You are on page 1of 33

MATS1101 - Materials - Bonding Between Atoms

Introduction to Materials Engineering


Why Study Materials? To understand, predict and control material behaviour. Enables us to use the correct materials for a given application and understand the limitations on their use.

To determine material behaviour we must understand the material properties, which are a result of the internal structure of the material. We need therefore to understand the structure/property relationship. Internal Structure-determines->Material Properties-dictates->Material Behaviour Atomic Bonding All materials are made up of atoms, and these atoms are held together by forces (called interatomic bonding). These forces act like springs, linking each atom to its neighbour. These atoms are arranged in different ways in different materials. The two important things to understand for atom packing are: Number of bonds per unit area. The angle of the bonds. There are two different ways that atoms can be bound together. Primary bonds. These are strong bonds and may be ionic, covalent or metallic. Secondary bonds. These are weak bonds and may be Van der Waals or hydrogen. Note: Although these different bonds are distinguished, in many materials there exists a mixture of different bonds. Primary Bonds All elements except inert gases have an unfilled outer shell of electrons (valence electrons). Primary bonding occurs when electrons are lost or gained so that the outer shell is filled. Ionic Bonding The ionic bond forms between two elements, one with a small number of electrons in the valence shell (metal) and one with an almost full outer shell (non-metals). Na and Cl form an ionic bond; with Na giving up an electron from its valance shell and donating it to the Cl atom to complete its valence shell. The non-metal atoms attract valence electrons from metal atoms to become negatively charged, e.g. chlorine accepts an electron to become Cl -. The metal atom releases electrons to become positively charged, e.g. sodium loses an electron to become Na+. These charged atoms (ions) are then held together by attraction of opposite charges creating an ionic bond between them. Ionic bonds are not directional, because the bonding is created by opposite charges on the ions. This means the ions have freedom in the way that they pack. However, ions of opposite sign must surround each other to retain the attraction between ions. Ionic bonds are: Strong and stiff Give a material high strength High elastic modulus High melting point Poor electrical conductivity

-1-

MATS1101 - Materials - Bonding Between Atoms

Ionic bonding is the dominant bonding in some ceramics such as magnesia (MgO), alumina (Al2O3) and cement. Covalent Bonding In covalent bonding electrons are shared between neighbouring atoms. E.g. in a diamond, each carbon atom shares its valence electrons with four other carbon atoms. A 3D structure is formed with covalent bonding with the bonds being very directional because the electrons are shared between atoms. Directionality of bonds dictates the atom packing. In a diamond, the carbon atoms build up into a 3D array with the bonds pointing towards the corners of tetrahedron. Covalent bonds are: Very stiff Give very high elastic modulus High (inherent) strength High melting point Low electrical conductivity Covalent bonding is the dominant bonding found in silicate ceramics and glasses. They also occur in the backbone of polymer chains and in the cross-links in thermosetting polymers. Metallic Bonding In metallic bonding the electrons are surrendered to common pool and become shared by all the atoms in the solid metal. When the atoms give up their electrons, they become positively charged and these ions are then held firmly in place by the attractive forces between the ions and the negatively charged electron cloud. Bonds are not directional, so atoms tend to pack to give simple, dense structures. The metallic bond is: Strong and stiff Giving high elastic modulus High strength Good electrical conductivity (because electrons have easy movement) Metallic lustre Good ductility Secondary Bonds Secondary bonds are created from weak attraction present due to asymmetrical charge distribution in adjacent atoms. Van der Waals Bonding This is a weak attraction caused by a temporary dipole between unchanged atoms. Although overall there is no charge on an atom, as the electrons whiz around in their orbit the atom is instantaneously negative on the side where the electron is situated. The instantaneous dipole created by one atom creates a dipole on a nearby atom and an attractive force is created. This is the bonding present in N 2 molecules. Hydrogen Bonding This a weak bonding due to a dipole produced by hydrogen when it is covalently bonded to other atoms. The hydrogen becomes slightly more positive than the atom it is bonded to, and therefore a dipole is produced. An example of hydrogen bonding is ice, where H2O molecules has slight dipole because the oxygen atom is slightly negative compared to the hydrogen. The hydrogen bond is the strongest type of secondary bond.

-2-

MATS1101 - Materials - Bonding Between Atoms

Bond Energy Atoms experience repulsive forces as well as attractive forces, because as the atoms get closer together, the attraction only continues until the charge distribution starts to overlap. The equilibrium spacing of atoms occurs where the bond energy is a minimum, i.e. where the net curve of attraction and repulsion has the minimum value. The force between atoms, F, is given by:

dU dr
where U = bond energy and r = atomic separation Note that the force is zero at the equilibrium spacing; if atoms are pulled apart a small distance there is a resisting force. Bond stiffness, S, is given by:

S=

dF d 2U = dr dr 2

When the stretching is small, S is constant and therefore the bond is linear-elastic. This means that as the interatomic distances increases, the force required increases linearly. Bond force determines the: Elastic modulus (or Youngs modulus) of a material (how stiff a material is) Melting or softening temperature (stronger bonds will enable a material to withstand higher temperature before bonds break)

-3-

MATS1101 - Materials - Packing of Atoms in Solids

Atom Packing For most engineering materials, atoms pack in regularly repeating 3D arrays. These 3D arrays create crystalline grains which are assembled together to make up the solid. Atom->3D Arrays->Grains->Solid Different materials have different types of packing, depending on the: Type of bonding The sizes of atoms involved and The different elements present. Metal Crystal Structures Bonds are non-directional in metals, atoms pack together efficiently into close packed or nearly close packed structures. Body Centred Cubic (BCC) Each atom is surrounded by 8 other atoms, each of which is at the corner of a cube. The number of nearest neighbours, which is termed the co-ordination number, is 8. E.g. iron.

Face Centred Cubic (FCC) There is one atom at each corner of the cube and an atom at the centre of each face of the cube. The co-ordination number is 12. E.g. aluminium, copper and nickel.

Close Packed Hexagonal (CPH) Atoms are arranged at corners of a hexagonal prism with an atom at the centre of the top and bottom face and three atoms within the prism. The co-ordination number is 12. E.g. titanium, zinc and magnesium.

Ceramic Crystal Structures There are more than one species of atom in ceramics, e.g. aluminium and oxygen in alumina, which makes most ceramic crystal structures quite complicated.

-4-

MATS1101 - Materials - Packing of Atoms in Solids

-5-

MATS1101 - Materials - Packing of Atoms in Solids

Polymer Structure Polymers consist of huge molecules, forming spaghetti-like chains, held together by covalent bonding. The chains are generally arranged randomly, i.e. the chains are not arranged in a regular repeating pattern. This means that polymers are non-crystalline. The term amorphous is used to describe non-crystalline materials.

Glass Structure Glasses are also amorphous since they also have no regular arrangement of the atoms.

Density Density of materials depends on the size, mass and efficiency of packing of the atoms. Metals are made up of heavy atoms that are densely packed; therefore they have a high density. Polymers and ceramics are made up of light atoms that are generally less closely packed; therefore they tend to have lower density.

-6-

MATS1101 - Materials - Youngs Modulus

The Elastic Moduli The elastic modulus measures the resistance of the material to elastic or springy deformation. Low modulus materials are floppy they suffer large deflections if they are loaded. Low modulus materials are desirable for springs, cushions or vaulting poles. For most engineering applications, a high modulus material is desired to withstand high loads with minimum deflection. Hookes Law is a description of the observation from experiments that when strain is small, the strain is proportional to the stress. From this law, the elastic modulus is defined as the ratio of stress, , to strain, , at small strains, i.e. = E where E = Youngs Modulus. This law holds for both tension and compression loading. In the same way, shear strain, , is related to shear stress, , by shear modulus, G, i.e. = G Pressure causes a shrinkage of volume, and so the negative of dilatation is proportional to the pressure. Bulk modulus is therefore defined as: p = -K where K = bulk modulus, p = pressure and = dilatation Poissons ratio, u, is the ratio of lateral shrinkage strain to longitudinal tensile strain, i.e. lateral strain = - tensile strain E, G, K and are the commonly used elastic constants. Physical Basis of Elastic Behaviour Bonding between Atoms Atoms in crystals are held together by bonds that behave like little springs. Atoms experience both attractive and repulsive forces. Atoms displaced from the equilibrium position, r0, create restoring forces, pulling the atom back to the equilibrium position. This is the origin of elasticity. Stiffness, S, of a bond is given by: S =

dF dr dF dr r =r0

When stretching is small, S is constant at S0, where S 0 =

This means that the bond behaves in a linear elastic way and for small displacements: F = S0 (r-r0) Consider unit area of material containing N bonds, stretched as shown:

The total force exerted across the unit area is the stress, . N is the number of bonds per unit area. Since the atom spacing = r 0, the average area per atom is r02. Therefore:

N=

1 r0
2

-7-

MATS1101 - Materials - Youngs Modulus

Stress = force / unit area = (no. of bonds / unit area) x stiffness x r

= NS0 (r r0 ) = S 0
As strain E =

(r r0 ) r0
2

r r0 S0 = , Youngs Modulus E = r0 r0

S0 can be calculated from theoretical bond energies, allowing calculations of E.

These calculations of E agree well with the moduli of metals and ceramics. However polymers generally have a lower modulus. This is because for polymers, the elastic behaviour is dominated by the secondary Van der Waal's bonds, between the spaghetti-like polymer chains rather than the covalent bonds in the backbone, and polymers therefore have low stiffness. In some polymers, such as polyethylene, the secondary bonds melt below room temperature and the stiffness (at room temperature) is even lower. The same is true for rubbers. In some polymers, some covalent bonds form between the polymer chains as well as the secondary bonds. Like the secondary bonds, these also contribute to stiffness, so that the stiffness increases as the amount of covalent cross-linking increases.

-8-

MATS1101 - Materials - Yield and Tensile Strength

Yield and Tensile Strength There are many engineering situations where yield strength becomes an important design consideration. Engineers generally aim to design against permanent deformation in structures (yield limited design). All solids have an elastic limit, beyond which the following occurs, depending on the material: 1. Brittle materials undergo fracture, e.g. glass (fractures suddenly) and concrete (cracks gradually). 2. Ductile materials undergo plastic deformation (i.e. change shape permanently), e.g. most metal alloys and elastomers will deform permanently when stretched too far. This is also known as plastic collapse. The strength at the elastic limit is the Yield Strength. Characterisation of Strength The tensile test is used to measure a materials strength. The test piece has a measured gauge length marked off on the reduced section. The test piece is usually circular in crosssection, but rectangular specimens are also used. A force is applied longitudinally to the test piece in the tensile test and measurements are made based on the elongation of the material under the applied force. Parameters are given below: L = Gauge length D = Diameter of the test piece F = Force applied to the test piece A = Cross sectional area of the test piece = D 2 n = Nominal stress = F / A0 (A0 is the original cross sectional area) n = Nominal strain = (L-L0) / L0 = L / L0 (L0 is the original gauge length, L is the distance between gauge marks after the load has been applied. The force is plotted against the extension of the test piece to give a load-extension curve.

As the test progresses, increasing force is applied to the test piece. Initially the test piece exhibits linear elastic behaviour, i.e. the extension increases linearly with the load, until the elastic limit. After this point, the test piece yields and continues to elongate uniformly until the load reaches a maximum point. This load is used to define the Ultimate Tensile Strength (UTS), which is the nominal stress at the maximum load. After the maximum load, the test piece necks (called plastic instability) and there is non-uniform elongation (the piece elongates in a localised region) until the test piece breaks.

-9-

MATS1101 - Materials - Yield and Tensile Strength

The nominal stress and strain are determined from the load-elongation curve and are plotted (n vs. n) to give an engineering stress-strain curve. This curve gives a number of parameters used for material characterisation. This curve is used because the results of the engineering stress-strain curve apply to all sizes and shapes of test pieces.

y = Yield Strength = Stress at the elastic limit (at the onset of plastic flow). It is the stress that divides the elastic and plastic behaviour of the material. To design a part that will not deform plastically in service, we must use a material with a high yield strength or make the component large enough that the force produces a stress below the yield strength. 0.1% = 0.1% Proof Stress = Stress at a permanent strain of 0.1%. Often 0.2% Proof Stress is also used. Proof Stress is used for materials that yield gradually and therefore do not exhibit a distinct elastic limit or yield point. A line is drawn parallel to the elastic region of the stressstrain curve, offset by 0.001 (or 0.002). The proof stress is the stress at which this line intersects the stress-strain curve. TS = Tensile Strength or Ultimate Tensile Strength = Stress at maximum load or onset of necking. E = Young's Modulus or Modulus of Elasticity = slope of the stress-strain curve within the elastic region. The modulus gives an indication of a material's stiffness or resistance to elastic deformation. f = Plastic Strain After Fracture or Tensile Ductility = Permanent elongation resulting from the tensile test. The broken pieces are put together and final gauge length l f is measured. Strain f = (lf-l0)/l0, often called percent elongation (if multiplied by 100). The ductility of the material is important in design, where ductility is needed so that the part deforms before fracturing, and in manufacturing, where ductility is required to form complicated shapes without breaking the material. The energy expended in deforming a material per unit volume is given by the area under the stress-strain curve.

- 10 -

MATS1101 - Materials - Yield and Tensile Strength

Linear Elasticity Most solids show linear elastic behaviour at small strains (<0.1%). In this region of the stress-strain curve, the slope of the line is Youngs Modulus, E. The material will behave like a linear spring as an elastic solid and therefore the area shaded is the elastic energy stored that will be returned if we unload the solid.

Non-linear Elasticity Rubbers show elastic behaviour to very large strains (~500%), but the stress-strain curve is not linear. This means that if the rubber is unloaded, all the energy stored is recovered.

Typical Mechanical Behaviour Yield Strength Ceramics have the highest yield strengths, but in a tensile test will usually fail well below the yield strength due to low fracture toughness. For ceramics, the yield strength is obtained using hardness measurements. Pure metals have low yield strengths, and a high ductility. This enables them to be formed into shapes. Yield strength is increased by alloying and special processing, but alloyed metals still have much lower yield strengths than ceramics. Polymers have lower yield strength than metals, but can be strengthened by fibre or particle reinforcement. Fibre or particle reinforced materials are referred to as composites.

- 11 -

MATS1101 - Materials - Dislocations

Dislocations Ideal Strength of Materials The slope of the interatomic force-distance curve at equilibrium spacing is proportional to Youngs Modulus, E. Interatomic forces become negligible for r > 2r0.

The maximum in the force-distance curve occurs at ~1.25r 0, (where F = Fmax). If applied stress is greater than Fmax per bond, bonds between atoms are broken and fracture occurs. Ideal strength, , corresponds to bond rupture at F max. Calculation of Ideal Strength From the force-distance curve, where r = 2r0, = 0.25 and 2, ( is ideal strength).

A better estimate using interatomic potential gives Glasses and some ceramics have a yield strength of For other ceramics and polymers, For metals, Actual vs. Ideal Strength Glasses and some ceramics show strengths close to their ideal strength. Polymers are also close to their ideal strength - although their strength is low, their modulus is also low. Why is y < for metals and some ceramics? Crystals are not perfect and contain defects. One of these defects is a dislocation. In these materials that do not exhibit ideal strengths, plastic (permanent) deformation occurs by shear stress at stresses much less than the ideal strength, via a dislocation mechanism.

Yield occurs by shear stress because tensile and compressive stresses create shear stresses. The shear stress is at a maximum at 45 to the axis of tension or compression.

- 12 -

MATS1101 - Materials - Dislocations

Dislocation Mechanism of Plastic Deformation Shear deformation occurs by creation and motion of dislocations on crystal planes. Dislocations are carriers of deformation as electrons are carriers of charge. They exist in almost all crystalline solids, typically introduced during solidification or deformation. Dislocations are line defects in the crystal structure, creating lattice distortion around the dislocation. Atoms are rearranged as the dislocation moves through the crystal. A dislocation is shown below:

Dislocations move easily (glide) at low stress because: 1. Atom displacements are small 2. Atom displacements are localised near the dislocation line (i.e. only a small fraction of atoms are displaced at any time). Dislocations form by mis-stacking of atoms during crystallisation. Typical engineering alloys contain ~100 000 km of dislocation lines per cm 3. Dislocation glide is easy in metals due to the non-specific nature of metallic bonding.

However it is difficult in ceramics due to the specific nature of covalent or ionic bonding. With covalent bonding the strength and directionality of the bonds prevent dislocations moving. With ionic bonding, movement of the dislocation disrupts the charge balance around surrounding atoms.

- 13 -

MATS1101 - Materials - Strengthening Methods

Strengthening Methods Forces on Dislocation Force, f, resisting motion in slip plane: f = .b = shear stress b = burgers vector Force, T, along dislocation line (line tension): T = Gb2 G = shear modulus

A crystal yields when force on dislocation, b, is greater than the force opposing motion, f. Thus, at yielding, dislocation yield strength, y, is related to lattice resistance to dislocation motion, f, by:

Intrinsic Lattice Resistance to Dislocation Glide The basic component of lattice resistance f is fi, the intrinsic lattice resistance - this is caused by the bonds between atoms that have to be broken and reformed as the dislocation moves. In metals, fi is low due to the non-specific nature of metallic bonding, e.g. pure single crystals. In ceramics, fi is high due to the specific nature of covalent or ionic bonding, e.g. diamond. As metals have a low f i, it is useful to be able to increase f in metals by strengthening. The overall lattice resistance, f, of metal crystals may be increased by: solution strengthening: fss precipitates (obstacles): fo work hardening: fwh because the effect of these adds to fi. The upper limit on strengthening is always less than or equal to the ideal strength. In practice, the ideal strength is rarely approached.

- 14 -

MATS1101 - Materials - Strengthening Methods

Solid Solution Strengthening Impurity atoms dissolve in solid metal to give an alloy. An example is brass, where zinc atoms replace copper atoms randomly, up to approximately 30% zinc. Impurity atoms distort the slip planes, because of atomic size mismatch and therefore dislocation motion is impeded. y is increased by fss/b The degree of solid solution strengthening depends on two factors: 1. The larger the size difference between the two atoms, the greater the strengthening. 2. The amount of impurity atoms added. fss is increased as the impurity concentration increases. Since impurity atom spacing on the slip plane is proportional to C -; y C

Precipitation and Dispersion Strengthening Impurity atoms dissolved at high temperature may precipitate on cooling. This creates fine, closely spaced precipitates that impede dislocation motion. An example is Al-Cu (2000 Aluminium Series) alloys, HSLA steels and precipitation hardened (p-h) stainless steels. Small precipitates or particles or particles obstruct the motion of dislocations, as shown below:

For the critical situation (c), the force bL on the segment of dislocation between two adjacent particles = force 2T due to line tension, i.e.

Obstacles exert resistance on dislocations Maximum effect is found with strong, closely-spaced particles (high T, low L).

- 15 -

MATS1101 - Materials - Strengthening Methods

Work Hardening Metal crystals (grains) have several slip planes. When a stress is applied to a material, dislocations are created in the material. The dislocations moving along intersecting slip planes interact - they obstruct one another and accumulate. The more dislocations there are, the more likely they are to interfere with each other and the stronger the material becomes. They are only removed by annealing (heating).

The stress-strain curve rises after the yield point and the increase can be described by: y m where m is the strain hardening index The lattice resistance increases by f wh (usually substantial). Examples of work hardening alloys are Al-Mn (3000 Aluminium alloys), brass, bronze and stainless steel.

Strengthening contributions are additive. Other strengthening mechanisms are also possible, for example, transformation strengthening in hardened steels and order strengthening in superalloys. Strong materials either have high f i, e.g. diamond, or they use other mechanisms, e.g. high-strength alloys. y is yield strength of isolated grains in shear, but we need the yield strength of polycrystalline materials in tension. Grains have different orientations within a material. Yielding occurs progressively from grain to grain and is constrained by adjacent grains. Slip begins in grains where there are slip planes nearly parallel to , e.g. grain 1. Slip will then spread to grains like grain 2 and lastly to grains like grain 3. The yielding takes place progressively in polycrystalline materials, which is why they don't have a sharp yield point on the stress-strain curve. The yield strength is therefore higher in polycrystalline materials than if the material was just a single grain. For polycrystalline materials y increases by Taylor factor of 1.5. The stress conversion factor is = 2. Therefore polycrystalline materials yield when y = 2 1.5 y. But grain size also affects y.

- 16 -

MATS1101 - Materials - Strengthening Methods

Grain Size Strengthening Grain boundaries (GBs) are surface defects in the crystal structure.

Grain boundaries impede slip dislocations and dislocations pile up at grain boundaries, creating a back-stress on dislocations. Therefore the dislocation yield stress, y, increases. Grain boundary spacing in slip planes decreases as the grain size decreases, therefore as grain size decreases, y increases. In low-carbon steels, the Hall-Petch equation holds: y = o + const x d- 0 is related to y d = average grain diameter

- 17 -

MATS1101 - Materials - Fast Fracture

Fast Fracture Role of Cracks Catastrophic failure by fast fracture may occur in structures even though they are properly designed to avoid yielding and excessive elastic deformation, e.g. welded ships, pressure vessels, welded gas pipelines, welded bridges. In situations where unexpected failure occurs, the common feature present is usually unstable cracks. Energy for Crack Growth When cracks are present, there is a critical stress that must be exceeded for the crack to grow. To calculate this critical stress, we examine the work required for crack growth. The criterion for crack advancement is that the work done by the load must be greater than the difference between the change of elastic energy and the energy absorbed at the crack tip. For a crack of length a, which advances a distance a, in a plate of thickness t, under a load F:

Work done by load = W Change of elastic energy = Uel Energy absorbed at crack tip = Gc.t.a W Uel + Gc.t.a where Gc is energy absorbed per unit area of crack (not only surface energy) and t. a is new crack area Toughness Gc is a material property that indicates the strain energy release rate, or "toughness" of a material. It is measured in energy per unit area, with units of J.m -2. A high Gc value indicates that crack propagation is difficult in that material, e.g. copper is tough and has a Gc of approximately 106 J.m-2. A low Gc value indicates that crack propagation is easy in that material, e.g. glass is not tough and has a Gc of approximately 10 J.m-2.

- 18 -

MATS1101 - Materials - Fast Fracture

Stress for Crack Growth Consider a plate with fixed ends, under tension, e.g. welds in steel structures.

The ends cannot move and therefore the forces acting do no work i.e. W = 0 and therefore: -Uel = Gcta As the crack advances, the material relaxes - elastic strain energy is lost, i.e. U el is negative, and Gc is positive. Inside the plate, for a unit cube element under stress, , and strain, , strain energy is:

Strain energy released per unit volume in the shaded region due to the crack is:

Overall energy change in the shaded region due to crack is:

If crack grows by a:

The critical condition for fixed displacement: -Uel = Gcta then gives:

However, the assumption about relaxation under-estimates Uel by about 2, so

Rearranging gives: at onset of fast fracture

- 19 -

MATS1101 - Materials - Fast Fracture

Consider a plate with a fixed load. As the crack grows, the plate relaxes and becomes less stiff. Applied forces move and do work, i.e. W is positive. However U el is also positive because some of W is used to increase Uel and the final result is the same.

The equation derived for both these situations is: at onset of fast fracture The right hand side is constant for a given material because these are material properties only. The left hand side gives either: the critical value of at which the crack of length a will propagate or the critical crack length a, which will become unstable under a given stress, , and propagate with fast fracture. Fracture Toughness Defining a stress intensity factor, K (MN.m-3/2), where

Fast fracture occurs when K = Kc, where Kc is critical stress intensity factor, or fracture toughness, given by:

Combining this with the definition of stress intensity factor, the fast fracture condition becomes:

It can be seen that: Ceramics have the lowest Gc and the lowest Kc. Metals have the highest Gc and the highest Kc. Polymers have an intermediate Gc and a low Kc (because of a low E). Composites have a higher Gc than polymers and an intermediate Kc, also greater than polymers. Note: some materials have a transition in behaviour with temperature, eg. bcc metals like steel becomes quite brittle if cooled sufficiently.

- 20 -

MATS1101 - Materials - Micromechanisms of Fracture

Micromechanisms of Fracture Crack Propagation Resistance The toughness of materials varies: Metals: high Gc and Kc, which means that cracks propagate slowly Ceramics: low Gc and Kc, which means that cracks propagate fast Polymers: intermediate Gc and low Kc Composites: Gc and Kc greater than for polymers These materials are different because they have different stress concentration effects at the crack tip and different micromechanisms of fracture. Stress Concentration In a material near the crack, local > , where is the average stress. As distance from the crack tip, r, decreases, local increases, because:

Approaching the crack tip, a distance, r y, is reached where local approaches y. Plastic flow then occurs, forming a plastic zone at the crack tip. To find ry, we set local = y and assume ry is much smaller than the crack length a:

since: The crack will grow when K = Kc and the size of the plastic zone is given by the definition of ry above. From the above equation, as y increases, the plastic zone size, ry, decreases. Therefore strong materials will have a small plastic zone, or none, e.g. ultra-high strength alloys or ceramics. Low strength materials will have large plastic zones, e.g. pure metals and ductile alloys. Ductile Rupture Ductile metals deform readily by plastic flow at low stresses, which means they require large strains before fracture will occur, e.g. pure copper and mild steel (at room temperature). If a crack is present, it will grow when the stress is sufficiently high, leading to fracture, but with the absorption of a large amount of energy, by ductile rupture, which is different to fast fracture. Alloys contain fine inclusions or impurities. Cracks and impurities both concentrate the stress and produce triaxial tension.

- 21 -

MATS1101 - Materials - Micromechanisms of Fracture

Plastic flow occurs, producing cavities around the inclusions. These cavities link and the crack advances by ductile tearing, which absorbs much energy per unit volume. The energy absorbed also increases as the plastic zone size, r y, increases. Therefore materials with low y have a large plastic zone size and absorb much energy before fracture, i.e. ductile materials have a high fracture toughness, Kc.

Fast Fracture - Cleavage Brittle materials have relatively flat featureless surfaces, e.g. ceramics. Cracks grow rapidly, with little or no plastic flow. This is because plastic flow is much more difficult than in metals. Also, as y is high, ry is very small, and little or no crack tip blunting occurs due to the plastic zone, i.e. the cracks remain very sharp.

Sharp cracks cause very high local:

Even with a small degree of blunting, the ideal strength of the material can be exceeded at the crack tip, ie. atomic bonds break to create an atomically flat surface by cleavage. Ductile to Brittle Transition At low temperatures, BCC and HCP metals are brittle (undergo fracture by cleavage), yet they are tough above room temperature. Dislocation motion is assisted by thermal agitation. At low temperatures, dislocation motion is more difficult, so as temperature decreases, y increases and the plastic zone size decreases. Eventually, the mechanism changes from ductile tearing to cleavage.

In steels, the ductile-brittle transition temperature is between -70C and -40C. Polymers also have a ductile-brittle transition temperature at the glass-rubber transition, TG.

- 22 -

MATS1101 - Materials - Micromechanisms of Fracture

Toughening in Composites The Gc of composites is greater than the Gc of polymers. In composites the Gc is increased by reinforcement - the fibres act as crack stoppers. The mechanism is that before the crack reaches the fibre, the local results in debonding (where fibres lose bonding with the matrix). This causes crack blunting and arrests the motion of the crack. (However this mechanism will only be successful if the crack is perpendicular to the fibres). Thus fibres in composites create high toughness and high stiffness.

Rubber-Toughened Polymers The Gc of polymers may be increased by 'filler' particles. Rubber-toughened polymers use rubber particles, e.g. ABS. The mechanism here is that the crack intersects and stretches the rubber particles, which clamp the crack shut. The load required for crack motion is therefore increased, which increases Gc.

Avoiding Fast Fracture Metals are the most important materials for highly stressed applications because of their fracture toughness. Pure metals have a high G c above room temperature, but alloying usually decreases Gc, since alloying increases the lattice resistance to dislocations, i.e. y increases and therefore ry decreases. Practical considerations: the extent of strengthening must be balanced with the loss of toughness avoid brittle phases in microstructures, because they create easy crack paths. Common examples of where easy crack paths are produced are: formation of brittle plate-like phases such as sigma phase in stainless steels or graphite in cast irons formation of brittle continuous grain boundary films, such as Fe 3C in annealed high carbon steels formation of brittle matrix structures, such as martensite in as-quenched steels, before tempering

- 23 -

MATS1101 - Materials - Fatigue Failure

Fatigue Failure Fast Fracture and Fatigue As we have seen, a crack becomes unstable if K=K c, and this leads to fast fracture. Conversely, if K is less than Kc, the crack is stable and therefore, if the maximum crack size is known, a safe load can be calculated and if the maximum load is known, the maximum crack size allowable can be calculated. However, cracks may form and grow at lower loads if the stress is cyclic. Eventually the crack reaches the critical size for that stress and catastrophic fast fracture will occur.

Fatigue of Uncracked Components Fatigue tests are used to determine the likelihood of fatigue failure. Tests are carried out using cyclic stresses, either in tension-compression or in rotating bending.

Three stress parameters are defined for the fatigue test:

Stress range Mean stress m Stress amplitude a

- 24 -

MATS1101 - Materials - Fatigue Failure

High Cycle Fatigue For high cycle fatigue, neither max nor |min| are above the yield stress and m = 0. Under these conditions, Basquin's law holds: where Nf = no. of cycles to failure a = constant (0.07<a<0.15) C1 = constant Basquin's law only holds where Nf > 104

Low Cycle Fatigue For low cycle fatigue, max or |min| are above the yield stress and m = 0. Under these conditions, Basquin's law no longer holds, but the Coffin-Manson law holds: where pl = plastic strain range b = constant (0.5<a<0.6) C2 = constant Coffin-Manson law only holds where Nf < 104

Goodman's Rule Where m is not equal to zero, the stress range must be decreased as the mean stress increases, according to Goodman's Rule:

where m = at mean stress m 0 = at mean stress 0 TS = Ultimate Tensile Strength

- 25 -

MATS1101 - Materials - Fatigue Failure

Miner's Rule When varies through the lifetime of the component, the cumulative damage is summed, according to Miner's Rule. Regions, i, are defined for each different Ds, each region uses a fraction of the life, Ni/Nfi. Failure occurs when:

where Ni = no. of cycles in region i Nfi = no. of cycles to failure in region i

Fatigue of Cracked Components Large structures always contain cracks. To find the safe life of the structure, we must know how many cycles there will be before one of these cracks grows to the critical size for fast crack propagation. This is done by fatigue testing of pre-cracked test pieces. A cyclic load is applied to a pre-cracked test piece. The rate of crack growth per cycle is measured. We define a number of parameters: Stress intensity range, K Mean stress intensity, Km Stress intensity amplitude, Ka

Stress intensity range is defined as: K increases with time at constant load, because the crack grows. Above a threshold K, the crack growth per cycle is:

where A and m are material constants

- 26 -

MATS1101 - Materials - Fatigue Failure

If the initial crack length, a0, is known and the final crack length, af, at which the crack becomes unstable is calculated using the fracture toughness K C and the maximum stress max, then the safe number of cycles, Nf, can be estimated from:

where

Fatigue Mechanisms Pre-Cracked Structures For pure metals and polymers, the tensile cycle stretches the crack tip to . The plastic zone creates a new surface. The compressive cycle then folds the new surface and the crack tip grows by ~ , i.e. da/dN ~

For alloys, which contain inclusions, the tensile cycle stretches the crack tip as in pure metals, but the inclusions form holes in the plastic zone, increasing its size. This means that increases and the crack growth rate, da/dN, is greater than that of pure metals.

Uncracked Components Under low-cycle fatigue, the plastic flow roughens the surface quickly and the crack nucleates at the rough surface. Initially the crack grows on the inclined slip plane. Later, the crack grows normal to the tensile axis.

Under high cycle fatigue, max is less than y, so no general plastic flow occurs. Most of the fatigue life is spent initiating the crack. A stress concentration (such as at a notch or change of section) leads to local plastic flow (because of the localised increased stress), which then leads to crack initiation.

- 27 -

MATS1101 - Materials - Creep

Creep For most metals and ceramics at room temperature, strain, which is a function of stress, i.e. = f(), is independent of time. These are known as elastic/plastic solids. As the temperature increases, they may creep at a stress which causes no permanent deformation at room temperature. Creep is slow, continuous deformation with time. Strain, , now depends on temperature, T, and time, t, as well as stress, , i.e. = f(,t,T). This is a creeping solid.

The temperature at which creep occurs is dependent on the melting temperature, T m (K). Creep starts in metals when the temperature is greater than 0.3-0.4 T m. Creep starts in ceramics when the temperature is greater than 0.4-0.5 T m. Creep is rapid at temperatures above 0.8 Tm (K). Creep temperature can be raised by alloying, for example, superalloys operate at 0.6-0.8 Tm. Most polymers are not crystalline, and therefore have no melting temperature. In these materials, creep is related to the glass transition temperature, T G (temperature at which Van der Waals bonds melt). At temperatures higher than T G, the polymer will creep, at temperatures less than TG, the polymer is hard and shows little creep. For many polymers, T G ~ room temperature, and significant creep occurs at room temperature. Creep Testing A sample is tested for creep under uniaxial tension at constant load and temperature. The strain is measured against time, t. Metals, polymers and ceramics all show similar creep curves.

There are three stages in the creep behaviour of materials: 1. Primary Creep. This occurs quickly, and is usually allowed for in design, as is elastic deformation. 2. Secondary Creep. This occurs at a steady state - the strain rate, (= ss) is constant. 3. This is usually of greatest concern in design. Tertiary Creep. Creep rate accelerates until fracture.

- 28 -

MATS1101 - Materials - Creep

Secondary Creep A). Dependence on Stress. Strain rate, , increases as n increases as given by: where n = constant called the creep exponent (3<n<8) B = material constant This is known as "power law creep". At low stresses, a different creep behaviour occurs, and n=1.

B). Dependence on Temperature. Strain rate, , increases exponentially as temperature increases, as given by:

where R = gas constant Q = activation energy

Combining both stress and temperature effects gives: where "creep constants": A, n & Q vary with the material. Consequences of Power Law Creep A). At constant load (stress), strain accumulates with time. This causes glaciers to flow and polymers to distort at temperatures around room temperature. Metals and ceramics in high temperature equipment undergo slow, but increasing deformation with time. B). At constant displacement (strain), stress relaxes with time. This is termed "stress relaxation". Bolts in high temperature equipment must be re-tightened periodically. Note: If the initial torque is increased, strain rate also increases and therefore stress relaxation can't be remedied by over-tightening.

- 29 -

MATS1101 - Materials - Creep

Tertiary Creep Internal cavities appear during the tertiary creep stage. As the holes grow, the effective cross-sectional area decreases, therefore the stress increases even when the load is kept constant. As a result, the creep rate increases. Since strain rate is proportional to n, creep rate increases faster than the stress increases.

Eventually failure occurs. The time to failure, t f, depends on both stress and temperature. Data are normally given as creep-rupture diagrams:

Creep Resistant Materials For a material to be resistant to creep: It should have a high melting temperature It should be used below 0.3Tm Alloying can increase creep temperature to substantially above 0.3T m E.g. Ni-base superalloys in aircraft gas turbines operate continuously at ~0.6 T m or intermittently at ~0.8Tm.

- 30 -

MATS1101 - Materials - Creep Mechanisms

Creep Mechanisms Dislocation Creep (Power Law Creep)

Below the elastic limit, dislocation motion is resisted by: a) intrinsic lattice resistance and b) obstacles, such as:

If dislocations can be made to climb over these obstacles, plastic flow can occur. Most pinned dislocations have a climb force (a component of reaction force).

Dislocations only move easily by glide:

For climb to occur, the dislocation must move out of its glide plane. Atoms must be removed from the edge of the half plane. This occurs by diffusion: if it occurs below ~0.5T m it occurs by core diffusion (atoms diffuse along the dislocation), if it occurs above ~0.5T m it occurs by bulk diffusion. This enables the dislocation to climb until it is free of the obstacle.

- 31 -

MATS1101 - Materials - Creep Mechanisms

Once the dislocation is "unlocked" by dislocation climb, it can glide, producing slip, until it meets the next obstacle, which pins it. The climb process is then repeated, and therefore dislocation creep is progressive and continuous. Effect of Temperature Climb occurs by diffusion, and diffusion rate increases as temperature increases. Therefore the rate at which dislocations become unlocked increases with temperature and creep rate increases with increasing temperature. Effect of Stress Increased stress results in increased driving force for dislocation climb, therefore more dislocations climb per second and more dislocations glide per second. Therefore the creep rate increases with increasing stress. Resisting Power Law Creep at High Stresses 1. Choose a material with a high melting point (these have a low diffusion rate) 2. Maximise obstruction to dislocations (with solute atoms, precipitates) 3. Choose, when possible, solids with high inherent lattice resistance to dislocation motion, e.g. ceramics and ordered alloys such as superalloys Diffusional Creep (Linear-Viscous Creep) At low stress, in the secondary creep domain, an alternative mechanism takes place. The observed creep exponent, n~ 1.

Creep occurs by atoms diffusing from the sides of grains to the ends to relieve the applied stress, i.e. bulk diffusion. Grain boundary sliding is also necessary for this creep to occur. Creep rate increases as the stress increases, as the temperature increases and as the grain size decreases. Resisting Diffusional Creep at Low Stresses 1. Choose a material with a high melting point (these have a low diffusion rate) 2. Choose materials with a coarse grain size (these have less grain boundary sliding) 3. Choose materials with precipitates along the grain boundaries (these impede grain boundary sliding).

- 32 -

MATS1101 - Materials - Creep Mechanisms

Creep Mechanisms for Polymers For temperatures greater than the glass transition temperature, T G, polymer chain sliding occurs, giving Newtonian viscous flow, which is a type of creep. The creep rate increases linearly with stress and increases exponentially with temperature: where C, Q & R are constants Q is the activation energy for viscous flow (the energy required to push a lump in one molecule past that in another) For temperatures much less than TG, polymers are elastic solids and show no creep. Many polymers are used at temperatures around the glass transition temperature, where most show a mixed visco-elastic behaviour. This can be modelled by coupled spring and dashpot elements (rheology). Applying a load will lead to creep, but at an ever-decreasing rate because the spring takes up the tension. Releasing the load will lead to slow reverse creep due to the extended spring.

Resisting Polymer Creep 1. Choose polymers with a high TG. TG increases with increased molecular weight, increased cross-linking (e.g. epoxies vs. polyethylene) and increased crystallinity (high density polyethylene vs. low density polyethylene) 2. Choose polymers with fillers like silica or glass, because the creep rate decreases as the volume fraction of filler increases (e.g. PTFE (Teflon) and polypropylene in autos) 3. Choose materials that are a polymer matrix reinforced with strong continuous fibres because most of the load is carried by the fibres and hence the creep resistance increases greatly e.g. GFRP (glass fibre reinforced plastic) and CFRP (carbon fibre reinforced plastic) composites.

- 33 -

You might also like