Vous êtes sur la page 1sur 27

ME 831: Fracture Mechanics

5
Fracture
Mechanisms
in Metals

Dr. Atta ur Rehman Shah

atta.shah@hitecuni.edu.pk
Assistant Professor
Department of Mechanical Engineering
HITEC University, Taxila - Pakistan
1
INTRODUCTION
The following figure schematically illustrates three of the most
common fracture mechanisms in metals and alloys.
Ductile materials usually fail as the result of nucleation, growth,
and the coalescence of microscopic voids that initiate at
inclusions and second-phase particles.
Cleavage fracture involves
separation along specific
crystallographic planes.
Note that the fracture
path is transgranular.

Intergranular fracture as
its name implies, occurs
when the grain boundaries
are the preferred fracture
path in the material.
2
DUCTILE FRACTURE
The figure schematically illustrates the uniaxial tensile behavior
in a ductile metal.

In high purity materials, specimens fracture after nearly 100%


reduction in area. Materials that contain impurities, however, fail
at much lower strains.
3
DUCTILE FRACTURE
The commonly observed stages in ductile fracture are as follows:
1. Formation of a free surface at an inclusion or second-
phase particle by either interface decohesion or particle
cracking.
2. Growth of the void around the particle, by means of
plastic strain and hydrostatic stress.
3. Coalescence of the growing void with adjacent voids.

In materials where the second-phase particles and inclusions are


well-bonded to the matrix, void nucleation is often the critical
step; fracture occurs soon after the voids form.
When void nucleation occurs with little difficulty, the fracture
properties are controlled by the growth and coalescence of voids;
the growing voids reach a critical size, relative to their spacing,
and a local plastic instability develops between voids, resulting in
failure.
4
DUCTILE FRACTURE
VOID NUCLEATION
A void forms around a second-phase particle or inclusion when
sufficient stress is applied to break the interfacial bonds between
the particle and the matrix.

The most widely used continuum model for estimating void


nucleation stress is due to Argon et al. They argued that the
interfacial stress at a cylindrical particle is approximately equal to
the sum of the mean (hydrostatic) stress and the effective (von
Mises) stress.
c  e m
where σe is the effective stress, given by
e 
1
2

 1   2    2   3    3   1 
2 2

2 12

1   2   3
σm is the mean stress, defined as m 
3
σ1 , σ2 and σ3 are the principal normal stresses. 5
VOID NUCLEATION
The Beremin research group applied the Argon et al. criterion to
experimental data for a carbon manganese steel, but found that
the following semiempirical relationship gave better predictions
of void nucleation at MnS inclusions that were elongated in the
rolling direction:

 c   m  C  e   YS 
where σYS is the yield strength and C is a fitting parameter that is
approximately 1.6 for longitudinal loading and 0.6 for loading
transverse to the rolling direction.

6
VOID NUCLEATION
Goods and Brown have developed a dislocation model for void
nucleation at submicron particles. They estimated that
dislocations near the particle elevate the stress at the interface
by the following amount:
α = constant that ranges from 0.14 to 0.33

 1b
µ = shear modulus
 d  0.5 ε1 = maximum remote normal strain
b = magnitude of Burger’s vector
r r = particle radius

Void nucleation occurs when the sum of the maximum principal


stress plus ∆σd reaches a critical value:
 c   d   1
The Goods and Brown dislocation model indicates that the local stress
concentration increases with decreasing particle size; void nucleation is more
difficult with larger particles.
The continuum models, which apply to particles with r > 1 µm, imply that σc is
independent of particle size. Experimental observations usually differ from
both continuum and dislocation models 7
VOID GROWTH AND COALESCENCE
Once voids form, further plastic strain and hydrostatic stress
cause the voids to grow and eventually coalesce.
The following figures are scanning electron microscope (SEM)
fractographs that show dimpled fracture surfaces that are
typical of microvoid coalescence.

Scanning electron microscope (SEM) High magnification fractograph of the steel


fractograph which shows ductile fracture in a ductile fracture surface. Note the spherical
low carbon steel. Photograph courtesy of Mr. inclusion which nucleated a microvoid.
Sun Yongqi. Photograph courtesy of Mr. Sun Yongqi. 8
VOID GROWTH AND COALESCENCE
The following figure schematically illustrates the growth and
coalescence of microvoids.
If the initial volume fraction of voids is low (<10%), each
void can be assumed to grow independently; upon further
growth, neighboring voids interact.

Void nucleation, growth, and coalescence in ductile metals: (a) inclusions in a ductile matrix, (b) void nucleation, (c)
void growth, (d) strain localization between voids, (e) necking between voids, and (f) void coalescence and fracture. 9
VOID GROWTH AND COALESCENCE
‘‘Cup and cone’’ fracture surface is commonly observed in
uniaxial tensile tests.
The neck produces a triaxial stress state in the center of the
specimen, which promotes void nucleation, growth and coalesce,
resulting in a penny-shaped flaw.
The penny-shaped flaw produces deformation bands at 45° from
the tensile axis, which may be referred to as “shear fracture.”
The 45° angle between the fracture plane and the applied stress
results in a combined Mode I/Mode II loading.

10
VOID GROWTH AND COALESCENCE
The outer surface of the specimen contains relatively few voids,
because the hydrostatic stress is lower than in the center.
The central portion of the
specimen exhibits a typical
dimpled appearance, but the At somewhat higher
outer region appears to be magnification, a few widely
relatively smooth, particularly at spaced voids are evident in the
low magnification outer region.

Cup and cone fracture in an austenitic stainless steel. Photographs courtesy of P.T. Purtscher. Taken from Purtscher, P.T.,
‘‘Micromechanisms of Ductile Fracture and Fracture Toughness in a High Strength Austenitic Stainless Steel.” Ph.D.
Dissertation, Colorado School of Mines, Golden, CO, 1990. 11
VOID GROWTH AND COALESCENCE
Rice and Tracey considered a single
void in an infinite solid, subject to
remote normal stresses σ1, σ2, σ3, and
remote normal strain rates 𝜀1ሶ , 𝜀2ሶ , 𝜀3ሶ .

The initial void is assumed to be


spherical, but it becomes ellipsoidal as
it deforms.
They showed for incompressible conditions:

R   eq
P
 1.5 m  P
ln   0.283 exp d eq
 R0  0
  YS 
where 𝑅ത = 𝑅1 + 𝑅2 + 𝑅3 Τ3 , Ro is the radius of the initial spherical void.
𝑃
𝜀𝑒𝑞 is the equivalent (von Mises) plastic strain.

This model is based on a single void, it does not take account of


interactions between voids, nor does it predict ultimate failure. A
separate failure criterion must be applied to characterize microvoid
coalescence. 12
VOID GROWTH AND COALESCENCE
Thomason developed a simple limit
load model for internal necking
between microvoids.
This model states that failure occurs
when the net section stress between
voids reaches a critical value σn(c) .
d
The row of voids illustrated  n c   1
in figure is stable if d b
d
and fracture occurs when  n c   1
d b
The Thomason model is of limited practical value. The void
interactions leading to ductile failure are far too complex to be
captured by a simple area reduction model.

Once the void fraction reaches 10 to 20%, failure occurs with


only a minimal increase in the nominal strain. 13
DUCTILE CRACK GROWTH
As the cracked structure is loaded, local strains and stresses at
the crack tip become sufficient to nucleate voids.

These voids grow as the crack blunts, and they eventually link
with the main crack. As this process continues, the crack grows.

Mechanism for ductile crack growth: (a) initial state, (b) void growth
at the crack tip, and, (c) coalescence of voids with the crack tip

14
DUCTILE CRACK GROWTH
When an edge crack in a plate grows
by microvoid coalescence, the crack
exhibits a tunneling effect, where it
grows faster at the center of the plate,
due to the higher stress triaxiality.

The through-thickness variation of


triaxiality also produces shear lips,
where the crack growth near the free
surface occurs at a 45° angle from the
maximum principal stress

The shear lips are very similar to the


cup and cone features in fractured
tensile specimens.

15
DUCTILE CRACK GROWTH
For a crack subject to plane strain Mode
I loading, the maximum plastic strain
occurs at 45° from the crack plane, this
angle is the preferred path for void
coalescence

But global constraints require that the


crack propagation remain in its
original plane. One way to reconcile
these competing requirements is for
the crack to grow in a zigzag pattern.

The zigzag crack appears flat on a


global scale, but oriented ±45° from
the crack propagation direction when
viewed at higher magnification.

16
DUCTILE CRACK GROWTH
The following figure shows a metallographic cross-section of a
growing crack that exhibits zigzag behavior.

Optical micrograph (unetched) of ductile crack growth in an A 710 high-strength low-alloy


steel. Photograph courtesy of J.P. Gudas. Taken from McMeeking, R.M. and Parks, D.M.,
‘‘On Criteria for J-Dominance of Crack-Tip Fields in Large-Scale Yielding.” ASTM STP 668,
American Society for Testing and Materials, Philadelphia, PA, 1979, pp. 175–194.
17
DUCTILE CRACK GROWTH

18
CLEAVAGE
Cleavage fracture can be defined as the
rapid propagation of a crack along a
particular crystallographic plane.

Cleavage may be brittle, but it can be


preceded by large-scale plastic flow and
ductile crack growth

The preferred cleavage planes are those with the lowest packing
density, since fewer bonds must be broken and the spacing
between planes is greater.

The propagating crack changes direction each time it crosses a


grain boundary; the nominal orientation of the cleavage crack is
perpendicular to the maximum principal stress.

19
CLEAVAGE
FRACTOGRAPHY
Fractography is used to
determine the causes of
fracture in engineering
materials.

The multifaceted surface is


typical of cleavage in a
polycrystalline material; each
facet corresponds to a single
grain.

As the multiple cleavage cracks


propagate, they are joined by
tearing between planes; there is
a tendency for the multiple
cracks to converge into a single
crack. These markings are
called ‘‘river patterns’’.
20
FRACTOGRAPHY

River
Patterns

River patterns in an A 508 Class 3 steel. Note the tearing


(light areas) between parallel cleavage planes.
Photograph courtesy of Mr. Sun Yongqi.

21
MECHANISMS OF CLEAVAGE INITIATION
Since cleavage involves breaking bonds, the local stress must be
sufficient to overcome the cohesive strength of the material
(Approximately E/π).

Generally the maximum stress achieved ahead of a blunted crack


tip is not sufficient to exceed the bond strength of a material.

In order for cleavage to initiate, there must be a local discontinuity


ahead of the macroscopic crack that is sufficient to exceed the
bond strength.

A sharp microcrack is one way to provide sufficient local stress


concentration.

22
THE DUCTILE-BRITTLE TRANSITION
The fracture toughness of ferritic steels can change drastically
over a small temperature range, as illustrated by the Figure.

At low temperatures, steel


is brittle and fails by
cleavage. At high
temperatures, the material
is ductile and fails by
microvoid coalescence.

In the transition region between


ductile and brittle behavior,
both micromechanisms of
fracture can occur in the same
specimen.

23
THE DUCTILE-BRITTLE TRANSITION
Heerens and Read performed a large number of fracture toughness
Tests on a quenched and tempered alloy steel at several temperatures
in the transition region, and found scattered data.
They examined the fracture surface of each specimen to determine the
site of cleavage initiation.

In specimens that exhibited


low toughness, the distance
(rc) was small; a critical
nucleus was available near
the crack tip. In specimens
that exhibited high
toughness, there were no
critical particles near the
crack tip; the crack had to
grow and sample additional
material before a critical
cleavage nucleus was
found.
(rc)
24
THE DUCTILE-BRITTLE TRANSITION
Cleavage propagation in the upper transition region often displays
isolated islands of ductile fracture.
A propagating cleavage crack in
the upper transition region
encounters barriers, such as
highly misoriented grains or
particles, through which the crack
cannot propagate, and unbroken
ligaments are discovered behind
the arrested crack tip.

The concentration of ductile ligaments on a fracture surface


increases with temperature, which may explain why crack-arrest
toughness (KIa) exhibits a steep brittle-ductile transition, much like
KIc and Jc.

25
INTERGRANULAR FRACTURE
Under some special circumstances, cracks
can form and propagate along grain
boundaries, which is called intergranular
fracture.

The situations that can lead to cracking on grain boundaries,


include:
1. Precipitation of a brittle phase on the grain boundary.
2. Environmental assisted cracking.
3. Intergranular corrosion.
4. Grain boundary cavitation and cracking at high
temperatures.

26
INTERGRANULAR FRACTURE
The following Figure shows an intergranular fracture surface in a
steel weld that was in contact with an ammonia environment.

Photograph courtesy of W.L. Bradley.


27

Vous aimerez peut-être aussi