We use cookies to distinguish you from other users and to provide you with a better experience on our websites. Close this message to accept cookies or find out how to manage your cookie settings.
To save content items to your account,
please confirm that you agree to abide by our usage policies.
If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account.
Find out more about saving content to .
To save content items to your Kindle, first ensure [email protected]
is added to your Approved Personal Document E-mail List under your Personal Document Settings
on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part
of your Kindle email address below.
Find out more about saving to your Kindle.
Note you can select to save to either the @free.kindle.com or @kindle.com variations.
‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi.
‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.
Here we give a brief review of the principles of stress, strain, and isotropic elasticity that will be needed in the study of defects. Readers familiar with this elementary material can skip on to the next chapter. The review is given here as a reference that will be used from time to time in the remainder of this book. A more thorough treatment of this subject can be found in many elasticity textbooks, such as [10, 11].
Stress
Stress is a measure of the intensity of force transmitted through a surface separating different parts of a body. The basic definition of stress is force per unit area. There are two types of stress: axial and shear, as shown in Fig. 2.1a, b, respectively. The axial stress is σ = P/A, where the force P is perpendicular to the surface area A. In other words, the force P acts along the surface normal. Hence the axial stress is also called the normal stress. The shear stress is τ = P/A, where the shear force P is parallel to the surface area A. The dimension of stress is f/l2, where f is the dimension of force and l is the dimension of length. The unit of stress is N/m2= Pa (pascal).
Stress as a second-rank tensor
To completely specify the stress state at a point, we need to consider a small cube around this point and specify the traction forces per unit area on all faces of this cube. The edges of the cube are chosen to be parallel to the axes of a given coordinate system. The positive faces of the cube are defined as the three faces whose outward normal vectors are along the positive x, y, and z axes, respectively. Because the size of the cube is vanishingly small, it suffices to specify the forces on the positive faces of the cube. The forces on the negative faces must be opposite to the forces on the corresponding positive faces.
The main purpose of this chapter is to introduce the geometrical properties of dislocations, the rules governing dislocation reactions, and the directions of dislocation motion in response to applied stress. The goal is to develop an intuitive understanding of the basic behaviors of dislocations without obtaining their stress field (which is the subject of the next chapter).
We start with Section 8.1 on why dislocations are necessary for plastic deformation of crystals. In Section 8.2, we introduce Volterra dislocations in an elastic continuum, and then describe the differences between them and dislocations in a crystal. In Section 8.3, we define the Burgers vector of a dislocation, and describe the geometric rule for Burgers vectors that must be satisfied when dislocations react. Section 8.4 shows which direction a dislocation should move on its glide plane under an applied stress. It also introduces cross-slip and climb, as alternative modes of dislocation motion. Section 8.5 describes where crystal dislocations come from. Severalmechanisms are presented in which the motion of existing dislocations can lead to multiplication, i.e. an increase of total dislocation length.
Role of dislocations in plastic deformation
We begin our study of dislocations by first thinking about plastic deformation in crystals – a problem that first led to the concept of crystal dislocations in the early 1930s. Although one's common experience with plastic deformation usually involves the continuous bending or stretching of a soft metal wire, the fundamental mechanism of plastic deformation is a shear process, as shown in Fig. 8.1.
The crystal, represented as a rectangular box, is plastically deformed in tension by sequential slip on various crystal planes. The bold lines indicate the active slip plane for that particular strain increment. Notice that the cumulative effect of these events is to make the crystal permanently longer and narrower. So themacroscopic shape change associated with ordinary tensile deformation is actually the cumulative effect of a large number of shear events. This can be confirmed by observing the surface of a plastically deformed metal crystal under an optical microscope, which reveals lots of surface steps, called slip traces. These are the intersection lines between slip planes and the sample surface.
The geometry of grain boundaries (GBs) can be specified with five degrees of freedom: three for the relative misorientation of the crystals and two for the direction of the boundary plane normal. Grain boundaries can be characterized as twist, tilt, or mixed depending on the relative orientation between the axis of rotation and the boundary plane normal.
The coincidence site lattice (CSL) theory describes special orientations between two lattices for which a fraction, 1/Σ, of the lattice points coincide. This leads to the designation of grain boundary misorientation by the Σ number. Special boundaries with low energies usually have low Σ numbers and appear as cusps in the plot of energy versus angle of misorientation.
The CSL theory predicts the vectors by which one lattice can be translated relative to the other while keeping the periodic coincidence pattern unchanged. These displacement vectors also form a lattice, called the displacement shift complete (DSC) lattice. The smallest repeat vectors of the DSC lattice are the Burgers vectors of GB dislocations. A crystal dislocation with an appropriate Burgers vector can spread out in the GB by dissociating into many GB dislocations with much shorter Burgers vectors and lower energies.
This chapter reviews the fundamental principles of thermodynamics and statistical mechanics, which are needed to derive the equilibrium distribution of point defects in a solid under external or internal stresses.
The first law defines the change in energy, E, a state variable, as the sum of the work and heat entering the solid. The second law introduces another state variable, the entropy, S, which may only increase in isolated systems and reaches a maximum at equilibrium. While the entropy can be described by considering heat into a solid, its physical meaning is clarified by Boltzmann’s entropy expression. The number of atoms in the solid, N, and the volume, V, are also state variables. The relation between these state variables, E(S,V,N), is called an equation of state.
The equation of state, E(S,V,N), can be rewritten into more convenient forms by Legendre transform, through which other thermodynamic potentials are defined, such as the enthalpy, H, the Helmholtz free energy, F, and the Gibbs free energy G. Intensive state variables, such as temperature, T, pressure, p, and chemical potential, μ, are defined as partial derivatives of the thermodynamic potentials.
Our study of grain boundaries to this point has focused on their geometry and special misorientations that lead to periodic patterns in the GB structure.We now turn to another important aspect of grain boundaries: their energies and possible elastic fields. A planar grain boundary usually does not have a long range stress field by itself. However, certain grain boundaries contain periodic dislocation arrays as part of their structures. In such cases, there is an appreciable stress field around the GB at distances comparable to the inter-dislocation spacing in the GB. The GB model based on dislocation arrays, combined with the theory of coincidence site and DSC lattices, provides a way to understand the GB energy as a function of its misorientation angle.
We have seen that the GB energy as a function of the misorientation angle has a complex structure, as shown in Fig. 13.2 and Fig. 13.8. Nonetheless, such plots suggest a classification of grain boundaries broadly into three types: singular, vicinal, and general [129]. The singular GBs correspond to the sharp minima on the energy plots. Their misorientations usually correspond to low-Σ CSLs. The singular GBs are usually special in other properties as well, such as mobility and point defect segregation. The vicinal GBs have both misorientation and GB plane direction sufficiently close to the singular GBs, and they can be considered as singular GBs superimposed with one or more GB dislocation arrays. The spacing between the nearest dislocations in the array reduces as the misorientation deviates further away from that of the singular GB. The general GBs are those boundaries that are sufficiently different from the singular GBs that the dislocation array is no longer a useful model as the necessary dislocation density would be so large that the dislocation cores would overlap.
In this classification of GBs, the case of zero misorientation and its vicinal range deserves extra attention. On the one hand, when the misorientation angle θ equals zero, the grain boundary disappears and the GB energy is zero, because the two crystals are perfectly aligned with each other.
A qualitative understanding of the behaviors of point defects can be established by considering atoms as hard spheres packed together to form the crystal. Crude as the hard sphere model may seem, it can be used to explain many of the observations made about point defects. In Section 4.1, we define the hard sphere radius of an atom and show its influence on the site preference of solute atoms. In Section 4.2, we use the hard sphere model to show the type of the distortions (spherically symmetric or not) in the host crystal around a solute atom. This allows us to explain why certain solutes have a much stronger solid solution hardening effect than others.
We then need to go beyond the hard sphere model in order to be more quantitative. In Section 4.3, we define the Seitz radius, which is more useful than the hard sphere radius for keeping track of the volume occupied by atoms of different kinds in solid solutions. We will see that atoms often appear to take on a different radius as a solute atom in another crystal compared to the radius it takes in its own crystal. In Section 4.4, we apply elasticity theory to predict the elastic fields around a solute atom. For simplicity, the size of the point defect is shrunk to zero and is modeled as force dipoles acting on a point in an elastic medium. In Section 4.5, a more realistic model is developed, in which the solute atom is modeled as an elastic sphere to be inserted into a hole inside an elastic medium. Elastic fields arise because the initial size of the sphere is larger than the initial size of the hole. Even though many atomistic and electronic details concerning point defects are ignored, the models developed in this chapter are increasingly more quantitative and can be used to explain a large number of behaviors of point defects.
Hard sphere model
Hard sphere radius
It is common to treat atoms in a crystal as undeformable spheres and to calculate the atomic sizes from the lattice parameters (measured using X-ray diffraction).We call this the hard sphere approach.
In our treatment of dislocations thus far, we have avoided the dislocation core. For example, in Volterra's dislocation model, the stress–strain fields diverge on the dislocation line, so that a cylindrical region of material is usually removed around the dislocation line to avoid the singularity. In the line tension model, the dislocation is modeled as a string that carries a line energy per unit length, but is otherwise featureless. In Chapter 11, we have seen that perfect dislocations in close-packed metals tend to dissociate into partial dislocations, but the partial dislocations were still treated as Volterra's dislocation lines. In reality, every (perfect or partial) crystal dislocation has a core region, which possesses a specific atomistic structure, called the core structure. The core structure is determined by non-linear interatomic interactions and the crystal structure, and, in turn, strongly influences the energetics and mobility of the dislocations. In this chapter, we discuss typical dislocation core structures and their effects on dislocation properties in several crystal structures.
In Section 12.1, we start our discussion with the classical Peierls–Nabarro (PN) model, which was the first physical model for the dislocation core and naturally predicts that the dislocation core should have a finite width. In Section 12.2, we generalize the original PN model to account for the presence of stacking faults in FCC metals. Consistent with the hard sphere model in Chapter 11, the generalized PN model also predicts dissociation of perfect dislocations into partials, except that each partial now has a finite width.
For crystals whose structures are sufficiently different from close-packed, hard spheres are no longer a good model for the atoms. Nonetheless, the geometry of the stacking of atomic layers is still useful for understanding the dislocation core structures in these crystals, as discussed in Section 12.3 (diamond cubic crystals) and Section 12.4 (BCC crystals). Finally, in Section 12.5 we discuss the interaction between dislocations and point defects, which usually leads to segregation of point defects around the dislocation core.
Peierls–Nabarro model
The classical model by Peierls and Nabarro [111, 112] considers the spreading of the dislocation over the glide plane.
Having discussed the elastic field around a single point defect, we now apply the thermodynamics principles (Chapter 3) to obtain the equilibrium concentration of point defects in crystals under a given temperature and pressure. The fundamental principle used repeatedly is that the Gibbs free energy of the crystal is minimized when the point defects reach the equilibrium concentration.
We start by discussing the equilibrium concentration of extrinsic point defects, i.e. substitutional and interstitial solutes, in Section 5.1. The approach is then applied, in Section 5.2, to vacancies, which are intrinsic point defects. In Section 5.3 we discuss the experimental methods to measure the equilibrium concentration and thermodynamic properties of vacancies, and compare the experimental data with theoretical estimates. The chemical potential of point defects is defined in Section 5.4.
Equilibrium concentration of solutes
We consider a dilute substitutional solution of B atoms in an A matrix. Let NA (which is fixed) be the number of A atoms in the system and let NB be the number of B atoms dissolved in the A-rich crystal. The total number of atomic sites in the A-rich crystal is N = NA + NB. Thus χ = NB/N is the fraction of atomic sites where the “wrong” kind of atom is located. χ is also the molar fraction of B atoms in the crystal.We follow the regular solution/quasi-chemical approach in which the formation energy of the point defect is dominated by the energies of the chemical bonds associated with the impurity defect (quasi-chemical, Eq. (5.5)) and where the mixing entropy is that for an ideal solution (regular solution, Eq. (5.19)).
Let the A-rich crystal be in contact with a large B crystal, which acts as an infinite supply of B atoms. For simplicity, we only allow B atoms to enter the A-rich crystal as solutes, but forbid A atoms to enter the B crystal as solutes. Each time a B atom is dissolved in the lattice, the A atom it replaces takes up a site at the A/B interface and extends the A lattice by one atomic volume, as shown in Fig. 5.1a. Note that the number of A atoms NA is conserved, while the number of B solute atoms NB and the total number of atomic sites N for the A-rich crystal are not conserved.
Our treatment of dislocations in previous chapters has focused on perfect dislocations, line defects that are surrounded by perfect crystal and have Burgers vectors equal to the shortest complete lattice translation vector. When perfect dislocations glide in a crystal they cause the atoms on either side of the slip plane to be displaced relative to each other by exactly a lattice translation vector, so that the crystal is perfect both ahead of and behind the gliding dislocation. Here we study the atomic motions associated with slip in more detail and observe that the sliding of atomic planes relative to each other often does not go immediately from one perfect state to the next. Instead, the slipping of atomic planes from one perfect state to the next may be broken up into two or more steps by the movement of partial dislocations separated by faults in the atomic stacking. We will see that partial dislocations and stacking faults in different crystal structures can be anticipated from the atomic packing arrangements in the crystal by treating the atoms as hard spheres and taking account of the bonding between them.
In Section 11.1, we consider the dissociation of perfect dislocations in FCC metals into Shockley partials, and obtain the equilibrium separation between the partials from a forcebalance analysis. We introduce Thompson's notation to conveniently label the Burgers vectors of various perfect and partial dislocations in FCC metals. We also examine non-planar dislocation structures such as the transient core structure during cross-slip of a screw dislocation and the stable core structure of a Lomer–Cottrell dislocation. Finally, we consider the Frank partial dislocation loop formed from the condensation of vacancies, and the transformation of the Frank partial loop into a perfect dislocation loop or a stacking fault tetrahedron.
In Section 11.2, we discuss the types of dislocations in HCP metals. Because the atomic arrangements on the basal planes of HCPmetals are very similar to those on the ﹛1 1 1﹜ planes in FCCmetals, perfect dislocations on the basal planes are also dissociated into Shockley partials in HCPmetals. In Sections 11.3 and 11.4, we discuss dislocations inCrCl3 andNi3Al, respectively, which are crystals formed by more than one chemical species.
In Table B.1 we present elastic constants, coefficient of thermal expansion, and melting temperature for a set of common crystalline solids of pure elements. Because for most materials each crystal grain is elastically anisotropic, isotropic elasticity is only an approximation. The values listed here correspond to the averaged elastic properties of a polycrystal consisting of a large number of randomly oriented single-crystalline grains.
There is a significant scatter in the averaged isotropic elastic constants in the literature. The scatter in the anisotropic elastic constants of single crystals is comparably less. Therefore, the values in the table are computed values from anisotropic elastic constants of single crystals. Specifically, the bulk modulus B and shear modulus μ are computed from the self-consistent averaging [151, 152] of single crystal values. Given B and μ, the Young's modulus is computed as E = 9Bμ/(3B + μ) and the Poisson's ratio as ν = (3B - 2μ)/(6B + 2μ).
The anisotropic elastic constants of single crystals, together with the thermal expansion coefficients and melting temperature are obtained from [153]. Both the elastic constants and coefficients of thermal expansion are room-temperature values. Given the experimental scatter and temperature dependence of these properties, an uncertainty of a few percent is to be expected for the values reported here. This should be sufficiently accurate for most calculations and estimates within the isotropic approximation. If very precise calculations are required, the original literature should be consulted and the temperature dependence of these properties should be taken into account. Three or more significant digits are often reported to provide self-consistency among B, E, μ, and ν, even though the data themselves are probably not accurate down to the last significant digit. The melting points are accurate to within 1 K.
A grain boundary (GB) is the interface between two single crystals (i.e. grains) of the same material with different orientations. We mentioned in Chapter 1 that most engineering metals and alloys are polycrystals, which are aggregates of a large number of single crystal grains separated by grain boundaries. Grain boundaries play an important role in a wide range of material behaviors and properties. For example, grain boundaries can increase the strength of metals by blocking the motion of lattice dislocations, leading to the Hall–Petch behavior, in which the strength of crystalline materials increases with decreasing grain size. We have mentioned in Chapter 7 that grain boundaries act as sources and sinks of vacancies, and that vacancy diffusion along GBs is the mechanism of Coble creep. Because impurities tend to segregate at grain boundaries, the chemical environment is often different at grain boundaries, which can be preferential sites for crack initiation or chemical attack. Grain boundary engineering [127], i.e. the keeping of “good” GBs and removal of “bad” GBs through processing, has led to significant improvements in material strength and corrosion resistance.
In Chapter 13 (Grain boundary geometry), we first define the five orientation variables that specify the misorientation between the two grains and the direction of the boundary plane normal.We then introduce the coincidence site lattice (CSL) and the associated Σ number that are very useful to characterize special (low energy) grain boundaries. We will see that special grain boundaries usually have low Σ numbers but low Σ numbers do not necessarily mean the boundaries are special. The displacement shift complete (DSC) lattice, of which the CSL is a sublattice, and prescribes the allowable Burgers vectors of GB dislocations, which can be much shorter than those of lattice dislocations.
In Chapter 14 (Grain boundary mechanics), we explore the relationship between grain boundaries and dislocations. First, low angle grain boundaries are equivalent to arrays of lattice dislocations, as described by the famous Read–Shockley model. Second, a grain boundary with misorientation vicinal to a low- Σ GB can be considered as a superposition of the low- Σ GB and an array of GB dislocations. Finally, we discuss disconnections, which are steps on the GBs with a non-zero Burgers vector content, i.e. they are simultaneously GB steps and GB dislocations.