You are on page 1of 12
Precipitation Hardening Related terms: Aluminum Alloys, Austenite, Heat Treatment, Microstructure, Hardness, Alloying Element Precipitation hardening is a strengthening mechanism in alloys with special characteristics (two phase or multiphase, in which the solubility limit is increased with temperature) to form small particles in the range of ~100nm which have three stages: solution treatment (ST), quenching, and aging. From: Severe Plastic Deformation, 2018 View all Topics > as A ® Precipitation Hardening REE, Smallman, A.H.W. Ngan, in Modern Physical Metallurgy (Eighth Edition), 2014 13.4.1 The signif cance of particle deformability The strength of an age hardening alloy is governed by the interaction of moving dislocations and precipitates. The ‘obstacles in precipitation hardening alloys which hinder the motion of dislocations may be either (i) the strains around GP zones, (i) the zones or precipitates themselves, or both. Clearly, if itis the zones themselves which are important, it will be necessary for the moving dislocations either to cut through them or go round them. Thus, merely from elementary reasoning, it would appear that there are at least three causes of hardening, namely: (i) coherency strain hardening, (i) chemical hardening, ie. when the dislocation cuts through the precipitate or (ii) dispersion hardening, ie. when the dislocation goes round or over the precipitate. The relative contributions will depend on the particular alloy system but, generally there is a critical dispersion at which the strengthening is a maximum, as shown in Figure 13.7. In the small-particle regime the precipitates, or particles, are coherent and deformable as the dislocations cut through them, while in the larger-particle regime the particles are incoherent and non-deformable as the dislocations bypass them. For deformable particles, when the dislocations pass through the particle, the intrinsic properties of the particle are of importance and alloy strength varies only weakly with particle size. For non-deformable particles, when the dislocations bypass the particles, the alloy strength is independent of the particle properties but is strongly dependent on particle size and dispersion strength decreasing as particle size or dispersion increases. The transition from deformable to non- deformable particle-controlled deformation is readily recognized by the change in microstructure, since the ‘laminar’ undisturbed dislocation f bw for the former contrasts with the turbulent plastic f bw for non-deformable particles. The latter leads to the production of a high density of dislocation loops, dipoles and other debris which results in a high rate of work hardening, This high rate of work hardening is a distinguishing feature of al dispersion-hardened systems. Gaseous processes for low temperature surface hardening of stainless steel MAJ. Somers, T.L. Christiansen, in Thermochemical Surface Engineering of Steels, 2015 15.5.3 Precipitation hardening stainless steel Precipitation hardening stainless steels are steels that are subjected to an annealing treatment in order to improve ‘mechanical properties (strength). The amount of bulk hardening depends on annealing timejtemperature (and degree of deformation). The temperature/time applied for low temperature surface hardening can coincide with the temperature/time for maximum bulk hardness which makes it possible to carry out simultaneous bulk hardening and surface hardening in a single process step. The microstructures of nitrided Sandvik Nanof bx® and carburised Uddeholm Corrax® are depicted in Figure 15.21 [38]. Nitriding of Nanof kx gives a layer thickness of more than 20 um. The surface layer actually consists of two zones, nitrogen expanded austenite, Yy, and nitrogen expanded martensite dy. The applied nitriding potential determines the phases formed; a high nitriding potential favours the formation of Yj, while a low nitriding potential favours the formation of @'y. The formation of @’y is likely to occur prior to the formation of Yw. At some threshold nitrogen content, the @'y transforms into Vw due to the strong austenite stabilising ef ft of nitrogen. This nitrogen threshold value is only reached for high nitriding potentials. A similar transformation mechanism prevails during carburising, which is clearly observed for carburising of Corrax (Figure 15.21(b)). The bulk hardness of Nanof Ex yields a signif tant increase from an as- delivered hardness of 400 HV to 684 HV after nitriding; this increase in bulk hardness is caused by the simultaneously occurring precipitation hardening of the bulk during the nitriding treatment. The bulk hardness of Corrax increases from the as-delivered hardness of 330 HV to 540 HV as a result of precipitation hardening during carburising. Examples of nitrided and nitrocarburised austenitic precipitation hardening stainless steel A286 (see Table 15.1) are shown in Figure 15.22 [39]. The precipitation hardening temperature for this material is 720°C, which is considered far beyond the temperature range for low temperature surface hardening. Indeed, simultaneous bulk and surface hardening at 720°C resulted in development of CrN at grain boundaries, albeit modest [39]. Nevertheless, the presence of the strong nitride formers (as compared to Cr) Ti, Al and V, makes this 2 most interesting material, as N atoms will be ‘bound’ preferably by the strong nitride forming elements rather than Cr. Prolonged nitriding at temperatures as high as 500°C and low ammonia contents in the gas produced expanded austenite zones without the presence of CrN. It should be noted that low temperature surface hardening of stainless steels is a surface engineering treatment in its infancy as compared to nitriding, nitrocarburising and carburising. For the latter treatments nitriding and carburising steels were designed which have an optimal response towards the performance improvement that can be achieved with thermochemical treatment. No such optimisation of the steel composition has so far been initiated for stainless steels that are to be surface hardened. Rather, more or less standardised treatments are applied to commercially available steels. It is anticipated that alloy development in parallel with further process optimisation of low temperature surface hardening will pave the way for wider applicability of this promising family of surface engineering treatments. For this purpose gaseous processes are preferred as they have the largest exibility [Poe bok | Strengthening of metal alloys In Introduction to Aerospace Materials, 2012 4.4.5 Precipitation hardening Precipitation hardening is a strengthening process involving the formation of hard precipitate particles in the host alloy and these restrict dislocation slip. Precipitation-hardened alloys are used extensively in aircraft structures and engines, and include aluminium, magnesium, titanium and nickel alloys as well as several types of steel. Precipitation hardening is a remarkably ef €ctive strengthening process; for instance, the strength of aluminium alloy can be increased from about 150 to over 500 MPa (or more than 300%). Precipitation hardening is achieved by heat treating the metal alloy to form a di impede the movement of dislocations. Heat treatment involves the following stages: persion of f he precipitates that 1. Heating the metal at high temperature within the single-phase region to dissolve and disperse the alloying ‘elements in the matrix of the host metal. This stage is known as solution treatment. 2. Rapid cooling using a cool solution (e.g. water, oil) from the solution treatment temperature to obtain a supersaturated solid solution of alloying elements in the host metal. This is known as quenching. 3. Reheating to an intermediate temperature to convert the supersaturated solid solution to finely dispersed precipitate particles. This is called the age-hardening stage. The heat-treatment temperatures depend on the composition of the metal alloy and the desired amount of strengthening, For example, solution treatment of aluminium alloys used in aircraft is usually performed in the range 450-500 °C, quenched and then age-hardened at 160-220 °C. Titanium alloys are solution treated at about 750 °C, quenched and aged at 450-650 °C. Precipitation hardening is only possible when one or more of the alloying elements is completely soluble in the host metal at elevated temperature (i.e. solution treatment) and their solubility decreases when the metal is cooled. As the solubility falls with temperature, the host matrix rejects the excess of atoms of the alloying element from the lattice sites. These atoms cluster into small precipitate particles that induce high lattice strains that resist dislocation slip and thereby increase strength, Precipitation hardening involves a complex series of physical transformations in the metal alloy at the atomic and microstructural levels that induces dif Grent strengthening mechanisms over time. Figure 4.22 shows the hardening mechanism with each successive transformation in the age-hardening process. The mechanisms change with increasing age-hardening time in the order: solid-solution strengthening (SSS), Guinier-Preston (GP) strengthening, coherent precipitate strengthening, and f nally incoherent precipitate strengthening. Guinier—Preston zone strengthening Precipitation hardening in the age-hardening stage of the heat-treatment process starts with the formation of GuinierPreston (GP) zones. The formation of GP zones which occurs fistly by GP1 zones and then GP2 zones is illustrated in Fig. 4.23. After solution treatment and quenching, the alloying elements are dissolved and dispersed 4 a supersaturated solid solution in the host metal. The material in this condition is relatively soft with strengthening achieved by solid solution hardening. Upon age hardening the supersaturated solid solution undergoes a transformation with the alloying atoms dif fising through the host lattice to form solute-rich clusters. The driving force for this process is the concentration of the alloying elements in solid solution being higher than their equilibrium solubility limit. Atomic dif fision and clustering of the alloying elements is assisted by vacancies in the lattice. The dif Gision process is accelerated by heat treating the metal at elevated temperature; this is called thermal age hardening, The process occurs at a slower rate at room temperature with some metal alloys, and this is known as natural ageing. The atoms of the alloying elements form a large number of tiny clusters (up to 10! per em®) that are 10-50 nm long and only one or two atomic planes thick. These clusters are called GP zones. Figure 4.24 shows copper-rich GP zones in an aluminium alloy used in aircraft structures. The solute- ich GP zones are coherent with the crystal structure of the host metal. In other words, the arrangement of alloying atoms in the GP zone and matrix phase match at the interface plane and the two lattices are continuous across the interface. GP zones induce high strain in the surrounding lattice thus impeding dislocation slip. Dislocations are able to shear through GP zones, although the lattice strain imposes resistance against slip which strengthens the metal. Coherent precipitate strengthening The formation of coherent precipitate particles is the next step in the sequence of age-hardening processes. The GP zones transform into coherent precipitates in the matrix phase. Coherent precipitates also form by heterogenous nucleation at lattice defect sites such as grain boundaries and dislocations. A precipitate develops when the concentration of alloying elements clustered within a small region gives the composition of another phase. For example, copper in aluminium forms precipitate particles with a dif €rent composition (Al,Cu) to the matrix phase. When precipitation fist occurs, the particles are coherent with the host lattice (Fig. 4.25). That is, the lattice planes to the precipitate particles are continuous with the planes to the matrix, and there is no distinct interface between the particles and matrix. An abrupt change in composition occurs across the precipitate-matrix interface, but there is no change to the crystal structure. For instance, coherent Al;Cu precipitates have an fec ceystal structure similar to the host aluminium metal Coherent precipitates distort the surrounding matrix which resists dislocation slip. Once the dislocation has ‘overcome the lattice strain in the matrix and reached the precipitate, it can move through the particle because the slip planes are coherent with the matrix. However, the particles resist the slip process, thus increasing strength. ‘The strength gained by coherent precipitates increases with time during the ageing process as their population increases and they become larger. Incoherent precipitate strengthening The fal transformation in the age-hardening process is the formation of incoherent precipitate particles. These particles are created by coherent precipitates changing their crystal structure such that the lattice planes are no longer continuous with the planes of the host metal lattice. Incoherent precipitates are truly distinct second-phase particles with their own crystal structure and separated from the surrounding matrix by 2 well-def ned interface (Fig. 4.26) Incoherent precipitates are ef €ctive obstacles against dislocation slip because the large distortion of the matrix surrounding the particles interacts strongly with the stress f eld of the dislocations. The precipitates also obstruct dislocation movement because slip cannot occur across the particle-matrix interface owing to the change in ceystal structure. Strengthening by incoherent precipitates occurs by the Orowan hardening mechanism whereby dislocations must bend and loop around the particles (Fig. 4.21). The combination of high lattice strain and Orowan hardening results in incoherent precipitates being very ef €ctive in strengthening age-hardenable alloys. ‘The maximum improvement in strength gained by precipitation hardening occurs immediately upon the coherent precipitates transforming into incoherent particles. During the ageing process the particle size increases with a corresponding reduction in the particle spacing over time, as shown in Fig. 4.27. The spacing increases because the precipitates grow via a process called particle coarsening. The precipitates are not all the same size at any point during the ageing process; instead they occur over a range of sizes. Larger particles and particles located at high energy lattice sites (e.g. grain boundaries) are thermodynamically more stable than small particles located in the core of the grain. As ageing proceeds, the smaller, less stable particles dissolve and thereby release the solute atoms back into the matrix. These atoms migrate through the lattice to the larger particles, thus expanding, their size. As a result, the particles become larger but fewer in number and therefore more widely spaced. Incoherent precipitates are more ef Ective than coherent particles at resisting dislocation slip because they create higher lattice strains and promote Orowan hardening. Strengthening by Orowan hardening is most ef Gctive when the particles are closely spaced; a high density of tightly packed incoherent particles being the best condition to resist the bending and looping of dislocations. When the spacing increases, the particles are more ‘easily by-passed by moving dislocations, and therefore the strengthening ef €ct diminishes. For these reasons, maximum strengthening occurs with closely spaced incoherent precipitates. The strengthening ef €ct diminishes with prolonged ageing because the incoherent particles coarsen and thereby Orowan hardening becomes less cf Ective. The age-hardenable alloys used in aircraft structures are usually heat-treated to close to maximum strength. More information on the age hardening of aerospace alloys is provided in later chapters. Purchase book Failure of 17-4PH stainless steel components in of shore platforms Sérgio S.M. Tavares, .. Juan M. Pardal, in Handbook of Materials Failure Analysis with Case Studies from the Oil and Gas Industry, 2016 Abstract Precipitation hardening is a strengthening mechanism very common in many classes of metallic materials, from Aland Cu alloys to high-strength steels. In special, precipitation hardening (PH) stainless steels may allow corrosion resistance and mechanical strength desired to special applications. Despite this, there are many cases of failures involving PH stainless steels reported in the literature. Incorrect material selection or wrong heat treatment procedures are usually associated to these cases of failure. This chapter is focused on three cases of failure of 17-4 PH steel, a common precipitation hardenable martensitic stainless steel used in oil and gas production and ref ning industries. earl cos Purchase book Mj i Microstructure control in creep-age forming of aluminium panels L. Zhan, ... D. Balint, in Microstructure Evolution in Metal Forming Processes, 2012 11.4.4 Finite element (FE) modelling of precipitation hardening Precipitation hardening is a key feature of CAF. The evolution of precipitates af €cts not only the creepjstress relaxation behaviour, but also the achievement of the material's Fnal strength. As discussed before, CAF is based on a complicated combination of stress relaxation, creep and age hardening. Creep deformation takes place at low stress levels, and the amount of plastic deformation is directly related to the ageing temperature and time. This is signif eantly dif €rent from other metal forming processes, where elastic-plastic deformation of the material is dominant. For industrial applications, numerical studies on springback and precipitation hardening in CAF need to be available. For the past 20 years, most of the fhite element (FE) methods for simulation or numerical calculation have taken only macroscopic parameters, such as stress relaxation, creep strain, springback and so on, into consideration. ‘The research carried out has concentrated mainly on the prediction of springback in CAF processes using simple creep or stress relaxation models. The coupling of ageing or precipitation hardening and creep or stress relaxation has not been considered. Stress-ageing and creep-ageing phenomena in aluminium alloys have been understood further in recent years, and industry needs to be able to know and predict the mechanical-property distribution of creep-age-formed aircraft wing parts. Thus Ho et al (2004) proposed a set of mechanism-based unif ed creep-ageing constitutive equations for AA7010, which were implemented into ABAQUS through the user-def ned subroutine CREEP. In addition to creep deformation and stress relaxation, the precipitate growth during ageing and the evolution of the yield stress increment during CAF were predicted. Similar work has also been done by Zhan et al. (20116) for AATOSS at 120°C, which will be discussed below. r Purchase book

You might also like