Vous êtes sur la page 1sur 106

Natural Sciences Tripos Part IA

MATERIALS SCIENCE Course D: Mechanical Behaviour of Materials

Pl view Plan i of f a dislocation di l ti l loop

Name............................. College..........................
Dr Howard Stone Lent Term 2013-14 2013 14

IA

DH1

Course D: Mechanical Behaviour of Materials

DH1

Contents
1 Synopsis 2 Reading list 2.1 Text books . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2 Web-based resources . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3 Elastic deformation 3.1 Normal stress and normal strain . . . . . . . . . 3.2 Shear stress and shear strain . . . . . . . . . . . 3.3 Elastic deformation . . . . . . . . . . . . . . . . . 3.3.1 Youngs modulus . . . . . . . . . . . . . . 3.3.2 Shear modulus . . . . . . . . . . . . . . . 3.3.3 Poissons ratio . . . . . . . . . . . . . . . 3.3.4 Multiaxial stress states . . . . . . . . . . . 3.4 Elastic strain energy . . . . . . . . . . . . . . . . 3.5 Atomic picture of elasticity . . . . . . . . . . . . . 3.6 Thermal expansion . . . . . . . . . . . . . . . . . 3.7 Youngs modulus of composites: Rule of mixtures 3.7.1 Axial modulus . . . . . . . . . . . . . . . . 3.7.2 Transverse modulus . . . . . . . . . . . . 3.8 Materials in structures under stress . . . . . . . . 3.8.1 Deformation of an elastic beam . . . . . . 3.8.2 Second moments of area . . . . . . . . . 3 5 5 5 7 7 9 10 10 10 12 13 14 15 17 19 20 21 22 22 24 25 26 28 28 31 32 32 34 36 37 38 39 40 40 42 44 46 48 49 50 51 52

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

4 Plastic deformation: An introduction to dislocations 4.1 Estimate of the yield stress in a perfect crystal . . . . . . . . . . . . . . . . 4.2 Dislocations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2.1 Edge dislocations . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2.2 Screw dislocations . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2.3 Mixed dislocations . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2.4 Dislocation loops . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.3 Motion of dislocations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.4 Shear stress required to move a dislocation . . . . . . . . . . . . . . . . . . 4.5 Force on a dislocation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.6 Slip systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.7 Geometry of single crystal slip and Schmids law . . . . . . . . . . . . . . . 4.8 Determining the operative slip systems . . . . . . . . . . . . . . . . . . . . . 4.8.1 The OILS rule . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.9 Geometry as slip proceeds . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.10 The energy associated with a dislocation . . . . . . . . . . . . . . . . . . . 4.11 Dislocation interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.12 Dislocation reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.12.1 Interaction of dislocations on different slip systems: The Lomer lock 4.13 Dislocation generation: Frank-Read sources . . . . . . . . . . . . . . . . . . 4.14 Jogs and kinks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.14.1 The absorption of vacancies/ atoms . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . .

DH2

Course D: Mechanical Behaviour of Materials

DH2

4.15 Climb and cross slip of dislocations . . . . . . 4.16 Plastic deformation of metallic single crystals 4.16.1 hcp metals . . . . . . . . . . . . . . . 4.16.2 fcc metals . . . . . . . . . . . . . . . . 4.16.3 Polycrystalline metals . . . . . . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

53 55 55 56 57 58 58 59 60 60 62 66 67 68 69 72 74 76 80 81 82 83 83 84 85 87 87 88 89 90 91 93 94 95 97

5 Strengthening mechanisms 5.1 Forest hardening . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.2 Dislocation pile ups and the effect of grain size . . . . . . . . . . . 5.3 Solid solution strengthening . . . . . . . . . . . . . . . . . . . . . . 5.3.1 Substitutional solute atoms . . . . . . . . . . . . . . . . . . 5.3.2 Interstitial solute atoms . . . . . . . . . . . . . . . . . . . . 5.4 Precipitate hardening . . . . . . . . . . . . . . . . . . . . . . . . . 5.4.1 Stress required to bow dislocations between precipitates . 5.4.2 Transition from cutting to bowing and maximum hardening . 5.4.3 Changing strengthening mechanisms during age hardening 5.5 Partial dislocations and stacking faults . . . . . . . . . . . . . . . . 5.6 Order hardening . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.7 Plastic deformation by cooperative shear - Twinning . . . . . . . . 6 Fracture 6.1 Estimate of ideal fracture stress . . . . . . . . . . . . . . . 6.2 Grifth criterion . . . . . . . . . . . . . . . . . . . . . . . . 6.2.1 The strain energy release rate . . . . . . . . . . . 6.2.2 Fracture energy (crack resistance) . . . . . . . . . 6.2.3 Critical aw size and toughening by aw removal 6.3 Ductile fracture . . . . . . . . . . . . . . . . . . . . . . . . 6.3.1 Effect of plasticity on fracture energy . . . . . . . . 6.3.2 Stress intensity factor . . . . . . . . . . . . . . . . 6.3.3 Uniting the stress and energy approaches . . . . . 6.3.4 Ductile rupture . . . . . . . . . . . . . . . . . . . . 6.4 The Ductile-Brittle Transition Temperature (DBTT) . . . . 6.5 Toughness of composites . . . . . . . . . . . . . . . . . . 6.6 Pressurised pipes . . . . . . . . . . . . . . . . . . . . . . 6.7 Aircraft stresses and materials . . . . . . . . . . . . . . . 7 Appendix: Observing dislocations 8 Glossary

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . . . .

. . . . . . . . . . . . . .

. . . . . . . . . . . . . .

. . . . . . . . . . . . . .

. . . . . . . . . . . . . .

. . . . . . . . . . . . . .

. . . . . . . . . . . . . .

. . . . . . . . . . . . . .

. . . . . . . . . . . . . .

. . . . . . . . . . . . . .

. . . . . . . . . . . . . .

. . . . . . . . . . . . . .

DH3

Course D: Mechanical Behaviour of Materials

DH3

Synopsis

Elastic deformation
Denitions of stress and strain, both normal and shear. Denitions of elastic deformation, Youngs modulus, shear modulus, Poissons ratio. Multiaxial stress states. Elastic strain energy. The atomic picture of elasticity: bond stretching. Calculation of Youngs modulus from binding energy curve for a crystal. Typical values of elastic moduli. Thermal expansion and bimetallic strips. Elastic properties of composite materials: axial and transverse moduli using the slab model. Deformation of an elastic beam. Bending moments. Second moments of area. Effect of increasing stress on materials: the elastic limit, plastic ow, brittle and ductile fracture.

Plasticity
Why and how does plastic ow occur in crystals? Yield stress for different materials. Frenkel calculation of theoretical shear strength. Difference between ideal and actual yield stress explained in terms of dislocations. Geometry of an edge dislocation in a simple cubic structure. Denitions of line vector, Burgers vector and slip plane. Screw dislocations, mixed dislocations, dislocation loops. Observation of dislocations using transmission electron microscopy. Movement of dislocations. Atom movement at the core of a dislocation due to glide. Concept of lattice potential (Peierls-Nabarro energy) leading to Peierls-Nabarro stress. Width of a dislocation determined by bonding type (metallic, covalent, ionic). Expression for Peierls stress in terms of Burgers vector. Force on a dislocation due to shear stress. Slip in single crystals. Denition of slip systems. Slip systems for fcc, bcc, hcp metals and NaCl structures. Geometry of single crystal slip. Schmid factor. Example of calculation of system with the highest Schmid factor for fcc. The OILS rule. Geometry of a single crystal as slip proceeds. Sample elongation in single crystal slip. Rotation of the tensile axis. Dislocation interactions. Repulsion and attraction of dislocations. Energies of dislocations. Franks rule. Formation of a Lomer lock. Dislocation generation (Frank-Read sources). Jogs and kinks and their consequence on dislocation mobility. Cross slip of screw dislocations. Climb of edge dislocations. Intersection of propagating dislocations. Stress-strain curves for single crystals of hcp and fcc metals. Explanations for stages I, II and III. Duplex slip. Geometrical softening. Differences in polycrystalline metals. Taylor factor.

DH4

Course D: Mechanical Behaviour of Materials

DH4

Strengthening mechanisms
Factors which control yield stress: Forest hardening; typical values of dislocation density; dislocation pile ups and the effect of grain size (Hall-Petch relation); interactions between dislocations and solute atoms. Cottrell atmosphere formation with interstitial solutes. Luders bands and the Portevin-Le Chatelier Effect. Precipitates as obstacles to dislocations. Coherency strains and precipitate cutting. Orowan bowing. Stress to by-pass obstacles. Age hardening and precipitation sequences. Mechanisms of hardening. Partial dislocations and stacking faults. Order hardening and anti-phase boundaries. Deformation twinning as a mechanism of plastic deformation. Twinning elements and the twinning shear. Factors favouring Twinning. Effects of temperature and strain rate. Tin Cry. Morphology of deformation and annealing twins.

Fracture
Estimate of theoretical cleavage stress. Explanation of difference from observed values in terms of propagation of pre-existing defects. Grifth criterion for fracture. Strain energy release rates and fracture energies. Fracture of ductile solids. Concept of stress concentration. Stress intensity factor. Crack tip plasticity. Fracture toughness. Ductile tearing vs cleavage fracture. Designing against fast fracture. Pressurised thin walled tubes. Stresses in aircraft fuselages. Toughness of engineering materials and composites. Ductile/brittle transition temperature.

DH5

Course D: Mechanical Behaviour of Materials

DH5

2
2.1

Reading list
Text books

1. Hull, D. and Bacon, D.J., Introduction to Dislocations, Butterworth, 4th ed., (2001). 2. Hosford, W.F. Mechanical Behaviour of Materials, Cambridge University Press, (2010). 3. Ashby M.F. and Jones D.R.H., Engineering Materials : An Introduction to their Properties and Applications, Pergamon, 2nd ed., (1996). 4. Hull, D. and Clyne, T.W., Introduction to Composite Materials, Cambridge University Press, (1996). 5. Hamley, I.W., Introduction to Soft Matter, Wiley, (2000). 6. Callister, W.D., Materials Science and Engineering: An Introduction, Wiley, 6th ed., (2002).

2.2

Web-based resources

Most of the material associated with the course (handouts, question sheets, practical scripts etc) can be viewed on the web (www.msm.cam.ac.uk/Teaching) and also downloaded. This includes model answers, which are released after the work concerned should have been completed. In addition to this text-based material, resources produced within a project based in the Materials Science Department, called DoITPoMS (Dissemination of Information Technology for the Promotion of Materials Science) are also available (www.doitpoms.ac.uk). These include libraries of Micrographs, Videos and Teaching and Learning Packages (TLPs). The following TLPs are relevant to course D: Thermal Expansion and the Bimaterial Strip Introduction to Dislocations Slip in Single Crystals Introduction to Mechanical Testing Hardness and Indentation Brittle Fracture Mechanics of Fibre-Reinforced Composites Bending and Torsion of Beams

DH6

Course D: Mechanical Behaviour of Materials

DH6

DH7

Course D: Mechanical Behaviour of Materials

DH7

3
3.1

Elastic deformation
Normal stress and normal strain

Consider the effect of forces acting on the surfaces of a cuboid. If a force acts perpendicular to a surface we have a normal stress, denoted by . With equal and opposite forces acting over two opposing faces, as shown in Fig. 1, the cuboid will remain stationary (will not accelerate).
Area (A)

Force (F)

Force (F)

Stress () =

Force (F) Area (A)

Figure 1: A cuboid subjected to forces acting perpendicular to two opposite faces generating a normal stress, .

Tensile stress (by convention positive). Compressive stress (by convention negative). The stress is dened as the force acting per unit area. true = F A (3.1)

in which F is the magnitude of the force and A is the area of the face. This denition of stress is referred to as the true stress. As the cuboid elongates, conservation of volume would suggest that the area of the cuboid faces over which the forces act will decrease. To avoid having to account for this effect, it is often more convenient to dene the force as acting over the original area, A eng = F A (3.2)

This is referred to as the nominal stress or engineering stress. In practice, the difference between the nominal stress/engineering stress and the true stress is usually small and is often ignored.

DH8

Course D: Mechanical Behaviour of Materials

DH8

The extension of the cuboid under the action of the stress is described by the strain, denoted by . The normal strain is the relative change in linear dimension in the direction of a normal force. If the cuboid extends by a length L in the direction of the applied force, from its initial length of L0 to L1 , as shown in Fig. 2, then the normal strain, is dened by
L0 Force (F) Force (F)

L1

Strain () =

L1 - L0 L0

Figure 2: Extension of a cuboid subjected to forces acting perpendicular to two opposite faces .

L (3.3) L If the strain is nite, this may be integrated to obtain the true strain or logarithmic strain. =
L1

true =
L0

L1 dL = ln L L0

(3.4)

For small strains, the strain is reasonably approximated by the nominal strain or engineering strain , which is taken to be L1 L0 L1 eng = = 1 (3.5) L0 L0 The distinction between nominal strain and true strain only becomes important for strains in excess of a few percent. Small Strains (up to a few %) True Stress Nominal Stress Large Strains True Stress > Nominal Stress

DH9

Course D: Mechanical Behaviour of Materials

DH9

3.2

Shear stress and shear strain

If a force acts parallel to the surface of the cuboid we have a shear stress, denoted by . As before, this is considered to act over the surface of the cuboid. To ensure that the cuboid does not accelerate (rotate) we must dene two pairs of opposing forces acting over 4 faces of the cuboid, as shown in Fig. 3.
Area (A0) original section deformed section y0 Force (F) x0 rotated deformed section

shear stress () =

Force (F) Area (A0)

shear strain =

y0 x0

= tan =

Figure 3: A cuboid subjected to forces acting parallel to two opposite faces of the cuboid generating a shear stress, .

As before, the shear stress is dened as the force acting per unit area. = F A (3.6)

The distortion of the cuboid under the action of the shear stress is shown in Fig. 3. There is an elongation of the cuboid along one face diagonal accompanied by a contraction along the other. This distortion can be described by the shear strain, denoted by , which is the ratio of the distances, y and x . The associated angle through which the cuboid has been sheared is typically referred to as the angle of shear, . = y x = tan (3.7) (3.8)

Normal Stresses & Strains +ive (tensile) or ive (compressive) physically different Shear Stresses & Strains +ive or ive, but depends only on convention used for +ive sense of reference directions no physical difference General stress states more complicated - but can always be resolved into shear stresses and normal stresses. Stresses & Strains Two directions associated with each value. Hence they are NOT vectors (one direction only). Stresses & strains cannot be resolved as if they were vectors (ie with a single cosine term). They need two cosine terms - see www.doitpoms.ac.uk/tlplib/tensors/index.php.

DH10

Course D: Mechanical Behaviour of Materials

DH10

3.3
3.3.1

Elastic deformation
Youngs modulus

For an ideal, elastic, isotropic material (no orientational dependence of the material properties) the normal strain, , that we measure will be proportional to the applied normal stress, such that = E (3.9) in which the constant, E , is called the Youngs modulus.

3.3.2

Shear modulus

Similarly, the shear strain, , will be proportional to the applied shear stress, , such that = G in which G is the Shear modulus. It should be noted that G and E are different as they describe the response of a material to different types of deformation. However, for many materials G 0.4E . Youngs modulus is related to Hookes law - Force is proportional to displacement (F = kL) Stiffness, k , is an extensive property (depends on the material properties and the shape) Youngs modulus, E , is an intensive property (depends on the material only) More generally useful. k can be related to E = E F L = E A L EA F = L F = kL (3.10)

DH11

Course D: Mechanical Behaviour of Materials

DH11

Figure 4: Youngs modulus values for various materials, sub-divided according to type (from: M.F. Ashby and D.R.H. Jones, Engineering Materials 1, Butterworth-Heinemann, 2001).

DH12

Course D: Mechanical Behaviour of Materials

DH12

3.3.3

Poissons ratio

For most materials, the application of a tensile normal stress would be expected to lead to a contraction in the perpendicular directions, as shown in Fig. 5.
Force (F)

Force (F)

Figure 5: Lateral contraction of a cuboid in directions x and y under the action of a tensile normal stress applied along direction z .

For an isotropic material the strains in the two directions perpendicular to the tensile normal stress will be the same. We dene the relationship between the tensile normal stress (acting along direction z ) and the normal strains (in directions x and y ) to be x = y = z = in which is the Poissons ratio. Poissons ration is dimensionless. It can take values between -1 and 0.5. The conservation of volume would suggest that Poissons ratio should be 0.5 (K=). In practice, Poissons ratios between 0.2 and 0.5 are common (most metals have Poissons ratios around 0.3) z E (3.11)

DH13

Course D: Mechanical Behaviour of Materials

DH13

3.3.4

Multiaxial stress states

If an elastic, isotropic material is subjected to normal stresses in three perpendicular directions, x , y and z , as shown in Fig. 6, the resultant normal strains in the directions, x, y and z will be given by x = y = z = 1 (x (y + z )) E 1 (y (x + z )) E 1 (z (x + y )) E (3.12) (3.13) (3.14)

x
z

y x

Figure 6: A cuboid subjected to three perpendicular normal stresses.

Some special cases: Plane stress - One of the normal stresses is zero (e.g. stretching a thin sheet.) Plane strain - One of the normal strains is zero Hydrostatic stress - All three normal stresses have the same magnitude.

DH14

Course D: Mechanical Behaviour of Materials

DH14

3.4

Elastic strain energy

When a body is elastically deformed the work done is stored as elastic strain energy. If we consider a cylinder extended by a length L under the action of opposing forces, F , the work done on the cylinder is F L.
L F L F

A
Figure 7: A cylinder extended under the action of opposing forces.

From our denition of normal stress, = F/A. Hence, F = A Similarly, from our denition of normal strain, = L/L. Hence, L = L Hence the work done on the cylinder is work done = F L = (A) (L) (3.15)

Assuming the volume of the cylinder, V = AL is invariant and substituting = E, Equation 3.15 may be integrated between 0 and max
max

work done =
0

V E d 1 V E2 max 2

(3.16) (3.17)

= Hence

2 the work done per unit volume under a normal stress is 1 2 E

This energy is stored as potential energy in strained interatomic bonds and is also referred to as the strain energy density. It can be similarly shown that
2 the work done per unit volume under a shear stress is 1 2 G 1 As = E, elastic strain energy can also be written as 2 .

This is the area under a curve.


1 . The equivalent for shear strain energy is 2

DH15

Course D: Mechanical Behaviour of Materials

DH15

3.5

Atomic picture of elasticity

The variation of potential energy, U , with distance, r, between two atoms, and the resultant force, is shown schematically in Fig. 8.
Energy, U

Inter-atomic spacing, r

ro
F (=dU/dr)

Figure 8: Variation of potential energy with inter-atomic spacing.

At very large distances the potential energy is approximately zero. As the atoms approach one another the potential energy falls, reaching a minimum of Umin at a distance of r , after which the potential energy rises, ultimately becoming positive at small distances. This variation in potential energy with distance can be reasonably approximated by the Lennard - Jones potential , ULJ ULJ = Umin r r
12

r r

(3.18)

As the atoms are displaced from their equilibrium separation of r they will experience a restoring force. F = dU dr (3.19)

For small displacements from r the variation in the potential energy is approximately linear (see Fig. 8). Hence, the force will be approximately proportional to the displacement - this is the origin of Hookes law. We can therefore estimate the Youngs modulus from the LennardJones potential.

DH16

Course D: Mechanical Behaviour of Materials

DH16

Consider a simple cubic crystal. If we stretched the crystal parallel to [100] we will be stretching a square array of atomic bonds as shown in Fig. 9

z y x

Figure 9: A simple cubic crystal - consider the square array of bonds identied by the dotted lines.

Prior to loading, we may reasonably expect the atoms to be r apart. The area occupied by 2. each bond in the plane perpendicular to the applied normal stress is r The stress on each bond, , is therefore, = F 1 dU = 2 2 r r dr (3.20)

By denition (from Eq. 3.3), the strain in the bond, , is = The Youngs modulus is therefore E= d dr d = = d dr d 1 d2 U 2 dr 2 r (r )
r

r r

(3.21)

(3.22)

The Youngs modulus can therefore be obtained from the interatomic potential by 1 d2 U r dr2

E=

(3.23)
r

DH17

Course D: Mechanical Behaviour of Materials

DH17

3.6

Thermal expansion

Temperature becomes a quantity denable either in terms of macroscopic thermodynamic quantities such as heat and work, or, with equal validity and identical results, in terms of a quantity which characterised the energy distribution among the particles in a system. (Quinn, T. J. Temperature 1990 Academic Press ISBN 0-12-569681-7) The temperature of any system, T , can be related to the mean kinetic energies of the atoms it contains according to 3 1 kb T = mv 2 (3.24) 2 2 where kb is Boltzmanns constant At higher temperatures, the greater kinetic energy of the atoms will allow them to move further away from their equilibrium positions. The asymmetry of the potential energy vs distance curve results in the mean position increasing for oscillations of higher energy. This is shown schematically in Fig. 10. This give rise to thermal expansion.
Energy, U

Inter-atomic spacing, r

rhigh T rlow T ro

Thermal Expansion

Figure 10: Variation of potential energy with inter-atomic spacing, illustrating the effect of temperature on the mean atomic position.

Macroscopically we dene the relationship between a temperature change T and the accompanying increase in thermal strain to be thermal = T where is the coefcient of thermal expansion or thermal expansivity. In general, materials with stronger interatomic bonds (e.g. ceramics) show low coefcients of thermal expansion as well as high Youngs moduli. The converse is true of weakly bonded materials (e.g. polymers). Examples of the temperature dependence of the coefcients of thermal expansion of a range of materials are given in Fig. 11 (3.25)

DH18

Course D: Mechanical Behaviour of Materials

DH18

linear coefficient of thermal expansion, ([strain]/C)

10

polyester 10
2

epoxy aluminium

10

titanium SiC Al2O3

10

SiO2

10

50

100

150

200

temperature/C

250

300

350

400

450

500

Figure 11: Temperature dependence of the coefcients of thermal expansion of selected materials.

DH19

Course D: Mechanical Behaviour of Materials

DH19

3.7

Youngs modulus of composites: Rule of mixtures

Composites Metals typically combine both high strength and damage tolerance. However, they have a number of drawbacks. In particular, they are often quite dense and may corrode. Ceramics and polymers are generally lighter and more corrosion-resistant. Regretably, ceramics and polymers do not have much damage tolerance. However, excellent properties can be obtained by combining different materials to produce composites, usually with one constituent in bre form. Ceramics and polymers are often much stronger when produced as ne bres. For ceramics, this strengthening arises from removal of aws. This is discussed in greater detail in Section 6. In polymers it is achieved by molecular alignment. Three main types of bre are used in composites: glass, carbon and aramid (aromatic polyamides, such as KevlarTM ). These bres are embedded in a matrix - usually a polymer and commonly a thermosetting resin, such as an epoxy or a polyester. In some types of composite, the bres are oriented randomly within a plane, while in others the material is made up of a stack of differently-oriented plies, each containing an aligned set of parallel bres.

Stiffness of Composites The materials used for the bres and matrices of composites typically have very different Youngs moduli. As such, the composite will exhibit highly anisotropic elastic behaviour depending upon whether it is loaded parallel or perpendicular to the bres. The Youngs moduli in these two cases can be reasonably predicted using a simple slab model and the rule of mixtures. These are shown in Fig. 12.

1-Vf

f
(a) (b)

Vf

1-Vf Vf

m f

(c)

Figure 12: A unidirectional long bre composite material and simplied slab models for (b) axial deformation and (c) transverse deformation.

DH20

Course D: Mechanical Behaviour of Materials

DH20

3.7.1

Axial modulus

The axial modulus may be obtained assuming the slabs shown in Fig. 12(b) are subjected to equal strain. This is known as the Voigt Model. Using the subscripts, c, f and m to refer to the composite, bres and matrix respectively. The stress experienced by the composite will be equal to the sum of the stresses experienced by the bres and matrix weighted by their area/volume fractions

c = f Vf + m Vm c = f Vf + m (1 Vf ) Given that = E Ec c = Ef f Vf + Em m (1 Vf ) Using the equal strain assumption, c = f = m Ec = Ef Vf + Em (1 Vf ) Hence, the axial stiffness of a unidirectional bre composite is

(3.26) (3.27)

(3.28)

(3.29)

Ec = Ef Vf + Em (1 Vf ) Axial - Rule of mixtures provides a reasonable estimate of the composite stiffness

(3.30)

DH21

Course D: Mechanical Behaviour of Materials

DH21

3.7.2

Transverse modulus

The transverse modulus may be obtained assuming the slabs shown in Fig. 12(c) are subjected to equal stress. This is known as the Reuss Model. The strain experienced by the composite will be equal to the sum of the strains experienced by the bres and matrix weighted by their area/volume fractions c = f Vf + m Vm c = f Vf + m (1 Vf ) Given that = /E f c m = (1 Vf ) Vf + Ec Ef Em (3.31) (3.32)

(3.33)

Using the equal stress assumption, c = f = m 1 Ec 1 Ec = = 1 1 (1 Vf ) Vf + Ef Em Em Vf + Ef (1 Vf ) Ef Em (3.34) (3.35)

Hence, the transverse stiffness of a unidirectional bre composite is Ef Em Em Vf + Ef (1 Vf )

Ec =

(3.36)

Transverse modulus / Inverse rule of mixtures is substantially lower that the axial value. Critically, it is not very accurate as parts of the matrix are shielded from the applied stress by being in parallel with the bres

DH22

Course D: Mechanical Behaviour of Materials

DH22

3.8

Materials in structures under stress

The effective design of structures capable of withstanding mechanical loads involves calculations of the stresses and strains as well as an understanding of how the materials will deform and fail. Today, the stresses that develop in engineering structures may be calculated using well established and powerful numerical models. These may be used without deep background knowledge. However, materials science is still a relatively new discipline and there have been, and continue to be, many examples of structures failing because of a poor choice of material, an inadequate understanding of material behaviour, or simply failing to understand the stresses and strains generated in the structure.

3.8.1

Deformation of an elastic beam

Consider a uniform beam bent under a bending moment, M , as shown in Fig. 13. In bending, the top surface has become longer (in tension) whilst the bottom surface has become shorter (in compression). Along the mid thickness of the beam, the length is unchanged - this is termed the neutral axis.

l y
neutral axis

R y
Figure 13: Strains induced during bending of a beam by the application of a moment, M .

The axial strain at a distance of y from the neutral axis is given by axial = extension (R + y ) R y = = original length R R (3.37)

DH23

Course D: Mechanical Behaviour of Materials

DH23

Hence, the axial stress is Ey (3.38) R The force on a strip of thickness dy and width b at a distance y from the neutral axis is axial = Eaxial = F = A = Ey R (b dy ) (3.39)

The bending moment about the neutral axis associated with this force is Bending moment = y Hence, the total bending moment is M= E R
h/2

Ey R

(b dy )

(3.40)

y 2 bdy
h/2 h/2

(3.41)

For any section, we can dene the second moment of area, I = h/2 y 2 bdy . Hence the total bending moment may be written in terms of the Youngs modulus, E , the second moment of area, I and the radius of curvature, R, or the curvature, = 1/R EI = EI R

M=

(3.42)

The product EI is termed the beam stiffness or sometimes as exural rigidity.

DH24

Course D: Mechanical Behaviour of Materials

DH24

3.8.2

Second moments of area

For a general shape, the second moment of area about a neutral axis is dened as I=
section

y 2 b{y }dy

(3.43)

Large values of I will give a stiff beam. This is most readily achieved with sections that have the greatest proportion of their sections at large distances from the neutral axis, hence the use of I-beams, box- and tube- sections. The formulae giving the second moment of area for a wide variety of common shapes are given in many books.

3 I = wh 12

4 I = D 64

4 4 I= (D -d ) 64

DH25

Course D: Mechanical Behaviour of Materials

DH25

Plastic deformation: An introduction to dislocations

If a stress below the yield stress is applied, the material deforms elastically, i.e. the material is able to return to its original shape after the stress is removed. With larger stresses, unrecoverable or plastic deformation occurs, i.e. it yields. The stresses at which materials yield vary widely. They are typically in the range 30-1000 MPa for most engineering materials but can be greater than 1 GPa or as low as 1 MPa.

Figure 14: Yield stresses for various materials, sub-divided according to type (from: M.F. Ashby and D.R.H. Jones, Engineering Materials 1, Butterworth-Heinemann, 2001).

DH26

Course D: Mechanical Behaviour of Materials

DH26

4.1

Estimate of the yield stress in a perfect crystal

Plastic deformation usually occurs by atomic planes sliding over each other, under the inuence of shear stresses. Note that when a normal stress is applied, shear stresses are generated on planes inclined at an angle to the stress axis, as shown in Fig. 15.

max /2

crit

Figure 15: Shear stresses generated in a material subjected to a normal stress and the associated slip of atomic planes over one another under the action of the shear stresses.

To estimate the yield stress of a perfect crystal, consider two close-packed planes with separation h and interatomic spacing b sliding over one another by a distance of u under the action of a shear stress, , as shown in Fig. 16. We can also imagine that, for small values of u, the deformation will be elastic. We need to understand how varies with u. It is clearly going to be periodic with the distance moved, u, and will be zero when u = 0, u = b/2, u = b, etc.
u=0 h b u = 0.5b u=b

Figure 16: Movement of a close packed plane past another close packed plane from positions, u = 0 to u = b.

DH27

Course D: Mechanical Behaviour of Materials

DH27

If we assume it is sinusoidal then we can write = C sin in which C is a constant. From the relationship between shear stress and shear strain given in Eqn. 3.10, = G . In this case, the shear strain is = u/h From Equation 4.44, for small u/b C Hence, C= and Equation 4.44 becomes = 2u Gb sin 2h b (4.48) Gb 2h (4.47) 2u 2 (h ) =C = b b 2Ch b (4.46) (4.45) 2u b (4.44)

For slip to occur we must apply a critical shear stress, crit , that will be sufcient to overcome the maximum resistance that will be encountered. Hence,

crit =

Gb 2h

(4.49)

For close packed spheres,

b 2 G G = and hence, crit = h 5 3 3

Note that for many materials G 0.4E . This is 1000 larger than the shear stresses actually required to deform real materials!

DH28

Course D: Mechanical Behaviour of Materials

DH28

4.2
4.2.1

Dislocations
Edge dislocations

A dislocation is a line defect in a crystal in which the atoms are systematically displaced from their ideal positions. Arguably the simplest dislocation to visualise is an edge dislocation. An example of an edge dislocation is shown in Fig. 17. This can be simply considered as an extra half plane of atoms in an otherwise perfect crystal structure. The line of the dislocation can easily be seen at the base of the extra half plane, going into the page.

Figure 17: Schematic representation of an edge dislocation in a simple cubic material.

If a dislocation is present in a crystal, slip can occur by the progressive movement of the dislocation. This motion is analogous to the movement of a caterpillar, as shown in Fig. 18. A better analogy may be moving a carpet by moving a ruck across it rather than trying to drag the whole carpet. As only a few bonds are being distorted at any point in time by the progressive motion of the dislocation, the shear stress required to move a dislocation is signicantly lower than that required to shear the whole crystal at once.

Figure 18: Movement of an edge dislocation under the action of a shear stress

An animation of the glide motion of an edge dislocation can be seen in a DoITPoMS TLP at www.doitpoms.ac.uk/tlplib/dislocations/dislocation glide.php.

DH29

Course D: Mechanical Behaviour of Materials

DH29

Geometry of a dislocation
For any dislocation we can dene a dislocation line vector, l. In an edge dislocation this corresponds to the base of the extra half plane of atoms. Perhaps more usefully, we can dene the dislocation line vector to be the line that separates the slipped from the unslipped regions of the crystal. Note that the dislocation line cannot end within the crystal (but it can form loops). The lattice displacement caused by the dislocation can be described by its Burgers vector, b. The Burgers vector may be obtained by constructing a Burgers circuit around the dislocation core. This is shown schematically in Fig. 19.

(a)
E A

(b)
b A
E

D B D B

C C

Figure 19: A Burgers circuit around an edge dislocation in a simple cubic crystal. The Burgers vector, b is dened by the closure failure, EA.

The Burgers vector is obtained by making integer steps between lattice points in the positive and negative directions as you go around the dislocation core in a right-handed sense. If the same number of steps in the same directions is performed in a perfect crystal you will NOT end up back where you started. This leads to a closure failure. The vector dening the closure failure, EA denes the Burgers vector, b. For an edge dislocation, the Burgers vector is perpendicular to its line vector. A dislocation can only slip in the plane containing both its Burgers vector and line vector. This is termed the slip plane. In the case of an edge dislocation, as the Burgers and line vectors are perpendicular to one another there is only ever one plane in which an edge dislocation can glide.

DH30

Course D: Mechanical Behaviour of Materials

DH30

b
Figure 20: Plan view of an edge dislocation in a simple cubic crystal.

DH31

Course D: Mechanical Behaviour of Materials

DH31

4.2.2

Screw dislocations

There are other types of dislocation. The other simple one is the screw dislocation, see Fig. 21. A screw dislocation can be considered as a helical spiral defect in the crystal. The axis of the helix denes the associated dislocation line vector, l, as well as the associated Burgers vector, b (i.e. b l).

(a)

(b)
A b G E D F A D E F B B C C G

Figure 21: (a) A screw dislocation in a simple cubic crystal. The Burgers circuit is shown. (b) Burgers circuit in a perfect crystal indicating the Burgers vector associated with the screw dislocation (GA).

Figure 22: Plan view of a screw dislocation in a simple cubic crystal.

Critically, because the Burgers vector and line vector of screw dislocations are parallel, screw dislocations may glide on a family of possible slip planes. In the same way that the movement of an edge dislocation caused shear of the crystal, (see Fig. 18), passage of a screw dislocation will shear a crystal.

DH32

Course D: Mechanical Behaviour of Materials

DH32

4.2.3

Mixed dislocations

Dislocations may also be intermediate between edge and screw and are termed mixed dislocations. The relative orientations of the line and Burgers vectors of edge, screw and mixed dislocations are shown in Fig. 23
l slipped unslipped b edge (a) screw (b) slipped l unslipped b mixed (c) slipped l unslipped b

Figure 23: Relative orientations of (a) edge, (b) screw and (c) mixed dislocations.

4.2.4

Dislocation loops

It is also possible to have a dislocation loop, as shown in Fig. 24. As before, the dislocation line separates the slipped from the unslipped crystal. The Burgers vector is the same all around the dislocation loop but as the line vector changes from being parallel to the Burgers vector to being perpendicular to the Burgers vector so the character of the dislocation changes from screw to mixed to edge and back again.
edge

mixed

mixed

unslipped
screw

slipped l b
mixed edge mixed screw

unslipped

Figure 24: A dislocation loop. The Burgers vector is constant all around the loop whilst the character changes between screw, mixed and edge.

DH33

Course D: Mechanical Behaviour of Materials

DH33

Figure 25: A plan view of a dislocation loop in a simple cubic crystal.

DH34

Course D: Mechanical Behaviour of Materials

DH34

4.3

Motion of dislocations

The motion of an edge and a screw dislocation are shown in Fig. 26. Note that in both cases the dislocation moves perpendicular to the dislocation line vector and the resultant slip of the crystal is equal to the Burgers vector of the dislocation. However, in the case of an edge dislocation the dislocation line lies perpendicular to the applied shear stress, whilst for a screw vector the dislocation line lies parallel to the applied shear stress

(a)

(b)

(c)

pe p i l S d

ipp l s n

ed

Figure 26: Schematic representations of the motion of (a) an edge dislocation and (b) a screw dislocation under an applied shear stress leading to shear of the crystal (c).

Note that for the crystal structure to be recovered after the passage of a dislocation the Burgers vector MUST correspond to a vector between lattice points. These are termed perfect dislocations. Other dislocations, not corresponding to lattice vectors, may exist but these will not recover the original crystal structure in their wake.

DH35

Course D: Mechanical Behaviour of Materials

DH35

Motion of dislocation loops Under the action of a shear stress applied parallel to the Burgers vector a dislocation loop, like that shown in Fig. 24, will expand (or contract) in all directions. This can be understood by considering that the edge segments of the dislocation loop will move out in the direction of the applied shear stress whilst the screw line segments will move out sideways, perpendicular to the applied shear stress. Note that the top and bottom (and left and right) segments move in opposite directions because, whilst they have the same Burgers vector dened with reference to the direction of the dislocation line marked on the diagram, they would have Burgers vectors of opposite sense if Burgers circuits were drawn around them in the same direction.

edge

unslipped
screw

slipped
screw

unslipped

b
edge

Figure 27: Schematic representations of the motion of a dislocation loop.

DH36

Course D: Mechanical Behaviour of Materials

DH36

4.4

Shear stress required to move a dislocation

As a dislocation is moved, bonds are required to break and reform. As such, a shear stress sufcient to overcome the associated energy barrier must be applied. This must move a dislocation beyond its highest energy conguration.

Figure 28: (a) lowest and (b) highest energy congurations of an edge dislocation moving in a simple cubic crystal.

The energy barrier to move a dislocation from one position to the next is called the PeierlsNabarro energy and the associated shear stress is called the Peierls-Nabarro stress, P . The Peierls-Nabarro stress cannot be easily calculated as it is very sensitive to the nature and directionality of the interatomic forces. However, a simple model gives P 3G exp 2w b (4.50)

where w is the dislocation width. This function is strongly dependent upon w/b. The width of a dislocation is the distance over which atoms are signicantly displaced from their perfect crystal positions. This can be taken to be the region in which the atoms are displaced by b/4. Note that at the core of the dislocation, the atoms are displaced by b/2. Schematic illustrations of wide and narrow edge dislocations are shown in Fig. 29.

0.25b

0.25b

0.25b

0.25b

Figure 29: Atomic disregistry in (a) wide and (b) narrow edge dislocations.

DH37

Course D: Mechanical Behaviour of Materials

DH37

4.5

Force on a dislocation

The force acting on a dislocation as a result of an applied shear stress may be obtained by considering the work done in moving the dislocation - see Fig. 30

Figure 30: Movement of an edge dislocation with Burgers vector, b, under the action of an applied shear stress.

If an edge dislocation of length, L experiences a force acting per unit length, F , then the total force acting over the dislocation will be F L. The work done in moving the dislocation through the crystal a distance d is therefore F Ld. This must be equal to the work done by the shear stress, , which acts over the entire area of the slip plane, Ld. Thus the work done in shearing the crystal a distance b is Ldb. Hence, Ldb = F Ld (4.51) Therefore, the force per unit length acting on a dislocation as a result of an applied shear stress is F = b Applies to all types of dislocation NB dot product = b only if acts to b (4.52)

Force always acts normal to the dislocation line ( to l)

DH38

Course D: Mechanical Behaviour of Materials

DH38

4.6

Slip systems

If we deform a single crystal of a material, we usually nd that it deforms on a specic set of parallel planes and in a specic set of directions within these planes. The combination of the plane in which slip occurs (slip plane) and the direction (slip direction) together make up the slip system. The slip system that operates in a given material is governed by the crystal structure and the nature of the interatomic bonds. These are most commonly, but not always, the closest packed planes and the closest directions within those planes. For example, in a cubic close-packed metal, slip invariably occurs on the close-packed planes, {111}, and in the close-packed directions in those planes, 1 10 .
Table 1: Slip systems in some common materials

Material fcc metals bcc metals C, Si, Ge (diamond structure) NaCl* CsCl* hcp metals * and other ionic crystals with this structure

Slip systems {111} 1 10 {110} 111 (others occur at high T) {111} 1 10 {110} 1 10 {110} 001 {001} 100 (others can also occur)

Figure 31: Slip systems and Burgers vectors of common crystal structures

Burgers vector of a (perfect) dislocation must be a lattice vector Other defects may exist but these lead to defects

DH39

Course D: Mechanical Behaviour of Materials

DH39

4.7

Geometry of single crystal slip and Schmids law

If a stress is applied at an arbitrary angle to a single crystal, as shown in Fig. 32, each available slip system will experience a resolved shear stress acting in the associated slip plane in the slip direction.
Area (A) Force (F)

slip plane normal (n)

slip direction

Force (F)

Figure 32: Geometry of slip during tensile testing of a single crystal.

If the normal to the slip plane makes an angle with the tensile axis then the area of the plane is A/ cos . If the slip direction is at an angle of to the tensile axis then the resolved component of the applied force, F , parallel to the slip direction is F cos . The resolved component of the shear stress on the slip plane acting parallel to the slip direction, R , is therefore R = = F cos A/ cos F cos cos A (4.53) (4.54)

For a given material we nd experimentally that the value of R at which slip occurs is constant. This value is called the critical resolved shear stress, c . This is Schmids law. The quantity cos cos is called the Schmid factor. The relationship between the yield stress, y , at which plastic deformation initiates and the critical resolved shear stress is therefore (4.55)

c = y cos cos Schmid factor Note the two cosine terms

(Both the area of the slip plane & the component of force in the slip direction change as different slip systems are considered)

DH40

Course D: Mechanical Behaviour of Materials

DH40

4.8

Determining the operative slip systems

During deformation, all of the available slip systems will experience a resolved shear stress. As the critical resolved shear stress is the same for all slip systems, so we would expect the one with the greatest Schmid factor to yield rst. If we know the orientation of the tensile axis we can calculate the Schmid factor of all of the slip systems and identify the largest Schmid factor. This is effective but time consuming. For bcc and fcc structures there is a simple method that identies the slip system with the greatest Schmid factor very easily, called the OILS rule.

4.8.1

The OILS rule

For fcc metals, the slip systems are {111} 1 10 . For bcc metals, the slip systems are {110} 1 11 Using the OILS rule: 1. Write down the indices of the tensile axis [U V W ]. 2. Ignoring the signs, identify the Highest, Intermediate and Lowest valued indices. 3. The slip direction in a fcc crystal is the 110 direction with zero in the position of the Intermediate index and the signs of the other two preserved. or The slip plane in a bcc crystal is the {110} plane with zero in the position of the Intermediate index and the signs of the other two preserved from the tensile axis. 4. The slip plane in a fcc crystal is the {111} plane with the sign of the Lowest valued index reversed and the signs of the other two preserved from the tensile axis. or The slip direction in a bcc crystal is the 111 direction with the sign of the Lowest valued index reversed and the signs of the other two preserved. The mnemonic OILS stands for zerO Intermediate, Lowest Sign.

Example of OILS rule: If the tensile axis is parallel to the [214] direction of a fcc crystal, the sequence is [ILH]. As the slip systems for fcc metals are {111} 1 10 zerO Intermediate gives the slip direction as being [011] and Lowest Sign reversed gives the slip plane as being ( 1 11) Hence, in this case, the slip system with the largest Schmid factor is ( 1 11)[011].

DH41

Course D: Mechanical Behaviour of Materials

DH41

Schmid factor calculation for the example above cos can be obtained from the dot product of the normal to the slip plane and the tensile axis. 1 2 1 1 1 4 5 cos = = = 0.630 (4.56) 3 21 63 Similarly cos can be obtained from the dot product of the slip direction and the tensile axis. 0 2 1 1 1 4 5 cos = = = 0.772 (4.57) 2 21 42 Hence, in this case cos cos = 0.486. Note that the largest possible value of the Schmid factor is 0.5 (cos = cos = 1/ 2).

DH42

Course D: Mechanical Behaviour of Materials

DH42

4.9

Geometry as slip proceeds

Slip on a single system tends to cause lateral displacement (as well as axial extension). If the grips used to hold the sample are aligned so as to prevent this, then the tensile axis will rotate towards the slip direction, as shown in Fig. 33.
(a) (b) (c) (d)

bending

1
slip direction

rotation

bending

Figure 33: Deformation of a single crystal with one operative slip system (a) before loading, (b) during loading, (c) allowing rotation at the grips, (d) with xed grips.

The spacing between the planes remains constant through the deformation and the number of planes is conserved. Hence, l cos will remain constant as deformation proceeds. This is shown schematically in Fig. 34. Hence, l cos = l1 cos 1 (4.58)

Similarly, if we consider the section of the sample that contains both the slip direction and the tensile axis, by the same argument l cos(90 ) = l sin is constant. Hence, l sin = l1 sin 1 (4.59)

As slip proceeds and the slip direction rotates towards the tensile axis, the resolved shear stress on all available slip systems will change. It is often more convenient to refer to this movement from the reference frame of the sample and consider the tensile axis rotating towards the slip direction. The direction of the tensile axis relative to the crystal lattice vectors as slip proceeds can also be obtained by adding multiples of the slip direction to the tensile axis indices. Using the example we had earlier TA = [ 214] + n[011] (4.60)

DH43

Course D: Mechanical Behaviour of Materials

DH43

slip plane normal

tensile axis

tensile axis

slip direction

Figure 34: Geometry as slip proceeds showing the number of planes remains constant.

This continues until two of the indices have the same value (ignoring their signs). In the example we are working with this will occur when the tensile axis reaches [ 225]. At this point two slip systems now have the same Schmid factor and slip proceeds on both of them. The new slip system can also be obtained by the OILS rule and, in our example, is (111)[ 101]. The subsequent reorientation of the tensile axis will now proceed by adding equal multiples of both slip directions to the tensile axis indices. Again, using the example we had earlier TA = [ 225] + n[011] + n[ 101] (4.61)

As can be seen from this equation, the tensile axis will approach a 112 type direction, but will never quite get there. For deformation along a 112 type direction the rotations induced by the two slip systems exactly cancel.

DH44

Course D: Mechanical Behaviour of Materials

DH44

4.10

The energy associated with a dislocation

The energy associated with the elastic distortion of the crystal around a dislocation can be most readily estimated by examining a screw dislocation. Consider the annulus of material around a screw dislocation (see Fig. 35(b)). If this annulus is unwrapped we can see that the original cuboid of material has been sheared.

(a)

(b)

r dr

(c)
b l

dr 2r
Figure 35: Lattice distortion around a screw dislocation.

The associated shear strain of the annulus, , is = The volume of material in this annulus, V , is V = 2 r l dr (4.63) b 2 r (4.62)

2 The shear strain energy per unit volume is 1 2 G . Hence the strain energy in this element, dU , is 1 b 2 dU = G (2 r l dr) (4.64) 2 2r

The total strain energy can be obtained by integrating this expression between r = r , the radius at which the strain is too great to be treated as elastic, and some outer limit, r = r . It is also necessary to make allowance for the core energy (the energy associated with the region for which r < r ). U = = Gb2 l 4
r r

dr + core energy r + core energy

(4.65) (4.66)

Gb2 l ln 4

r r

The challenge is to choose suitable values for r , r and the core energy.

DH45

Course D: Mechanical Behaviour of Materials

DH45

Clearly as r approaches zero, Hookes law would suggest that the stress would become innite. However, the actual stress will remain nite (as dU/dr will remain nite). It has been suggested that taking a limit of integration of r = b/4 provides a slight overestimate of the actually stress and hence the strain energy density in the region immediately above r that is approximately equal to the core energy (see Fig. 36).

real

Figure 36: Strain energy density close to the dislocation core.

Similarly, r cannot be larger than the crystal. In practice taking a value equal to half the distance between dislocations is considered reasonable. With such assumptions, the overall result is 1 U Gb2 l 2 (4.67)

A similar result can be obtained for an edge dislocation, but is more difcult to derive. Importantly, the energy of an edge dislocation is always higher than that of a screw dislocation. It is often more useful to express the strain energy per unit length of dislocation, . 1 Gb2 2 (4.68)

DH46

Course D: Mechanical Behaviour of Materials

DH46

4.11

Dislocation interactions

The distortion induced in a crystal by the presence of a dislocation leads to local tensile, compressive and shear stresses. A screw dislocation only gives rise to localised shear stresses. As we saw earlier, the shear strain associated with a screw dislocation is = b/2r (from Eq. 4.62). Hence, the local shear stress acting along the axis of the dislocation is Gb 2r in which r is the radial distance from the dislocation core. = G = (4.69)

The local stress eld around an edge dislocation is much more complex and contains tensile, compressive and shear stresses. However, by just considering the hydrostatic tensile and compressive stresses around an edge dislocation it is easy to visualise the local stress elds. Specically, the extra half plane of atoms above the slip plane puts this region into compression whilst the region below the slip plane goes into tension. The stress elds associated with dislocations interact. The resulting stress at any point will then be the sum of the stress contributions from each dislocation. As the elastic strain energy is proportional to the square of the local strain, it is energetically favourable for the stress elds to congure themselves to minimise this strain. As a result, two dislocations of the same sign on the same slip plane will repel one another, whilst two dislocations of opposite sign on the same slip plane will attract one another. Similarly, dislocations of the same sign moving on different slip planes may be attracted to one another, and possibly form dislocation arrays.
(a) C C (b) C T

T (c) C

T (d)

T T

C T C

Figure 37: Interactions between the hydrostatic stress elds surrounding (a) two dislocations of the same sign moving on the same slip plane, (b) two dislocations of opposite sign moving on the same slip plane, (c) two dislocations of the same sign moving on different slip planes, (d) two dislocations of opposite sign moving on different slip planes.

Dislocation arrays may form that minimise the overall energy, as shown in Fig. 38.

DH47

Course D: Mechanical Behaviour of Materials

DH47

(a)

(b)

Figure 38: (a) original dislocation conguration. (b) dislocation array.

Figure 39: Dislocation arrays in a single-slip-oriented copper single crystal (from: Lepisto et al., Materials Science and Engineering, 81 (1986) 457-463).

DH48

Course D: Mechanical Behaviour of Materials

DH48

4.12

Dislocation reactions

If two dislocations combine, the new dislocation will have a Burgers vector, b3 , which will be the sum of the Burgers of the two dislocations from which it formed, b1 and b2 .

The line vector of the new dislocation, l3 , will be the intersection of the slip planes of the two original dislocations.

The plane in which the new dislocation may slip will be the plane containing both its Burgers vector, b3 , and its line vector, l3 . For such a dislocation reaction to occur, however, it must be energetically favourable. As the energy associated with a dislocation is proportional to the square of the magnitude of the Burgers vector, dislocations will combine if
2 2 b2 3 < b1 + b2

(4.70)

This is known as Franks rule.

Consider the example of slip in the fcc single crystal we considered earlier. The two slips systems that became active when the tensile axis rotated to [ 225] were ( 1 11)[011] and (111)[ 101]. a a [011] and b2 = [ 101]. 2 2 If dislocations with these two Burgers vectors were to combine then the resultant dislocation would have a Burgers vector The associated Burgers vectors are therefore b1 = b3 = b1 + b2 = a a a [011] + [ 101] = [ 112] 2 2 2 (4.71)

The square of the magnitude of the Burgers vectors of the two original dislocations are both a2 /2, whilst the square of the magnitude of the Burgers vectors of the resultant dislocation would be 3a2 /2
2 2 In this case, b2 3 > b1 + b2 , so we do not expect these dislocations to combine.

DH49

Course D: Mechanical Behaviour of Materials

DH49

4.12.1

Interaction of dislocations on different slip systems: The Lomer lock

Dislocation interactions can have profound effects upon the plastic deformation behaviour of materials. Consider the example of the interaction of the dislocations moving on the following two slip systems in a fcc metal. ( 111)[101] and (111)[ 110].

Figure 40: Dislocation interaction leading to the formation of a Lomer lock.

a a [101] and b2 = [ 110]. 2 2 If these dislocations combine then the resultant dislocation would have a Burgers vector The associated Burgers vectors are b1 = b3 = b1 + b2 = a a a [101] + [ 110] = [011] 2 2 2 (4.72)

In this case, the square of the magnitude of the Burgers vectors of the two original dislocations and the resultant dislocation are all equal to a2 /2. We therefore expect these dislocations to 2 2 combine as b2 3 < b1 + b2 . The line vector of the new dislocation, l3 , must lie at the intersection of both slip planes, i.e. along [0 11] (or [01 1]), which is perpendicular to b3 (you can obtain this result using the Weiss Zone law). The new dislocation is therefore an edge dislocation and can glide in the plane containing b3 and l3 , i.e. (100). Unfortunately, (100) is not a plane on which slip occurs in fcc metals. As such this dislocation will not move - it is sessile and will block the movement of further dislocations on both slip planes. This type of sessile dislocation is called a Lomer lock.

DH50

Course D: Mechanical Behaviour of Materials

DH50

4.13

Dislocation generation: Frank-Read sources

Once a dislocation has moved across a crystal and generated a step on the surface of the crystal it is no longer available to participate in plastic deformation. The gross plastic deformation we see in metals therefore requires sources of dislocations. In practice, dislocations can be created at free surfaces, grain boundaries and within grains. The most famous type of dislocation source is a Frank-Read source - named after the two people who postulated its existence. This is shown in Fig. 41

Figure 41: A schematic representation of the operation of a Frank-Read source

Imagine a dislocation with Burgers vector, b, of initial length, L, pinned at both ends. With no applied shear stress, the dislocation will lie in a straight line, as this minimises its elastic strain energy. When a shear stress, , is applied, the dislocation will experience a force (F = b) normal to the dislocation line. As the dislocation is pinned at both ends it will bow out, balancing the line tension (i.e. dislocation energy) and the force due to the applied shear stress. As the shear stress and hence the force on the dislocation is increased, it will eventually become unstable and bow outwards until it forms a loop. When the two segments at the bottom of the loop touch one another they annihilate each other as their line vectors are dened in opposite directions, implying that their Burgers vectors will also be opposite. Provided the shear stress is maintained, the dislocation loop formed may continue to propagate outwards whilst the remaining segment between the two pinning points may go on to produce further dislocations. An example of a Frank-Read source in silicon is shown in Fig. 42.

Figure 42: A Frank-Read source in Si (from: W.G. Dash, Dislocations and Mechanical Properties of Crystals, ed. J.C. Fisher, Wiley New York, 1957.)

DH51

Course D: Mechanical Behaviour of Materials

DH51

4.14

Jogs and kinks

Dislocations are rarely straight and may not lie entirely in a single slip plane. Steps which lie in the same slip plane are termed kinks.
(a) (b)

Figure 43: Schematic illustration of kinks in (a) an edge dislocation, and (b) a screw dislocation.

Steps in the dislocation from one slip plane to another are termed jogs.
(a) (b)

Figure 44: Schematic illustration of a jog in (a) an edge dislocation, and (b) a screw dislocation.

As kinks lie in the same slip plane they do not inhibit the movement of the dislocation. In fact they may assist its motion, as atoms or vacancies diffusing to them may enable the dislocation to move at stresses below the critical resolved shear stress! Although, in the case of the screw dislocation, the kink will have some edge character the screw dislocation may now be conned to a single slip plane. Kinks and jogs are formed by: The absorption of vacancies/ atoms Intersection of propagating dislocations

DH52

Course D: Mechanical Behaviour of Materials

DH52

4.14.1

The absorption of vacancies/ atoms

The absorption of vacancies or atoms into a dislocation may lead to the formation of kinks or jogs. This is a thermally activated process. Similarly, if kinks or jogs are already present along the dislocation, they may get larger or smaller (and possibly annihilate) as a result of this process. The mobility of atoms in most materials is sufcient that an equilibrium exchange of atoms is set up between the slipped and unslipped crystals either side of the dislocation. If it is energetically favourable for one process to dominate, for example under an applied stress, the kink or jog may move. This is most readily visualised with the migration of kinks in the extra half plane of an edge dislocation.
(a)

(b)

Figure 45: Movement of a kink in an edge dislocation by the transfer of vacancies/ atoms.

Spontaneous creation of double kinks by thermal motion A dislocation may move under an applied shear stress lower than the critical resolved shear stress by the spontaneous formation and growth of double kinks. As this is a thermally activated process it is favoured at higher temperatures.
(a) (b)

Figure 46: Formation and propagation of double kinks by thermal motion of atoms.

DH53

Course D: Mechanical Behaviour of Materials

DH53

4.15

Climb and cross slip of dislocations

If dislocation glide on a given slip plane is inhibited by an obstacle it may be possible for the dislocation to migrate onto another slip plane and continue gliding from there. Two important processes by which this can occur are: cross slip of screw dislocations and climb of edge dislocations.

Climb of edge dislocations As the Burgers vectors and line vectors of edge dislocations are perpendicular to one another, they are conned to a single slip plane and, as such, it is not possible for edge dislocations to undergo cross slip. Edge dislocations may, however, move onto other slip planes to avoid obstacles or to arrange themselves into energetically favourable congurations by the process of dislocation climb. If a vacancy in the crystal structure migrates to the dislocation core, the segment of the dislocation where the vacancy is absorbed will rise by one atomic spacing. This is termed positive climb. It is also possible for extra atoms to diffuse to the edge dislocation core, moving the dislocation core downward. This is termed negative climb.
(a) (b) (c)

Obstacle
Figure 47: Positive climb of a dislocation to bypass an obstacle.

Note that a lot of vacancies/ extra atoms may be required to raise or lower an entire edge dislocation to the next slip plane. The diffusion processes required for either positive or negative climb are thermally activated and therefore occur more readily at high temperatures.

DH54

Course D: Mechanical Behaviour of Materials

DH54

Cross slip of screw dislocations As the Burgers vector and line vector of screw dislocations are parallel, they may glide on the family of crystallographically related planes that contain these vectors. This enables these dislocations to change the slip system in which they move to detour past obstacles. Consider a screw dislocation with Burgers vector a/2[1 10] propagating along (111) as shown in Fig. 48. It will continue to slip on this slip plane until the local stress eld changes so that motion on the (11 1) becomes preferred. At which point it will begin gliding on this new slip system. Double cross slip may also be possible, as shown in Fig. 48(d). By this process, screw dislocations may move past obstacles in their initial slip plane.
[110] (111) (a) (b) (c) (d)
Figure 48: Schematic illustrations of cross slip in a fcc metal. (a) A screw dislocation is gliding along (111). (b) The dislocation may begin slipping on (11 1) if it is favourable for it to do so. (c) The dislocation may continue to propagate along (11 1). (d) Double cross slip back onto a (111) plane may also be possible.

(111)
b

DH55

Course D: Mechanical Behaviour of Materials

DH55

4.16

Plastic deformation of metallic single crystals

If we conduct a tensile test on a metallic single crystal we expect to see the normal strain increase approximately linearly with applied normal stress, according to Hookes law, until the stress reaches the yield stress, y . At this stress, the resolved shear stress on the slip system with the largest Schmid factor will be sufcient for dislocations to move. As the crystal plastically deforms, the tensile axis rotates towards the slip direction, with a progressive change in the Schmid factor.

4.16.1

hcp metals

During the deformation of hcp metals, only one slip system will operate. The critical resolved shear stress will not vary during deformation, as shown in Fig. 49. However, as the tensile axis

Initial elastic strain

Plastic deformation

Figure 49: Shear stress vs shear strain typically observed for hcp metal single crystals.

rotates towards the slip direction, its Schmid factor will change. Depending upon the initial orientation of the crystal the Schmid factor may decrease (Fig. 50 - A) or it may rst increase before decreasing (Fig. 50 - B). As a result, for some initial orientations, the tensile normal stress required to continue plastic deformation may actually decrease after yield (Fig. 50 - B). This effect is referred to as geometric softening.

A Initial elastic strain Plastic deformation

Figure 50: Normal stress vs strain observed for hcp metal single crystals. A - decreasing Schmid factor. B - initially increasing Schmid factor (Geometric softening).

DH56

Course D: Mechanical Behaviour of Materials

DH56

4.16.2

fcc metals

During the deformation of fcc metals, one or two slip systems will operate, depending upon the orientation of the crystal. The critical resolved shear stress will vary during deformation, as shown below.

Stage Stage

III

II
Initial elastic strain Stage

I
Plastic deformation

Figure 51: Shear stress vs shear strain typically observed for fcc metal single crystals.

Stage I: One operative slip system. The tensile axis rotates towards the slip direction. The critical resolved shear stress remains constant but the Schmid factors on all slip systems change. The extent of Stage I is determined by the initial orientation of the crystal. Stage I is often referred to as easy glide.

Stage II: Two operative slip systems (duplex slip). The tensile axis has rotated into a position in which two slip systems share the largest Schmid factor. Dislocations moving on these slip systems interact with each other producing jogs, locks and pile ups. This leads to a rapid increase in the critical resolved shear stress required to move further dislocations on these slip systems. The increase in strength with plastic strain that arises by this process is referred to as work hardening or strain hardening .

Stage III: The resolved shear stress becomes sufcient to activate other slip systems and allow dislocations to bypass obstacles. The transition to stage III correlates with stacking fault energy in fcc metals as cross-slip occurs more readily with higher stacking fault energies and therefore the transition in such materials occurs at lower applied stresses.

DH57

Course D: Mechanical Behaviour of Materials

DH57

4.16.3

Polycrystalline metals

In a polycrystalline metal, the orientation of the individual grains and hence also their Schmid factors are different. The stress-strain curve (Fig. 52) shows no Stage I deformation behaviour as duplex slip and work hardening initiates after different plastic strains in each grain. As a result, the stress-strain curve shows a continuous work hardening after yield.

Initial elastic strain Plastic deformation

Figure 52: Normal stress vs strain typically observed for polycrystalline fcc metals.

For randomly orientated grains the average value of the Schmid factor is 1/3. This average value is referred to as the Taylor factor. Hence for polycrystalline materials it might be expected that y 3c (4.73)

In practice, the yield stress of polycrystalline metals is often much higher than this due to effect of grain boundaries. In a polycrystal, the deformation of each individual grain has to be compatible with that of its neighbours i.e. there is a strong constraint effect. Multiple slip is normally required from the outset in virtually all grains in order to satisfy this requirement, and substantially higher stresses are needed for yielding and plasticity, compared with single crystals.

DH58

Course D: Mechanical Behaviour of Materials

DH58

5
5.1

Strengthening mechanisms
Forest hardening

The dislocation density, , is a measure of the total dislocation line length per unit volume or, equivalently, the number of dislocations intersecting unit area. It has units of m2 . The average spacing between dislocations, L, is related to the dislocation density. If we assume we have a simple cubic array of dislocations, then an area of L2 has 1 dislocation passing through it. Hence, = 1/L2 With higher dislocation densities, dislocation motion becomes increasingly difcult due to the presence of sessile dislocations blocking the easy glide of glissile dislocations. A greater shear stress is therefore required to push dislocations through the material resulting in higher yield strengths. This effect is known as forest hardening.

Figure 53: TEM micrograph of a creep deformed nickel alloy showing regions with high and low dislocation densities.

Dislocation densities in metals range from 1010 m2 in annealed metals to 1016 m2 in cold worked metals. These correspond to spacings of 10 m to 10 nm.

DH59

Course D: Mechanical Behaviour of Materials

DH59

5.2

Dislocation pile ups and the effect of grain size

Dislocations will move along a slip plane until they encounter an obstacle. A grain boundary forms such an obstacle as the dislocation cannot pass from one grain to the next. Instead the grain must exert a sufcient stress on the neighbouring grain to initiate slip in it. Dislocations travelling on the same slip plane will therefore accumulate at grain boundaries, forming a dislocation pile-up, see Fig. 54. Similar dislocations travelling on the same slip plane will repel each other. If there are n dislocations in a pile up, the stress at the obstacle will be n times the applied stress.

Figure 54: Dislocation pile up a a grain boundary.

The effect of grain size on yield strength can now be understood. For macroscopic yielding of the metal, slip must be initiated in all grains. If slip in a favourably orientated grain (A) is underway a dislocation pile-up must be established that is capable of exerting sufcient stress on the neighbouring grain (B) to initiate yield. With a large grain size it is possible to reach the necessary stress by forming a long pile-up. If the grain size is small only shorter pile-ups can be supported and hence a greater stress is required. Experimentally, the tensile yield stress, y , is seen to be related to the diameter of the grains, d by the Hall-Petch relation. k y = + d in which, is the intrinsic yield stress and k is a constant for the material. Neighbouring grains: constraint on deformation (back-stress on dislocations) yield stresses much higher than for single crystals ner grain size usually gives higher yield stress (5.74)

DH60

Course D: Mechanical Behaviour of Materials

DH60

5.3

Solid solution strengthening

The effect of solute atoms can best be understood by considering the interaction between the stress elds they produce and those of the dislocations with which they interact. Substitutional and interstitial solute atoms act differently and must be considered separately.

vacancy

substitutional defects

interstitial defect

Figure 55: Lattice distortions associated with point defects

5.3.1

Substitutional solute atoms

If the substitutional solute atom is larger than the atoms in the crystal it resides in it will introduce a spherically symmetric compressive stress eld. If the solute atom is smaller, it will introduce a spherically symmetric tensile stress eld. As the stress elds induced by either large or small solute atoms contain no shear stress component, they will not interact with screw dislocations, because the stress elds of screw dislocations involve shear only. Substitutional solute atoms may, however, interact with the stress elds generated by edge dislocations. It is favourable for the solute atoms to position themselves such that the overall strain energy is minimised.

DH61

Course D: Mechanical Behaviour of Materials

DH61

Larger solute atoms will position themselves in the tensile eld of the dislocation (below the extra half plane of an edge dislocation - see Fig. 56(a)). Smaller solute atoms will position themselves in the compressive eld (at the end of the extra half plane of an edge dislocation - see Fig. 56(b)). In practice, dislocations are much more mobile than substitutional solute atoms. As such, the presence of solute atoms in the slip plane can be considered on the basis of their relative size and where they sit relative to the gliding dislocation.

(a)

(b)

Figure 56: (a) Larger solute atoms in the tensile eld of an edge dislocation. (b) Smaller solute atoms in the compressive eld of an edge dislocation.

For example, if a small solute atom sits just above the slip plane of an edge dislocation, it will attract the dislocation as this will form an energetically favourable conguration (see Fig. 56(b)). Similarly, if it was located on the other side of the slip plane it would be expected to repel the dislocation. As it is energetically favourable for the dislocation to be in certain arrangements with substitutional solute atoms, the dislocation will spend more time in these energetically favourable congurations. Overall this results in dislocation motion being retarded with a greater shear stress being required to move dislocations.

DH62

Course D: Mechanical Behaviour of Materials

DH62

5.3.2

Interstitial solute atoms

Interstitial solute atoms typically have a stronger effect than substitutional solute atoms. This may be attributed to the larger strains they typically generate and, more importantly, they may also produce asymmetric distortions of the crystal structure. The best known example is carbon atoms in -iron.
Octahedral interstice

Figure 57: Carbon atom occupying an octahedral interstice in -iron.

The octahedral interstice occupied by the carbon atom has two iron atoms at a distance of a/2 from the centre of the interstice and four iron atoms at a distance of a/ 2 from the centre of the interstice. If the atom in the interstice has a diameter greater than a 2rFe it will induce an asymmetric distortion of the lattice. Such asymmetric distortions can interact with both edge and screw dislocations. The comparative effectiveness of solid solution strengthening with substitutional and interstitial additions is shown in Fig. 58
C,N

150
Si y (MPa)

100
Mn

50 0
Mo Ni

1.0

2.0

% alloying element
Figure 58: Comparative effect of solid solution strengthening in -iron. (from: F.B. Pickering and T. Gladman, ISI Special Report 81, Iron and Steel Inst., London, p109 (1963)).

DH63

Course D: Mechanical Behaviour of Materials

DH63

Interstitial diffusion rates can be very high (see Fig. 59). In steels carbon atoms may therefore migrate to dislocations very quickly (in a few hours at room temperature, a few seconds at 150 C, or virtually instantaneously at 300 C).
100

Diffusion distance (m)

10

300C 150C

0.1

0.01

20C

0.001
0 1 2 3 4 5

10

10

10

10

10

10

Time (s)
Figure 59: Diffusion distances for (interstitial) carbon in -iron.

These rapid diffusion rates can allow carbon atoms to locate themselves in every crosssectional plane containing the dislocation (see Fig. 60). These accumulations of interstitial carbon atoms at dislocations are said to form carbon atmospheres or Cottrell atmospheres.

Figure 60: Schematic depiction of the formation of a carbon atmosphere along the length of a dislocation in a steel.

Cottrell atmospheres can also have a profound effect upon the yield behaviour as a function of temperature.

DH64

Course D: Mechanical Behaviour of Materials

DH64

Luders bands The stress-strain curve of a low carbon steel is shown schematically in Fig. 61.

upper yield stress lower yield stress

Luders bands unyielded yielded

Figure 61: Schematic illustration of the occurrence of Luders bands associated with a localised drop in yield stress, observed with low-carbon steels.

The steel rst starts to deform plastically at the upper yield stress as the dislocations escape from their Cottrell atmospheres. During a tensile test, stress concentrations towards the ends of the sample will initiate yield in those regions. As the dislocations in these regions are no longer pinned by their Cottrell atmospheres, the local yield stress will now be lower than that of the rest of the sample. Plastic deformation is therefore concentrated in these regions. However, as they deform they work harden and the local yield stress will again rise. Plastic deformation may continue at the lower yield stress by the movement of the boundary between the yielding and unyielding regions of the sample. These boundaries, or Luders bands typically lie at 50 to the tensile axis and move together towards the centre of the sample. When they meet, continued plastic deformation will lead to work hardening and a progressive increase in the yield stress.

DH65

Course D: Mechanical Behaviour of Materials

DH65

Portevin-Le Chatelier effect As the temperature increases, the shape of the stress-strain curves changes markedly as the mobility of carbon atoms increases. At high temperatures (300 C) carbon atoms are sufciently mobile to move with the dislocations. As such, they no longer inhibit dislocation motion and provide no strengthening effect. At intermediate temperatures (150 C) the carbon atoms and dislocations move at comparable speeds. This enables dislocations to repeatedly escape their Cottrell atmospheres and for them to reform. This produces serrations in the stress-strain curve. This behaviour is known as the PortevinLe Chatelier effect.

room temperature

150 C 300 C

Figure 62: Schematic stress-strain plots for low carbon steel, showing the Portevin-Le Chatelier effect with increasing test temperature

DH66

Course D: Mechanical Behaviour of Materials

DH66

5.4

Precipitate hardening

A highly effective method of restricting dislocation motion is through the use of precipitates in the crystal matrix. Small coherent precipitates may be sheared by the passage of a dislocation.

(a)

matrix

(b)

(c)

r
precipitate

Figure 63: Shearing of a coherent precipitate by a dislocation. (a) the dislocation approaches the precipitate. (b) the precipitate has passed part way through the precipitate. (c) the dislocation has exited the precipitate, leaving it sheared by the Burgers vector.

For precipitate shearing in this fashion, the increase in shear stress required for dislocation motion, , is proportional to the square root of the precipitate radius, i.e. r.

Larger precipitates will normally have a different crystal structure from the matrix. This prevents dislocations from the matrix passing directly into the precipitate. However, they may bow around the precipitate, leaving a dislocation loop around the precipitate (Fig. 64). This is know as Orowan bowing.

(a)

(b)

(c)

(d)

(e)

Figure 64: Schematic depiction of Orowan bowing of dislocations around a pair of large, unshearable precipitates leaving dislocation loops around them.

DH67

Course D: Mechanical Behaviour of Materials

DH67

5.4.1

Stress required to bow dislocations between precipitates

Consider a dislocation bowing between two precipitates separated by a distance, L (Fig. 65). The peak curvature of the dislocation line, and hence also the peak energy, will occur when the dislocation forms a semi-circle between the precipitates.
dl

R r L

Figure 65: Force on a short segment of a dislocation to bow it between two precipitates.

For a segment of the dislocation of length, dl, the force acting normal to the segment is F dl. Thus the resolved component of the force acting in the vertical direction is

F sin dl =
0 0

F sin

L d 2

FL [ cos ] 0 = F L = bL 2

(5.75)

This force is balanced by the line tension of the dislocation in the downwards direction at each precipitate. The line tension is equivalent to the energy per unit length of the dislocation at each precipitate and is given by = Gb2 /2. Thus, bL = 2 Gb2 2 (5.76)

The bowing stress or Orowan stress required to bow a dislocation between precipitates separated by a distance, L, is therefore = Gb L (5.77)

Note that as the Orowan stress is inversely proportional to the precipitate spacing we need very ne arrays of precipitates to provide signicant strengthening.

DH68

Course D: Mechanical Behaviour of Materials

DH68

5.4.2

Transition from cutting to bowing and maximum hardening

If the precipitates are very small (and coherent) they may be sheared readily providing little hardening. As their radii increases (and number density decreases for constant volume fraction), so the resistance to dislocation cutting will increase with r. Eventually it becomes easier for the dislocations to bow between the precipitates rather than cut them. Further increases in precipitate radii will result in a decrease in the critical resolved shear stress with 1/r. A peak strength will be seen at the transition from precipitate cutting to Orowan bowing (Fig. 66).

Cutting

Bowing

r
Figure 66: Change in the critical resolved shear stress as a function of precipitate radius in precipitate hardened alloys showing the transition from cutting to bowing mechanisms.

DH69

Course D: Mechanical Behaviour of Materials

DH69

5.4.3

Changing strengthening mechanisms during age hardening

In nucleating and growing precipitates in an alloy from a supersaturated solid solution, several factors will contribute to the strength of the alloy and these will change with time (see Fig. 67). Solute hardening At short ageing times the majority of the strength of the alloy will come from solid solution strengthening in the supersaturated solid solution. This will decrease with time as solute is rejected from the matrix to form the precipitates. Coherency strains The coherent and semi-coherent precipitates that form in the early stages of ageing will distort the lattice in which they sit. This induces coherency strains in the matrix which extend the range of inuence of the precipitates and inhibit dislocation motion. At this stage the coherent precipitates may be readily sheared by dislocations. Precipitate cutting As the coherent and semi-coherent precipitates become larger signicantly larger stresses are required to shear the precipitates. Orowan bowing As the precipitates become large and the distance between precipitates increases, dislocation bypass by bowing becomes possible. This becomes easier as the precipitate array becomes coarser (over-ageing).

overall yield stress

yield strength, y

precipitate cutting

precipitate bypass (Orowan

solution strengthening

coherency strengthening

aging time
Figure 67: Origins of changes in yield stress of an age hardenable alloy as a function of heat treatment time.

DH70

Course D: Mechanical Behaviour of Materials

DH70

Precipitate hardening in Al-Cu alloys: If we quench an Al-Cu alloy from the single phase eld we obtain a supersaturated solid solution. Ageing the alloy will rst lead to coherent Gunier-Preston (GP) zones, then coherent , then semi coherent and nally incoherent precipitates . As the precipitates age their average size and spacing will increase. This will be accompanied an initial increase in yield stress with ageing time followed by a decrease in yield stress with longer ageing times, see Fig. 68.

Figure 68: Hardness of Al-Cu alloys (related to yield stress) as a function of heat treatment time at two different temperatures. (from: J. Silcock, et al., Inst. Metals, 82, 239, 1953-54.)

Higher peak strength at lower T Longer to reach peak strength at lower T (note the log scale) Sometimes get multiple peaks (optimum precipitate size of a given type)

DH71

Course D: Mechanical Behaviour of Materials

DH71

Images of the precipitates in an Al-Cu alloy are shown in Fig. 69.

Figure 69: TEM micrographs of Al-Cu alloys at different stages of ageing:


GP zones (from: R.B. Nicholson & J. Nutting, Phil. Mag., 3, 531, 1958), (from: R.B. Nicholson et al., J. Inst. Metals., 87, 431, 1958-59), (from: G.C. Weatherly & R.B. Nicholson, Phil. Mag., 17, 813, 1968), (from: G.A. Chadwick, Metallography of phase transformations, Butterworths, London, 1973).

The nature of the precipitates also has a pronounced effect upon the rate of strain hardening (see Fig. 70). Whilst GP zones may provide signicant strengthening, once slip starts the

Aged to peak hardness

Underaged to produce GP zones Overaged Solid solution

Pure Al

Figure 70: Schematic illustrations of the stress-strain curves of Al-Cu alloys after different stages of ageing.

precipitates may be repeatedly sheared by dislocations and the yield stress may actually decrease. As dislocations may pass through the entire crystals relatively unhindered the rate of strain hardening is low. In the peak hardened and overaged states, dislocations must bow around precipitates. The increase in dislocation debris (loops, tangles, etc) left by the process inhibits the passage of further dislocations and the material strain hardens rapidly.

DH72

Course D: Mechanical Behaviour of Materials

DH72

5.5

Partial dislocations and stacking faults

In practice, it is often energetically favourable for dislocations to dissociate into dislocations with smaller Burgers vectors. Consider the (111) plane of an fcc metal. The crystal structure is restored after the passage of a perfect dislocation that shifts atoms in a B position to the next B position. The Burgers vector required to do this is shown by b1 = a/2[ 101] in the gure below.

A b2 B A C C b1 A b3

A C B A C

Figure 71: The (111) plane of an fcc metal. Atoms in the A layer are drawn. The B and C layer atomic positions are marked.

An alternative path between B positions is to rst shift an atom in a B position to a C position and then on to the next B position. This requires the passage of two smaller dislocations that are not lattice vectors, and are shown by b2 = a/6[ 211] and b3 = a/6[ 1 12]. These are referred to as partial dislocations. In an fcc metal these partial dislocations will have Burgers vectors of a/6 11 2 After passage of the rst dislocation with Burgers vector b2 , the stacking sequence will change from A B C A B C A B C to A B C A C A B C A This generates a stacking fault. The passage of the second dislocation with Burgers vector b3 will restore the original structure. An illustration of a stacking fault formed between two partial dislocations is shown in Fig. 72. Consider the dissociation of the dislocation b1 into the two partial dislocations, b2 and b3 . a a a [101] [ 211] + [ 112] 2 6 6 The energies of these dislocations are proportional to a2 a2 a2 a2 + = 2 6 6 3 (5.78)

(5.79)

Hence, the dissociation of the dislocation into partial dislocations is energetically favourable.

DH73

Course D: Mechanical Behaviour of Materials

DH73

Figure 72: A stacking fault in a fcc (or hcp) metal.

Once a perfect dislocation has dissociated into two partial dislocations, the elastic stress elds will cause the partial dislocations to repel each other, generating an extended stacking fault between them. As there is an energy penalty associated with forming the stacking fault (stacking fault energy). The width of the stacking fault will be limited by the condition where Monday, 23 January 12 the net repulsive force (or reduction in elastic strain energy) between the partial dislocations equals the total stacking fault energy. Hence, if the stacking fault energy (per unit area) is low, the two partial dislocations may be well separated. If the stacking fault energy is high, then these two partial dislocations will be close together. The stacking fault energy has a strong inuence on the ease with which cross-slip may occur in fcc metals. A perfect screw dislocation may cross-slip readily. However, as partial dislocations have both screw and edge components they may not cross-slip directly. Instead they must recombine before cross-slip can occur. As a consequence, the onset of Stage III is delayed in metals with low stacking faults energies (e.g. brass and stainless steels) and they will work-harden rapidly.
Table 2: Stacking fault energies of selected fcc metals (in mJ m2 ).

Ag 16

Al 166

Au 32

Cu 45

Ni 125

Pd 180

Pt 322

Rh 750

Ir 300

DH74

Course D: Mechanical Behaviour of Materials

DH74

5.6

Order hardening

For those compounds that show an transition between ordered and disordered (related) structures an abrupt change in strength is often seen. For example, in brass a large change in the critical resolved shear stress is observed around the order-disorder transition at 470 C.

c (MPa)

Long range order 0.6 0.5 0.2 0

5 4 3 2 1

400

440 Tc 480 Temperature C

520

Figure 73: Effect of order/disorder on the critical resolved shear stress of brass [from: N. Brown in Mechanical Properties of Intermetallic Compounds, J.H. Westbrook (ed), John Wiley and Sons, (1960)]

slip In bcc metals, like high temperature brass, slip would be expected on the {110} 111 systems. The associated Burgers vectors are a/2 111 . However, these are not lattice vectors in the CsCl structure, like low temperature brass. The passage of such a dislocation through a crystal with the CsCl structure would therefore generate a layer with energetically unfavourable like-like bonds. This energetically unfavourable layer is called an anti-phase boundary and is a 2D planar defect. dislocation following the rst one will restore the original The passage of a second a/2 111 structure. The two dislocations (and their anti-phase boundary) collectively form a superdislocation and the individual dislocations are often referred to as superpartial dislocations. Superdislocations can be incredibly effective in strengthening materials at high temperature (more on this in Course F). Antiphase boundaries are therefore associated with dislocations at either end in a similar way that stacking faults are associated with partial dislocations, as described previously. As before, if the energy associated with the antiphase boundary is low these two dislocations

DH75

Course D: Mechanical Behaviour of Materials

DH75

[001]

[110]

Figure 74: (1 10) plan view of the CsCl structure showing the formation of an anti-phase boundary after a displacement of a/2[111].

may be well separated. If the energy associated with the antiphase boundary is high, these two dislocations will need to be close together. Note that in the CsCl structure, the shortest vectors between lattice points are a 110 . These are much longer than a/2 1 11 dislocations and therefore have a much greater energy associated with them.

DH76

Course D: Mechanical Behaviour of Materials

DH76

5.7

Plastic deformation by cooperative shear - Twinning

Whilst dislocation motion is the predominant deformation mechanism in most metals at room temperature, plastic deformation may also occur by other mechanisms. Another important deformation mechanism is the co-operative (simultaneous) shearing of successive atomic planes. If the crystal structure formed by this process differs from the original structure it is termed a martensitic phase transformation. If the crystal structure formed is the same as the original, but in a different orientation, it is termed deformation twinning or mechanical twinning. It is called twinning because the crystal either side of the twin boundary are mirror images of each other. Deformation twinning occurs through a combination of a twinning plane and a twinning direction. These are analogous to the slip plane and slip direction that dene deformation by dislocation slip.
Table 3: The twinning planes and twinning directions seen in selected crystal structures.

Crystal structure bcc bct fcc

Twinning plane {112} {011}, {102}, etc {111}

Twinning direction 11 1 100 , 201 , etc 11 2

As with other shear processes, the shear associated with twinning (twinning shear) is simply the displacement parallel to the plane divided by the interplanar separation.

Slip Lattice unaffected can generate large shear displacements on a single plane (by repeated passage of dislocations) Deformation Twinning can generate similar type of displacement, but homogeneously distributed elastic strain due to constraint from surrounding material (in polycrystals)

DH77

Course D: Mechanical Behaviour of Materials

DH77

Twinning in fcc metals Fcc metals rarely undergo signicant deformation twinning, except under specic conditions (e.g. Cu or Ag-Au alloys at cryogenic temperatures and alloys with low stacking fault energies, such as TWIP steels and Ni-Co alloys). However, the twinning process in fcc metals is perhaps the easiest to visualise. It requires the shear of each successive {111} plane. This is shown schematically in Fig. 75 below.
{111} in fcc Layer A Layer B Layer C A C B A C B A C B A B A C B A C C B A C B A C B A A C parent B A B twin C A C B parent A

(a/6)[112]

twin

parent

[001] [110] [112]

Figure 75: Stacking sequence changes during shearing of {111} planes in the fcc structure to form a twin.

Each successive atomic plane is displaced by a/6 11 2 . The magnitude of this displacement is = a/ 6. Hence, the twinning shear in fcc metals is a/ 6 1 = = (5.80) Twinning shear = d111 a/ 3 2

DH78

Course D: Mechanical Behaviour of Materials

DH78

Effect of temperature: Deformation twinning is athermal (does not require thermal activation). This is in contrast to dislocation glide, which does show temperature dependence, particularly if cross-slip or climb are required. Deformation twinning is therefore favoured at lower temperatures. The temperature dependence of dislocation glide and deformation twinning are shown in Fig. 76 below.

crit

Ambient T

Twinning Slip (high strain rate) Slip (low strain rate)

T
Figure 76: Effect of temperature and strain rate on dislocation slip and deformation twinning.

Effect of strain rate: Deformation twinning typically requires higher shear stresses than dislocation glide (at room temperature). However, it occurs very quickly, propagating at the speed of sound in the material. As such, deformation twinning may become favoured at higher strain rates.

Effect of crystal structure: Deformation twinning is more likely to occur in crystal systems in which limited slip systems are available (e.g. hcp metals) or where dislocation mobility is impaired.

DH79

Course D: Mechanical Behaviour of Materials

DH79

Morphology of twins Deformation twins typically display a lenticular (lens like) shape as it is necessary to maintain strain compatibility at grain boundaries. In contrast, annealing twins formed naturally at high temperature in the absence of stress typically display straight interfaces. Such twins are common in materials with very low twin boundary energies.

Annealing Twins

Deformation Twins

Figure 77: Schematic representation of the morphologies of annealing and deformation twins.

Figure 78: Annealing twins in an brass (DoITPoMS micrograph library number 430)

Figure 79: Deformation twins in commercially pure Ti. (Courtesy of N.G. Jones)

All twins crystallographically-related to parent grain Annealing twins formed by reconstructive (diffusional) processes during heat treatment (often as recrystallisation occurs) boundaries usually straight (fully coherent) common in materials with very low twin boundary energy Deformation twins formed by cooperative shear, accommodating imposed shear strain often lens-shaped, since anchored to grain boundaries at ends

DH80

Course D: Mechanical Behaviour of Materials

DH80

Fracture

The catastrophic extension of cracks throughout the thickness of a component is termed fracture. In brittle materials, where no plastic deformation occurs, brittle fracture is to be expected. However, it is also possible for materials that exhibit plastic deformation to fail in a similar catastrophic fashion by ductile fracture. In both cases, understanding the circumstances under which fracture may occur is therefore critical and has been addressed through the eld of fracture mechanics. In brittle materials, the path of the fracture surface may pass straight across grain boundaries in transgranular fracture or follow the path of grain boundaries in intergranular fracture. Examples of the fracture surfaces from these two modes of failure are shown below.

Figure 80: Example of a transgranular fracture surface on a notched low carbon ferritic steel (DoITPoMS micrograph library number 143).

Figure 81: Example of a intergranular fracture surface on HY100 forging steel (DoITPoMS micrograph library number 146).

DH81

Course D: Mechanical Behaviour of Materials

DH81

6.1

Estimate of ideal fracture stress

The ideal fracture stress of a material can be estimated by considering the interatomic forces between atoms. If the atomic bonds across the surfaces that are being pulled apart are separated by a dis2 bonds crossing the area, A. The force on each bond is theretance of r , there will be A/r fore 2 F = r (6.81) Given that the force on a bond may be related to the potential energy of the bond, U , by F = The stress across the bond is = dU dr (6.82)

1 dU 2 dr r

(6.83)

The Youngs modulus is related to the potential energy of the bond by E= 1 d2 U r dr2 (6.84)
r

Using the Lennard-Jones potential, the Youngs modulus can be approximated by E= Hence, 72Umin 3 r (6.85)

Substituting this expression for Umin tained = This has a maximum value at

3 Er (6.86) 72 into the Lennard-Jones potential, the stress can be ob-

Umin =

E 6

r r

r r
1 6

13

(6.87)

r = r

13 7

(6.88)

Substituting this value into the equation for the stress, the maximum stress is therefore max = 7 169 7 13
1 6

E = 0.0373E

(6.89)

For a material with a Youngs modulus of 300 GPa, a value typical of many ceramics, the expected fracture stress is therefore 11.2 GPa. In practice is it almost impossible to approach such theoretical fracture stresses due to the presence of cracks, holes or aws in the material. We will examine this effect in the next section.

DH82

Course D: Mechanical Behaviour of Materials

DH82

6.2

Grifth criterion

Alan Arnold Grifth, one of the pioneers of fracture mechanics, used energetic considerations to determine when a crack present in a material under stress will grow. For such a crack to grow, sufcient energy must be supplied to allow the crack to extend. This energy can come from the stored elastic strain energy in the material or work done by the applied stress. Following this approach, consider a plate subjected to an applied stress, , with a pre-existing crack of length 2c at its centre, as shown in Fig. 82. As the crack extends by a length, dc, at each end it relaxes the elastically strained material around it. This is shown by the dotted region.

dc c

0
Figure 82: Stress-free region shielded by a crack from the applied load.

The strain energy per unit volume in an elastically stressed material is given by 2 E2 = 2 2E (6.90)

The strain energy released by the extension of the crack by dc at each end is the product of this expression and the increase in the stress-free volume. The shape of the stress-free region is not well-dened, and the stress was not uniform within it before the crack advanced, but taking the relieved area to be twice that of the circle having the crack as the diameter gives a fair approximation. Thus, for a plate of thickness, t, the energy released during incremental crack advance under a stress of is given by dW =
2 2 2 c t dc 2 (2c t dc) = 2E E

(6.91)

DH83

Course D: Mechanical Behaviour of Materials

DH83

6.2.1

The strain energy release rate

The concept of stored elastic strain energy being released as the crack advances is central to fracture mechanics. The strain energy release rate (crack driving force) is usually given the symbol G (not to be confused with shear modulus or Gibbs free energy!). It is a rate with respect to the creation of new crack area (and so has units of J m2 ) and does not relate to time in any way. It follows that G=
2 c t dc /E 2 dW = = new crack area (2t dc) 2c E

(6.92)

The value of the constant ( in this case) is not well dened. It depends on specimen geometry, crack shape/orientation and loading conditions. In any event, the approximation used 2 c/E ) is more for the stress-free volume is simplistic. However, the dependence of G on ( general and has important consequences.
2c E

The strain energy release rate is therefore: * Larger for larger cracks (aws) * More energy released (per unit of new crack area) when longer cracks propagate * Larger under larger applied stresses

6.2.2

Fracture energy (crack resistance)

In order for crack propagation to be possible, the strain energy release rate must be greater than or equal to the rate of energy absorption (expressed as energy per unit area of crack). This energy requirement is sometimes known as the Grifth criterion. For a brittle material this fracture energy is simply given by 2 (where is the surface energy, with the factor of 2 arising because there are two new surfaces created when a crack forms). It can be considered as a critical strain energy release rate, Gc . It is a material property. It is sometimes termed the crack resistance or work of fracture. The fracture strength can therefore be written as G Gc = 2 (6.93)

Therefore, combining this inequality with the expression for the strain energy release rate we obtained earlier 2c Gc = 2 (6.94) E Rearranging this expression, the fracture stress, , is = 2E c (6.95)

DH84

Course D: Mechanical Behaviour of Materials

DH84

6.2.3

Critical aw size and toughening by aw removal

If a component is to be subjected to a particular stress level in service, then the concept emerges of a critical aw size, c ,which must be present in order for fracture to occur c 2E 2 (6.96)

Fig. 83 shows predicted values of the fracture stress for Al2 O3 obtained using this equation and data from the databook (Gc = 0.05 kJ m2 , E = 390 GPa). Note that the tensile strength of Al2 O3 given in the databook of 500 MPa corresponds to a critical crack size of 0.025 mm .

Fracture stress (MPa)

6 4 2

100
6 4 2

10 0.01 0.1 1 10 100 Critical crack length (m)

Figure 83: Fracture stress of Al2 O3 as a function of crack size.

If high stress levels (several hundred MPa) are to be sustained, then all sizeable aws ( 1 m) must be removed. This can often be done by various kinds of surface treatment, some of which put the surface into compression (close up cracks). However, it should be noted that, while these procedures can raise the maximum tolerable stress, they do not actually change the material toughness (fracture energy), which is still equal to 2 (for a brittle material). Its also worth noting that certain conditions, such as a corrosive environment or a repeatedly cycled load (fatigue), can cause cracks to elongate progressively with time, even if they are initially below the critical length for fast (unstable) fracture. Clearly this may eventually lead to unstable (energetically favoured) crack growth. Fortunately, many materials are not brittle, and so can sustain relatively high stress levels without fracturing, even in the presence of large aws. Can apparently toughen by reducing the size of the largest aw present. As most aws are likely to be found on the surface polish surface put surface into compression (eg shot peen, quench surface of glass plates) (BUT, doesnt raise the fracture energy that is real toughening of material)

DH85

Course D: Mechanical Behaviour of Materials

DH85

6.3

Ductile fracture

In the calculations in the preceding section we did not consider the effect of stress concentrations around the crack tip.

Figure 84: Schematic representation of the lines of force that lead to stress concentrations around crack tips.

Inglis examined this mathematically, focussing on the crack tip region, and derived the following expression for the peak stress, max , in terms of the applied stress, . max = 0 1 + 2 c r (6.97)

where 2c is the crack length (c for a surface crack) and r is the radius of curvature at the tip. (It follows that the stress concentration factor for a circular hole is 3.) The stress distribution ahead of a crack tip is shown schematically in Fig. 85.
(a)

0 c

(b)

2c

max 0

0
Figure 85: Stress concentration at a crack tip: (a) loading of a at plate containing an ellipsoidal crack and a surface crack and (b) schematic stress distribution in the vicinity of the ellipsoidal crack.

In ductile materials, the high stresses in the vicinity of the crack tip may result in local plastic deformation producing a zone of plasticity. This is shown in Fig. 86. This local plasticity limits the peak stress to approximately the yield stress of the material, y , and increases

DH86

Course D: Mechanical Behaviour of Materials

DH86

the radius of curvature, r, at the crack tip and thereby also reduces the stress concentration effect. Critically, this local plastic deformation requires work to be done, most of which is dissipated as heat. This demands a much higher strain energy release rate to propagate the crack.

0 max 0

0 max = y X Y 0

Figure 86: Crack tip shapes and stress distributions for (a) brittle and (b) ductile materials.

Stress concentration at crack tip stimulates plasticity blunts the crack tip reduces the peak stress (to Y ) increases energy absorption

DH87

Course D: Mechanical Behaviour of Materials

DH87

6.3.1

Effect of plasticity on fracture energy

The Grifth criterion is only truly valid for brittle materials. However, Irwin proposed (1948) that the Grifth condition could be modied to include the plastic work done during crack advance, p G Gc = 2 ( + p ) (6.98) As before, we can write
2c E

Gc = 2 ( + p )

(6.99)

which can be rearranged to give the associated fracture stress = 2 ( + p ) E c (6.100) Gc

The extra energy required by p make a big difference. Gc for metals ( 1-100 kJ m2 ) brittle materials (ceramics, glasses etc) ( 1-100 J m2 )

6.3.2

Stress intensity factor

In most cases, the Inglis stress concentration equation (Eqn. 6.97) can be approximated by c c max = 0 1 + 2 2 (6.101) r r Since a critical stress is needed at the crack tip to open up the atomic planes, it follows that crack propagation is expected when c critical value (6.102) where the critical value is expected to be constant for a given material, but to vary between materials (max r is constant for a given material). In the 1950s, Irwin proposed the concept of a stress intensity factor, K , such that K = c (6.103) The stress intensity factor (units of MPa m) scales with the level of stress at the crack tip. Fracture is expected when K reaches a critical value, Kc , the critical stress intensity factor, which is often termed the fracture toughness. K Kc {crack tip conditions} analogous to G Gc {global energy balance}

DH88

Course D: Mechanical Behaviour of Materials

DH88

6.3.3

Uniting the stress and energy approaches

There are clearly parallels between K reaching a critical value, Kc , and G reaching a critical value of Gc . The magnitude of K can be considered to represent the crack driving force, analogous to G. Consider again the Grifths energy criterion G Gc Therefore, , i.e.
2c E

=G

(6.104)

EGc = Kc

(6.105)

It can be seen that this is similar in form to Eqn. 6.103. It follows that K = EG and Kc = EGc Kc fracture toughness (critical stress intensity factor) Gc fracture energy (critical strain energy release rate)

(6.106)

While it is reassuring to be able to treat fracture from both stress and energy viewpoints, it is not immediately apparent what advantages are conferred by using a stress intensity criterion, rather than an energy-based one. However, in practice it is possible to establish K values, and corresponding Kc values, for various loading and specimen geometries, whereas this is not really possible with an energy-based approach. (The 3-D stress state at the crack tip, and hence the size and shape of the plastic zone, can be affected by specimen thickness and width.) Moreover, rates of sub-critical crack growth (progressive advance of a crack, due to a corrosive environment or cyclic loading) can often be predicted from the stress intensity factor, since it is directly related to conditions at the crack tip, whereas the strain energy release rate is a more global parameter.

DH89

Course D: Mechanical Behaviour of Materials

DH89

6.3.4

Ductile rupture

The preceding section suggests that the Grifth criterion may be applied to ductile materials and that they should be as sensitive to aws as brittle materials, although they would have higher fracture stresses for a given crack size. In practice, failure stresses are not systematically higher for ductile materials, although they certainly fracture less readily and require much more energy input. Furthermore, ductile materials show little or no sensitivity to initial aw size. The differences between these modes of fracture, which are illustrated in Fig. 87 below are clearly important.

Figure 87: Micrographs of fractured specimens, and schematic depictions, of different modes of failure, showing (a) ductile rupture, (b) moderately ductile fracture and (c) brittle fracture.

Very ductile materials often fail by ductile rupture (progressive necking down to a point), with little or no crack propagation as such. The failure stress in such cases can only be predicted by analysis of the plastic ow and Eqn. 6.100 is irrelevant. In practice, such highly ductile materials are too soft to be useful for most purposes. However, fracture of engineering metals is commonly preceded by signicant amounts of plastic ow and some necking, (Fig. 87(b)). Clearly, use of Eqn. 6.100 requires care in such cases. It should be noted that, even if the fracture stress is not very high, a good toughness will ensure that components are durable and robust under a range of service conditions. An example of a high magnication image of fracture surface following Figure 88: Example of a fracture surface on aluminium (DoITPoMS micrograph library ductile rupture is shown in Fig. 88.
number 83).

Often get extensive macroscopic plastic ow before fracture Necking True stress at fracture of the neck may be high, but the nominal stress (apparent strength) is often quite low

DH90

Course D: Mechanical Behaviour of Materials

DH90

6.4

The Ductile-Brittle Transition Temperature (DBTT)

Toughness is sensitive to plastic ow characteristics, which can change with T . As such, fracture behaviour can change on heating or cooling. In general, heating leads to increased plasticity and hence an improved toughness. There is thus less concern about fracture at higher T , and more about excessive plastic ow. Conversely, as T is reduced, fracture becomes of greater concern. There is a broad relationship between strain rate and T . Increasing the strain rate is often similar to reducing T , since both tend to inhibit dislocation motion. Hence, fracture is often more likely if a higher strain rate is imposed (although relatively large changes in strain rate may be needed to generate the same effect as a relatively small change in T ). It is conventional when investigating the onset of brittleness that a fairly high strain rate is used to reect the most demanding conditions likely to be encountered in service, e.g. by impacting a specimen with a pendulum. This also provides a convenient method of characterising the toughness, since the recovery height of the pendulum after fracturing the specimen gives an indication of the energy absorbed, which may be related to the specimen toughness.

Figure 89: Impact energy as a function of temperature for several materials, some of which exhibit ductile to brittle transitions on cooling.

The data shown in Fig. 89 illustrate how the toughness (measured by impact testing) of a few materials varies with T in the range from 100 C to 150 C. An fcc metal such as Cu remains tough down to very low T , whereas an hcp metal like Zn, in which dislocation motion occurs by basal slip, may be quite brittle even at ambient T . Interest centres mainly on steel (and other bcc metals). Dislocations are less mobile in bcc metals than in fcc and are more prone to pinning by interstitials. Grain boundary embrittlement by segregation of impurities, such as P and S, can also occur in steels. The transition T may be around ambient, or not far below, so that service conditions may be encountered (e.g. in polar regions) in which the toughness is much lower than expected. Steels have, however, been developed with good toughness down to very low T .

DH91

Course D: Mechanical Behaviour of Materials

DH91

6.5

Toughness of composites

In composite materials, both the matrix and reinforcing bres may be brittle but the composite may still exhibit considerable toughness. A component such as a vaulting pole will acquire many, relatively large, surface defects, so the fact that it can sustain such high stresses without fracturing indicates that, as a material, the composite has a high toughness. The main energy-absorbing mechanism raising the toughness of bre composites is the pulling of bres out of their sockets in the matrix during crack advance - see Fig. 90.

Figure 90: (a) Schematic of a crack passing through an aligned bre composite, showing interfacial debonding and bre pull-out, and (b) a SEM micrograph of a composite fracture surface.

Consider a bre with a remaining embedded length of x being pulled out an increment of distance dx. The associated work is given by the product of the force acting on the bre and the distance it moves dW = (2rx) i dx (6.107)

where i is the interfacial shear stress, taken as constant along the length of the bre. The work done in pulling this bre out completely is therefore given by
x

W =
0

2rxi dx = rx2 i

(6.108)

where x is the pull-out length. The number of bres per unit sectional area, M , is related to the bre volume fraction, f , and the bre radius, r, by M= so the pull-out work of fracture, Gcp , is given by f r2 (6.109)

DH92

Course D: Mechanical Behaviour of Materials

DH92

Gcp = M W =

f rx2 i = 4f s2 ri r2

(6.110)

where s is the pull-out aspect ratio (x /2r). Since there will be a range of values, an (RMS) average value should be used. The process can make large contributions to the fracture energy. For example, with f =0.5, s=20, r=5 m and i =20 MPa we get a fracture toughness of 80 kJ m-2 , a value typical of a tough metal. It might be expected that continuous bres would break in the crack plane, where the stress is highest, so there would be no pull-out. However, since the bres are brittle, aw-sensitive materials, they will often break some distance from the crack plane, where there is a aw, even though the stress is somewhat lower there. Pull-out aspect ratios of the order of 2050 (e.g. pull-out lengths of several hundred m for a 10 m diameter bre) are commonly observed - see Fig. 90(b). It may be necessary to carefully control certain bre and interfacial characteristics in order to maximise the pull-out toughness.

DH93

Course D: Mechanical Behaviour of Materials

DH93

6.6

Pressurised pipes

A common and important structural application where fast fracture should be avoided is in pressurised pipes. The material from which they are fabricated should be sufciently tough that aws and cracks can be detected before they cause fast fracture. Consider a cylindrical pipe of length, L, and radius, r, with an internal pressure, P , as shown in Fig. 91.

Figure 91: Derivation of the stresses in the wall of an internally pressurised cylinder.

The force associated with the hoop stress, , in the pipe wall must balance the force from the internal pressure acting over the projected area across the pipe. (2tL) = P (2rL) Hence the hoop stress in a pressurised pipe is = Pr t (6.112) (6.111)

Similarly, the force associated with the axial stress, z , in the pipe wall must balance the force from the internal pressure acting over the projected area along the pipe. z (2rt) = P r2 Hence the axial stress in a pressurised pipe is z = Pr 2t (6.114) (6.113)

It can be seen from these expressions that the hoop stress is twice the axial stress. Many components & structures in this condition Boilers, pressure vessels etc Hoses, pipelines etc Many biological structures (blood vessels etc) Submarines (external pressure compressive stresses) Aircraft & spacecraft

DH94

Course D: Mechanical Behaviour of Materials

DH94

6.7

Aircraft stresses and materials

Aircraft are subject to both bending moments and pressurisation stresses. During ight, wings experience bending moments, generating compressive stresses in the top surface and tensile stresses in the underside. (During landing, these stresses are reversed.) The bending moment, and hence the stress levels, are highest close to the fuselage. Air regulations require that the cabin pressure should be at least 0.8 bar. At typical long-haul operating heights, atmospheric pressure is 0.2 bar, so the over-pressure 0.6 bar. Resultant stresses in the fuselage skin clearly depend on cabin diameter and skin thickness, but often reach signicant levels. Parts of the fuselage are also subject to bending moments, since it is being supported where it is attached to the wings, but gravity acts on the parts fore and aft of this region. A schematic depiction of stress distributions during ight is shown in Fig. 92.

Figure 92: Schematic depiction of the stress distributions in a large aircraft during ight at high altitude

NB Airframes are made of either Al alloy or composite material Both have low density, but neither has a very high fracture toughness Must design to avoid both fast fracture and fatigue crack growth Create circular aw (with needle, diameter 100 m) c < 1 m (if rubber ideally brittle) fast fracture predicted? NB Sticky tape constrains relaxation of rubber reduces energy release rate

DH95

Course D: Mechanical Behaviour of Materials

DH95

Appendix: Observing dislocations

This is non-examinable and has been included for reference only.

High resolution transmission electron microscopy With modern transmission electron microscopes, it is possible to image the atoms in the vicinity of the dislocation core directly. From such images it is possible to determine the Burgers vector of the dislocation from a Burgers circuit. However, imaging dislocations in this way is challenging as it requires the most advanced electron microscopes and extremely careful sample preparation. A useful and effective method of determining the character of dislocations is through diffraction contrast.

Diffraction contrast imaging of dislocations As shown in Fig. 93, the local distortion of the lattice planes in the vicinity of a dislocation core may result in it fullling diffraction criteria. The diffraction of electrons away from the transmitted beam in such regions will show the dislocations as dark lines in bright eld images.
ee-

Strong transmitted beam away from the dislocation core

Strong diffracted beam in the vicinty of the dislocation core

Figure 93: Schematic illustration showing the origin of diffraction contrast in transmission electron microscopy in the vicinity of dislocation cores.

Similarly, if a dark eld image is taken using the diffraction spots from the crystal, the region in the vicinity of the dislocation will not diffract at the same angle as the surrounding crystal and will again appear dark in the image. However, as those lattice planes with normals perpendicular to the Burgers vector of the dislocation remain undistorted, they will not diffract at an angle different to the perfect crystal. If a bright eld image is formed from one of these spots, the dislocation will be invisible.

DH96

Course D: Mechanical Behaviour of Materials

DH96

We can therefore use this phenomenon to determine the Burgers vectors of dislocations by taking bright eld images from a selection of spots with different reciprocal lattice vectors and seeing in which ones the dislocation lines are invisible. When they are invisible gb=0 (7.115)

in which g is the reciprocal lattice vector of the diffraction spot used to generate the bright eld image and b is the Burgers vector of the invisible dislocation line segment. An example of a g b analysis is shown in Fig. 94.

Figure 94: A series of transmission electron images of intersecting dislocation segments. The operative strongly diffracting beam g is shown on each image. The Burgers vectors of the various segments have been indexed using the g b criterion.

DH97

Course D: Mechanical Behaviour of Materials

DH97

Glossary

Terms are listed alphabetically. Words in italics provide cross-references to other entries in the glossary. Beam stiffness () Product of Youngs modulus and moment of inertia, characterising the resistance to deection when subjected to a bending moment. Bending moment (M ) Turning moment generated in a beam by a set of applied forces. The bending moment is balanced at each point along the beam by the moment of the internal stresses. Brittle fracture Fracture not involving gross plastic ow, although some local plastic deformation may occur at the tip of the crack. Burgers vector (b) Vector giving magnitude and direction of lattice displacement generated by passage of a dislocation. It is a lattice vector for a perfect dislocation. Climb Movement of an edge dislocation by absorption or emission of a vacancy. Coefcient of thermal expansion () See thermal expansivity . Critical resolved shear stress (c ) Value of resolved shear stress at which slip occurs on a specied slip system. Cross-slip Movement of a screw dislocation from one permissible slip plane on to another, usually in order to bypass an obstacle. Curvature () Reciprocal of the radius of curvature adopted by a beam subject to a bending moment. Also equal to the through-thickness gradient of strain in the beam. Deformation twinning Mode of plastic deformation involving co-operative shear of atomic arrays into a new orientation of the same crystal structure (reecting the parent orientation across the twin plane). Also termed mechanical twinning. Dislocation line Boundary between slipped and unslipped regions of a crystal. Dispersion Strengthening Raising of yield stress via obstacles to dislocation glide that are thermally stable. Ductile fracture Fracture involving gross plastic ow. Edge dislocation Dislocation with a Burgers vector normal to the dislocation line. Engineering strain () Normal strain, given by ratio of change in length to original length. Also called nominal strain. See also true strain. Engineering stress ( ) Normal stress, given by ratio of applied force to original sectional area. Also called nominal stress. See also true stress. Fracture energy (Gc ) Energy per unit area required to extend a crack. Also termed the critical strain energy release rate. Fracture toughness (Kc ) Critical value of the stress intensity factor, characterising the resistance of a material to crack propagation.

DH98

Course D: Mechanical Behaviour of Materials

DH98

Franks rule A dislocation reaction will occur only if the energy of the product dislocation(s) is less than that of the reacting dislocation(s). The energy of a dislocation is proportional to the square of its Burgers vector. Frank-Read source A mechanism by which dislocations can multiply. Glide Motion of a dislocation on its slip plane by small co-operative movements of atoms close to the dislocation core. Glissile dislocation A dislocation which is able to glide. Grifth criterion Energy-based condition for fracture. Hookes Law Stress is proportional to strain during elastic deformation. Jog A step in a dislocation that does not lie in the slip plane. Kink A step in a dislocation that lies in the slip plane. Martensitic transformation Phase transformation occurring via co-operative shear of atomic arrays into a new crystal structure. Mechanical twinning See deformation twinning. Mixed dislocation Dislocation in which the Burgers vector is neither parallel nor perpendicular to the dislocation line. Neutral axis (of a beam) Axis (strictly, a plane) parallel to the length of a beam, along which there is no change in length on bending. Nominal strain () See engineering strain . Nominal stress ( ) See engineering stress . Normal strain () Deformation in which change in length is parallel to original length. Normal stress ( ) Stress induced by a force acting normal to the sectional area to which it is applied. OILS rule Non-graphical method to nd the slip system with the highest Schmid factor in a cubic crystal. Orowan bowing The bowing of dislocations between precipitates that cannot be cut. Partial dislocation A dislocation for which the Burgers vector is a smaller than a lattice vector. Perfect dislocation A dislocation for which the Burgers vector is a lattice vector. Primary slip system Slip system which rst becomes active. Normally the one with the highest Schmid factor. Resolved shear stress (R ) Component of shear stress acting on a slip plane and parallel to a slip direction in that plane. Schmid factor (cos cos ) Geometrical factor relating resolved stress to normal stress along the tensile axis in a single crystal under tension or compression. Schmids law Slip initiates at a critical value of resolved shear stress.

DH99

Course D: Mechanical Behaviour of Materials

DH99

Screw dislocation Dislocation with a Burgers vector parallel to its dislocation line. Second moment of area (I ) Parameter dependent on sectional shape, characterising the resistance a beam offers to deection under an applied bending moment. Also called the moment of inertia . Sessile dislocation A dislocation which is unable to glide. Shear modulus (G) Constant of proportionality between shear stress and shear strain. Shear strain ( ) Distortional deformation (an angle) arising from a shear displacement. Shear stress ( ) Stress induced by a force acting parallel to the sectional area to which it is acting. Slip system [UVW ] (hkl ) Combination of slip plane (hkl ) and slip direction [UVW ], which lies in the (hkl ) plane. Solution strengthening Increase in yield stress of a material through the addition of solute atoms, caused by the interaction of these solute atoms with dislocations. Stacking fault A two-dimensional defect in which the sequence of stacking of atomic planes is interrupted. Strain energy release rate (G) Elastic strain energy released per unit of created crack area. Strain hardening see work hardening. Stress intensity factor (K ) Parameter characterising the crack driving force, in terms of applied stress level and crack length. Surface energy ( ) Energy per unit area associated with a free surface. Thermal expansivity () Ratio of change in length to original length, per unit change in temperature. Also known as coefcient of thermal expansion. True strain () Ratio of change in length to current length. May differ from the nominal, or engineering, strain. True stress ( ) Ratio of force to current sectional area over which it is acting. May differ from the nominal, or engineering, stress. Twinning plane Plane within which shear takes place during deformation twinning, which also forms a mirror plane between parent and twin structures). Work hardening Increase in plastic ow stress with strain. Associated with an increase in dislocation density, formation of entanglements, locks etc. Also known as strain hardening. Yield point Point (stress level = y ) at which plastic deformation rst occurs. Youngs modulus (E ) Constant of proportionality between normal stress and normal strain .

DH100

Course D: Mechanical Behaviour of Materials

DQ1

Question Sheet 13
1. A titanium wire of circular section has a diameter of 200 m. What would cause it to exhibit the greater elongation: (a) suspending a weight of 1 kg from it or (b) heating it up by 300 C? (Property data are supplied in the Data Book and the coefcient of thermal expansion can be obtained from DH18. State any assumptions.) 2. The potential energy, U , between two carbon atoms joined by a single covalent bond and separated by a distance, x, can be approximated by U = A B + 8 2 x x

where A = 9.0 10-39 J m2 and B = 3.0 10-98 J m8 . Calculate the equilibrium separation of the atoms. If solid carbon had a simple cubic structure (primitive cubic lattice, with one atom per lattice point) estimate the Youngs modulus, E . For diamond, E is actually 1200 GPa, while for graphite it is 27 GPa (averaged over all directions). Compare these values with the value you have estimated for the simple cubic structure, and discuss the reasons for any discrepancies. Polyethylene ([-CH2 -]n ) contains long hydrocarbon chains but exhibits a much lower value of E , typically 0.2 to 0.7 GPa, why is this? 3. A composite (60% aligned glass bres (7 m diameter) in an epoxy matrix) is in the form of thin sheets. Using equal strain and equal stress models respectively, calculate the axial and transverse Youngs moduli. A cross-ply is produced by bonding together several sheets, each with its bre direction normal to that of its neighbours. Using an equal strain assumption, calculate the Youngs modulus of such a material, parallel to either of the two bre directions. (Youngs moduli: for glass, E = 76 GPa: for epoxy, E = 3 GPa) 4. A cantilever is loaded with a static force, F , at its tip, as shown in the gure below.
L x z F

Obtain an expression for the deection, z , as a function of position along the cantilever, x. You may take the relationship between the curvature, (1/R), and the deection, z , to be 1/R = d2 z/dx2 .

DH101

Course D: Mechanical Behaviour of Materials

DQ1

Hence show that the deection at the tip of the cantilever, , is = F L3 3EI

At what position does the cantilever experience the largest tensile stress? At this position, what stress will be experienced by a solid, square sectioned bar of thickness, b?

DH102

Course D: Mechanical Behaviour of Materials

DQ2

Question Sheet 14
1. Determine whether the following dislocations in aluminium are edge, screw or mixed. Identify in each case the slip plane(s) on which the dislocation will be able to glide. .) (Al has an fcc crystal structure and its slip system is {111} 110 Burgers vector, b, (a) (b) (c) [1 10] [101] [1 10] to Line vector [ 110] [112] [ 1 12] to

directions. List all the physically dis2. Aluminium (fcc) slips on {111} planes in 110 tinct slip planes in this structure, and for each plane list all the physically distinct slip directions. How many physically distinct slip systems are there? A single crystal of aluminium is subjected to tensile load along [123]. For each of the distinct slip systems, evaluate the Schmid factor (cos cos ), and hence determine which slip system will operate rst. Show that the OILS rule leads to the same prediction. 3. A cylindrical single crystal of copper (fcc) is tested in tension along the axis parallel to Which slip system has the highest Schmid factor? [125]. As the specimen is deformed, how does the orientation of the tensile axis change? Deformation is continued until the Schmid factors for two slip systems become equal. What is the orientation of the tensile axis at this point, and what are the two systems? 4. A bar of polycrystalline aluminium is plastically deformed in tension, to a strain of 10%. It is found that the ow stress (yield stress) is initially 50 MPa, rising linearly with increasing strain to 150 MPa at 10% (plastic) strain. Estimate the mechanical work done on the bar per unit volume (= area under stress-strain plot). While this plastic deformation is taking place, the dislocation density in the bar increases by a factor of 10, from 1013 to 1014 m2 . Using the standard expression for the stored elastic strain energy per unit length of dislocation line, and assuming that it applies to all types of dislocation, estimate the increase in stored energy per unit volume in the bar, associated with this increase in dislocation density. Explain why the mechanical work done is much greater than the energy associated with the increased dislocation density and describe the fate of the lost energy. (Al is fcc, with a = 0.405 nm and shear modulus G = 28 GPa)

DH103

Course D: Mechanical Behaviour of Materials

DQ3

Question Sheet 15
1. The following refer to dislocations in a fcc metal. In each case, determine whether the dislocation will move under the action of the stress, and if so, in what direction will its line move? (a) a screw dislocation with b (b) a screw dislocation with b (c) an edge dislocation with b ). [110] subjected to a shear stress along [110] on (111 ] on (111 ). [110] subjected to a shear stress along [112 ] subjected to a shear stress along [011 ] on (111). [011

2. The peak yield stress, y , of an age hardening Al-Cu alloy is 430 MPa. When heavily over-aged (precipitates too widely spaced to affect dislocations), y falls to 100 MPa. Assuming that the difference between these two yield stress values corresponds to the Orowan bowing stress, estimate the spacing between precipitate surfaces in the peakaged alloy. In the peak-aged condition, the precipitate morphology is thin cylindrical discs (thickness 20% of diameter). Estimate the volume fraction of precipitates present, assuming that they are mono-sized, with a disc diameter of 20 nm, and distributed in a simple cubic array (The average peak shear stress on slip planes in a polycrystal (with randomly-oriented grains) is 1/3 (= Taylor factor) of the applied uniaxial stress. Al is fcc, with a = 0.405 nm and shear modulus G = 28 GPa) 3. During deformation of an aluminium alloy, a dislocation with Burgers vector b1 = a 2 [110] a moving on (111) interacts with a second dislocation with Burgers vector b2 = 2 [101] moving on (111). Using Franks rule, show that it is energetically favourable for these dislocations to add. Describe the character of the resultant dislocation. What is the signicance of the new slip system in terms of continued plastic deformation? 4. Outline the factors favouring deformation twinning during imposed plastic straining. Calculate the twinning shear (= displacement of an atom in the twinning direction / distance of the atom from the twin-parent boundary plane) for the fcc structure. direcIf the same shear strain were to be generated by dislocation glide (in a 110 tion), with slip occurring on ({111}-type) planes spaced 1 m apart, estimate how many dislocations would need to glide across each slip plane. (Assume a 0.4 nm.) 5. The following table lists the values of the tensile yield stress for four copper-based materials. In each case, explain briey the dominant mechanism which gives rise to the value of the yield stress. (a) 99.99% Cu, single crystal, y = 10 MPa (b) 99.99% Cu, polycrystalline, annealed, grain size 10 m, y = 60 MPa (c) material (b) after deformation to 50% plastic strain, y = 325 MPa (d) Cu - 2 at.% Be, polycrystalline, annealed, grain size 10 m, y = 185 MPa

DH104

Course D: Mechanical Behaviour of Materials

DQ4

Question Sheet 16
1. Explain why the strengths of bulk ceramics and inorganic glasses are much lower in tension than in compression, and outline methods by which the tensile strengths of brittle materials may be increased. A material with tensile yield stress, y , Youngs modulus, E , and toughness, Gc contains an internal crack of half-length, a. What condition must be obeyed for the material to yield in tension rather than failing in a brittle manner? Use this condition and information from the databook to rank the following materials in order of increasing aw tolerance: polyethylene, high strength steel, aluminium alloy, and alumina (Al2 O3 ). 2. A small stone weighing 30 g, thrown up from the tyre of a car, strikes the windscreen of a following car with an impact velocity of 50 km h1 . Calculate its kinetic energy. The windscreen, which has an area of 2 m2 and is 8 mm thick, fractures throughout in a mosaic of through-thickness cracks, creating segments approximating to squares of side 5 mm. If the surface energy, , of the glass is 2 J m2 , estimate the energy associated with the newly-created surface area. Your calculations should indicate that the impact energy is insufcient to create the observed new surface area, even if the fracture is ideally brittle. Its suggested that the extra energy could have been supplied by release of residual stresses in the glass, which were deliberately created during manufacture by quenching the surface layers from high temperature with air jets (in order to toughen the glass). This process imparts a residual compressive stress of 30 MPa into 1 mm-thick surface layers on both sides. Taking the Youngs modulus, E , of the glass to be 70 GPa, estimate the associated stored energy (= 1/2 stress strain volume) and check whether it is sufcient to account for the observed fracture behaviour. Comment on the nature of toughening of windscreens achieved in this way. 3. Explain carefully why the Grifth fracture criterion cannot be used to obtain the failure stress of a typical engineering metal, such as mild steel, being subjected to a conventional tensile test. What information would be needed in order for this tensile strength to be estimated by analysing the fracture behaviour? 4. The air pressure in a large aircraft is 0.8 atm (0.08 MPa) and it ies at an altitude where atmospheric pressure is 0.2 atm. The fuselage approximates to a cylinder of diameter 7 m and skin thickness 3 mm, with a length behind the wings of 40 m. The weight distribution in the rear half of the plane, taking account of lift provided by the tailplane, is equivalent to cantilever loading with a weight of 25 tonnes (25,000 kg) acting on the end of the fuselage. Where is the peak tensile stress in the fuselage and what is its orientation and magnitude? An Air Marshall on the plane inadvertently res his weapon. Despite prior assurances that this is impossible, the bullet penetrates the fuselage skin, leaving a jagged hole 10 mm in diameter (in the region of peak tensile stress). Neglecting any effects of depressurisation, will major disaster follow in the form of rapid crack propagation around

DH105

Course D: Mechanical Behaviour of Materials

DQ4

the fuselage? As cabin pressure drops, does the likelihood of such fracture increase or decrease? (The fuselage is made of an aluminium alloy (2024), which has a fracture toughness, Kc 40 MPa m. The moment of area, I , of the section of a hollow, thin-walled cylinder, diameter D, wall thickness t, is approximately D3 t/8)

Vous aimerez peut-être aussi