries are preferred sites for the onset of corrosion and for the precipitation of new phases from the solid. They are also important to many of the mechanisms of creep. On the other hand, grain boundaries disrupt the motion of dislocations through a material, so reducing crystallite size is a common way to improve mechanical strength, as described by the Hall–Petch relationship.

<span class="mw-page-title-main">Crystal twinning</span> Two separate crystals sharing some of the same crystal lattice points in a symmetrical manner

Crystal twinning occurs when two or more adjacent crystals of the same mineral are oriented so that they share some of the same crystal lattice points in a symmetrical manner. The result is an intergrowth of two separate crystals that are tightly bonded to each other. The surface along which the lattice points are shared in twinned crystals is called a composition surface or twin plane.

<span class="mw-page-title-main">Superalloy</span> Alloy with higher durability than normal metals

A superalloy, or high-performance alloy, is an alloy with the ability to operate at a high fraction of its melting point. Key characteristics of a superalloy include mechanical strength, thermal creep deformation resistance, surface stability, and corrosion and oxidation resistance.

Dynamic recrystallization (DRX) is a type of recrystallization process, found within the fields of metallurgy and geology. In dynamic recrystallization, as opposed to static recrystallization, the nucleation and growth of new grains occurs during deformation rather than afterwards as part of a separate heat treatment. The reduction of grain size increases the risk of grain boundary sliding at elevated temperatures, while also decreasing dislocation mobility within the material. The new grains are less strained, causing a decrease in the hardening of a material. Dynamic recrystallization allows for new grain sizes and orientation, which can prevent crack propagation. Rather than strain causing the material to fracture, strain can initiate the growth of a new grain, consuming atoms from neighboring pre-existing grains. After dynamic recrystallization, the ductility of the material increases.

<span class="mw-page-title-main">Nanocrystalline material</span>

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 to 500 nm are typically considered "ultrafine" grains.

Nickel aluminide typically refers to one of the two most widely used compounds, Ni3Al or NiAl, but can refer to most aluminides from the Ni-Al system. These alloys are widely used due to their corrosion resistance, low density and ease of production. Ni3Al is of specific interest as the strengthening γ' phase precipitate in nickel-based superalloys allowing for high temperature strength up to 0.7-0.8 of its melting temperature. Meanwhile, NiAl displays excellent properties such as low density (lower than that of Ni3Al), good thermal conductivity, oxidation resistance and high melting temperature. These properties make it ideal for special high temperature applications like coatings on blades in gas turbines and jet engines. However, both these alloys do have the disadvantage of being quite brittle at room temperature, with Ni3Al remaining brittle at high temperatures as well. However, it has been shown that Ni3Al can be made ductile when manufactured in single-crystal form rather than in polycrystalline form.

Damage mechanics is concerned with the representation, or modeling, of damage of materials that is suitable for making engineering predictions about the initiation, propagation, and fracture of materials without resorting to a microscopic description that would be too complex for practical engineering analysis.

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.

<span class="mw-page-title-main">Grain boundary strengthening</span> Method of strengthening materials by changing grain size

In materials science, grain-boundary strengthening is a method of strengthening materials by changing their average crystallite (grain) size. It is based on the observation that grain boundaries are insurmountable borders for dislocations and that the number of dislocations within a grain has an effect on how stress builds up in the adjacent grain, which will eventually activate dislocation sources and thus enabling deformation in the neighbouring grain as well. By changing grain size, one can influence the number of dislocations piled up at the grain boundary and yield strength. For example, heat treatment after plastic deformation and changing the rate of solidification are ways to alter grain size.

<span class="mw-page-title-main">SRAS</span>

SRAS a non-destructive acoustic microscopy microstructural-crystallographic characterization technique commonly used in the study of crystalline or polycrystalline materials. The technique can provide information about the structure and crystallographic orientation of the material. Traditionally, the information provided by SRAS has been acquired by using diffraction techniques in electron microscopy - such as EBSD. The technique was patented in 2005, EP patent 1910815.

Nabarro–Herring creep is a mode of deformation of crystalline materials that occurs at low stresses and held at elevated temperatures in fine-grained materials. In Nabarro–Herring creep, atoms diffuse through the crystals, and the creep rate varies inversely with the square of the grain size so fine-grained materials creep faster than coarser-grained ones. NH creep is solely controlled by diffusional mass transport. This type of creep results from the diffusion of vacancies from regions of high chemical potential at grain boundaries subjected to normal tensile stresses to regions of lower chemical potential where the average tensile stresses across the grain boundaries are zero. Self-diffusion within the grains of a polycrystalline solid can cause the solid to yield to an applied shearing stress, the yielding being caused by a diffusional flow of matter within each crystal grain away from boundaries where there is a normal pressure and toward those where there is a normal tension. Atoms migrating in the opposite direction account for the creep strain. The creep strain rate is derived in the next section. NH creep is more important in ceramics than metals as dislocation motion is more difficult to effect in ceramics.

Geometrically necessary dislocations are like-signed dislocations needed to accommodate for plastic bending in a crystalline material. They are present when a material's plastic deformation is accompanied by internal plastic strain gradients. They are in contrast to statistically stored dislocations, with statistics of equal positive and negative signs, which arise during plastic flow from multiplication processes like the Frank-Read source.

<span class="mw-page-title-main">W. Craig Carter</span> American materials scientist, engineer and academic

W. Craig Carter is an American materials scientist, a POSCO Professor of Materials Science and Engineering at Massachusetts Institute of Technology. He is also a co-founder of the 24M Technologies Company.

Liquid phase sintering is a sintering technique that uses a liquid phase to accelerate the interparticle bonding of the solid phase. In addition to rapid initial particle rearrangement due to capillary forces, mass transport through liquid is generally orders of magnitude faster than through solid, enhancing the diffusional mechanisms that drive densification. The liquid phase can be obtained either through mixing different powders—melting one component or forming a eutectic—or by sintering at a temperature between the liquidus and solidus. Additionally, since the softer phase is generally the first to melt, the resulting microstructure typically consists of hard particles in a ductile matrix, increasing the toughness of an otherwise brittle component. However, liquid phase sintering is inherently less predictable than solid phase sintering due to the complexity added by the presence of additional phases and rapid solidification rates. Activated sintering is the solid-state analog to the process of liquid phase sintering.

Electroplasticity, describes the enhanced plastic behavior of a solid material under the application of an electric field. This electric field could be internal, resulting in current flow in conducting materials, or external. The effect of electric field on mechanical properties ranges from simply enhancing existing plasticity, such as reducing the flow stress in already ductile metals, to promoting plasticity in otherwise brittle ceramics. The exact mechanisms that control electroplasticity vary based on the material and the exact conditions. Enhancing the plasticity of materials is of great practical interest as plastic deformation provides an efficient way of transforming raw materials into final products. The use of electroplasticity to improve processing of materials is known as electrically assisted manufacturing.

<span class="mw-page-title-main">Precipitate-free zone</span> Region around a material grain boundary free of solid impurities

In materials science, a precipitate-free zone (PFZ) refers to microscopic localized regions around grain boundaries that are free of precipitates. It is a common phenomenon that arises in polycrystalline materials where heterogeneous nucleation of precipitates is the dominant nucleation mechanism. This is because grain boundaries are high-energy surfaces that act as sinks for vacancies, causing regions adjacent to a grain boundary to be devoid of vacancies. As it is energetically favorable for heterogeneous nucleation to occur preferentially around defect-rich sites such as vacancies, nucleation of precipitates is impeded in the vacancy-free regions immediately adjacent to grain boundaries

References

  1. Hu, J.; Shen, Z. (2012-10-01). "Grain growth by multiple ordered coalescence of nanocrystals during spark plasma sintering of SrTiO3 nanopowders". Acta Materialia. 60 (18): 6405–6412. Bibcode:2012AcMat..60.6405H. doi:10.1016/j.actamat.2012.08.027. ISSN   1359-6454.
  2. Hu, Jianfeng; Shen, Zhijian (March 2021). "Grain growth competition during sintering of SrTiO3 nanocrystals: Ordered coalescence of nanocrystals versus conventional mechanism". Scripta Materialia. 194: 113703. doi:10.1016/j.scriptamat.2020.113703. S2CID   233644242.
  3. Bhattacharya, Aditi; Shen, Yu-Feng; Hefferan, Christopher M.; Li, Shiu Fai; Lind, Jonathan; Suter, Robert M.; Krill, Carl E.; Rohrer, Gregory S. (2021-10-08). "Grain boundary velocity and curvature are not correlated in Ni polycrystals". Science. 374 (6564): 189–193. Bibcode:2021Sci...374..189B. doi:10.1126/science.abj3210. ISSN   0036-8075. OSTI   1866159. PMID   34618565. S2CID   238474395.
  4. Hu, Jianfeng; Wang, Xianhao; Zhang, Junzhan; Luo, Jun; Zhang, Zhijun; Shen, Zhijian (September 2021). "A general mechanism of grain growth ─I. Theory". Journal of Materiomics. 7 (5): 1007–1013. arXiv: 1901.00732 . doi: 10.1016/j.jmat.2021.02.007 . ISSN   2352-8478.
  5. Hu, Jianfeng; Zhang, Junzhan; Wang, Xianhao; Luo, Jun; Zhang, Zhijun; Shen, Zhijian (September 2021). "A general mechanism of grain growth-II: Experimental". Journal of Materiomics. 7 (5): 1014–1021. doi: 10.1016/j.jmat.2021.02.008 .
  6. Rios, P. R.; Zöllner, D. (2018). "Critical assessment 30: Grain growth – Unresolved issues". Materials Science and Technology. 34 (6): 629–638. Bibcode:2018MatST..34..629R. doi:10.1080/02670836.2018.1434863.
  7. Hu, Jianfeng; Wang, Xianhao; Zhang, Junzhan; Luo, Jun; Zhang, Zhijun; Shen, Zhijian (February 2021). "A general mechanism of grain growth ─I. Theory". Journal of Materiomics. 7 (5): 1007–1013. arXiv: 1901.00732 . doi:10.1016/j.jmat.2021.02.007. ISSN   2352-8478.. Other explanations
  8. Zhang, L.; Han, J.; Srolovitz, D.J. (2021). "Equation of motion for grain boundaries in polycrystals". npj Comput Mater. 7 (64). Bibcode:2021npjCM...7...64Z. doi:10.1038/s41524-021-00532-6.
  9. Novikov, Vladimir Yu. (2004). "Triple Junction Controlled Grain Growth". Materials Science Forum. 467–470: 1093–1098. doi:10.4028/www.scientific.net/MSF.467-470.1093. S2CID   137134066.
  10. Darvishi Kamachali, Reza (2013). "Grain boundary motion in polycrystalline materials, PhD thesis" (PDF). Archived from the original (PDF) on 2018-10-25.
  11. Acta Materialia 60 (2012). "3-D phase-field simulation of grain growth: Topological analysis versus mean-field approximations".{{cite web}}: CS1 maint: numeric names: authors list (link)
  12. Brown, L.C. (1992-06-15). "Answer to the rebuttal of Hillert, Hunderi and Ryum". Scripta Metallurgica et Materialia. 26 (12): 1945. doi:10.1016/0956-716X(92)90065-M. ISSN   0956-716X.
  13. Coughlan, S.D.; Fortes, M.A. (1993-06-15). "Self similar size distributions in particle coarsening". Scripta Metallurgica et Materialia. 28 (12): 1471–1476. doi:10.1016/0956-716X(93)90577-F. ISSN   0956-716X.
  14. Rios, P.R (1999-02-19). "Comparison between a computer simulated and an analytical grain size distribution". Scripta Materialia. 40 (6): 665–668. doi:10.1016/S1359-6462(98)00495-3. ISSN   1359-6462.
  15. Acta Materialia 90 (2015). "Geometrical grounds of mean field solutions for normal grain growth".{{cite web}}: CS1 maint: numeric names: authors list (link)
  16. Hanaor, D.A.H; Xu, W; Ferry, M; Sorrell, CC (2012). "Abnormal grain growth of rutile TiO2 induced by ZrSiO4". Journal of Crystal Growth. 359: 83–91. arXiv: 1303.2761 . Bibcode:2012JCrGr.359...83H. doi:10.1016/j.jcrysgro.2012.08.015. S2CID   94096447.