Fundamentals of Fatigue and Fracture Mechanics: RVC E
Fundamentals of Fatigue and Fracture Mechanics: RVC E
Author:
Benjamin R OHIT
A compilation of notes available on fatigue and fracture mechanics. Only for use at
the department of Aerospace Engineering, RVCE
“Thanks to my solid academic training, today I can write hundreds of words on virtually any
topic without possessing a shred of information, which is how I got a good job in journalism.”
Dave Barry
v
Acknowledgements
The acknowledgments and the people to thank go here, don’t forget to include your
project advisor. . .
vii
Contents
Acknowledgements v
3 • 29
4 Fatigue in Structures 31
4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
4.2 S-N curves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
4.2.1 Fatigue Ratio . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
4.3 Stress Life Approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
4.3.1 Mean Stress Effects on Fatigue Life . . . . . . . . . . . . . . . . . 33
General Observations . . . . . . . . . . . . . . . . . . . . . . . . 35
4.4 Strain Life Approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
Low cycle and high cycle fatigue lifes . . . . . . . . . . . . . . . 36
4.4.1 Cyclic Stress-Strain Behavior . . . . . . . . . . . . . . . . . . . . 37
Bauschinger Effect . . . . . . . . . . . . . . . . . . . . . . . . . . 37
Cyclic Strain Hardening and Softening . . . . . . . . . . . . . . 37
4.4.2 Mean Stress Effect . . . . . . . . . . . . . . . . . . . . . . . . . . 38
4.5 Stress Life vs Strain Life . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
4.6 Notched Stress - Neubers Stress Concentration . . . . . . . . . . . . . . 40
List of Figures
1.1 • . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2 a) Transcrystalline crack, b) intercrystalline crack . . . . . . . . . . . . . 5
1.3 Stress-Strain curves for Brittle and Ductile materials . . . . . . . . . . . 7
1.4 Stress-Strain curves for Brittle and Ductile materials . . . . . . . . . . . 7
1.5 Stages of crack growth in Ductile failure. . . . . . . . . . . . . . . . . . 8
1.6 True stress-strain curves at room and elevated temperatures(Image
Courtesy-ICME) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.7 Fracture behaviour of different materials with temperature, including
some that don’t have a transition zone.)) . . . . . . . . . . . . . . . . . . 10
1.8 Constant Displacement and Constant Load)) . . . . . . . . . . . . . . . 11
1.9 Load vs Displacement graphs for Constant Load and constant dis-
placement type crack growth)) . . . . . . . . . . . . . . . . . . . . . . . 11
1.10 Fracture Modes: Mode I is a normal-opening mode and is the one we
shall emphasize here, while modes II and III are shear sliding modes. . 12
1.11 crack growth rate as a function of stress intensity . . . . . . . . . . . . . 14
1.12 The damage tolerance approach to design. . . . . . . . . . . . . . . . . 15
2.1 Plate with centrally cracked hole and two crack edges . . . . . . . . . . 18
2.2 Griffiths energy graph . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.3 Energy Release Rate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
2.4 Load vs Displacement graph for constant displacement and constant
load . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.5 Compliance as a function of crack length . . . . . . . . . . . . . . . . . 23
2.6 Driving force & R curve as a function of crack length . . . . . . . . . . 23
2.7 R-curve for various crack lengths . . . . . . . . . . . . . . . . . . . . . . 24
2.8 K is measured as a function of specimen thickness . . . . . . . . . . . . 27
2.9 K is measured as a function of specimen thickness . . . . . . . . . . . . 27
4.1 • . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
4.2 • . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
4.3 • . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
4.4 • . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
4.5 • . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
4.6 • . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
4.7 • . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
4.8 • . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
4.9 Strain Hardening . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
4.10 Strain Softening . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
5.1 • . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
5.2 • . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
5.3 • . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
5.4 • . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
x
5.5 • . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
5.6 Time compressed representation of the load history . . . . . . . . . . . 47
5.7 • . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
5.8 Two different sequences of blocks in a simple VA load history (R = 0)
applied to a notched specimen . . . . . . . . . . . . . . . . . . . . . . . 48
xi
List of Tables
xiii
List of Abbreviations
Physical Constants
List of Symbols
a distance m
P power W (J s−1 )
ω angular frequency rad
xix
Chapter 1
Fundamentals of Fracture
Mechanics
1. Comet: Three comet air planes crashed in a short period of 12 months, leading
to the aircraft being grounded. The crashes were found to be caused by cracks
growing from corners of the square fuselage windows which served as stress
concentration points leading to accelerating crack formation.
2. Aloha Airlines Flight 243: The famous failure of the upper fuselage of the air-
craft is one of the landmark fatigue failures in aviation history. The failure was
the result of fatigue cracking of the skin panel adjacent to rivet holes at a lap
joint which was accompanied by corrosion.
2 Chapter 1. Fundamentals of Fracture Mechanics
3. Unexpected failures of ships and aircraft called for an expansion in the field of
fracture mechanics with the early era of research being limited to linear elastic
mechanics. In the present day research has advanced to transient crack growth
and crack behaviour even in exotic materials.
3. Prediction of the growth rate of cracks when they initiate from damage or fa-
tigue cycling.
If the structure and the operating environment could be perfect, then the struc-
ture would be designed to have no cracks during its operating life. However, given
the random variations in the real environment, it is necessary to deal with this three-
requirement problem using the concept of ’fail safe’ design, whose aim is not so
much to completely avoid the appearance of cracks due to defects in the material
itself, or manufacturing method, or to accidental damage in service., but to under-
stand the growth rate of the cracks. If this is achieved, the cracks can be found and
repaired before they become catastrophic.
The central difficulty in designing against fracture in high-strength materials is
that the presence of cracks can modify the local stresses to such an extent that the
elastic stress analyses done so carefully by the designers are insufficient. When a
crack reaches a certain critical length, it can propagate catastrophically through the
structure, even though the gross stress is much less than would normally cause yield
or failure in a tensile specimen . The term "fracture mechanics" refers to a vital spe-
cialization within solid mechanics in which the presence of a crack is assumed, and
we wish to find quantitative relations between the crack length, the material’s inher-
ent resistance to crack growth, and the stress at which the crack propagates at high
speed to cause structural failure.
The conventional approach to the design against fatigue failure (fatigue strength-
endurance limit approach) involves cyclic-life predictions based on nominal stress
(or strain) vs. elapsed-cycles data developed from endurance tests on smooth labora-
tory specimens. However, such data do not distinguish between the crack-initiation
and crack-growth phases of fatigue. Consequently, smooth-specimen endurance-
test fatigue data do not provide information regarding the effect of pre-existing flaws
on component life.
a b
F=− m
+ n (1.1)
r r
1.3. Micro Mechanics of Fracture 3
Here, the first term represents attractive forces, while the second term describes re-
pulsive ones. The parameters a, b, m, and n (m < n) are constants which depend on
the bond type.
F IGURE 1.1: •
For small displacements from the equilibrium configuration d0, the bonding
force F(r) can be approximated by a linear function. This is equivalent to a material
behaviour which macroscopically is described by Hooke’s law. During the release of
bonds, i.e., the separation of elements, a negative material specific work WB is done
by the bonding force.
we consider in a somewhat simplified manner the separation process of two
atomic lattice planes of a crystal. For the separation stress σ(x) we assume a de-
pendence on the separation displacement x similar to the bonding force.
Equating the latter with Hooke’s law
x
σ = E·e = E· (1.2)
d0
yields for the so-called theoretical strength, i.e., the cohesive stress that has to be
overcome during separation.
a
σc ≈ E · (1.3)
π · d0
If we further assume that the bonds are completely broken i.e
d0 ≈ a (1.4)
then we get,
E
σc ≈ (1.5)
π
This can only be true for defect free crystals which is an ideal condition in nature.
For real polycrystalline materials, the fracture strength is 2-3 orders of magni-
tude less. In parallel, the energy necessary for the creation of new fracture surfaces
exceeds the value defined (σc ≈ πE ) by several orders of magnitude. The reasons for
this can be found in the inhomogeneous structure of the material and, first of all, in
its defect structure.
4 Chapter 1. Fundamentals of Fracture Mechanics
Size effects
With respect to experiments the stress needed to fracture bulk glass is around 100
MPa but the theoretical stress needed for breaking atomic bonds of glass is approx-
imately 10,000MPa as per the equation 1.5. Also, experiments on glass fibres that
Griffith himself conducted suggested that the fracture stress increases as the fibre
diameter decreases. Hence the uni-axial tensile strength, which had been used exten-
sively to predict material failure before Griffith, could not be a specimen-independent
material property. Griffith suggested that the low fracture strength observed in ex-
periments, as well as the size-dependence of strength, was due to the presence of
microscopic flaws in the bulk material.
A dislocation pile-up causes not only stress concentrations. It also can be the
responsible source for the formation of microscopic voids and cavities.
Irwins Modification
In the 1950s, G.R. Irwin incorporated γ p &γs into a single term, Gc, known as the
critical strain energy release rate:
Gc = 2(γs − γ p ) (1.9)
The strain energy release rate or the crack extension force (G) is the change in
potential energy (U) of the system per unit increase in crack area (G = dU/da).
The original Griffith criterion can now be written for both brittle and ductile
materials as:
πσ2 a
Gc = (1.10)
E
πσ2 a
Therefore, crack extension occurs when exceeds the value of Gc for the
E
material in question.
1. Dislocations
3. Precipitates
1. Sharp notches
2. Surface scratches
3. Cracks
It should also be noted that materials may have ductile or brittle behavior, depend-
ing on the temperature, rate of loading and other variables
8 Chapter 1. Fundamentals of Fracture Mechanics
of atomic bonds along crystallographic planes and hence also called cleavage frac-
ture. This fracture is also said to be transgranular as the crack propagates through
grains. Thus it has a grainy and faceted texture when the cross section of failure is
viewed. Crack propagation across grain boundaries is called transgranular, while
propagation along grain boundaries is termed intergranular. Think of the metal
micro-structure as a 3-D puzzle. Transgranular fracture cuts through the puzzle
pieces whereas intergranular fracture propagates along the puzzle pieces precut
edges. Another important point to notice is that the surface of the brittle fracture
tends to be almost perpendicular to the direction of load(uniaxial) or perpendicular
to the direction of principal tensile stress. Brittle fracture usually occurs at stress lev-
els well below those predicted theoretically.
Properties of a brittle fracture.
Fig 1.6 shows the true stress-strain curves at room and elevated temperatures.
As temperature increases, the atoms within the material absorb energy and vi-
brate with greater frequency and amplitude. This additional energy allows the
atoms under stress to break bonds and form new bonds. This behaviour is seen
10 Chapter 1. Fundamentals of Fracture Mechanics
From the figure 1.7 it can be concluded that there is no single criterion that de-
fines the DBTT. With decreasing temperature, the yield strength increases rapidly to
the point where it equals the tensile stress for brittle failure, and below this temper-
ature, fracture usually occurs in brittle/cleavage mode, hence we can say below the
DBTT,σy = σ f
Let us now consider the crack growth under two different conditions.
12 Chapter 1. Fundamentals of Fracture Mechanics
1. Constant Load: Under a constant load, half the work during fracture goes into
elastic strain, and half goes into fracture. So we can say that the strain energy
increases. Because some amount of external work is being done.
By looking at the constant displacement graph in fig 1.9 we can say that the
strain energy decreases when the crack grows during constant displacement and the
strain energy increases when the grows under constant load. We can now see that
the elastic strain energy increases when work is done by increasing displacement at
constant load. So hence we can say that the work done can be the area AEFB, and the
amount by which strain energy has increased is AEA, by simple math we can note
that 2×OEA = AEFB, i.e, Under a constant load, half the work during fracture goes
into elastic strain, and half goes into fracture. A detailed explanation along with the
mathematics will be discussed in the next unit.
The stress intensity approach states that fracture occurs due to stress concentra-
tion at flaws, like surface scratches, voids, etc. If ’c’ is the length of the crack and
ρ the radius of curvature at crack tip, the enhanced stress (σm ) near the crack tip is
given by:
0.5
c
σmax = 2 · σnom (1.11)
ρ
1.12. Time dependent crack growth and damage tolerance 13
The above equation states that smaller the radius, higher is the stress enhancement.
Another parameter, often used to describe fracture toughness is known as critical
stress concentration factor, K, and is defined as follows for an infinitely wide plate
subjected to tensile stress perpendicular to crack faces: In traditional engineering
design such stress concentrations are denoted by the stress concentration factor kt
σmax
Kt = (1.12)
σnom
Do not confuse the stress concentration factor kt with the stress intensity factor
K or KIc . Typical situations where stress concentrations arise are changes of section
or point loads. For such situations the use of appropriate kt values in design is
essential, particularly in fatigue design.
For a sharp crack the radius of the tip is extremely small and hence we can say
that as σmax reaches infinity as ρ equals zero. This suggests that for a sharp crack,
any applied stress will cause infinitely high stresses at the tip. Also for very sharp
cracks, this approach cannot distinguish between long and short cracks whereas we
know that failure stress depends on crack length. Hence we can conclude that the
concept of stress concentration factor breaks down as crack tip radius tends to zero.
Hence we introduce a stress intensity factor which is more finite and often used to
describe fracture toughness denoted by K and is defined as follows for an infinitely
wide plate subjected to tensile stress perpendicular to crack faces.
√
K = σ cρ (1.13)
This relation holds for specific conditions, and here it is assumed that the plate
is of infinite width having√a through-thickness crack.It is worth noting that K has
the unusual units of MPa m. It is a material property in the same sense that yield
strength is a material property. The stress intensity factor K is a convenient way
of describing the stress distribution around a flaw. For the general case the stress
intensity factor is given by, √
K = ασ cπ (1.14)
where α is a parameter that depends on the specimen and crack sizes and ge-
ometries, as well as the manner of load application.
da
= C (∆K )m (1.15)
dN
where da/dN is the crack growth per cycle, ∆K is the stress-intensity range, and
C and m are material constants. In static loading, the stress intensity factor for a
small crack in a large specimen can be expressed as eqn 1.13. If the stress is kept
14 Chapter 1. Fundamentals of Fracture Mechanics
constant, we will get fracture for a certain crack length, a=ac , which will give KI =
KIc where ac is the critical crack length and KIc is the critical stress intensity factor
in mode I.
For a < ac (KI < KIc ) the crack will not propagate. In dynamic loading, however,
for (KI < KIc ) , the crack may still propagate. This means that a (and KI) will increase,
we will eventually obtain fracture when a = ac.
Chapter 2
2.1 Introduction
Linear Elastic Fracture Mechanics (LEFM) first assumes that the material is isotropic
and linear elastic. Based on the assumption, the stress field near the crack tip is cal-
culated using the theory of elasticity. When the stresses near the crack tip exceed the
material fracture toughness, the crack will grow. The concepts provide an analytical
method based upon the stress intensity factor, which characterises the stress distri-
bution in the vicinity of the crack tip, and are valid in design applications provided
that gross yielding does not occur.
Linear elastic fracture mechanics can be used to describe ultimate static failure of
low toughness high strength materials used in aerospace and other specialised ap-
plications.
Under fatigue loading, crack growth rates can be correlated by the stress intensity
factor for a wide range of materials, because even in the more ductile materials the
amount of plastic flow that can occur under fatigue loading is restricted. It should
not be used in cases where the fatigue cycles involve extensive plastic deformation,
often referred to as "low cycle" fatigue. In Linear Elastic Fracture Mechanics, most
formulas are derived for either plane stresses or plane straines, associated with the
three basic modes of loadings on a cracked body: opening, sliding, and tearing.
Again, LEFM is valid only when the inelastic deformation is small compared to the
size of the crack, what we called small-scale yielding. If large zones of plastic defor-
mation develop before the crack grows, Elastic Plastic Fracture Mechanics (EPFM)
must be used.
KI
a, B, (W − a) ≥ 2.5 (2.1)
σy
Where a is half the crack length, B is the thickness of the specimen and W is
the width of the specimen. These requirements are to assure that the plastic zone is
18 Chapter 2. LEFM - Linear Elastic Fracture Mechanics
sufficiently small (i.e. nearly plane strain conditions (B) and elastic conditions (W-a))
for the assumption of LEFM to be valid
F IGURE 2.1: Plate with centrally cracked hole and two crack edges
Griffith made the important connection in recognising that the driving force for
crack extension is the energy which can be released and that this is used up as the
energy required to create the two new surfaces. This thermodynamic description of
the fracture process has the huge advantage of removing attention from the small
area at the crack tip and the precise micro mechanism of fracture. Let us now con-
sider the plate with hole. The strain energy in the presence of the crack is given
by
σ2
Ua = Volume of Triangles × (2.2)
2E
σ2
2E is the stain energy within elastic limit i.e the area under the stress strain curve
upto elastic limit(Area of the Triangle). The two surfaces formed are the triangles
whose height is 2a λ where 2a is the crack length and λ is the proportionality factor.
For thin plates, λ = π/2,
2.2. Griffith’s Energy balance criterion 19
σ2
Ua = 2 · (0.5 × 2a × 2aλ) × (2.3)
2E
hence we get,
σ 2 · a2 π
Ua = (2.4)
E
which is the strain energy in the presence of a crack with two crack edges per unit
thickness of the plate. Now because the new surfaces have been formed, the total
energy of the system can be thought of as being the sum of the potential energy term,
U plus the surface energy of the crack, S, because no work is being done externally,
The surface energy required for thickness B is given by,
Us = 4aν (2.5)
where ν is the surface energy (e.g., Joules/meter2 ) and the factor 4 is needed
since two free surfaces have been formed at each crack tip. But we know that Ua is
negative as no external work is being done, hence we get total energy as
hence we get,
σ 2 · a2 π
Ua + Us = − + 4aν (2.6)
E
Here we should note that the line which describes the total energy reaches a
peak and drops, if we draw a line perpendicular to the x-axis from the max point
that would indicate the critical crack length. After differentiating equation 2.6 and
equating it to zero to obtain maximum value, we get,
σ2 · aπ
= 4ν (2.7)
E
which is nothing but
σ2 · aπ
G= = R = 4ν (2.8)
E
20 Chapter 2. LEFM - Linear Elastic Fracture Mechanics
where G is the energy release rate and R is the surface energy per unit extension.
hence by rearranging we get the following for a plane stress condition,
r
2Eν
σf = (2.9)
πa
and for a plane strain condition we get,
s
2Eν
σf = (2.10)
πa(1 − µ2 )
In reality these relations are only the basis for further extrapolation since these two
equations are really only valid for truly brittle materials such as glass.Thus we need
to modify these relations before we can apply them to the problem of fracture in
materials, which are not classically brittle.
Fracture is deemed to occur when the potential energy release rate, G exceeds
the surface energy per unit crack extension which must be provided to the system if
crack growth is to occur. Where these two lines meet the crack length is the critical
or Griffith crack length, ac. The problem is to determine G either analytically or by
experiment such that we have an input value in order to predict the fracture stress of
a supposed cracked body. Again if we look at our elastic loading diagram for crack
lengths a and a+δa we can hopefully see an experimental method by which we could
determine G.
will decrease as the crack extends. As we know for crack length a the elastic strain
energy is given by
1
P1 U1 (2.11)
2
and this changes to
1
P2 U1 (2.12)
2
Hence under fixed grip conditions the extension of the crack from a to a+δa re-
sults in the release (decrease) of elastic strain energy to
1
( P1 − P2 )U1 (2.13)
2
and this loss of energy is because energy is being consumed in the work of frac-
ture required to create the two new crack surfaces.
Now we should consider what happens under constant load condition since this
represents the other end of the spectrum. The strain when the crack is of length a is
given by
1
P1 U1 (2.14)
2
and the strain energy when the crack is developed to a+δa is given by
1
P1 U2 (2.15)
2
which is greater, and from the graph we can say that the work done externally is
P1 (U2 − U1 ) (2.16)
from the graph we can note that the triangle O,P1(U1),P1(U2) is the overall change
in energy which can be expressed as,
22 Chapter 2. LEFM - Linear Elastic Fracture Mechanics
1 1
P1 (U2 − U1 ) + P1 U1 − P1 U2 (2.17)
2 2
simplifying it we get
1
P1 (U2 − U1 ) − P1 (U2 − U1 ) (2.18)
2
hence from the above equation we can say that half the work done goes into
creating new cracks and other half goes into stored strain energy. Now we can note
that energy released in constant load condition is
1
P1 (δU ) (2.19)
2
and energy released during a constant displacement condition is
1
(δP)U (2.20)
2
Also we need to consider the relationship between load and displacement in the
general case. As for any elastic system the displacement and load are related through
a simple linear equation such that for any given crack length we can write that
δU = C · δP (2.22)
and substituting this into Eqs. 2.19 and 2.20 we get,
1
P · C (δP) (2.23)
2
1
P · C (δP) (2.24)
2
hence we can conclude that There is no difference in the energy released when an
infinitesimally small increment of crack growth occurs under conditions of constant
load or constant displacement condition.
We know that the strain or potential energy release for an increment of crack growth
δa is given by G δa per unit thickness and if we define B as the thickness of the plate
we can say that:
1
G · δa · B = P(δU ) (2.25)
2
invoking the compliance relationship,
1 2
G · δa · B = P (δC ) (2.26)
2
rearranging,
2.3. Stability of Crack growth and the R Curve 23
P2 (δC )
G= (2.27)
2δa · B
σ2 · aπ
G= = R = 4ν (2.28)
E
as shown in figure 2.6. But crack growth may be stable or unstable, depending
on how G and R vary with crack size. A plot of R vs. crack extension is called a
resistance curve or R curve. The corresponding plot of G vs. crack extension is the
driving force curve. Consider a wide plate with a through crack of initial length 2a.
At a fixed remote stress σ , the energy release rate varies linearly with crack size.
The first case, shows a flat R curve, where the material resistance is constant
with crack growth. When the stress is σ1 , the crack is stable. Fracture occurs when
the stress reaches σ2 ; the crack propagation is unstable because the driving force
increases with crack growth, but the material resistance remains constant. In the
second case, the crack grows a small amount when the stress reaches σ2 , but cannot
24 Chapter 2. LEFM - Linear Elastic Fracture Mechanics
grow further unless the stress increases. When the stress is fixed at σ2 , the driving
force increases at a slower rate than R. Stable crack growth continues as the stress in-
creases to σ3 . Finally, when the stress reaches σ4 , the driving force curve is tangent
to the R curve. The plate is unstable with further crack growth because the rate of
change in the driving force exceeds the slope of the R curve. Imagine testing a series
of thin (plane stress conditions) centre-cracked panels to instability with different
initial crack lengths. The results of such test would appear schematically as shown
below.
The graph(figure 2.7) shows a locus of points of initial crack length and critical
crack length.
G=R (2.29)
and
dG dR
≤ (2.30)
da da
unstable crack growth occurs when
dG dR
≥ (2.31)
da da
take on a variety of shapes. For example, ductile fracture in metals usually results
in a rising R curve; a plastic zone at the tip of the crack increases in size as the
crack grows. The driving force must increase in such materials to maintain the crack
growth. If the cracked body is infinite (i.e., if the plastic zone is small compared to
the relevant dimensions of the body) the plastic zone size and R eventually reach
steady-state values, and the R curve becomes flat with further growth. The size and
geometry of the cracked structure can exert some influence on the shape of the R
curve. A crack in a thin sheet tends to produce a steeper R curve than a crack in a
thick plate because there is a low degree of stress tri-axiality at the crack tip in the
thin sheet, while the material near the tip of the crack in the thick plate may be in
plane strain. The R curve can also be affected if the growing crack approaches a
free boundary in the structure. Thus, a wide plate may exhibit a somewhat different
crack growth resistance than a narrow plate of the same material. Ideally, the R
curve, as well as other measures of fracture toughness, should be a property only
of the material and not depend on the size or shape of the cracked body. Much
of fracture mechanics is predicated on the assumption that fracture toughness is a
material property.
G · E = πaσ2 (2.34)
√ √
G · E = σ πa (2.35)
This parameter is called the stress intensity factor(K) which is the crack driv-
ing force, and its critical value is a material property known as fracture toughness,
which, in turn, is the resistance force to crack extension. Hence we can say,
√ √
K= G · E = σ πa (2.36)
26 Chapter 2. LEFM - Linear Elastic Fracture Mechanics
Which is a material property and a function of the crack length and stress. The
stress intensity factor, K , is a single parameter which completely specifies the ampli-
tude of the stress field in the vicinity of the crack tip. In general, the stress intensity
factor depends on the geometry of the cracked body (including the crack length) and
it is usual to express it as
√
K = Yσ a (2.37)
where Y is called the shape factor and is a function of body geometry and crack
length.
K = G · E = πaσ2 (2.38)
√ √
K = G · E = σ πa (2.39)
K2 = G · E (2.40)
This equation applies only for a plane stress condition, extending it to a plane
strain condition we get,
K 2 = G · E (1 − µ2 ) (2.41)
we can note that µ is 0.3 for metals and hence(1-µ2 )=0.91 which is not a very
big change, however, the numerical values of G or KI are very different in plane
stress(thin specimens) or plane strain(thick specimens) situations. This is a direct
consequence of an effective stiffness increase experienced when an object is pulled
in tension, but with one lateral plane constrained from contracting under Poisson
effects.
Above a certain thickness Bmin , the fracture toughness has a minimum value
which is independent of thickness. This minimum value of KIc is known as the
plane strain fracture toughness and is denoted KIc.
Particular attention is paid to KIc because this is the minimum toughness that can
be achieved under the most severe conditions of loading. The changes in KIc with
thickness are accompanied by corresponding changes in fracture geometry. In the
plane strain regime the fracture surface is oriented at 90o to the direction of loading
(ie “square” fracture). As the thickness decreases, 45 deg “shear lips” appear on
either side of a flat central regime. At and below the thickness corresponding to the
maximum KIc position, the shear lips occupy the full thickness and one has a 45 deg
“shear” or plane stress fracture.
29
Chapter 3
•
31
Chapter 4
Fatigue in Structures
4.1 Introduction
Fatigue, as understood by materials technologists, is a process in which damage
accumulates due to the repetitive application of loads that may be well below the
yield point. The process is dangerous because a single application of the load would
not produce any ill effects, and a conventional stress analysis might lead to a as-
sumption of safety that does not exist. In one popular view of fatigue in metals, the
fatigue process is thought to begin at an internal or surface flaw where the stresses
are concentrated, and consists initially of shear flow along slip planes. Over a num-
ber of cycles, this slip generates intrusions and extrusions that begin to resemble a
crack. A true crack running inward from an intrusion region may propagate initially
along one of the original slip planes, but eventually turns to propagate transversely
to the principal normal stress. The modern study of fatigue is generally dated from
the work of A. Wöhler, a technologist in the German railroad system in the mid-
nineteenth century. Wohler was concerned by the failure of axles after various times
in service, at loads considerably less than expected. A railcar axle is essentially a
round beam in four-point bending, which produces a compressive stress along the
top surface and a tensile stress along the bottom. After the axle has rotated a half
turn, the bottom becomes the top and vice versa, so the stresses on a particular re-
gion of material at the surface varies sinusoidally from tension to compression and
back again. This is now known as fully reversed fatigue loading.
F IGURE 4.1: •
The fatigue life is usually split into a crack initiation period and a crack growth
period. The initiation period is supposed to include some microcrack growth, but
the fatigue cracks are still too small to be visible. In the second period, the crack
is growing until complete failure. It is technically significant to consider the crack
initiation and crack growth periods separately because several practical conditions
have a large influence on the crack initiation period, but a limited influence or no
influence at all on the crack growth period.
From such an experiment, the stress amplitude σa for fully reversed loading
(equal to one-half of the stress range from the maximum tension to maximum com-
pression), is plotted against the number of fatigue cycles to failure. For materials
which harden by strain-ageing, under constant amplitude loading conditions, these
alloys exhibit a plateau in the stress-life plot typically beyond about 106 fatigue cy-
cles. Below this plateau level, the specimen may be cycled indefinitely without caus-
ing failure. This stress amplitude is known as the fatigue limit or endurance limit.
The value of at is 35% to 50% of the tensile strength UTS for most steels and copper
alloys.
of the endurance limit Se to the ultimate strength Su of a material is called the fatigue
ratio. It has values that range from 0.25 to 0.60, depending on the material. For steel,
the endurance strength can be approximated by:
Se = 0.5Su (4.1)
In addition to this relationship, for wrought steels the stress level corresponding
to 1000 cycles, S1000 , can be approximated by:
∆σ
= σa = σ0f (2N f )b (4.3)
2
where σ0f is the fatigue strength coefficient (which, to a good approximation,
equals the true fracture strength σ, corrected for necking, in a monotonic tension
test for most metals) and b is known as the fatigue strength exponent or Basquin
exponent.
behavior of engineering materials. In this case, the stress range, the stress amplitude
and the mean stress, respectively, are defined as
F IGURE 4.2: •
The mean stress is also characterized in terms of the load ratio, R = σmin /σmax .
With this definition, R = -1 for fully reversed loading, R = 0 for zero-tension fatigue,
and R = 1 for a static load.
When the stress amplitude from a uniaxial fatigue test is plotted as a function of
the number of cycles to failure, the resultant S-N curve is generally a strong function
of the applied mean stress level.
Figure 4.5 shows the typical S-N plots for metallic materials as a function of four
different mean stress levels. One observes a decreasing fatigue life with increasing
mean stress value.
Mean stress effects in fatigue can also be represented in terms of constant-life
diagrams, as shown.
Here, different combinations of the stress amplitude and mean stress providing
a constant fatigue life are plotted. Most well known among these models are those
due to Gerber (1874), Goodman (1899), and Soderberg (1939).
F IGURE 4.3: •
F IGURE 4.4: •
General Observations
• Most actual test data tend to fall between the Goodman and Gerber curves.
• For most fatigue situations R<1 ( i.e. small mean stress in relation to alternat-
ing stress), there is little difference in the theories
• In the range where the theories show large differences (i.e. R values approach-
ing 1) there is little experimental data. In this case the yield stress may set the
design limits.
• The Soderberg line is very conservative and seldom used
F IGURE 4.5: •
∆e p
= e0f (2N f )c (4.8)
2
where e0f is the fatigue ductility coefficient and c is the fatigue ductility exponent.
In general, e0f is approximately equal to the true fracture ductility e f in monotonic
tension.
∆e ∆ee ∆e p
= + (4.9)
2 2 2
from hookes law,
∆ee ∆σ σa
= + (4.10)
2 2E E
4.4. Strain Life Approach 37
F IGURE 4.6: •
∆ee σ0f
= (2N f )b (4.11)
2 E
combining Coffin manson’s relation and Basquin relation,
∆e σ0f
= (2N f )b + e0f (2N f )c (4.12)
2 E
The first and second terms on the right hand side, are the elastic and plastic
components, respectively, of the total strain amplitude. Equation forms the basis for
the strain-life approach to fatigue design and has found widespread application in
industrial practice.
Bauschinger Effect
The stress-strain behavior obtained from a monotonic test can be quite different from
that obtained under cyclic loading. This was first observed by Bauschinger. His ex-
periments indicated the yield strength in tension or compression was reduced after
applying a load of the opposite sign that caused inelastic deformation. Thus, one
single reversal of inelastic strain can change the stress -strain behavior of metals.
F IGURE 4.7: •
F IGURE 4.8: •
by Manson that if
σult
> 1.4 (4.13)
σy
is satisfied then the material will cyclically harden. And if
σult
< 1.2 (4.14)
σy
occurs and mean stress tends to zero. (This is not cyclic softening.) Mean stress
relaxation occurs only in materials that are cyclically stable. Modifications to strain
life equation have been made to account for mean stress effects.
∆ee σ0f − σo
= (2N f )b (4.15)
2 E
where σo is the mean stress.
K f = 1 + q ( K t − 1) (4.18)
Notch sensitivity in fatigue decreases as the notch radius decreases and as the
grain size increases. A larger part will generally have greater notch sensitivity than
a smaller part with proportionally similar dimensions. This variation is known as
the scale effect. Larger notch radii result in lower stress gradients near the notch,
and more material is subjected to higher stresses.
In almost all cases, the fatigue notch factor is less than the stress concentration
factor, and is less than 1. That is: 1≤ K f ≤ Kt
Neuber has developed the following approximate formula for the notch factor
for R= -1 loading, and the Neuber Equation for notch factor sensitivity is:
1
q= q (4.19)
ρ
1+ r
where r is the notch root radius and ρ is a material constant that is related to the
grain size of the material. The notch factor K f is usually used to correct the fatigue
strength for the notched member.
Neuber also established a rule that is useful beyond the elastic limit relating the
effective stress and strain concentration factors to the theoretical stress concentration
factor. Neuber’s rule contends that the formula
Kσ Ke = Kt2 (4.20)
which can be rewritten as,
Chapter 5
1. Life
2. Frequency
4. Stress Amplitude
strength increases. For example, high-strength steels heat treated to over 1380 MPa
(200 ksi) yield strengths have much higher fatigue strengths than aluminum alloys
with 480 MPa (70 ksi) yield strengths. For a large number of steels, there is a di-
rect correlation between tensile strength and fatigue strength; that is, higher-tensile-
strength steels have higher endurance limits. The endurance limit is normally in the
range of 0.35 to 0.60 of the tensile strength. This relationship holds up to a hard-
ness of approximately 40 HRC ( 1240 MPa, or 180 ksi, tensile strength), and then the
scatter becomes too great to be reliable.
This does not necessarily mean it is wise to use as high a strength steel as pos-
sible to maximize fatigue life because, as the tensile strength increases, the frac-
ture toughness decreases and environmental sensitivity increases. In addi- tion, the
endurance limit of high-strength steels is extremely sensitive to surface condition,
residual stress state, and the presence of inclusions that act as stress concentrations.
Fatigue cracking can occur quite early in the service life of the member by the for-
mation of a small crack, generally at some point on the external surface. The crack
then propagates slowly through the material in a direc- tion roughly perpendicular
to the main tensile axis.
Optimizing the overall fatigue properties thus inevitably requires a judicious bal-
ance between strength and ductility.
Mean stress effects have also been incorporated into the uniaxial strain-based
characterization of fatigue life in a simple manner. Assuming that a tensile mean
stress reduces fatigue strength.
Taking cyclic relaxation into account, Cyclic hardening reduces the plastic strain
range and increases the stress range for a fixed total strain.
F IGURE 5.1: •
Peak Counting
Peak counting identifies the occurrence of a relative maximum or minimum load
value. Peaks above the reference load level are counted, and valleys below the ref-
erence load level are counted. Results for peaks and valleys are usually reported
separately. A variation of this method is to count all peaks and valleys without re-
gard to the reference load. To eliminate small amplitude loadings, mean-crossing
peak counting is often used. Instead of counting all peaks and valleys, only the
largest peak or valley between two successive mean crossings.
44 Chapter 5. Statistical Aspects of Fatigue Behavior
F IGURE 5.2: •
Simple-Range Counting
For this method, a range is defined as the difference between two successive rever-
sals, the range being positive when a valley is followed by a peak and negative when
a peak is followed by a valley. Positive ranges, negative ranges, or both, may be
counted with this method. If only positive or only negative ranges are counted, then
each is counted as one cycle. If both positive and negative ranges are counted, then
each is counted as one-half cycle. Ranges smaller than a chosen value are usually
elimi- nated before counting.
F IGURE 5.3: •
da
= A∆K m (5.1)
dN
5.6. Damage rule for irregular loads 45
where da/dN is the fatigue crack growth rate per cycle, ∆K = Kmax - Kmin is
the stress intensity factor range during the cycle, and A and m are parameters that
depend the material, environment, frequency, temperature and stress ratio. This is
sometimes known as the “Paris law,”.
The crack growth rate (da/dN) can be determined from the slope of the curve.
Initially, the crack growth rate is slow but increases with increasing crack length.
Of course, the crack growth rate is also higher for higher applied stresses. If one
can characterize the crack growth, it is then possible to estimate the service life or
inspection intervals required under specific loading conditions and service environ-
ment. In the fracture mechanics approach to fatigue crack growth, the crack growth
rate, or the amount of crack extension per loading cycle, is correlated with the stress-
intensity parameter (K ). This approach makes it possible to estimate the useful safe
life and inspection intervals.
F IGURE 5.4: •
F IGURE 5.5: •
n1 n2
+ =1 (5.2)
N1 N2
Note that absolute cycles and not log cycles are used here.
F IGURE 5.7: •
n1
∑ N1 = 1 (5.3)
Sequence Effect
Two simple VA load sequences are shown in Figure 10.3. These sequences are ap-
plied to a notched specimen. The same two amplitudes are used in both sequences,
but in the first sequence the test starts with the low amplitude (low-high sequence,
or LoHi), and in the second one with the high amplitude (high-low sequence, or
HiLo). The stress ratio is supposed to be zero (S min = 0, or R = 0).
The peak stress at the root notch exceeds the yield stress only in the block with
the high amplitude, whereas this does not occur in the block with the low amplitude.
It implies that notch root plasticity did not occur in the LoHi sequence during the
low-amplitude cycles of the first block of the LoHi sequence.
However, in the other sequence (HiLo), notch root plasticity occurs immediately
in the first block with the high amplitude. In this case, compressive residual stresses
at the root of the notch are present at the beginning of the second block with the
low-amplitude cycles. This is favorable for fatigue in the second block. Although
5.8. Miner’s Rule 49
the number of high-amplitude cycles (n2 ) as observed in the LoHi test is applied
in HiLo test, the fatigue life will be larger in this test due to the favorable residual
stress. In other words, the sequence of the two blocks is significant for the fatigue
life. This sequence effect is not predicted by the Miner rule because the rule ignores
any change of residual stresses induced by previous cycles.