nus me2151-chp6

22
6-1 CHAPTER 6 DISLOCATIONS, DEFORMATION AND STRENGTHENING IN METALS 6.1 PLASTIC DEFORMATION BY SLIP 6.2 DISLOCATIONS AND SLIP 6.3 STRENGTHENING MECHANISMS 6.3.1 Dislocation Stress Fields and Strain Energies 6.3.2 Strain Hardening 6.3.3 Grain Size Strengthening 6.3.4 Solid Solution Strengthening 6.3.5 Dispersion Hardening 6.3.6 Combined Strengthening Mechanisms 6.4 ANNEALING 6.4.1 Recovery 6.4.2 Recrystallization 6.4.3 Grain Growth

Upload: ahmad-azhar

Post on 09-Nov-2015

38 views

Category:

Documents


2 download

DESCRIPTION

Lecture notes

TRANSCRIPT

  • 6-1

    C H A P T E R 6

    D I S L O C A T I O N S , D E F O R M A T I O N A N D S T R E N G T H E N I N G I N M E T A L S

    6.1 PLASTIC DEFORMATION BY SLIP

    6.2 DISLOCATIONS AND SLIP

    6.3 STRENGTHENING MECHANISMS 6.3.1 Dis locat ion Stress F ie lds and

    Stra in Energies 6.3.2 Stra in Hardening 6.3.3 Gra in S ize Strengthening 6.3.4 Sol id Solut ion Strengthening 6.3.5 Dispers ion Hardening 6.3.6 Combined Strengthening

    Mechanisms

    6.4 ANNEALING 6.4.1 Recovery 6.4.2 Recrysta l l i zat ion 6.4.3 Gra in Growth

  • 6-2

    6.1 PLASTIC DEFORMATION BY SLIP

    Plastic deformation in a crystal mostly involves the sliding

    of one plane of atoms over another under the action of a

    shear stress (Fig. 6.1-1); this process is known as slip.

    Fig. 6.1-1 Plastic deformation by slip in an ideal crystal occurs when one plane of atoms slides over another, producing a step of one atomic spacing.

    The plane and direction in which slip occurs are the slip

    plane and slip direction. The slip direction always lies

    within the slip plane. The combination of a slip plane and a

    slip direction forms a slip system.

    Slip does not take place in any arbitrary plane or direction.

    The preferred slip planes and directions are those in which

    the atoms are most densely packed. This is because slip

    occurs in steps of one atomic spacing, so moving atoms

    from one stable site to the next would involve the least

    energy when the atoms are closest together (Figs. 6.1-2 & 3).

    6-3

    !

    "max #ba

    Fig. 6.1-2 The maximum shear stress, !max, required to move one plane of atoms

    over another by one atomic spacing is a function of the interatomic distances, such that smaller stresses are necessary for closely-spaced atoms.

    Fig. 6.1-3 Slip requires less energy on (a) a close-packed plane in a close-

    packed direction than on (b) a less closely-packed plane and direction.

    Since most engineering alloys are polycrystalline, the

    change in orientation from grain to grain means that each

    grain is strained differently by an applied stress (Fig. 6.1-4).

  • 6-4

    Fig. 6.1-4 Resolving a uniaxial tensile stress " into shear stress

    !

    " = F sin#A/cos# = $ sin# cos#

    Slip will begin on a slip system when the resolved shear

    stress acting on the slip plane in the slip direction reaches a

    critical value. If two or more slip systems have the required

    shear stress acting on them, they all slip together (Fig. 6.1-5).

    Fig. 6.1-5 Slip lines on the surface of polycrystalline copper that has been deformed. Slip lines are actually a series of fine steps on the surface.

    Note that the slip lines change direction at grain boundaries. Note also intersecting sets of lines within the same grain, indicating the operation of more than one slip system.

    6-5

    In a polycrystalline solid, the deformation in each grain

    must be compatible with its neighbours to maintain

    mechanical integrity and coherency along the grain

    boundaries. This requires the grains of various orientations

    to slip on 5 independent systems simultaneously.

    Metals with FCC and BCC structures are ductile because

    they possess a relatively large number of slip systems (12 in

    FCC; up to 48 in BCC) (Table 6.1-1).

    The slip systems in FCC and BCC are also well-distributed

    in space, such that at least one slip system would be

    favourably oriented for slip at low applied stresses.

    Table 6.1-1 Primary slip systems in the common metal structures. BCC and HCP contain secondary slip systems, which are not shown.

  • 6-6

    Furthermore, slip systems in FCC and BCC intersect, so if

    one system be constrained, slip can continue on a different

    intersecting slip system; this is known as cross-slip.

    However, unlike the FCC structure, the BCC structure does

    not contain close-packed planes, so slipping atoms must

    move greater distances from one equilibrium lattice

    position to another. Higher shear stresses are thus

    necessary for slip in BCC than in FCC metals (Fig. 6.1-2 & Table

    6.1-2), which translates into higher strengths for BCC metals.

    [Note: critical shear stress is the shear stress required to move a dislocation in its slip system.

    Although the HCP structure contains both close-packed

    planes and directions, its geometry gives rise to fewer slip

    systems. Furthermore, the slip systems, being parallel, do

    not intersect, so cross-slip is not possible (Fig. 6.1-6). Most

    polycrystalline HCP metals are relatively brittle. 6-7

    Fig. 6.1-6 Comparison of the slip systems in

    (a) an FCC structure, and (b) an HCP structure.

  • 6-8

    6-9

    6.2 DISLOCATIONS AND SLIP

    In a perfect crystal, slip would involve a whole plane of

    atoms sliding over another in a single movement, which

    would require the simultaneous stretching, breaking, and

    remaking of all atomic bonds in the slip plane. The

    theoretical shear strengths of metals have been roughly

    estimated to be in the order of 1010 N/m2 (10 GPa).

    However, the actual measured yield strengths of bulk

    metals are at least 1,00010,000 times lower than this

    value (Table 6.2-1). This is because slip in real metal crystals

    occurs via the movement of dislocations, during which only

    a small fraction of atomic bonds are broken at any one

    time, with minimal disruption to the crystal lattice.

    Table 6.2-1 Comparison of theoretical and experimental yield strengths of some metals.

    Dislocations are linear or one-dimensional crystal defects

    where local faults in the atomic arrangement lie along a

    straight line, curve or loop through the crystal.

  • 6-10

    Dislocations can be introduced into a crystal in a number

    of ways: during solidification, during plastic deformation,

    or as a result of thermal stresses arising from rapid cooling.

    All bulk crystalline materials (metals and ceramics) contain

    dislocations.

    There are two fundamental types of dislocations: edge and

    screw. An edge dislocation may be thought of as an extra

    half-plane of atoms inserted into the crystal (Fig. 6.2-1). The

    bottom edge of half-plane that ends within crystal is the

    edge dislocation line. [Note: the extra half-plane of atoms itself is not the dislocation.]

    Fig. 6.2-1 An edge dislocation, showing the extra half planes of atoms.

    Note the regions of compression and tension around the dislocation line.

    6-11

    A screw dislocation may be thought of as making a cut

    half-way through the crystal, and then skewing the two

    halves by one atomic spacing (Fig.6.2-2).

    Fig. 6.2-2 A screw dislocation, showing the spiral screw-like arrangement of atoms above and below the plane of the cut.

    Most dislocations, however, are mixed dislocations, which

    contain both edge and screw dislocation components with

    a transition region in between (Fig. 6.2-3).

    Fig. 6.2-3 A mixed dislocation. Fig. 6.2-4 Transmission electron micrograph of a titanium alloy in which dark lines are dislocations.

  • 6-12

    When a shear stress is applied to the edge dislocation

    shown in Fig. 6.2-5, the extra half-plane of atoms, plane A,

    will be forced to the right; this in turn pushes the top

    halves of planes B, C, D, etc., to the right.

    If the shear stress is high enough, plane A eventually

    becomes closer to the bottom half of plane B than the top

    half of plane B itself. It is then more favourable

    energetically for the atomic bonds across the two halves of

    plane B to be severed and for plane A to bond with the

    bottom half of plane B.

    The extra half-plane moves by discrete steps through the

    crystal and ultimately emerges from the surface, forming a

    slip step that is one atomic distance wide (~10-10m).

    Macroscopic plastic deformation is the cumulative effect of

    the motion of large numbers of dislocations.

    Fig. 6.2-5 The step-by-step movement of an edge dislocation

    under a low shear stress produces a unit step of slip.

    6-13

    Before and after the movement of a dislocation through a region of the crystal, the atomic arrangement is perfect and ordered; it is only during the passage of the dislocation that the lattice structure is disrupted. Only a relatively small shear stress is required to operate in the immediate vicinity of the dislocation in order to produce a step-by-step shear.

    Fig. 6.2-6 Heimlich

    the caterpillar illustrating (a) the difficulty of

    moving without (b) a dislocation mechanism.

    Although the edge, screw and mixed dislocation move in different directions, the result is the same shear (Fig. 6.2-7).

    Partially sheared Totally sheared

    Fig. 6.2-7 Shear produced by motion of (a) edge, (b) screw and (c) mixed dislocations. The dark arrows indicate the direction in which the dislocations move.

  • 6-14

    While bulk ceramics and other crystalline compounds contain dislocations, the shear stress required for dislocation motion is at least 2-4 times that in metals. Not only are covalent and ionic bonds stronger, but ions in the more complex ceramic structures must also move greater distances between equilibrium lattice positions.

    In addition, ceramics in which the bonding is predominantly ionic contain very few slip systems. If slip were to occur in some directions, ions of the same charge would be brought close together (see also Sec. 3.8.4), generating strong electrostatic repulsion that would resist slip (Fig. 6.2-8).

    Fig. 6.2-8 (a) Before slip; (b) like charges repel in this slip direction; (c) slip possible.

    For ceramics with highly covalent bonding, the directional nature of the bonds makes the displacement of atoms from their lattice sites extremely difficult.

    The shear stress that must be applied to activate slip in bulk ceramics is higher than that required to cause fracture (Chp. 7). Ceramics are therefore hard and brittle, and do not generally undergo plastic deformation by slip, except at high temperatures (~ 0.5-0.7 TM [Note: TM is the melting temperature]).

    6-15

    6.3 STRENGTHENING MECHANISMS IN METALS

    Because plastic deformation in metals corresponds to the

    movement of large numbers of dislocations, the capacity

    of a metal for plastic deformation depends on the ability of

    dislocations to move.

    Since the hardness and strength of a metallic alloy are

    related to the stress at which plastic deformation can be

    made to occur (and thus, the stress at which dislocations

    are able to move), there are two possible methods of

    hardening or strengthening a metal:

    ! Eliminating all crystal defects, including dislocations

    this has only been achieved in whiskers (very thin

    single crystals only a few m in diameter), in which

    strengths approaching theoretical values are possible.

    ! Creating so many crystal defects that they restrict or

    hinder the passage of dislocations this is the method

    used to strengthen bulk metals, but yield strengths are

    still much lower than theoretical levels.

    In the second approach, strengthening is achieved through

    interactions of the stress fields of moving dislocations with

    those created by other crystal defects.

  • 6-16

    6.3.1 Dislocation Stress Fields and Strain Energies

    Atoms surrounding a dislocation are displaced from their

    equilibrium lattice positions. Such elastic strain produces

    an elastic stress field around the dislocation.

    In an edge dislocation, the presence of the extra half plane

    of atoms above the dislocation line means that atoms in its

    vicinity are squeezed together, resulting in compressive

    stresses. Conversely, the atoms below the dislocation line

    experience tensile stresses due to an increase in

    interatomic separation in this region (Fig. 6.3-1).

    Fig. 6.3-1 (a) Regions of compression and tension located around an edge dislocation. (b) Detailed stress state of an edge dislocation showing

    compressive, tensile and shear stresses.

    In a screw dislocation, the lattice spirals around the centre

    of the dislocation. The stress field is one of pure shear and

    is symmetrical about the dislocation line (Fig. 6.3-2).

    6-17

    Fig. 6.3-2 Shear stress and strain associated with a screw dislocation.

    The distortion of atomic bonds around any dislocation

    increases potential energy because of non-equilibrium

    interatomic separations (see also Sec. 3.7). This energy is known

    as strain energy, since it is associated with the strain or

    distortion of the crystal lattice.

    When a dislocation is in close proximity to another, the

    stress fields surrounding each dislocation will interact. For

    example, if the compressive and tensile stress fields of two

    edge dislocations lie on the same sides of the slip plane (Fig.

    6.3-3a), the overall strain energy will be raised if the two

    fields overlap; this gives rise to mutual repulsion as the

    dislocations approach each other.

    Conversely, if the compressive and tensile stress fields were

    on opposite sides (Fig. 6.3-3b), the dislocations would

    annihilate each other when they meet, with a lowering of

    the overall strain energy; the dislocations would thus be

    attracted to each other.

  • 6-18

    Fig. 6.3-3 (a) The interaction of two edge dislocations of the same sign causes repulsion, (b) while that of different signs causes attraction and annihilation.

    C and T denote compressive and tensile regions, respectively.

    Two dislocations can attract and annihilate each other only

    if they meet exactly on the same slip plane, and the

    components of their stress fields match exactly and are of

    opposite signs (Fig. 6.3-3b); i.e. tension cancels compression,

    but tension/compression does not interact with shear.

    Since most dislocations are randomly curved mixed

    dislocations, there is a only a very low probability that all

    the conditions for dislocation annihilation will be fulfilled

    simultaneously. Thus, dislocation interactions with one

    another tend to be mutually repulsive.

    6-19

    These repulsive interactions obstruct the motion of those

    interacting segments of different dislocations, while non-

    interacting segments continue to move, creating many

    dislocation tangles (Figs. 6.3-4&5) during plastic deformation.

    Dislocations are therefore obstacles to the movement of

    other dislocations.

    Fig. 6.3-4 An edge dislocation (wavy black line) moving through a forest of other dislocations (red verticle lines). Intersecting segments that are mutually obstructive

    tangle with one another, distorting and lengthening the original dislocation.

    Fig. 6.3-5 Tangling dislocations marked with a b.

  • 6-20

    6.3.2 Strain Hardening

    Strain hardening, or work hardening is the phenomenon whereby a ductile metal becomes harder and stronger as it is plastically deformed. It is also known as cold working because the temperature at which deformation takes place is cold relative to the melting temperature of the metal.

    During plastic deformation, dislocations move under the action of a shear stress and encounter other dislocations. Since their interactions are generally repulsive (Sec. 6.3.1), a higher applied stress is necessary to overcome this mutual repulsion such that dislocation movement can continue; i.e. the metal has become stronger/harder.

    Furthermore, many new dislocations are continuously

    created during plastic deformation (Fig. 6.3-6), significantly increasing the dislocation density. The average distance between dislocations decreases, and the mutual resistance to motion becomes more pronounced, requiring an increasingly higher applied stress for continued plastic deformation; thus, the metal strengthens until fracture.

    Crystals that have intersecting slip systems, e.g. FCC and

    BCC, often strain-harden rapidly because slip tends to occur in more than one slip system, causing dislocations on different systems to intersect, impeding mutual motion.

    6-21

    Fig. 6.3-6 The sequence of events for the multiplication of a dislocation from a Frank-Read source. A segment of dislocation pinned at two points bows out into a loop.

    Continued stress will cause the loop to expand and the residual segment to bow out again into another loop. This process repeats over and over, sending out a set of concentric loops away from the source, creating many new dislocations. This is

    analogous to the ripples generated when a pebble is dropped into a quiet pond.

    The yield strength and tensile strength of a metal increases

    with increasing cold work, but ductility decreases (Figs. 6.3-7 &

    6.3-8). Physical properties such as thermal and electrical

    conductivity are also reduced due to the scattering of

    electrons and phonons by dislocations.

    Strain hardening in metals explains why the true stress-

    strain curve obtained during a tensile test shows a rising

    stress from the start of yielding to fracture (Sec. 2.2.5).

  • 6-22

    Fig. 6.3-7 Stress-strain diagram showing the effects of strain hardening.

    (a) Initially, yielding beings at A; (b) upon unloading and re-loading, yielding now occurs at the higher stress B.

    Fig. 6.3-8 The effects of cold work on the mechanical properties of copper.

    Metals may be shaped and strengthened at the same time

    by cold working (Fig. 6.3-9).

    6-23

    Fig. 6.3-9 Common metalworking processes: (a) rolling, (b) forging

    (open and closed die), (c) extrusion (direct and indirect), (d) wire drawing, (e) stamping.

  • 6-24

    6.3.3 Grain Size Strengthening

    In a polycrystal, each grain is has a different orientation to

    its neighbours. Since slip occurs only on specific planes,

    dislocations cannot move from one grain into another in a

    straight line (Fig. 6.3-10). Furthermore, the atomic disorder at

    grain boundaries interrupts the continuity of the slip

    planes, and acts as a barrier to dislocation motion.

    Fig. 6.3-10 Slip planes are discontinuous and change directions across the grain

    boundary. Dislocations cannot move through the

    grain boundary.

    A dislocation can move only within the grain in which it

    was created. Dislocations pile up at the grain boundary,

    causing strain energy to increase locally, creating a back

    stress that repels other dislocations approaching the pile-

    up (Fig. 6.3-11). A higher applied stress is needed to overcome

    this repulsion for continued dislocation movement.

    Fig. 6.3-11 Dislocation pile-up

    at a grain boundary.

    6-25

    The more grain boundaries there are (i.e. the smaller the

    grain size), the more obstacles there are to dislocation

    motion, and the higher the stress needed to cause plastic

    deformation; i.e. the metal becomes stronger/harder.

    The relationship between yield strength and grain size is

    expressed by the Hall-Petch equation: "y = "0 +

    kyd

    where "y = yield strength

    d = average grain diameter

    "0, ky = material constants

    Generally, polycrystals are stronger than single crystals (Fig.

    6.3-12); fine-grained metals are stronger than coarse-grained

    (Fig. 6.3-13). Effective strengthening can be realized only when

    the grain size is of the order of 5 m or less.

    .

    Fig. 6.3-12 Stress-strain curves for single crystal and polycrystalline copper.

    Fig. 6.3-13 Hall-Petch plot for brass, showing the effects of grain size on yield strength.

  • 6-26

    Grain size strengthening is one of the major reasons for the

    interest in nanocrystalline materials, in which grain sizes are

    less than 100 nm. However, as grain size is reduced below

    ~ 20 nm and becomes comparable to the width of grain

    boundaries, a reverse Hall-Petch effect is observed, where

    decreasing grain size causes softening (Fig. 6.3-14).

    Fig. 6.3-14 Hardness of a metal as a function of grain size.

    Grain size may be refined by cooling quickly from the

    molten state, inoculation of the melt (i.e. adding numerous

    impurity particles to the liquid to encourage solidification

    on the particles), or by extensive plastic deformation

    followed by rapid annealing (Sec. 6.4). Other special

    techniques are required to obtain grain sizes in the

    nanometre range.

    6-27

    6.3.4 Solid Solution Strengthening

    All materials contain small amounts of foreign atoms

    (element or compound). These impurities may arise

    unintentionally from raw materials and processing, or may

    be added intentionally to obtain specific properties.

    Impurities added intentionally are also known as alloying

    elements in metals, additives in polymers and ceramics,

    and dopants in semiconductors.

    Within a crystal, impurities (solute) may occupy interstitial

    sites or substitute for atoms of the host material (solvent),

    depending on the relative sizes of solute and solvent

    atoms. The incorporation of solute atoms without altering

    the crystal structure of the host results in a solid solution.

    Solute atoms distort the surrounding lattice and increase

    the strain energy of the crystal (Fig. 6.3-15).

    Fig. 6.3-15 Compressive strain imposed on host atoms by

    (a) an interstitial solute atom, and (b) a large substitutional atom. (c) Tensile strain imposed by a small substitutional atom.

  • 6-28

    The solute stress field could interact with that of an

    approa-ching dislocation such that repulsion arises (Fig. 6.3-

    16), similar to repulsion between dislocations of like signs

    (Sec 6.3-1). A higher applied stress is needed to overcome this

    repulsion.

    Fig. 6.3-16 Repulsion between compressive stress fields of solute and dislocation.

    On the other hand, if an interstitial or large substitutional

    solute atom with a compressive stress field were to be

    located in the tensile region around a dislocation, lattice

    strain is reduced (Fig. 6.3-17). A similar reduction in strain is

    seen for a small substitutional atom with tensile strain field

    located in the compressive region of a dislocation (Fig. 6.3-18).

    Once such a configuration of low strain are established

    between a solute atom and dislocation, further movement

    of the dislocation (i.e. away from the solute) would again

    raise strain energy (Fig. 6.3-19). This increase in energy is met

    by applying a higher stress; i.e. the metal strengthens. 6-29

    Fig. 6.3-17 (a) Compressive strains imposed by a large substitutional solute atom.

    (b) Possible locations of large solute atoms relative to an edge dislocation, leading to a reduction in overall lattice strain.

    Fig. 6.3-18 (a) Tensile strains imposed by a small substitutional solute atom.

    (b) Possible locations of small solute atoms relative to an edge dislocation, leading to a reduction in overall lattice strain.

    Fig. 6.3-19 Interaction with a suitable solute lowers dislocation strain energy. Continued plastic deformation requires the movement of dislocation away from the

    solute, which returns the dislocation and solute to their states before interaction. A higher stress is needed to restore the original strain energies of dislocation and solute.

  • 6-30

    Higher stresses are thus required for dislocation movement in the presence of solute atoms, which act as obstacles. The more solute added (without exceeding the solubility limit), the greater the strengthening (Fig. 6.3-20). Metals are seldom used pure, but are usually alloyed for strength.

    The degree of solid solution strengthening depends on the

    relative sizes of the solute and solvent atoms. The larger the size difference, the greater the distortion of the surrounding lattice, and the stronger the strengthening effect (Fig. 6.3-20). Too large a size difference, however, would lower solute solubility in the host lattice (Sec. 9.3.1).

    Fig. 6.3-19 The effects of several alloying elements on the yield strength of copper.

    6-31

    The strengthening effect further depends whether the

    solute is substitutional or interstitial. The stress field of a

    substitutional solute atom is spherically symmetric, without

    any shear component, and as such, does not interact with

    the shear stress fields of screw dislocations. Conversely,

    interstitial solute atoms in BCC crystals cause a tetragonal

    distortion, generating a stress field that can interact with

    both edge and screw (and thus, mixed) dislocations.

    Interstitial carbon solute atoms, having sufficient diffusivity

    in iron even at room temperatures, tend to move to

    favourable locations around dislocations in iron that would

    lead to mutual lowering of strain energy, thus pinning

    down these dislocations. During the initial stages of plastic

    deformation, a higher stress is needed to tear away the

    dislocations from their interstitial carbon solute atoms,

    after which only a lower stress is necessary to keep the

    dislocations moving. This gives rise to the yield point

    phenomenon in the stress-strain behaviour of steel (Fig. 6.3-21).

    Fig. 6.3-21 Dislocations in iron pinned down by interstitial carbon artoms require a higher stress to begin moving, resulting in the yield point phenomenon in the stress-strain curve of steel during tensile testing.

  • 6-32

    6.3.5 Dispersion Hardening

    Small, hard particles of a second phase dispersed in a

    softer, ductile matrix are effective obstacles to dislocation

    motion, and lead to significant strengthening. Such

    particles may be introduced intentionally, or arise naturally

    from precipitation reactions in an alloy, the latter

    producing a precipitation or age hardening effect.

    The interaction between dislocation and dispersed particle

    depends on the nature of the particle-matrix boundary. For

    particles that are intentionally incorporated, and for many

    precipitates, the particle-matrix interface is non-coherent

    and disordered (Fig. 6.3-22a); i.e. there is no atomic matching

    between the crystal lattice of the particle and matrix. Such

    particles do not distort the surrounding lattice.

    Fig. 6.3-22 (a) A particle that has no relationship with the crystal structure of the

    surrounding matrix forms a non-coherent interface with the matrix. (b) When there is a definite relationship between the crystal structures of the precipatate

    and matrix, a coherent or semi-coherent interface exists.

    6-33

    At a non-coherent interface between particle and matrix

    (Fig. 6.3-22a), there is a discontinuity of slip planes, much like

    that at grain boundaries (Sec. 6.3.3). A dislocation would be

    unable to move through such a particle.

    The dislocation may be forced to keep on moving by

    extruding or bowing between the particles (Fig. 6.3-23). Since

    a curve between two points is longer than a straight line,

    the bowed dislocation introduces greater lattice distortion

    and higher strain energy than the original, straight

    dislocation. A larger shear stress must now be applied to

    cause such bowing and continued plastic deformation.

    Fig. 6.3-23 A view looking down on a slip plane showing the bowing of a dislocation past particles having a non-coherent interface with the matrix.

    [Note that the circles represent particles, not single atoms.]

    After a dislocation has bowed past, dislocation loops are

    left around the particles (Fig. 6.3-23). The stress fields of these

    loops would interact with subsequent dislocations, and

    add resistance to their motion, leading to further

    strengthening.

  • 6-34

    Fine particles that are precipitated from an alloy often have

    planes of atoms in their crystal structures that are related

    to, or even continuous with, planes in the matrix lattice;

    such precipitate-matrix interfaces are said to be coherent

    (Fig. 6.3-22b).

    Since a coherent precipitate does not usually share the

    same lattice parameters as the matrix, this results in lattice

    strain. The stress field thus generated would interact with

    passing dislocations in a manner analogous to that of solid

    solution strengthening (Sec. 6.3.4).

    Because the stress field generated by a coherent precipitate

    is relatively wide, interactions with dislocations would

    occur wherever the stress fields impinge upon one

    another. The precipitate does not need to be on the slip

    plane of a dislocation to have a strengthening effect.

    When a coherent precipitate lies directly in the path of a

    dislocation, the coherency at the interface would allow the

    slip plane of the dislocation to pass from the matrix into

    the precipitate and shear the precipitate (Fig. 6.3-24).

    However, the creation of new matrix-precipitate surface

    area after shearing raises interfacial energy. For such

    shearing to occur, a higher stress must be applied to fund

    the increase in energy, thus strengthening the alloy.

    6-35

    Fig. 6.3-24 The formation of new precipitate-matrix interfaces when a

    dislocation cuts through a coherent precipitate.

    The degree of strengthening depends on the number and

    distribution of dispersed particles and precipitates: these

    should be as numerous as possible and uniformly

    distributed, so that they are closely spaced.

    Dispersion hardening is the principle behind metal matrix

    composites (MMCs), in which an alloy is strengthening by a

    dispersion of fine, hard particles, usually of a ceramic

    material. Ceramics retain their shape, distribution and

    superior hardness when heated, and are more effective at

    strengthening an alloy at high temperatures than

    precipitates, which tend to agglomerate (and hence,

    reduce in number) and re-dissolve into the matrix.

  • 6-36

    6.3.6 Combined Strengthening Mechanisms

    Two or more strengthening mechanisms may operate

    simultaneously to improve strength and hardness in metals

    (Tables 6.3-1&2 & Fig. 6.3-25).

    Table 6.3-1 The effectiveness of the various strengthening mechanisms on copper.

    6-37

    Fig. 6.3-25 Strengthening mechanisms in copper alloys and the variation in properties.

    Table 6.3-2 Metal alloys with typical applications, and the strengthening mechanisms used.

    Alloy Typical uses Solution hardening Precipitation hardening

    Work hardening

    Pure Al Kitchen foil !!! Pure Cu Wire !!! Cast Al, Mg Automotive parts !!! !

    Bronze (Cu-Sn), Brass (Cu-Zn) Marine components !!! ! !!

    Non-heat-treatable wrought Al Ships, cans, structures !!! !!!

    Heat-treatable wrought Al Aircraft, structures ! !!! !

    Low-carbon steels Car bodies, structures, ships, cans !!! !!!

    Low alloy steels Automotive parts, tools ! !!! !

    Stainless steels Pressure vessels !!! ! !!! Cast Ni alloys Jet engine turbines !!! !!! Symbols: !!! = Routinely used. ! = Sometimes used.

  • 6-38

    6-39

    6.4 ANNEALING

    There is a limit to which metals may be plastically

    deformed, beyond which fracture occurs.

    During forming operations, it is sometimes necessary to

    restore the ductility of work-hardened metals to their state

    prior to deformation, in order to carry out further plastic

    deformation.

    Work hardening also lowers the thermal and electrical

    conductivity of metals, which might require restoring; e.g.

    copper electrical wires.

    The effects of work hardening can be removed by heating

    the metal to a sufficiently high temperature in a process

    called annealing. Annealing replaces the highly distorted

    work-hardened grains with new, equiaxed grains

    containing few dislocations.

    The driving force for annealing is the reduction of strain

    energy associated with the high density of dislocations in a

    severely work-hardened metal.

    In annealing, there are three temperature ranges (from low

    to high) in which different phenomena occur: recovery,

    recrystallization and grain growth.

  • 6-40

    6.4.1 Recovery

    When heated sufficiently, dislocations in a strain hardened

    metal rearrange themselves into configurations with lower

    strain energy, forming the cell boundaries of a subgrain

    structure within the old grains (Fig. 6.4-1c), in a process called

    polygonization.

    Dislocation density is lowered slightly through mutual

    annihilation, but because the reduction is not significant,

    hardness, strength and ductility are almost unchanged (Fig.

    6.4-2). However, thermal and electrical conductivity are

    restored close to their pre-cold-worked states.

    6.4.2 Recrystall ization

    After recovery is complete, the strain energy of the crystal

    is still relatively high due to the large number of

    dislocations remaining. If the temperature were raised

    further, recrystallization will follow.

    New, small, dislocation-free grains nucleate at the high-

    energy cell boundaries of the polygonized subgrain

    structure (Fig. 6.4-1d), eliminating most of the dislocations as

    they grow and replace the strain hardened grains (Fig. 6.4-1e).

    6-41

    Fig. 6.4-1 Microstructural changes in cold working and annealing: (a) original state with few dislocations; (b) high density of dislocations after

    cold working; (c) recovery; (d) recrystallization; (e) fully recrystallized structure of new relatively strain-free grains with few dislocations.

    Since recrystallized grains are relatively free of dislocations,

    the hardness, strength and ductility of the metal are

    restored to their pre-cold-worked values; i.e. hardness and

    strength decrease while ductility increases (Fig. 6.4-2).

    The low strength and high ductility of a recrystallized

    metal are exploited in hot working, in which plastic

    deformation of a metallic alloy is carried out at

    temperatures above its recrystallization temparature

    (usually above 0.6 TM). The continually recrystallizing metal

    allows extensive deformation without strain hardening.

  • 6-42

    The main disadvantage of hot working is the poor surface

    finish as a result of oxidation of the metal surface at high

    temperatures. Dimensional accuracy is also an issue due to

    the elastic recovery (springback) and thermal contraction

    that occur when the component is cooled subsequently.

    Fig. 6.4-2 The effects of annealing temperature on mechanical properties and grain size.

    6-43

    6.4.3 Grain Growth

    If heating were to continue after complete recrystallization

    has occurred, the new grains will grow in size. [Note that grain growth occurs in all polycrystalline materials at sufficiently high temperatures; it is not related

    to cold-working. Only in cold-worked metals do recovery and recrystallization take place

    before grain growth.]

    The driving force for grain growth is the reduction of the

    interfacial energy associated with the atomic disorder at

    grain boundaries (Sec. 4.7). Grain growth results in fewer

    grains, thereby decreasing the total area of grain

    boundaries and lowering the interfacial energy.

    Grain growth involves the diffusion of atoms across the

    grain boundary from one grain to another, such that some

    grains grow at the expense of others (Fig. 6.4-3).

    Fig. 6.4-3 Grain growth occurs as atoms diffuse across the brain boundary

    from one grain to another.

    Grain growth reduces the strength and hardness of

    metallic alloys (Sec. 6.3.2), because the number of grain

    boundaries, which are barriers to dislocation motion, are

    now fewer.