You are on page 1of 18

2

LEFM - Linear Elastic Fracture Mechanics

2.1.
Griffith’s Energy balance criterion
When A.A. Griffith (1893–1963)1 began his pioneering studies of fracture in glass in the years just prior to 1920,
he was aware of Inglis work in calculating the stress concentrations around elliptical holes, and naturally
considered how it might be used in developing a fundamental approach to predicting fracture strengths.
However, the Inglis solution poses a mathematical difficulty: in the limit of a perfectly sharp crack, the stresses
approach infinity at the crack tip. This is obviously nonphysical and using such a result would predict that
materials would have near-zero strength: even for very small applied loads, the stresses near crack tips would
become infinite, and the bonds there would rupture. Rather than focusing on the crack-tip stresses directly,
Griffith employed an energy-balance approach that has become one of the most famous developments in
materials science.

2.1.1.
Theoretical Stress Approach to Fracture
we know that two atoms (or ions) can attract or repel each other, but they remain two entities. Attractive
forces exist at long range and short range, but repulsive forces are only significant at short range.
Consider the simplest case of 2 layers of atoms, packed in square, on planes perpendicular to the applied
force. Consider what happens when we try to pull apart two atoms, one in each layer, so that the interatomic
spacing increases from r o to r o + dr.

Figure 2.1: Lattice image of atoms

1 A.A. Griffith, The phenomenon of rupture and flow in solids, Philosophical transactions of the Royal Society of London, A221, 163-197,
1921

11
12 2. LEFM - Linear Elastic Fracture Mechanics

The theoretical strength from this kind of approach is given by


s

σf = (2.1)
ro

In reality γ is of the order of r o /100 such that σ f has a value of E/10 which is at least one to two orders of
magnitude higher than we observe in practice. The theoretical approach is somehow fatally flawed.

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 approximately 10,000 MPa as per the equation 2.1.
Also, experiments on glass fibers that Griffith himself conducted suggested that the fracture stress in-
creases as the fiber diameter decreases. Hence the uniaxial tensile strength, which had been used extensively
to predict material failure before Griffith, could not be a specimen-independent material property. Griffith
suggested that the low fracture strength observed in experiments, as well as the size-dependence of strength,
was due to the presence of microscopic flaws in the bulk material.

2.1.2.
Energy Balance Approach
Now we know that the strain energy changes decreases when crack grows when no external work is done
during constant displacement and during constant load. But how do we find out the strain energy in a cracked
component. Let us now consider a infinite plate with a central crack under uniaxial tension. This crack now
has two tips.

Figure 2.2: 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 micromechanism of fracture.
Let us now consider the plate with hole. The strain energy in the presence of the crack is given by

σ2
U a = V ol umeo f t r i ang l es × (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 2aλ where 2a is the crack length and
λ is the proportionality factor.
For thin plates λ = π/2
Hence, we get,
σ2
U a = 2 · (0.5 × 2a × 2aλ) × (2.3)
2E

σ2
U a = 2 · (0.5 × 2a × 2a × π/2) × (2.4)
2E
2.1. Griffith’s Energy balance criterion 13

we get,

σ2 a 2 π
Ua = (2.5)
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,

U s = 4aν (2.6)

but we know that U a is negative as no external work is being done, hence we get total energy as

σ2 a 2 π
U a +U s = − + 4aν (2.7)
E
if we differentiate the above with respect to ’2a’ and equate to zero we can get maximum total energy. Now
if we plot equations 2.5, 2.6 and 2.7 we get the following.

Figure 2.3: Griffiths energy graph

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 2.7 and equating it to zero to obtain maximum value, we get,

σ2 aπ
= 2ν (2.8)
E
which is nothing but
G =R (2.9)

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,
s
2E ν
σf = (2.10)
πa
and for a plane strain condition we get,
s
2E ν
σf = (2.11)
πa(1 − µ2 )
and for a volume with a penny shaped crack we get,
14 2. LEFM - Linear Elastic Fracture Mechanics

s
πE ν
σf = (2.12)
2a(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.

2.1.3.
Energy Release Rate
The next step in the development of Griffith’s argument was consideration of the rates of energy change with
crack extension, because the critical condition corresponds to the maximum point in the total energy curve.
Hence to obtain the energy release rate ’G’ we differentiate equation 2.5 with respect to ’2a’ and to obtain
thesurface energy per unit extension we differentiate equation 2.6 with respect to ’2a’ and plot the two on a
graph of energy vs crack length.

δU a σ2 aπ
=G = (2.13)
δ(2a) E
and
δU s
= R = 2ν (2.14)
δ(2a)

Figure 2.4: Energy Release Rate

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, a c .
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.

Compliance Change
produce a load-displacement diagram for the condition where the crack length is a and then superimpose
the diagram for the condition when the crack length is a + δa.
let us consider fixed grip i.e constant displacement condition. As the crack extends the stiffness of the
plate will decrease such that because the grips are fixed (equivalent to fixed displacement, i.e. constant U1
) the load applied by the grips will decrease as the crack extends. As we know for crack length a the elastic
strain energy is given by

1
P 1U 1 (2.15)
2
and this changes to
1
P 2U 1 (2.16)
2
as the crack extends to a + δa.
2.1. Griffith’s Energy balance criterion 15

Figure 2.5: Load vs Displacement graph for constant displacement and constant load

Hence under fixed grip conditions the extension of the crack from a to a + δa results in the release (de-
crease) of elastic strain energy from the plate equivalent to

1
(P 1 − P 2 )U1 (2.17)
2

and this loss of energy is because energy is being consumed in the work of fracture 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
P 1U 1 (2.18)
2
and the strain energy when the crack is developed to a + δa is given by

1
P 1U 2 (2.19)
2
which is greater, and from the graph we can say that the work done externally is

P 1 (U2 −U1 ) (2.20)

from the graph we can note that the triangle O,P 1 (U1 ),P 1 (U2 ) is the overall change in energy which can be
expressed as
1 1
P 1 (U2 −U1 ) + P 1U1 − P 1U2 (2.21)
2 2
simplifying it we get
1
P 1 (U2 −U1 ) − P 1 (U2 −U1 ) (2.22)
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
P · δU (2.23)
2

and energy released during a constant displacement condition is

1
δP ·U (2.24)
2
16 2. LEFM - Linear Elastic Fracture Mechanics

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 rela- ted through a simple linear equation such that for any
given crack length we can write that
U = CP (2.25)

where C is a constant referred to as the compliance of the system. (Note that C has the inverse units to
stiffness since compliance is in effect the inverse of stiffness) also we can write

δU = C · δP (2.26)

and substituting this into Eqs. 2.23 and 2.24 we get,

1
P · C · δP (2.27)
2

1
P · C · δP (2.28)
2
hence we can conclude that There is no difference in the energy released when an infinitesimally small incre-
ment 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 (2.29)
2

invoking the compliance relationship,


1
G · δa · B = P 2 · δC (2.30)
2
rearranging,
P 2 · δC
G= (2.31)
2δa · B

Figure 2.6: Compliance as a function of crack length

2.1.4.
Stability of Crack growth and the R Curve
Crack growth occurs when
σ2 aπ
G= = 2ν = R (2.32)
E
as shown in figure 2.4. 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.
2.1. Griffith’s Energy balance criterion 17

Figure 2.7: Driving force & R curve as a function of crack length

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 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 increases 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.

Figure 2.8: R-curve for various crack lengths

The graph(figure ??) shows a locus of points of initial crack length and critical crack length. 2

Stability Criteria
The conditions for stable crack growth can be expressed as,

G =R (2.33)

and
dG d R
≤ (2.34)
da da
2 Try to comprehend what happens to the distance between a and a and their respective stresses
i c
18 2. LEFM - Linear Elastic Fracture Mechanics

unstable crack growth occurs when


dG d R
≥ (2.35)
da da

Shape of the R-curve


Some materials exhibit a rising R curve, while the R curve for other materials is flat. The shape of the R curve
depends on the material behavior and, to a lesser extent, on the configuration of the cracked structure. The R
curve for an ideally brittle material is flat because the surface energy is an invariant material property. When
nonlinear material behavior accompanies fracture, however, the R curve can 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 triaxiality 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
behavior 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.

2.2.
Stress Intensity Factor
2.2.1.
Stress Intensity Factor vs Stress Concentration
Stress concentration factors are due to geometrical changes of cross sections and regardless of the load con-
dition such as bending,tension,compression or shearing.
Therefore, you can find these factors mentioned in tables or figures in handbooks. No matter of what is
the kind of load condition, you can choose the one corresponding to your geometry during your design.
The theoretical stress concentration factor is
σmax
Kt = (2.36)
σ
However , when there is a crack in your model, the stress intensity factor comes into design which not only
is dependent to geometry also extremely to load condition and that’s why you can’t find these factors easily
in handbooks. They are determined experimentally according to each part geometry and load condition.

2.2.2.
Stress Intensity
Let us now look at the equation,

πaσ2
G = 2ν = (2.37)
E
let us rearrange all material property terms,

G · E = πaσ2 (2.38)

p p
G · E = σ πa (2.39)
This parameter is called the stress intensity factor(K) which is the crack driving 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, p
K = σ πa (2.40)
2.2. Stress Intensity Factor 19

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 amplitude 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
p
K =Yσ a (2.41)

where Y is called the shape factor and is a function of body geometry and crack length.

Stress Intensity factor(SIF) for different geometries


The above expression for K is applicable to an ideal situation of an infinite plate containing a centre crack of
length 2a . In practice the presence of finite boundaries and the way in which the crack is loaded affects the
value of the stress intensity factor.
Stress intensities can be defined for each type of displacement and are designated K I , K I I and K I I I with
respect to the three modes of fracture as discussed previously. However, the great majority of failures occur
under mode I loading and hence most fracture toughness data relate to this mode of loading. For mode I
loading, the critical stress intensity factor is designated K I c .

1. Center cracked plate specimen

Figure 2.9: Center cracked plate specimen

p
KI = Y σ a (2.42)

where Y is,
πa
r
Y = sec (2.43)
W

2. Single edge notched plate geometry

Figure 2.10: Single edge notched plate geometry

p
KI = Y σ a (2.44)
20 2. LEFM - Linear Elastic Fracture Mechanics

where Y is, 1.99 for short cracks or,


a a a a
Y = 1.99 − 0.41 + 18.70( )2 − 38.48( )3 + 53.85( )4 (2.45)
W W W W
upto a/W=0.6
2.2. Stress Intensity Factor 21

3. Double edge notched plate specimen

Figure 2.11: Double edge notched plate specimen

p
KI = Y σ a (2.46)
where Y is, 1.99 for short cracks or,
a a 2 a 3 a 4
1.99 − 0.994 W + 0.363( W ) − 0.835( W ) + 0.34( W )
Y = q (2.47)
a
1− W

4. Embedded penny shaped crack geometry

Figure 2.12: Embedded penny shaped crack geometry

p
KI = Y σ a (2.48)

2
Y =p (2.49)
π
22 2. LEFM - Linear Elastic Fracture Mechanics

5. Compact tension (CT) geometry

Figure 2.13: Compact tension (CT) geometry

P a
KI = ·f( ) (2.50)
BW 1/2 W
where,
a a a a a a
f( ) = 29.6( )1/2 − 185.5( )3/2 + 655.7( )5/2 − 1017( )7/2 + 638.9( )9/2 (2.51)
W W W W W W
6. Single edge notched bend geometry

Figure 2.14: Single edge notched bend geometry

PS a
KI = ·f( ) (2.52)
BW 3/2 W
where,
a a a a a a
f( ) = −37.6( )7/2 + 38.7( )9/2 + 2.9( )1/2 − 4.6( )3/2 + 21.8( )5/2 (2.53)
W W W W W W
7. Double cantilever beam geometry

Figure 2.15: Double cantilever beam geometry

p Pa
KI = 2 3 (2.54)
Bh 3/2
2.2. Stress Intensity Factor 23

2.2.3.
Relationship between K and G
Stress intensity factors are used in design and analysis by arguing that the material can withstand crack tip
stresses up to a critical value of stress intensity, termed K I c , beyond which the crack propagates rapidly. This
critical stress intensity factor is then a measure of material toughness. The failure stress σ is then related to
the crack length a and the fracture toughness from equation 2.40.
p
K = σ πa (2.55)

p p
K = σ πa = G · E (2.56)

hence we can say that

K2 =G ·E (2.57)

This equation applies only for a plane stress condition, extending it to a plane strain condition we get,

K 2 = G · E (1 − µ2 ) (2.58)

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 K I 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.

Role of Material Thickness


Specimens having standard proportions but different absolute size produce different values for K I . This re-
sults because the stress states adjacent to the flaw changes with the specimen thickness (B) until the thickness
exceeds some critical dimension. Once the thickness exceeds the critical dimension, the value of K I becomes
relatively constant and this value, K I c , is a true material property which is called the plane-strain fracture
toughness. The relationship between stress intensity, K I , and fracture toughness, K I c , is similar to the rela-
tionship between stress and tensile stress. The stress intensity, K I , represents the level of “stress” at the tip
of the crack and the fracture toughness, K I c , is the highest value of stress intensity that a material under very
specific (plane-strain) conditions that a material can withstand without fracture. As the stress intensity fac-
tor reaches the K I c value, unstable fracture occurs. As with a material’s other mechanical properties, K I c is
commonly reported in reference books and other sources.

Figure 2.16: K is measured as a function of specimen thickness

Above a certain thickness B mi n , the fracture toughness has a minimum value which is independent of
thickness. This minimum value of K I c is known as the plane strain fracture toughness and is denoted K I c .
24 2. LEFM - Linear Elastic Fracture Mechanics

Figure 2.17: K is measured as a function of specimen thickness

Particular attention is paid to K I c because this is the minimum toughness that can be achieved under the
most severe conditions of loading.
The changes in K I c 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, 45o “shear lips” appear on either side of a flat central regime. At and below the
thickness corresponding to the maximum K I c position, the shear lips occupy the full thickness and one has a
45o “shear” or plane stress fracture.
This has been represented in figure 2.18.

2.2.4.
Mixed Mode Fracture and Crack Propagation
Consider a through crack in an infinite plate where the normal to the crack plane is oriented at an angle β
with the stress axis as shown in the figure.

Figure 2.18: Through crack in an infinite plate for the general case where the principal stress is not perpendicular to the crack plane.

If β 6= 0 , the crack experiences combined Mode I and Mode II loading; K I I I = 0 as long as the stress axis
and the crack normal both lie in the plane of the plate. If we redefine the coordinate axis to coincide with
the crack orientation we see that the applied stress can be resolved into normal and shear components. The
stress normal to the crack plane, σ y 0 y 0 , produces pure Mode I loading, while τx 0 y 0 applies Mode II loading to
the crack. Which can be related by Mohrs circle,

p
K I = σ y 0 y 0 πa (2.59)
2.2. Stress Intensity Factor 25

p
K I = σcos 2 (β) πa (2.60)

and similarly,
p
K I I = τx 0 y 0 πa (2.61)

p
K I I = σsi n(β)cos(β) πa (2.62)

The above equations reduce to the pure Mode I solution when β = 0. The maximum K I I occurs at β = 450 ,
where the shear stress is also at a maximum.

Direction of Crack Propagation


A propagating crack seeks the path of least resistance (or the path of maximum driving force) and need not
be confined to its initial plane. If the material is isotropic and homogeneous, the crack will propagate in such
a way as to maximize the energy release rate. What follows is an evaluation of the energy release rate as a
function of propagation direction in mixed-mode problems. Only Mode I and Mode II are considered

Figure 2.19: Typical propagation from an initial crack that is not orthogonal to the applied normal stress. The loading for the initial
angled crack is a combination of Mode I and Mode II, but the crack tends to propagate normal to the applied stress, result- ing in pure
Mode I loading.

Figure 2.20: Optimum propagation angle for a crack oriented at an angle β from the stress axis.
26 2. LEFM - Linear Elastic Fracture Mechanics

2.3.
Standard Test Method for Linear-Elastic Plane-Strain Fracture Toughness
K I c of Metallic Materials
This test method covers the determination of fracture toughness K I c of metallic materials under predomi-
nantly linear-elastic, plane-strain conditions using fatigue precracked specimens. The property K I c deter-
mined by this test method characterizes the resistance of a material to fracture in a neutral environment in
the presence of a sharp crack under essentially linear-elastic stress. Variation in the value of K I c can be ex-
pected within the allowable range of specimen proportions, a/W and W/B. The best would be to follow ASTM
Standard No. E399.

Figure 2.21: Specimen with varying thickness

Figure 2.22: Load vs Displacement

the stress intensity factor can be calculated using the following formula.
All standardized fracture toughness protocols include the same elements which are described as follows.

Specimen Geometry
Acceptable specimen geometries are those for which the compliance calibration curve as a function of nor-
malised crack length is well documented and accepted. Given that most practical or potential brittle fracture
problems involve fairly large pieces of metal the ASTM and BS both include the Compact Tension (CT) and
Single Edged Notched Bend specimen (SENB) which will cater for most available material configurations.
Since the test procedure requires that a sharp starter crack in the form of fatigue crack is present in the spec-
imen the standards also dictate the geometry of the starter notch from which the fatigue crack has to be
initiated and grown to a required length. Specifying the geometry of the starter notch ensures that the fatigue
crack initiates and remains in the mid-plane of the specimen and that the crack front is symmetric.

Specimen Size
Plane strain fracture toughness testing is almost unique in requiring that the experimenter has a good idea of
the result before carrying out the test! Since one of the main validity restriction for testing of this type is that
2.4. Crack-tip plasticity 27

fully plane strain conditions are developed at the crack tip the specimen must have certain minimum dimen-
sions. Typical geometries which are or have been used for plane strain fracture toughness determination are
shown below together with their stress intensity factor solutions and/or G solutions.

2.3.1.
Typical K I c test procedure
As previously noted it is a great advantage in trying to carry out a fracture toughness test to have some reason-
able idea of the likely result since it is the result which effectively determines whether or not the procedure
will have to be repeated.

• Determine the critical dimensions of the specimen in terms of the required, a , W - a and B to ensure
true plane strain conditions. This can only be achieved by having a good idea of the probable fracture
toughness or by ‘guessing’ the likely value of KQ since it is this value which will determine the physical
dimensions required of the specimen to obtain valid plane strain conditions at the crack tip. Quite
often a Charpy or Izod toughness value is available and these values can be used in conjunction with
empirically developed relations to predict a stress intensity fracture toughness.

• Select an appropriate specimen geometry consistent with available material, testing machine and crack
orientation.

• Machine specimen to suit chosen geometry and fatigue pre-crack ensuring that fatigue pre-cracking
load does not lead to a large crack tip plastic zone. Maximum applied K in the final stages of fatigue
pre-cracking must be less than 60% of subsequently determined KQ .

• Set up specimen in testing machine with appropriate instrumentation to be able to record the load-
displacement diagram with sufficient accuracy.

• Test specimen recording a continuous graph of load versus displacement up to eventual fracture of the
specimen.

• Analyse load-displacement record.

• Calculate KQ on the basis of the chosen protocol method.

• Check for validity of plane strain conditions and if satisfied then KQ =K I c .

2.4.
Crack-tip plasticity
2.4.1.
Stress ahead of the crack tip
So far we have considered the elastic stress normal to the crack plane σ y y = p K Now we need to look at the
2πr
x and y directions. At a blunted crack tip we have a small area of free surface normal to the x direction, hence
at this position σxx = 0 . However, as we enter the crack tip zone σxx tends to σ y y .

Figure 2.23: Variation of Stresses ahead of the crack tip


28 2. LEFM - Linear Elastic Fracture Mechanics

Now let us look at how the stresses in the Z-direction behave. For thin sheet there is zero stress in the
thickness direction, but there will be a thickness strain which manifests itself as a local contraction. This
condition is known as plane stress.
For very thick material there will be no strain in the thickness direction due to the constraining influence
of the surrounding lightly stressed material. This condition is known as plane strain. Although there is no
thickness strain in plane strain, there must be a positive thickness stress to resist the Poisson contraction.

2.4.2.
Crack Tip Plasticity
LEFM assumes that the material is elastically loaded, i.e. there is negligible plastic deformation of the mate-
rial. The K-field associated with the crack predicts that the local stress σ y y tends to infinity as the distance
ahead of the crack tip tends to zero.
the stresses immediately ahead of the crack tip must exceed the yield strength of the material and a local
region of plasticity must develop. The extent of the plastic zone depends on the yield strength of the material
and the state of stress at the crack tip.

Plane stress plastic zone size


In plane stress σzz = 0. Hence yield can occur when σ = σ y (Tresca).

K
σy = σy y = p (2.63)
2πr p
where r p is the radius of the crack tip plastic zone.

1 K 2
rp = ( ) (2.64)
2π σ y

Figure 2.24: Yield region and Plastic region near the crack tip

With respect to the above figure the material in the plastic zone carries less stress than it would otherwise
carry if the material remained elastic. Irwin accounted for the softer material in the plastic zone by defining
an effective crack length that is slightly longer than the actual crack size. The effective crack length is defined
as the sum of the actual crack size and a plastic zone correction.

ae f f = a + r y (2.65)
The effective stress intensity is obtained by inserting a e f f into the K expression for the geometry of inter-
est.

K e f f = Y σ πa e f f
p
(2.66)
where Y is also affected by a e f f .

You might also like