You are on page 1of 93

Dislocations

NFPL135: Physics of Materials I

Faculty of Mathematics and Physics, Charles University


Prague, Czech Republic

Jana Šmilauerová
Plastic deformation in single crystals – slip
I plastic deformation in single crystals –
translation of one atomic plane over
another, i.e. slip or glide
I sufficiently high resolved shear stress
→ yielding of the material
I critical resolved shear stress –
minimum shear stress for slip
I slip in a certain crystallographic
direction and plane, slip direction and
slip plane
I slip plane – the most densely packed
plane (the largest planar density of
atoms, the largest interplanar spacing) Schematic image of slip lines; slip
plane 45◦ to the stress axis (side
I slip direction – the most densely
and front view) [1]
packed direction in the slip plane

slip system = slip plane + slip direction


Slip systems

W. D. Callister: Materials Science and Engineering, 2010


Slip systems

FCC: four non-parallel slip planes, each with three close-packed directions
→ at least one slip system is favourably oriented to the applied
stress → ductile behaviour
BCC: slip in {110} planes most frequent – the most densely packed plane;
however, slip in {112} and {123} also possible (the slip direction is
always h1̄11i) → total of 48 slip systems → wavy character of slip
bands, so-called pencil slip, very ductile
HCP: metals with a low c/a ratio or higher temperatures – activation of
other slip systems (prismatic or pyramidal)
Dislocations

I existence of dislocations was postulated at least a quarter of century


before the methods to visualize them were available
I 1905 - 1907 – the concept of dislocations in continuum elasticity
theory (Volterra, Timpe)
I 1926 – estimation of the theoretical shear stress for crystals (Y.
Frenkel)
I 1934 – papers which laid the foundations of dislocation theory,
postulating edge dislocations in crystals:
I E. Orowan, Zeitschrift für Physik 89 (1934) p. 634
I M. Polanyi, Zeitschrift für Physik 89 (1934) p. 660
I G. I. Taylor, Proccedings of the Royal Society A145 (1934) p. 362
I 1939 – postulation of screw dislocations by J. Burgers
I 1950 – multiplication mechanisms of dislocations (C. Frank, T. Read)
I 1956 – direct observation of dislocations (P. Hirsch et al.)
The concept of dislocation
I Frenkel∗ – an estimate of shear stress
required for moving one atomic plane
over another
I assumptions: two neighbouring planes
with the distance a and interatomic
spacing b, planes do not distort
I shear stress needed to move one atomic
plane over another has a roughly
sinusoidal shape:
2πx 2πx
τ = τmax sin ' τmax ,
b b
where τ is the shear stress and τmax the maximum shear stress
occurring for x = 4b


J. Frenkel, Zeitschrift für Physik 37 (1926) p. 572
The concept of dislocation
I for small displacements, Hooke’s law
applies:
x
τ = Gγ = G ,
a
where G is the shear modulus and γ is
the shear strain
I combining the two equations and
considering that b ≈ a yields the
theoretical shear stress for single
crystals:
G
τmax =

I metals: G ∼ 20 − 150 GPa ⇒ τmax ∼ 3 − 25 GPa
I real shear stresses for metals are τreal ∼ 0, 5 − 10 MPa ⇒ slip cannot be
realized only by shearing of atomic planes
I the discrepancy between the theoretical and observed shear stresses
⇒ dislocations
The concept of dislocation
I nano-sized single-crystalline fibres – whiskers – may approach the
theoretical strength

Tensile testing of Cu whiskers (G. Richter et al. Nano Letters 9 (2009) 3048)
Dislocations

J. F. Shackelford: Introduction to Materials Science for Engineers, 2015


(adapted from: W. C. Moss: Ph.D. Thesis, University of California, Davis, 1979)
Dislocations
I line defect, 1D
I displacement of nearby atoms – lattice
distortions
I can move with high velocities under
external stress
I edge dislocation
I dislocation line – along the edge of the Edge dislocation [2]
“extra” half-plane of atoms
I representation by a ⊥ symbol (or > when
the half-plane is inserted in the lower part
of the crystal)
I screw dislocation
I may be thought of as produced by
shearing distortion
I atomic distortion along a line – atomic
planes trace a helical path around the
dislocation line
I sometimes, a symbol is used
Screw dislocation [3]
Dislocations
I mixed dislocation
I most dislocations are not pure screw or edge, but a combination of both
I transition region exhibits varying degrees of screw and edge character

Top-view of a mixed dislocation. Empty circles


Mixed dislocation [3] are atoms above the slip plane, full circles are
below the slip plane. Adapted from [1]
Dislocations

I provide a mechanism for plastic deformation – slip of a large number


of dislocations
I NB: plastic deformation = irreversible changes in shape due to
applied stress (× elastic deformation – reversible, stretching of
atomic bonds)
I explain why strength of materials is lower than theoretical values –
due to dislocations, only a small number of bonds is broken at any
given time during slip
I dislocations provide ductility in metals
I controlling mechanical properties of metals and alloys by interfering
with dislocation movement (obstacles → stronger material)
I dislocation density = the total length of dislocations per unit volume;
units: m/m3 = m−2
I ∼ 1010 m−2 for soft metals
I ∼ 1012 m−2 deformed material
Dislocations

I dislocation line marks the


boundary between sheared and
unsheared parts of the crystal
I dislocation cannot end inside
the crystal – always at surface,
grain boundary etc., or it Schematics of a closed dislocation loop: a) positive
edge, b) right-hand screw, c) negative edge,
creates a loop or branches into d) left-hand screw. The area inside the dislocation
other dislocations loop is sheared by one atomic distance [1]

I left-/right-hand screw dislocation – left/right thumb pointing along the


dislocation line, fingers showing advance of the spiral towards the
thumb
I positive/negative edge dislocation – extra plane of atoms above/below
the slip plane – only arbitrary (up and down are relative in the crystal)
Dislocations

I direction and magnitude of


shear indicated by Burgers
vector, ~b
I although the dislocation varies
in direction, shear (and ~b) are Schematics of a closed dislocation loop: a) positive
everywhere the same → ~b is edge, b) right-hand screw, c) negative edge,
d) left-hand screw. The area inside the dislocation
characteristic for the dislocation loop is sheared by one atomic distance [1]

I Burgers vector + orientation of dislocation line → complete description


of a dislocation
I the same line sense and opposite Burgers vectors (or vice versa) →
annihilation

“Opposite” dislocations annihilate [4]


Burgers vector

Burgers circuit for edge dislocation: a) perfect lattice, b) lattice with dislocation [1]
I Burgers vector determination by a circuit method
I the circuit must be drawn in the positive direction of a right-hand
screw dislocation
I the circuit must close in a perfect crystal
I Burgers vector = vector closing the loop in an imperfect crystal
I arbitrary assumptions – not always the same convention in literature
Burgers vector

Burgers circuit for screw dislocation: a) perfect lattice, b) lattice with dislocation [1]
I length of Burgers vector = the shortest lattice translation vector
(perfect dislocation)
I Burgers vector lies in the slip direction (usually the closest packed
direction)
Burgers vector

I denoted by Miller indices


preceded by a fraction of
the lattice parameter a,
e.g.:
SC: ~b = a[100]
FCC: ~b = a [110]
2
BCC: ~b = a [111]
2

Spacing between atoms in the most densely packed


directions in SC, BCC and FCC [1]
Burgers vector

Edge dislocation
I lies perpendicular to the Burgers vector
I moves (in its slip plane) in the direction of the Burgers vector
Screw dislocation
I lies parallel to the Burgers vector
I moves (in its slip plane) in the direction perpendicular to the Burgers
vector
Slip plane = plane containing both the Burgers vector and the dislocation
line
⇒ slip plane of edge dislocation is unambiguously defined → edge
dislocation is confined to move (glide) in a single plane
⇒ slip plane of screw dislocation can be any plane containing the
dislocation → screw dislocations can glide in any direction as long as
they move parallel to their original orientation
Dislocation motion

Slip plane, Burgers vector and dislocation line for a) edge and b) screw dislocation [5]
I dislocation motion in the slip plane (slip) requires only a small
rearrangement of atoms → a small force is needed
I Peierls-Nabarro stress = stress required to move a dislocation
 
2G −2πw
τPN = exp ,
(1 − ν) b

where G is the shear modulus, ν the Poissons ratio, b the Burgers


vector and w is the width of the dislocation
Dislocation motion
I dislocation width w – the
distance over which the
displacement difference ∆u
is greater than one half of its
maximum value, i.e.
−b/4 ≤ ∆u ≤ b/4
I w – measure of the size of
dislocation core, i.e. the
region in which
displacements and strains
cannot be described by
elasticity theory
I Peierls & Nabarro:
w = a/(1 − ν) for an edge
dislocation, a) Displacement of atoms at an edge dislocation. Open
w = a for a screw and full circles are atom positions before and after the
dislocation; a is the extra plane was inserted. b) Displacement difference ∆u
and the width of the dislocation w [6]
interplanar spacing
Dislocation motion
 
2G −2πw
τPN = exp ,
(1 − ν) b

I dislocations move in the system in which the lowest energy is needed


I Peierls-Nabarro stress
I increases exponentially with Burgers vector length → close-packed
directions
I decreases exponentially with lattice spacing of the slip planes → easier
slip in “smooth” planes that are further apart
I wide, planar core → low τPN → edge dislocations are generally more
mobile than screw
I materials resistant to slip: covalent bonds (e.g. Si), ionic bonds (e.g.
Al2 O3 – charge balance disruption, repulsive forces between same
charges)
I note – slip associated with a single dislocation is small (an atomic
step on the surface) – slip lines on the surface of crystals are created
by many dislocations which moved on the same plane
I slip band – a group of slip lines
Dislocations in FCC

Perfect edge dislocation in FCC lattice (top view) [1]


I FCC slip plane: {111} (octahedral plane)
I movement in the direction of ~b – difficult – each atom would need to
“climb over” two atoms in the bottom plane
Dislocations in FCC

Partial edge dislocations in FCC lattice (top view) [1]


I easier – zig-zag motion through “valleys” between atoms –
dislocation breaks into two partial dislocations

~b1 = ~b2 + ~b3 ; a a a


[1̄10] = [1̄21̄] + [2̄11]
2 6 6
I decrease in strain energy: E ∝ b2 ; b21 > b22 + b23
Dislocations in FCC

Extended dislocation, stacking fault: ABCACABCA instead of ABCABCA [1]


I repulsive forces between partial dislocations – partials forced apart
I extended dislocation – dissociated into a pair of partials
I stacking sequence discontinued between the partials – stacking fault
I Shockley partial dislocation – ~b lies in the fault plane (depicted)
I Frank type – ~b normal to the fault plane
Shockley partial dislocation
I atoms in different
positions than in
a perfect lattice →
stacking fault has
a surface energy
I 1st partial passes →
fault in the stacking
sequence, 2nd partial →
stacking is restored
I ↑ separation between
partials ⇒ ↓ repulsive
force
I ↑ separation ⇒ ↑
surface energy
I separation distance – equilibrium between the two effects
cos 60◦
I repulsive force: F ≈ Gb2 b32πd , G is the shear modulus in the slip
plane, d is the separation between partials
I equilibrium separation: d0 = Gb 2 b3
2πγ , γ is the force per unit length of line
provided by stacking fault energy
Shockley partial dislocation

I high stacking fault energy ⇒ small separation between partials


I low stacking fault energy ⇒ large separation

Typical values of stacking fault energy


Metal Stacking fault energy (mJ/m2 )
Brass <10
303 stainless steel 8
304 stainless steel 20
310 stainless steel 45
Silver ∼25
Gold ∼50
Copper ∼80
Nickel ∼150
Aluminium ∼200
G. E. Dieter: Mechanical Metallurgy, 1988
Frank partial dislocation

I stacking fault formed by inserting or removing a part of a


close-packed {111} plane
I Burgers vector: 13 [111] (normal to the {111} slip plane – Frank type)
I Burgers vector is not contained in one of the {111} glide planes →
sessile dislocation – cannot glide, only climb

Intrinsic stacking fault can be created by Extrinsic stacking fault can be created by
clustering of vacancies on an octahedral precipitation of interstitials. Stacking
{111} plane. Stacking ABCACABCA, same ABCACBCABC. Adapted from [1].
as for Shockley partial.
Extended dislocations in HCP

I basal plane of HCP has the same close-packed arrangement as


{111} planes in FCC → extended dislocations also occur in HCP
metals
I dissociation of the perfect dislocation:
a a a
[1̄21̄0] = [011̄0] + [1̄100],
3 3 3
which are exactly the same Burgers vectors as in the FCC system
(try the conversion between the three- and four-index notations!)
I passage of 1st partial dislocation: ABABABABAB... →
ABACBCBCBC...
I passage of 2nd partial: restored to ABABABABAB...
Dislocations in BCC
I slip planes {110}, {112} and {123}; slip directions h111i; Burgers
vector 12 h111i
I three {110}, three {112} and six {123} planes intersect along the
same h111i direction ⇒ many possible slip planes for screw
dislocations → wavy slip lines
I asymmetry of slip – when a single crystal yields, different slip
systems are active in tension and compression
I stacking faults not observed experimentally

I dislocation reaction leading


to formation of immobile
dislocation → crack
nucleus for brittle fracture
a a
[1̄1̄1] + [111] → a[001]
2 2
Climb of edge dislocations
I edge dislocation: ~b ⊥ ~l → only one
possible slip plane
I screw dislocation ~b k ~l → slip/glide in
any direction perpendicular to the
dislocation line
I climb of edge dislocation – motion
perpendicular to the slip plane
I positive climb – removing atoms from
the crystal to make the extra half-plane
smaller → the crystal shrinks →
positive climb is promoted by a) - b) Positive climb of edge dislocation
compressive stress perpendicular to with annihilation of a vacancy; b) - c)
the extra plane Negative climb with formation of a
vacancy.

I negative climb – the extra plane grows and produces vacancies →


expansion of the crystal → negative climb promoted by tensile stress
I ↑ temperature ⇒ ↑ vacancy motion ⇒ ↑ climb
I NB: slip occurs as the result of shear stress
Cross slip
I movement of screw dislocations on a slip plane – when it encounters an
obstacle, it can easily shift to other intersecting slip system (~b k ~l) →
cross slip

I cross slip
possible in FCC
and BCC metals
I cross slip not
possible in many
HCP metals (slip
planes – basal –
not intersecting)
until alloyed or
heated
I cross slip
produces
non-planar slip a) - c) Cross slip in FCC. The [1̄01] direction is common to the two slip
surfaces planes → a screw dislocation at point S can move in either slip plane.
d) Double cross slip [6]
Dislocation intersections

I real material - many dislocations on


any given slip plane – dislocation on
one slip plane intersects other slip
planes
I movement of a dislocation – passing
through dislocations on other
intersecting slip planes
Dislocation moved across slip plane
I intersection – additional work needed ABCD and created two jogs in the
dislocation loop [1]

I e.g. shearing of a dislocation loop (see figure) – both dislocations


acquire steps equal to the Burgers vector of the other one
Dislocation intersections

Edge and screw dislocations with kinks lying in the slip plane [1]
I kink – a step lying in the slip plane
I jog – a step normal to the slip plane
I kink in an edge dislocation has a screw orientation
I kink in a screw dislocation has an edge or orientation
I kinks can be easily eliminated by slip
I eliminating kinks lowers energy → kinks tend to disappear
Dislocation intersections

Edge and screw dislocations with jogs normal the slip plane [1]
I kink – a step lying in the slip plane
I jog – a step normal to the slip plane
I jog at edge dislocation has an edge character → dislocation free to
move
I jog at screw dislocation (intersection of two screw dislocations) has
an edge orientation – jog movement by climb only producing row of
interstitials or vacancies
Dislocation intersections
I What will happen in each of the following cases? Will a jog, a kink or
nothing form on dislocation lines?
I What is the Burgers vector for each formed jog/kink?
I Does the jog/kink have an edge or a screw orientation?
I What is the length of each formed jog/kink?
Dislocation intersections
I a dislocation moving on a slip plane
has to intersect dislocations crossing
the slip plane (i.e. lying in different slip
planes), so-called forest dislocations
I combination of different dislocation
types – different result – jog/kink
I for simplicity, let us consider orthogonal
slip planes
two edge dislocations with perpendicular
Burgers vectors
I jog forms on dislocation AB
I Burgers vector of the jog is ~b2
I jog length is |~b1 |
I no jog forms on dislocation XY, since
Burgers vector of AB (~b2 ) is parallel to
Intersection of two edge dislocations with
Burgers vectors perpendicular to each
the XY dislocation line
other [6]
Dislocation intersections

two edge dislocations with parallel Burgers


vectors
I both dislocations obtain steps which lie
in the respective slip planes – kinks
I both kinks have a screw orientation
I length of kink QQ’ is |~b2 |
I length of kink PP’ is |~b1 |

Intersection of two edge dislocations with


Burgers vectors parallel to each other [6]
Dislocation intersections

edge and screw dislocations


I jog on the edge dislocation with an
edge orientation
I kink on the screw dislocation with an
edge orientation

Intersection of an edge with


a right-handed screw dislocation [6]
Dislocation intersections

Intersection of two screw dislocations [6]

Two screw dislocations


I jogs with an edge orientation produced on both screw dislocations
I these jogs are very immobile (only thermally activated climb) →
important in plastic deformation
Movement of dislocations containing jogs
Edge dislocations
I jogs on edge dislocations have an edge character and Burgers
vector always lies in the slip plane → jog can glide with the edge
dislocation
I kinks on edge dislocations are easily removed by glide

steps in pure edge dislocations do not impede their glide

Screw dislocations
I kinks on screw dislocations have an edge character, Burgers vector
of the kink lies in the slip plane → it can move by slip along the axis
of the screw dislocation
I jogs on screw dislocations have an edge character with Burgers
vector lying normal to the slip plane → these jogs cannot move by
glide → pinning points of screw dislocations
I screw dislocations with jogs can move only by climb – temperature
dependent
Movement of dislocations containing jogs
jogs

b
a)

b)

c)
vacancies

a) dislocation with no applied stress


b) dislocation segments pinned by jogs bowing under stress in the slip
plane
c) glide of jogged screw dislocation producing trails of point defects –
vacancies or interstitials depending on the direction of movement
Superjogs
I superjogs – jogs more than
one interplanar (slip plane)
spacing high
I a few atomic planes high –
dragging of jogs (trails of
point defects) may still be
possible
I jogs of intermediate height
→ dislocation dipole – two
edge lines of the same
Burgers vector and opposite
signs – they interact and
cannot pass each other
(except at high stresses)
I long jogs → two edge a) small jog is dragged along creating point defects, b)
“arms” behave long jog – independent dislocations, c) intermediate jog –
dislocation dipole [6]
independently –
single-ended dislocation
sources
Jogs and prismatic loops

I trails of defects and prismatic loops


often produced during plastic
deformation behind jogged screw
dislocations – debris
I formation of loops
I coalescence of vacancies (higher
temperatures) or interstitials (easier
diffusion at lower temperatures)
generated during movement of jogs
I pinch-off of dislocation dipoles
I due to multiple cross slip (interaction
with obstacles)

a) dislocation dipole,
b) elongated loop and
c) row of small loops and a
jogged dislocation [6]
Interaction of dislocations with obstacles

Movement of dislocation around a particle a) without cross slip (Orowan mechanism),


b) – d) with cross slip of screw segments (Hirsch mechanism) [6]
Intersections of extended dislocations

I complex problem – many variants of


intersections
I extended dislocations can cut through
each other only after constricting to
perfect dislocations
I to form jogs on extended dislocations,
work needs to be done to constrict
dislocations → energy of the jog
depends on the stacking fault energy
I jogs/kinks are produced (which also
may separate into partials)

Intersection of extended
dislocations [6]
Cross slip of extended dislocations
I individual Shockley
partial cannot cross slip,
because a 6a h112i lies in
only one {111} plane
(FCC) ⇒ extended
dislocation can glide
only in the {111} plane Constriction of extended dislocation in FCC [6]
of the stacking fault
I cross slip after
constriction to a perfect
dislocation
I energy required for
constriction → easier in
metals with a high
stacking fault energy
(i.e. small distance Stages of cross slip of extended dislocation, constriction is
necessary for the cross slip [6]
between partials)
Lomer-Cottrell sessile dislocation
I FCC: when two dissociated
dislocations gliding on different {111}
planes meet, leading partials repel or
attract each other depending on the
particular 6a h112i vectors and their
orientations (36 possibilities)
I the most favourable interaction:
a a a
[1̄21̄] + [11̄2] → [011]
6 6 6
I the product partial – stair-rod
dislocation whose Burgers vector is
perpendicular to the dislocation line
and does not lie in either of the {111}
planes – sessile dislocation
I the three partials form a stable
arrangement – Lomer-Cottrell lock
→ an important obstacle for dislocation
movement Formation of a Lomer-Cottrell lock [6]
Stacking fault tetrahedra
I metals and alloys of low stacking fault energy
I supersaturation of vacancies
I stacking fault tetrahedron – tetrahedron of
intrinsic stacking faults on {111} planes with
1
6 h110i stair-rod dislocations along its edges
I formation by Silcox-Hirsch mechanism:
dissociation of Frank partial into a stair-rod
dislocation and a Shockley partial on an
intersecting slip plane

1 1 1
[111] → [101] + [121]
3 6 6
I Shockley partials are repelled by stair-rod
dislocations and bow out on their slip planes
→ they attract each other and create stair rod
dislocations along remaining edges of the
tetrahedron ( 16 h110i)
Stacking fault tetrahedra

TEM of stacking fault tetrahedra in quenched Au; shape of tetrahedra depends on the
orientation of the foil (here: (110)). The high contrast arises from overlapping of stacking
faults in different faces of tetrahedrons. (Cottrell, Phil. Mag 6, 1351, 1961)
Multiplication of dislocations

I dislocation density in well-annealed materials ∼ 104 − 106 m−2


I dislocation density after large deformation ∼ 1012 − 1015 m−2
I for large plastic strains in metals and alloys – regenerative
multiplication sources are neccessary
I important mechanisms: Frank-Read source, multiple cross glide,
Bardeen-Herring source
Frank-Read source
Frank-Read source
I segment dislocation line AB is
anchored at both ends by
obstacles (forest dislocations,
precipitates, jogs, nodes, etc.)
I applied resolved shear stress τ
exerts force τ b per unit length of
the line ⇒ dislocation segment
bows out
I as τ increases, radius of the
curvature R decreases until
R = L/2, where L is the length
of segment AB
I the maximum τ for a semicircle
dislocation: τ ≈ Gb Gb
2R ≈ L

Principle of Frank-Read source [6]


Frank-Read source
Frank-Read source
I further increase in τ again
increases R and the dislocation
becomes unstable → forms a
kidney-shaped loop
I annihilation of the segment
which come into contact (same
Burgers vector, opposite line
sense)
I the outer loop continues to
expand, dislocation between AB
is regenerated and the process
repeats

Principle of Frank-Read source [6]


Frank-Read source

Frank-Read source in silicon crystal in {111} plane. Anchoring points are dislocations in another slip
plane (out of focus). Dislocation lines lie along h110i directions in which their energy is the lowest.
(Dash, Dislocation and Mechanical Properties of Crystals, Wiley, 1957)
Single-ended Frank-Read source

I a dislocation anchored at one end in the slip


plane → moves only by rotation around the
anchoring point
I each revolution → displacement of the crystal
above the slip plane by b
I regenerative: n revolutions → displacement nb
→ large slip step may be produced at the
surface

Single-ended
Frank-Read source [6]
Multiple cross slip
I screw dislocation along
AB can cross glide onto
position CD, long jogs
(AC, BD) relatively
immobile
I segments on primary
slip planes expand and
operate as Frank-Read
sources
I one continuous
dislocation loop lying in
many parallel slip planes
can be formed (i.e. the
Frank-Read source does
not complete a cycle)
I more effective than
simple Frank-Read
source
Bardeen-Herring source

I only high temperatures – climb due to


vacancy diffusion
I edge dislocation – mechanism similar
to the Frank-Read source, but the
dislocation bows out not in the slip emits
plane, but in the one containing the vacancies
extra half-plane
I regenerative only if anchor points are absorbs
screw dislocations – otherwise a full vacancies

revolution would remove (or complete)


the extra half-plane without creating a
new anchored dislocation

Schematics of a Bardeen-Herring
source - edge dislocation bows out by
climb and emits/absorbs vacancies
Thompson’s tetrahedron
I notation for describing dislocations and
dislocation reactions in FCC metals
I four sets of {111} form a regular tetrahedron
whose edges are h110i slip directions
I corners of the tetrahedron: A, B, C, D
I mid-points of faces: α, β, γ, δ
I Burgers vectors of dislocations denoted by two
end points on the tetrahedron
I perfect dislocations a h110i are AB, BC, etc.
2
I Shockley partials a h112i are Aβ, Aγ, etc.
6
I dissociation
a a a
h110i → h211i + h121̄i
2 6 6
AB = Aδ + δB (on ABC slip plane)
AB = Aγ + γB (on ABD slip plane)

I formation of Lomer-Cottrell lock from two Shockley partials:


αD + Dβ → αβ
Thompson’s tetrahedron
Stress fields of dislocations
I displacements of atoms around dislocations from their ideal lattice
sites ⇒ dislocations are sources of internal stress
I displacement of a point

~u = [ux , uy , uz ]

I nine components of strain (tensor):

 
∂ux 1 ∂uy ∂uz
εxx = εyz = εzy = +
∂x 2 ∂z ∂y
 
∂uy 1 ∂uz ∂ux
εyy = εzx = εxz = +
∂y 2 ∂x ∂z
 
∂uz 1 ∂ux ∂uy
εzz = εxy = εyx = +
∂z 2 ∂y ∂x
Stress fields of dislocations
I stress – force per unit area of surface (i.e. the orientation of the
surface needs to be specified along with the magnitude and direction
of force)

Stress components on cube faces [6]


I shear components of stress – with i 6= j; shear stress acting on a slip
plane denoted τ
I due to moments of the forces and the rotational equilibrium of the
element:
σyz = σzy , σzx = σxz , σxy = σyx
I normal components: σxx , σyy , σzz
I positive normal stress → tension, negative → compression
Stress fields of dislocations
I Hooke’s law – relationship between stress and strain, linear elasticity
I isotropic solids – two proportionality constants (Lamé constants G, λ)

σxx = 2Gεxx + λ(εxx + εyy + εzz )


σyy = 2Gεyy + λ(εxx + εyy + εzz )
σzz = 2Gεzz + λ(εxx + εyy + εzz )
σxy = 2Gεxy σyz = 2Gεyz σzx = 2Gεzx
I G ... shear modulus
I since only two material constants are needed in Hooke’s law, other
elastic constants are interrelated, e.g.:
I Young’s modulus:
E = 2G(1 + ν)
I Poisson’s ratio:
λ
ν=
2(λ + G)
I bulk modulus:
E
K=
3(1 − 2ν)
I a full list of conversion formulas can be found here
Stress fields of dislocations

I sometimes cylindrical coordinates (r, θ, z) are more convenient


I definitions of strain and stress tensors are the same

Stress components in cylindrical coordinates [6]


Stress fields of dislocations
Screw dislocation
I infinitely long screw dislocation – cylinder of
elastic material (Volterra model) which has
been sheared along a longitudinal cut
I no displacements in x and y directions:

ux = uy = 0
I displacement along z from 0 to b (length of
Burgers vector) as θ increases from 0 to 2π:


uz = , θ = arctan(y/x)

I from the relations for components of strain and stress
εxx = εyy = εzz = εxy = εyx = 0 σxx = σyy = σzz = σxy = σyx = 0
b y b sin θ Gb y Gb sin θ
εxz = εzx = − =− σxz = σzx = − =−
4π x2 + y2 4π r 2π x2 + y2 2π r
b x b cos θ Gb x Gb cos θ
εyz = εzy = = σyz = σzy = =
4π x2 + y2 4π r 2π x2 + y2 2π r
Stress fields of dislocations
Screw dislocation
I in cylindrical coordinates – using relations for shear stress (and
analogous for shear strain)
σrz = σxz cos θ + σyz sin θ
σθz = −σxz sin θ + σyz cos θ

the only non-zero elements are found to be


b Gb
εθz = εzθ = σθz = σzθ =
4πr 2πr
⇒ elastic distortion consists of pure shear (no compressive/tensile
components): σzθ acts parallel to z axis in radial planes of constant θ;
σθz acts as a torque on planes normal to z axis
I stress field is radially symmetric
I for a dislocation of opposite sign (i.e. left-handed), the signs of all
components are reversed
I stress and strain ∼ 1r ⇒ divergence for r → 0 ⇒ core of the dislocation
r0 ∼ 1b − 4b, where the linear elasticity theory ceases to be valid
Stress fields of dislocations
Edge dislocation
I faces of a cut through a cylinder displaced by b
in x direction
I displacement and strains in z direction are zero
I stresses are

3x2 + y2 sin θ(2 + cos 2θ)


σxx = −Dy = −D
(x2 + y2 )2 r
x2 − y2 sin θ cos 2θ
σyy = Dy =D
(x2 + y2 )2 r
σzz = ν(σxx + σyy )
x2 − y2 cos θ cos 2θ
σxy = σyx = Dx 2 2 2
=D
(x + y ) r
σxz = σzx = σyz = σzy = 0
Gb
where D = 2π(1−ν)
Stress fields of dislocations
Edge dislocation
I stress field has both normal and shear components
I the maximum normal stress is σxx , compressive above the slip plane
(σxx < 0 for y > 0), tensile below (σxx > 0 for y < 0)
I dislocation of opposite sign – reversed signs of stress components
I similar to screw ∼ 1r – the solution is valid only outside a core of
radius r0

Left: stresses on elementary cube; right: iso-stress contours for normal and shear components [7]

Mixed dislocation
I stress and strain fields – sum of screw and edge constituents with
corresponding Burgers vectors ~bscrew + ~bedge = ~b
Strain energies of dislocations
I distortion around a dislocation → material is not in its lowest energy
state → increase by strain energy
I energy added by a dislocation = energy of dislocation core + elastic
strain energy
Edisl = Ecore + Eel
I elastic strain energy of volume dV is dEel = 12
P P
σij εij dV
i=r,θ,z j=r,θ,z

Screw dislocation
b Gb
εθz = εzθ = σθz = σzθ =
4πr 2πr
ZL Z2π ZR
1 Gb2 L R
Eel = dz dθ (σθz εθz + σzθ εzθ )r dr = ln
2 4π r0
0 0 r0

I NB: in cylindrical coordinates dV = r dz dθ dr


I r0 is the core radius, R is the outer radius (due to crystal/grain size)
Strain energies of dislocations
Edge dislocation
I integration much more difficult, but the result is rather simple
Gb2 L R
Eel = ln
4π(1 − ν) r0
In general:

Gb2 L R
Eel = ln , K = 1 for screw, K = (1 − ν) for edge
4πK r0

I the total energy of a dislocation is proportional to its length L


I often the equations for Eel are normalized to unit length (i.e. without L)
I a dislocation tends to be straight between two points
I the (normalized) energy of an edge dislocation is always higher than
that of a screw
(edge)
I ν ≈ 0.33 → E(screw)
el ≈ 0.66Eel
I a dislocation tends to have its screw component as large as possible →
can be in contradiction with length minimization
I logarithmic dependence on core radius r0
Strain energies of dislocations
Gb2 L R
Eel = ln , K = 1 for screw, K = (1 − ν) for edge
4πK r0

I elastic energy weakly depends (logarithmically) on R, i.e. the size of


crystal/grain or distance to other dislocations in heavily deformed
materials (dislocations with different Burgers vectors → strain fields
cancel out)

1
Rmin ∼ √ ∼ 10−7 m for ρ ∼ 1014 m−2
ρ
Rmax ∼ 1 cm ∼ 10−2 m sample size

assuming r0 = 0.5 nm

Rmin 10−7

ln = ln = 5.3 

r0 5 · 10−10

factor of ∼ 3 variation
Rmax 10−2 
ln = ln = 16.8


r0 5 · 10−10
Strain energies of dislocations
I mixed dislocation – sum of edge and screw energies with b replaced
by b sin θ and b cos θ, respectively (normalized to dislocation length)
 2 2
Gb2 cos2 θ Gb2 (1 − ν cos2 θ) R

(mixed) Gb sin θ R
Eel = + ln = ln
4π(1 − ν) 4π r0 4π(1 − ν) r0
I approximation of dislocation energy (normalized to L)

Gb2 Gb2 R
Ecore = Z, Z ≈ 1 − 2 Eel = ln
4π 4πK r0
Gb2
 
R
Edisl = Ecore + Eel = ln + Z
4πK r0
Edisl = αGb2
I α ∼ 0.5 − 1.5

Edisl ≈ Gb2

I note that the elastic energy is higher than the core energy
I the total energy of a dislocation is proportional to b2 ⇒ dislocations
tend to have the smallest possible Burgers vector
Strain energies of dislocations

Edisl ≈ Gb2

I Frank’s rule for dissociation reactions of


b1 b2 b3
the type ~b1 = ~b2 + ~b3 :

if b21 > b22 + b23 favourable = +


(0 < φ < π/2)
if b21 < b22 + b23 unfavourable
(π/2 < φ < π)
b2 ϕ b3
I recall dissociation into Shockley partials
and formation of Lomer-Cottrell lock
b1

I NB: dislocations increase the crystal energy (TD equilibrium – no disl.)


I crystal with dislocations – TD unstable system × crystal with point
defects – can be TD stable (i.e. certain equilibrium vacancy
concentration at a given T)
Forces on dislocations

I movement of a dislocation on its glide plane – only the shear stress


on this plane needs to be considered → resolved shear stress τ
I a dislocation moves in the slip plane due to a uniform resolved shear
stress τ
I when en element dl of a dislocation line of
Burgers vector ~b moves a distance ds, the
atomic planes above the slip plane will be
displaced by b relative to the atomic planes
below
I average shear displacement produced by
glide dl is
(ds dl) Displacement ds of
b, a dislocation line segment dl
A in its glide plane [6]
where A is the area of the slip plane
Forces on dislocations
I the force due to τ acting on the plane is Aτ
I the work done when the element of slip occurs is

(ds dl)
dW = Aτ b = τ b ds dl
A
I the force F on a unit length of dislocation = the work done when a
unit length of dislocation moves a unit distance

dW dW
F= =
ds dl dA

F = τb

... simple form of Peach-Koehler equation


I F is a magnitude of the force, not a vector
I F acts normal to the dislocation at every point of its length
I the sign of F depends on the Burgers vector, line vector and τ
I τ is the shear stress in the glide plane resolved in the direction of ~b
Forces on dislocations

To produce the same deformation, the same τ generates force on a screw dislocation that is
perpendicular to the force on an edge dislocation. Adapted from [6]
Forces on dislocations

I general form of Peach-Koehler equation

~F = (σ · ~b) × ~l

~F force per unit length at an arbitrary point along a dislocation line


σ local stress field
~l line tangent direction at the given point
σ · ~b local force per unit length acting on a plane normal to the
Burgers vector
Forces on dislocations
Example: edge dislocation along y axis, Burgers vector in the negative
direction of x, i.e. ~l = (0, 1, 0), ~b = (−b, 0, 0), and the glide plane is (001)
~F = (σ · ~b) ×~l
| {z }
~g=(gx ,gy ,gz )

gx = σxx bx + σxy by + σxz bz = −σxx b


gy = σyx bx + σyy by + σyz bz = −σyx b
gz = σzx bx + σzy by + σzz bz = −σzx b
gx = −σxx b ⊥ ~l ⇒ σxx produces a component of ~F perpendicular to the
glide plane → climb force
if σxx < 0 (compression), the edge dislocation climbs
upwards, and vice versa
gy = −σyx b k ~l ⇒ σyx – no contribution to ~F
gz = −σzx b ⊥ ~l ⇒ σzx produces a component of ~F along the x axis →
glide force
Forces on dislocations
Example: screw dislocation along x axis, Burgers vector in the negative
direction of x, i.e. ~l = (1, 0, 0), ~b = (−b, 0, 0), and the glide plane is (001)
~F = (σ · ~b) ×~l
| {z }
~g=(gx ,gy ,gz )

gx = σxx bx + σxy by + σxz bz = −σxx b


gy = σyx bx + σyy by + σyz bz = −σyx b
gz = σzx bx + σzy by + σzz bz = −σzx b
~
gx = −σxx b k l ⇒ σxx – no contribution to ~F
gy = −σyx b ⊥ ~l ⇒ σyx produces a component of ~F downwards/upwards
→ cross slip
gz = −σzx b ⊥ ~l ⇒ σzx produces a component of ~F along y axis → glide
force in the slip plane
Forces on dislocations
I the line energy (= energy per length) has the same dimension as force
(F = − dU/ dL) → expresses a line tension, i.e. a force in the direction
of the line vector which tries to shorten the dislocation (analogy to
surface tension of a bubble)
I the line energy is of the order of 5 eV per Burgers vector
I line tension – increase in energy per unit increase in length

T = αGb2
I outward force acting on
dislocation length dl due to
shear stress τ0 (needed to
maintain the curvature R)

τ0 b dl

I force due to line tension


straightening the line


2T sin ≈ T dθ for small dθ
2
Forces on dislocations
I the dislocation line is in equilibrium when

T dθ = τ0 b dl

I the stress τ0 required to bend the dislocation to a radius R is then

αGb
τ0 = ,
R
where we substituted for T and used dθ = dl/R
I recall Frank-Read source
I expression for τ0 is strictly valid only if edge, screw and mixed
segments have the same energy per unit length (i.e. ν = 0)
I if ν 6= 0, the line experiences a torque trying to rotate it towards the
screw segment, which has lower energy per unit length
I the true line tension of a mixed segment

(mixed) d2 Eel (θ)


T = Eel (θ) +
dθ2
Forces between dislocations
I consider two parallel edge dislocations
I the total elastic energy per unit length of dislocations far apart

αGb2 + αGb2
I if the dislocations are very close
– approximated by a single
dislocation with Burgers vector
magnitude 2b – the elastic
energy would be

αG(2b)2

which is twice the energy of


dislocations further apart
I dislocations of the same sign
repel each other
I similarly, dislocations with Edge dislocations with parallel Burgers vectors a)
same signs, same slip plane, b) opposite signs,
opposite signs will attract each same slip plane, c) opposite signs, neighbouring slip
other → annihilation planes [6].
Forces between dislocations
I consider two edge dislocations along z axis with parallel Burgers
vectors on different slip planes, i.e.
I ~b1 = (b1 , 0, 0) I II
~l1 = (0, 0, 1) ~F = (σ · b2 ) ×~l2
~
II ~b2 = (b2 , 0, 0)
| {z }
~g
~l2 = (0, 0, 1)

gx = σxx bx + σxy by + σxz bz = σxx b2


gy = σyx bx + σyy by + σyz bz = σyx b2
gz = σzx bx + σzy by + σzz bz = 0
~g = (σxx b2 , σyx b2 , 0) Forces on two edge dislocations [6]
I interaction force on dislocation II from dislocation I

~F = ~g × ~l2 = (σyx b2 , −σxx b2 , 0)

where σxy and σxx are the stresses of dislocation I evaluated at position
of dislocation II
Forces between dislocations

~F = ~g × ~l2 = (σyx b2 , −σxx b2 , 0)

I if dislocations have opposite signs, force is reversed


I equal and opposite forces act on dislocation I
I substituting for σxy and σxx

Gb1 b2 x(x2 − y2 ) Gb1 b2 y(3x2 + y2 )


Fx = Fy =
2π(1 − ν) (x2 + y2 )2 2π(1 − ν) (x2 + y2 )2
I the most important component is Fx (in the slip plane)
Forces between dislocations

Fx nature x range
negative repulsive −∞ < x < −y
positive attractive −y < x < 0
negative attractive 0<x<y
positive repulsive y<x<∞

Stable positions of two edge dislocations with a) the same and b) opposite Burgers vectors [6]
Forces between dislocations

Glide force per unit length between edge dislocations of parallel Burgers vectors. Full curve – the same
Burgers vectors, dashed – opposite Burgers vectors. Force units Gb2 /[2π(1 − ν)y] [6]
Forces between dislocations
I consider two parallel screw dislocations lying along z axis
I ~b1 = (0, 0, b1 )
~l1 = (0, 0, 1) gx = σxx bx + σxy by + σxz bz = σxz b2
gy = σyx bx + σyy by + σyz bz = σyz b2
II ~b2 = (0, 0, b2 )
~l2 = (0, 0, 1) gz = σzx bx + σzy by + σzz bz = 0
I II ~g = (σxz b2 , σyz b2 , 0)
~F = (σ · ~b2 ) ×~l2
| {z }
~g

I interaction force on dislocation II from dislocation I


~F = ~g × ~l2 = (σyz b2 , −σxz b2 , 0)

I substituting for σyz and σxz

Gb1 b2 x Gb1 b2 Gb1 b2 y Gb1 b2


Fx = = cos θ Fy = = sin θ
2π x2 + y2 2πr 2π x2 + y2 2πr

I repulsive for screws of the same sign and attractive for opposite
signs
Dislocation configurations
Dislocation wall
I stable configuration of straight edge dislocations
I strong long-range stress field
I due to superposition of compressive and tensile
stress fields → screening radius R (in the term
ln rR0 ) decreases to h ⇒ energy of a dislocation in a
dislocation wall can be several times smaller than
that of an isolated dislocation
I dislocation walls form during recovery (stored
internal energy due to plastic deformation
decreases while dislocations form low-energy
configurations)
I dislocation wall corresponds to a low-angle tilt
grain boundary – rearrangement of dislocations Dislocation wall [4]
into low-angle grain boundaries can lead to cellular
subgrain structure
Dislocation configurations

Pile-up of straight edge dislocations of opposite Dislocation dipoles.


signs. Weak long-range stress field [4] Weak long-range stress
field [4]

Dislocation pile-up of like dislocations in


front of an obstacle. These pile-ups are
unstable configurations and can be
formed during initial stages of plastic
deformation. At large distances from the
pile-up, the stress field is analogous to a
Chessboard structure. Weak long-range stress super-dislocation of ~ B = n~b [4]
field [4]
Image forces
I a dislocation near a surface experiences forces not present in bulk
I dislocations attracted to free surfaces – material more compliant –
reduction of dislocation energy × repelled by rigid surfaces
I normal and shear stress components at a free surface are zero
(“nothing” on one side cannot provide reaction forces)
I screw dislocation parallel to z axis, distance d from a free surface x = 0
I the condition of zero stresses at x = 0, i.e. σxx = σyx = σzx = 0, stress
field of an imaginary screw dislocation at x = −d of opposite sign added
I stress in the material
 
Gb y y
σzx = −
2π (x + d)2 + y2 (x − d)2 + y2
 
Gb x+d x−d
σzy = − −
2π (x + d)2 + y2 (x − d)2 + y2

Image screw dislocation [6]


Image forces
I a dislocation near a surface experiences forces not present in bulk
I dislocations attracted to free surfaces – material more compliant –
reduction of dislocation energy × repelled by rigid surfaces
I normal and shear stress components at a free surface are zero
(“nothing” on one side cannot provide reaction forces)
I screw dislocation parallel to z axis, distance d from a free surface x = 0
I the condition of zero stresses at x = 0, i.e. σxx = σyx = σzx = 0, stress
field of an imaginary screw dislocation at x = −d of opposite sign added
I stress in the material
 
Gb y y
σzx = −
2π (x + d)2 + y2 (x − d)2 + y2
 
Gb x+d x−d
σzy = − −
2π (x + d)2 + y2 (x − d)2 + y2

image real Image screw dislocation [6]


Image forces

I force per unit length acting on the screw dislocation from the surface
(from the image dislocation), evaluated at x = d, y = 0

Gb2
Fx = σzy b = −
4πd

I edge dislocation parallel to z axis, distance


d from a free surface x = 0
I adding an image edge dislocation at
x = −d cancels σxx but not σyx → an extra
term needs to be added to satisfy the
boundary condition
Image edge dislocation [6]
Image forces
I shear stress is then image real

Gb2 (x + d)2 − y2 (x − d)2 − y2



σyx = − (x + d) 2 2 2
− (x − d)
2π(1 − ν) ((x + d) + y ) ((x − d)2 − y2 )2
2d(x − d)(x + d)3 − 6x(x + d)y2 + y4

+
((x + d)2 + y2 )3

extra term to ensure σyx = 0 at x = 0

I force acting on the edge dislocation from the surface (from the image
dislocation and the extra term), evaluated at x = d, y = 0

Gb2
Fx = σyx b = −
4π(1 − ν)d

I contribution from the 3rd (extra) term is zero


I interactions of curved dislocations, loops, dipoles, etc. with free
surfaces can result in very complex stress fields
Questions

1. Does the concept of dislocations explain these experimental


observations? Why?
I The yield stress of a material is lower than that calculated from
assuming a perfect lattice.
I In some cases, the stress required to continue plastic deformation
increases as deformation proceeds.
I The yield stress of most metals decreases as the temperature
increases.
I Ceramics tend to be brittle, whereas metals tend to be ductile.
2. Consider a screw dislocation in a BCC metal having the Burgers
vector of 2a [111]. On which planes can this dislocation move? What
would the slip plane be if the dislocation had an edge character?
References
[1] R. Abbaschian, L. Abbaschian, and R. E. Reed-Hill. Physical
Metallurgy Principles. Cengage Learning, 2009.
[2] W. D. Callister and D. G. Rethwisch. Materials Science and
Engineering. Wiley, 2010.
[3] J. F. Shackelford. Introduction to Materials Science for Engineers.
Pearson, 2015.
[4] Leonid Zhigilei. Mse 6020: Defects and microstructure in materials.
http://people.virginia.edu/˜lz2n/mse6020/. University of
Virginia.
[5] D. R. Askeland, P. P. Fulay, and W. J. Wright. The Science and
Engineering of Materials. Cengage Learning, 2010.
[6] D. Hull and D. Bacon. Introduction to Dislocations.
Butterworth-Heinemann, 2001.
[7] Helmut Föll. Defects in crystals. https://www.tf.uni-kiel.de/
matwis/amat/def_en/kap_5/illustr/i5_2_1.html.
University of Kiel.

You might also like