You are on page 1of 15

10/18/2008

Lecturer : Assoc. Prof. Dr. Tich Thien TRUONG

Failure Mechanics
1 Elasto-Plastic
1. El t Pl ti F Fracture
t M
Mechanics
h i (EPFM)

2. Fatigue

3. Creep

1
10/18/2008

Elasto‐Plastic Fracture Mechanics (EPFM)
As it was discussed, due to finite strength of materials, there is always a small
damaged zone around the crack tip. For metals, this damaged zone is referred
to as the crack tip plastic zone. If the size of the plastic zone is small enough
that it can be contained within the K-dominant region, we may use K and G as
th LEFM parameters.
the t Thi condition
This diti i also
is l referred
f d to
t as the
th small-scale-
ll l
yielding condition (SSY).
On the other hand, if this zone is larger than the K-dominant region, then our
linear elastic assumptions are not correct, i.e., LEFM is not applicable and
nonlinear models must be used. Figure 1 shows three different situations
regarding the spread of crack tip plastic zone. The first one represents the
SSY condition. The second one shows the situation when the crack tip plastic
zone is large enough to cause some nonlinearity in the overall response of the
component. t However,
H if this
thi nonlinearity
li it is
i nott very significant,
i ifi t it can be
b
handled with a non-linear elastic model, for which we will introduce a non-
linear-elastic energy release rate called J, usually known as the J-Integral.
However, we should note that, similar to the LEFM, there is a limit to the
validity of J with regard to the size of the plastic zone compared to the J-
dominant region.

Elasto‐Plastic Fracture Mechanics (EPFM)

Figure 1: From left a) Linear Elastic,


Elastic b) Elastic-Plastic,
Elastic Plastic c) Fully Plastic
Plastic, d) Overall Plasticity.
Plasticity

For situations where the crack tip plasticity is so wide-spread that even plastic
ligaments may form within the component, we will show that the appropriate
parameter would be the crack-tip opening displacement (CTOD). Finally, when
the loading causes overall plastic deformation, even in presence of cracks, the
failure mode would be plastic collapse not fracture.

2
10/18/2008

Elasto‐Plastic Fracture Mechanics (EPFM)
J‐Integral
The J contour integral is extensively used in fracture mechanics as an energy-
based criterion for determining the onset of crack growth. However, it can also
be used as a stress based criterion. Referring g to Figure
g 2,, the original
g form of
the J- Integral for a line contour surrounding the crack tip can be written as:
⎛ ∂u ⎞
J = ∫ ⎜ wdy − Ti i ⎟ ds
Γ⎝
∂x ⎠
in which, w=∫σijdεij is the strain energy
density with σij and εij as stress and
strain tensors, Ti=σijnj are the
components of the traction vector which
acts on the contour, ui are the
displacement components, and ds is a
length increment along the contour Г.

Figure 2: An arbitrary contour around the


crack tip

Elasto‐Plastic Fracture Mechanics (EPFM)
J‐Integral
At first, the above integral might look unfamiliar and rather strange. However, it
should be noted that J- Integral is nothing but a non-linear energy release rate
defined by: y dΠ
d Π
J =− , Π =U−V
dA
in which П is the total potential energy, U is the strain energy, and V is the
external work.
Characteristics of the J‐Integral
In this section we will explore some interesting
characteristics of the J-Integral. For instance we will
show that the J-integral is zero over a closed path
and, as a result, it remains independent of the path
considered for its evaluation around the crack tip.
Referring to Fig. 5, we may define the J-Integral as an
area Integral:
⎡ ∂w ∂ ⎛ ∂ui ⎞ ⎤ Figure 5
J= ∫ ⎢⎢ ∂x − ∂x ⎜ σ ij
j ⎝
⎟ ⎥ dA
∂x ⎠ ⎥⎦
A* ⎣

3
10/18/2008

Elasto‐Plastic Fracture Mechanics (EPFM)
J‐Integral
We conclude that the J-Integral is zero
for a closed path. Now, if we consider a
closed path like the one shown in Fig.
Fig 6,
6
we may write:
J = J1 + J 2 + J 3 + J 4 = 0
As the crack surfaces are traction free and
perpendicular to the y axis, we may write:
Ti = dy = 0 → J 3 = J 4 = 0 → J1 = − J 2
which shows that J is p
path-independent.
p

Figure 6

Elasto‐Plastic Fracture Mechanics (EPFM)
J‐Integral
J as a stress intensity parameter
For an elastic-plastic material, the Ramberg-Osgood equation can be written as:
n
ε σ ⎛σ ⎞
= +α ⎜ ⎟
ε0 σ 0 ⎝ σ0 ⎠
in which σ0 and ε0 are the yield stress and yield strain respectively, and n is the
strain hardening coefficient.
For a linear elastic material n = 1. The following relations, known as HRR
singularities (due to Hutchinson, Rice, and Rosengren) show that J can in fact
represent the strength of the stress and strain distributions around the crack tip.
1 n
⎛ EJ ⎞ n +1 ασ 0 ⎛ EJ ⎞ n +1
σ ij = σ 0 ⎜ 2 ⎟ σ% ij (n,θ ) , ε ij = ⎜ ⎟ ε%ij (n, θ )
⎝ ασ 0 I n r ⎠ E ⎝ ασ 02 I n r ⎠
in which, In is an integration constant that depends on n, and σ% ij (n,θ) andε%ij(n,θ)
are geometric expressions.

4
10/18/2008

Fatigue
It has been known for a long time that a component subjected to fluctuating
stresses may fail at stress levels much lower than its monotonic fracture
strength. The underlying failure process involves a gradual cracking of the
component and is called Fatigue. Fatigue is an insidious time-dependent type
off failure
f il which
hi h can occur without
ith t any obvious
b i warning.
i It is
i believed
b li d that
th t
more than 95 percent of all mechanical failures can be attributed to fatigue.
There are normally three distinct stages in the fatigue failure of a component,
namely: Crack Initiation, Incremental Crack Growth, and Final Fracture.
Fatigue crack initiation usually occurs at free surfaces, because of the higher
stresses and the higher probability of the existence of defects at these
locations (existence of corroded or eroded areas, scratches, etc.).
Nevertheless, even at highly-polished defect-free surfaces, fatigue cracks can
initiate through repeated microplastic deformations which result in the
formation of the so called “intrusions” and “extrusions” on the surface. The
former can act as local stress concentration sites which may eventually lead
to the formation of microcracks (see Fig. 1).

Fatigue

Figure 1. Schematics of fatigue crack initiation


Fatigue crack propagation occurs through repeated crack tip blunting and
sharpening
h i effects
ff which
hi h are ini turn caused d by
b microplastic
i l i deformation
d f i
mechanisms operating at the crack tip (see Fig.2).

Figure 2. Schematics of fatigue crack propagation.

5
10/18/2008

Fatigue
A macroscopic examination of fatigue failures reveals several distinct fracture
surface markings. In general, the fracture surface is flat with no sign of
significant plastic deformation, except for the portion related to the final rupture.
For fatigue failures which occur over a long period of time, the fracture surface
may contain characteristic markings which are called “beach beach markings
markings” or
“clam shell markings”. These markings, which are recognizable even by naked
eye (see Fig.3, left), reflect the occurrence of different periods of crack growth.
On the other hand, there are extremely fine markings called “striations”, which
represent the crack growth due to individual loading cycles and can only be
seen at very high magnifications using electron microscopes (see Fig.3, right).

Figure 3. Left: fracture appearance of different stages of fatigue failure, including


the beach markings. Right: typical striations which form during the growth period.

Fatigue
Fatigue problems in engineering design are treated by three different
approaches as briefly described in sequel.
Classical Fatigue Approach
The classical approach to fatigue, also referred to as Stress Controlled Fatigue
or High
Hi h Cycle
C l Fatigue
F ti (HCF) through
(HCF), h h S/N or Wöhler
Wöhl diagrams,
di constitutes
i the
h
basis of the SAFE LIFE philosophy in design against fatigue. In order to
determine the strength of materials under the action of fatigue loads, specimens
with polished surfaces are subjected to repeated or varying loads of specified
magnitude while the stress reversals are counted up to the destruction point,
see Fig.4.

Figure 4: Rotating-Bending fatigue testing machine

6
10/18/2008

Fatigue
Classical Fatigue Approach
The number of the stress cycles to failure can be approximated by the
WOHLER or S-N DIAGRAM, a typical example of which is given in Fig. 5.

Figure 5: A typical S-N diagram

Fatigue
Classical Fatigue Approach
In HCF terminology, the following notions can be defined:
The Range of the stress cycle which will cause failure in (N) repetitions can
be defined as ∆S=Smax −Smin
S − S mini
Sa = the alternating stress, which is 1/2 the Range of Stress: S a = max
2
N = the Fatigue Life
Su = the Ultimate Tensile Strength of the material
SN = the Fatigue Strength to N Cycles
Sf = the Fatigue or Endurance Limit of the material, corresponding to the
median fatigue strength as the fatigue life becomes very large.
Other important definitions are:
S min
Load Ratio, R =
S max
S max + Smin
Mean or Static Stress, S mean =
2

7
10/18/2008

Fatigue
Classical Fatigue Approach
The original Wohler diagram defines the fatigue failure surface when the
Smean is zero and no fatigue strength reducing factors are involved. It is
usually constructed on either an arithmetic-logarithmic or a logarithmic-
logarithmic scale.
scale Attempts have been made to express the shape of the S-N
diagram in mathematical form, one simple form of these equations is:
log S N = − A log N + B
103 < N < 106
1 ⎛ S ⎞
A = log ⎜ 0.9 u ⎟⎟
3 ⎜ S
⎝ f ⎠
⎡ ( 0.9 Su )2 ⎤
B = log ⎢ ⎥
⎢⎣ S f ⎥⎦

Fatigue
Classical Fatigue Approach
Real components differ markedly from the laboratory specimens usually used
for generating the S-N Diagrams. Hence, the fatigue strength S-N curve, shown
in Fig. 5 for zero mean stress, should be adjusted for the effects of various
modifying factors,
factors
σ N = M f × SN
where Mf is the product of several fatigue strength modifying factors and may
be defined as:
M f = msur × msize × mrel × mload × mtemp × mconc × mmisc
These factors are attributed to the followings:
msurf = surface finish
msize = size
 size
mrel = reliability
mload = load
mtemp = temperature
mconc = stress concentration
mmisc = miscellaneous effects

8
10/18/2008

Fatigue
Low Cycle Fatigue Approach
Based on the LCF LOCAL STRAIN PHILOSOPHY, fatigue cracks initiate as a
result of repeated plastic strain cycling at the locations of maximum strain
concentration. It is also assumed that the most highly strained region can be
represented by a filament of material whose mechanical response is similar to
that of a smooth specimen. The basic assumptions of the LCF approach can
be summarized as:
a) The fatigue crack initiation of a notched member can be considered to occur
by the rupture of a filament of the material located nearest the surface in the
vicinity of the stress raiser.
b) Under appropriate control, a smooth
specimen can be used to reproduce the
stressstrain history of the filament (see Fig.6).
c) For identical stress-strain histories of the
filament and smooth specimen, the fatigue life
of the smooth specimen can be taken as the
fatigue life of the filament i.e., the crack
initiation life (see Fig.6) .
Figure 6: Schematic representation
of crack initiation according to LCF

Fatigue
Low Cycle Fatigue Approach
In LCF, the terminology crack initiation is used in the sense of the number of
fatigue cycles required to either fail the smooth specimen or the filament.
The necessaryy requirements
q for a LCF life assessment p program
g can be
summarized as below:
1. A mechanics analysis for the determination of the stress-strain behavior at the
critical location (notch).
2. A knowledge of the cyclic stress-strain
properties of the material to determine the
response of the material at the notch to
remotely applied stresses.
3. A knowledge of the low cycle fatigue
properties of the material for use in an
appropriate cumulative damage assessment
procedure (see Fig.7).
4. A cumulative fatigue damage rule to Figure 7: Schematic representation
accurately predict the LCF life for an arbitrary of cyclic strain-life curves
loading sequence

9
10/18/2008

Fatigue
Low Cycle Fatigue Approach
5. A method of combining 1-4 such that the LCF initiation life of a notched
member subjected to any arbitrary loading sequence can be calculated on a
reversal by reversal basis using computer simulation methods (see Fig.8).

Figure 8: Schematic representation of LCF life assessment activities

Fatigue
Fracture Mechanics Approach
If a crack exists in the component before it goes into service (for example due to
weld fabrication) the initiation stage is by-passed and the fatigue failure process
consists of the incremental crack growth and the final fracture stages. Crack
propagation normally occurs at right angles to the principal tensile stress direction.
In practice, however, most fatigue failures are in the low stress region (much less
than the yield stress) where the LEFM is likely to be valid. Hence, the LEFM
principles can be applied to predict incremental fatigue crack growth. In fact,
extensive fatigue tests on a wide variety of materials show that the stress intensity
factor is a much more effective parameter in describing fatigue propagation than
the stress amplitude. The key point of theses tests is that the rate of crack
propagation, measured in terms of incremental crack growth per cycle of loading,
depends primarily on the range of crack tip stress intensity,
intensity as follows:
da
= f (ΔK )
dN
The most widely used expression, proposed by Paris, is:
da
= C ( ΔK ) m
dN
in which C and m are material properties obtained from experiment.

10
10/18/2008

Fatigue
Fracture Mechanics Approach
The standard methods for fatigue crack growth tests can be found under
ASTM E647. The most commonly used specimen in fatigue crack propagation
studies is the C(T) or compact tension specimen (see Fig 9).

Figure 9

Fatigue
Fracture Mechanics Approach
Figure 11 shows a typical crack-growth-rate
versus stress-intensity-range diagram. Three
regions of different behavior can normally be
identified on such data presentations:
1. The threshold region, is attributed to very
low levels of ∆Ks, where the crack does not
propagate. The ‘threshold’ region is strongly
influenced by the mean stress.
2. The stable propagation region where the
crack grows incrementally according to the
Paris law
3 The final unstable region,
3. region where the crack
propagates more rapidly, often in a less
uniformly incremental manner. In the unstable
region, various mechanisms are responsible
for the increased growth rate.

Figure 11

11
10/18/2008

Fatigue
Fatigue Life Prediction
The useful aspect of fatigue crack growth laws is that they can be used to
calculate the number of cycles required to propagate a crack from a given
initial size to some final size which is critical for failure. Thus if the initial size is
ai and the final size af we may write:
da 1 ⎡⎣ a1f−m /2 − ai1− m /2 ⎤⎦
= C (ΔK ) ⇒ N =
m

dN C β m (Δσ ) m π m /2 1− m / 2
In the above equations, the geometric factor β is assumed to be constant, since
the inclusion of a function of a/W within the integral sign will usually lead to a
formulation which cannot be integrated analytically.
In practice, it is more straightforward and very often sufficiently accurate to solve
the fatigue life equation by splitting the crack growth history into a series of crack
increments. An average value within each step may then be used to calculate β
and hence an average K is considered during the step. The average propagation
rate within the step can then be calculated from the Paris Law. In the case of a
pressure vessel, af may simply be defined in terms of a crack big enough to
cause leakage, or one which results in the limiting fracture toughness being
reached. In sequel, we will briefly introduce two software tools commonly used
for fatigue crack growth studies.

Creep
Creep can be defined as a time-dependent deformation of materials under
constant load (stress). The resulting progressive deformation and the final
rupture, can be considered as two distinct, yet related, modes of failure. For
metals, creep becomes important at relatively high temperatures, i.e., above 0.3
off their
th i melting
lti point
i t in
i Kelvin
K l i scale.
l However,
H f polymers
for l substantial
b t ti l creep
can occur at room temperature. Creep tests are usually performed under
constant load, as shown in Fig. 17. The general procedures for these tests can
be found in ASTM E139-69.

Figure 17. Creep testing machine

12
10/18/2008

Creep
Figure 18 shows a typical creep curve which usually includes:
• initial elastic strain, which occurs immediately upon application of load
• primary creep, where the strain rate gradually decreases due to strain
hardening
• secondary or steady state creep, where the balance between strain
hardening and softening processes result in a constant creep rate
• tertiary creep, which includes material separation at micro level and leads to
final rupture.

Figure 18

Creep
As shown in Fig.19, the above mentioned processes can operate locally at the
tip of a pre-existing crack and lead to further crack growth.

Figure 19. Creep zones ahead of crack tip


It should be mentioned that for short-term
short term applications at high temperatures,
temperatures
one may use the pertinent time-independent mechanical properties such as
yield or tensile strength to define the design regime. However, for the long term
exposure to loadings at high temperatures, the design regime should be
adjusted according to the creep curves. The time-dependent constitutive
equations can be constructed using a variety of the combinations of the two
basic mechanical elements, i.e., the spring and the dashpot elements, which
represent linear-elastic and viscous behaviors respectively

13
10/18/2008

Creep
Creep Crack Growth
The structural components that are subjected to uniform loading and uniform
temperature distribution during service are vulnerable to widespread bulk
damage due to creep. On the other hand, for components that are subjected to
p
stress and temperature g
gradients it is likelyy that creep
p cracks initiate at critical
locations and propagate to cause failure.
Similar to fatigue crack growth modeling, we need appropriate crack-tip
parameters to correlate creep crack growth data. Depending on the material and
on the extent of creep deformation at the crack tip, various parameters have
been successfully used to correlate the rates of creep crack growth. In general,
three regimes of creep crack growth can be distinguished for materials
exhibiting power-law creep behavior, depending on the size of the crack-tip
creepp zone relative to the specimen
p dimensions.

Figure 20. From left: Small-Scale-Creep, Transitional Creep, and Steady-State-Creep conditions.

Creep
Creep Crack Growth
The two major parameters used for correlating creep crack growth data are the
stress intensity factor K and the integral C*. The time-dependent energy
Integral, C*, is similar to the J-Integral, but is written in terms of strain rates
instead of strain:
⎛ ∂u& ⎞
C* = ∫ ⎜ wdy
& − σ ij n j i ds ⎟
Γ⎝
∂x ⎠
ε&kl
&=
in which w ∫σ
0
ij d ε ij

The applicability of K is limited to situations where the size of the crack-tip creep
zone is small relative to the crack length and other geometric parameters of the
component.
p This is the so-called Small Scale condition ((SSC), ), as opposed
pp to
the Steady State condition (SS) in which the crack propagation is accompanied
by extensive creep deformation ahead of the crack tip. In the latter condition, the
pathindependent integral C* is usually used.

14
10/18/2008

Creep
Creep Crack Growth
The transition time for SSC condition to turn to SS condition can be estimated
by :
1 + 2β n K 2 (t1 )(1 −ν 2 )
t1 =
n +1 C * (t1 ) E
in which β is dependent on the waveform
of loading and is defined by:

K (t ) = K1t β
where K(t) is the applied stress intensity
parameter as a function of time and K1 is
a constant, ν is the Poisson’s ratio, E is
the elastic modulus, n is the Norton Law
exponent, and C* is the creep integral. If
the cycle time tC is less than t1 then K is
the correct crack tip parameter for
correlating creep crack growth.

Figure 21

15

You might also like