Sie sind auf Seite 1von 21

M.

Mirzaei, Fracture Mechanics

Fracture Mechanics
Lecture Notes: 4
Majid Mirzaei, PhD
Associate Professor
Dept. of Mechanical Eng., TMU
mmirzaei@modares.ac.ir
http://www.modares.ac.ir/eng/mmirzaei/FM.htm

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 of failure which can occur
without any obvious warning. It is believed that 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).

Extrusions

Slip Planes

microcrack
Intrusions

Figure 1. Schematics of fatigue crack initiation.

1
M. Mirzaei, Fracture Mechanics

Fatigue crack propagation occurs through repeated crack tip blunting and sharpening
effects which are in turn caused by microplastic deformation mechanisms operating at
the crack tip (see Fig.2).

Figure 2. Schematics of fatigue crack propagation.

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 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 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 Cycle Fatigue (HCF), through S/N or Wöhler diagrams, constitutes the basis of
the SAFE LIFE philosophy in design against fatigue. In order to determine the

2
M. Mirzaei, Fracture Mechanics

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

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

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 ΔS = S max − S min
Sa = the alternating stress, which is 1/2 the Range of Stress:
S − S min
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.

3
M. Mirzaei, Fracture Mechanics

Other important definitions are:

S min
Load Ratio, R =
Smax
S max + S min
Mean or Static Stress, S mean =
2

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. 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 ⎜ Sf (4-1)
⎝ ⎠
⎡ ( 0.9 Su )2 ⎤
B = log ⎢ ⎥
⎢⎣ S f ⎥⎦

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,

σ N = M f × SN (4-2)

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 (4-3)

These factors are attributed to the followings:

msurf = surface finish


msize = size
mrel = reliability
mload = load
mtemp = temperature
mconc = stress concentration
mmisc = miscellaneous effects

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

4
M. Mirzaei, Fracture Mechanics

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 stress-
strain 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) .

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.

Figure 6: Schematic representation of crack initiation according to LCF

The necessary requirements for a LCF life assessment program 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).

Figure 7: Schematic representation of cyclic strain-life curves

5
M. Mirzaei, Fracture Mechanics

4. A cumulative fatigue damage rule to accurately predict the LCF life for an
arbitrary loading sequence
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).

Stress Analysis

Stress-Strain Notch Stress-Life


Curve Sensitivity Curve

Fatigue
Σ1/Nf f
Life
Stress-Strain Local Cycle Cumulative
Model Stress-Strain Counting Damage

Loading History

Figure 8: Schematic representation of LCF life assessment activities

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, as follows:

da
= f (ΔK ) (4-4)
dN

The most widely used expression, proposed by Paris, is:

6
M. Mirzaei, Fracture Mechanics

da (4-5)
= C (ΔK ) m
dN

in which C and m are material properties obtained from experiment.

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

Figure 10 is a schematic presentation of a servo-hydraulic mechanical testing machine


which is usually used for various types of fatigue testing.

Figure 10

7
M. Mirzaei, Fracture Mechanics

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 ΔKs, 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, Eqn.(4.5).

3. The final unstable 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

Crack Tip Shielding


The stress intensity factor range is generally obtained from a complete analysis of the
applied loading and geometry. Although these analyses are based on global
considerations, they are assumed to characterize the local field. In practice, the local
near-tip driving force may differ from the nominal stress intensity factor due to some
local mechanical, microstructural, or environmental phenomena in the vicinity of the
crack tip. This process is generally referred to as crack tip shielding which causes the
so-called extrinsic toughening. The origins of the above argument go back to 1968
when it was discovered that a fatigue crack may remain closed, during a significant

8
M. Mirzaei, Fracture Mechanics

portion of a loading cycle, due to elastic constraints acting on the plastically stretched
material in the crack wake.
EXTRINSIC TOUGHENING MECHANISMS

1. CRACK DEFLECTION AND MEANDERING

2. ZONE SHIELDING
* transformation toughening
* microcrack toughening
* crack wake plasticity

* crack field void formation


* residual stress fields
* crack tip dislocation shielding
3. CONTACT SHIELDING
* wedging :
corrosion debris-induced closure
crack surface roughness-induced closure
* bridging :
ligament or fiber toughening
* sliding :
sliding crack surface interference
* wedging + bridging :
fluid pressure-induced closure

4. COMBINED ZONE AND CONTACT SHIELDING


* plasticity-induced closure

* phase transformation-induced closure

Figure 12: A general classification of extrinsic toughening mechanisms

The occurrence of crack closure, which is now recognized as being due to a variety of
mechanisms, significantly affects the local crack driving force (stress intensity factor)
and plays a crucial role in fatigue crack growth or arrest characteristics. While the
qualitative interpretation of the closure phenomena can be used to justify a large
number of fatigue crack growth anomalies, the quantitative knowledge of the crack
closure stress intensity level is required to correlate fatigue crack growth rate data.

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
= C ( ΔK ) m ⇒ dN = da
dN C (ΔK ) m
N af
1 1
∫0 dN = C a∫ ⎡ βΔσ (π a)1/ 2 ⎤ m da (4-6)
i ⎣ ⎦
1 ⎡⎣ a (1f − m / 2) − ai1− m / 2 ⎤⎦
N=
C β m (Δσ ) m π m / 2 1− m / 2

9
M. Mirzaei, Fracture Mechanics

plastic wake

Plastic Zone

(a) (b)

Plasticity-Induced Roughness-Induced
Crack Closure Crack Closure

Oxide Debris Transformed Zone

(c) (d)

Oxide-Induced Phase Transformation-Induced


Crack Closure Crack Closure

Viscous Fluid

(e)

Viscous Fluid-Induced
Crack Closure

Figure 13: A general classification of the mechanisms of crack tip closure.

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.

Assignment:
A rotating shaft is made of an alloy-steel for which we have:

K IC = 54 MPa m
da
(mm/cycle) = 7.72 × 10-8 ΔK eff (MPa m)
dN
⎛ 1 − br ⎞
ΔK eff = ⎜ ⎟ ΔK
⎝ 1− r ⎠
(r < 0, b = 0.219), (0 ≤ r , b = 1)

where, r is the stress ratio. The shaft is subject to an alternating stress range of
Δσ = 180 MPa and contains a crack of half length a=0.2 mm. Assuming a geometric
factor of β=1, calculate the number of cycles to fracture for stress ratios of 0 and -1.

10
M. Mirzaei, Fracture Mechanics

NASGRO
The NASGRO computer software was initially developed to provide an automated
procedure for fracture control analysis of NASA space flight hardware and launch
support facilities. It is also applicable to stress and fracture mechanics analysis of
aircraft and non aerospace structures and may be used as a learning and research tool
in fracture mechanics. NASGRO is a collection of computer programs comprised of
three modules: called NASFLA, NASBEM and NASMAT. The NASFLA program
uses fracture mechanics principles to calculate stress intensity factors, compute
critical crack sizes, or conduct fatigue crack growth analyses. Material properties for
crack growth can be obtained from the database supplied with the program, or entered
either as a 1-D table or a 2-D table. The fatigue loading spectra can be input from a
standard file or individual files. User-defined materials properties and fatigue spectra
may also be supplied manually. The second module NASBEM implements the
boundary element method for stress analysis and can be used to obtain stress intensity
factors and stresses for two-dimensional geometries with holes and cracks. The third
module NASMAT can be used to enter, edit and curve-fit fracture toughness and
fatigue crack growth data obtained in a laboratory.

Crack Growth Relationship


Crack growth rate calculations in NASGRO 3.0.20 use a relationship called the
NASGRO equation. Different elements of this equation were developed by Forman
and Newman of NASA, Shivakumar of Lockheed Martin, de Koning of NLR, and
Henriksen of ESA. It was first published by Forman and Mettu and is given by:
p
⎛ ΔKth ⎞
⎜1 −n ⎟
da ⎡⎛ 1 − f ⎞ ⎤ ⎝ ΔK ⎠
= C ⎢⎜ Δ
⎟ ⎥
K (4-7)
⎣⎝ 1 − R ⎠ ⎦ ⎛
q
dN Kmax ⎞
⎜1 − ⎟
⎝ Kc ⎠

where N is the number of applied fatigue cycles, a is the crack length, R is the stress
ratio, ΔK is the stress intensity factor range, and C, n, p, and q are empirically derived
constants. The crack opening function, f, for plasticity-induced crack closure has
been defined by Newman as:

Kop ⎧⎪max ( R , A0 + A1 R + A2 R 2 + A3 R 3 ) R≥0


f = =⎨ (4-8)
Kmax ⎪⎩ A0 + A1 R −2≤R<0

and the coefficients are given by:

[
A0 = ( 0.825 − 0.34α + 0.05α 2 ) cos( π2 Smax / σ 0 ) ]
1
α

A1 = ( 0.415 − 0.071α ) S max / σ 0 (4-9)

A 2 = 1 − A 0 − A1 − A 3

A3 = 2 A0 + A1 − 1

11
M. Mirzaei, Fracture Mechanics

In these equations, α is a plane stress/strain constraint factor, and S max / σ 0 is the ratio
of the maximum applied stress to the flow stress. The plane stress/strain constraint
factor, α, has been treated as a constant for the purposes of curve fitting the crack
growth data for each particular material system. Values range from 1, for plane
stress, to 3, for plane strain condition. The ratio of the maximum applied stress to the
flow stress, Smax/σ0, is assumed to be constant. Most materials that were curve fit for
NASGRO 3.0.20 use a value of Smax/σ0 = 0.3, which was chosen because it is close to
an average value obtained from fatigue crack growth tests using various specimen
types. The threshold stress intensity factor range in Eq (4.7), ΔKth, is approximated
by the following empirical equation:
1 (1+ Cth R )
⎛ a ⎞ ⎛ 1− f ⎞
2

ΔK th = ΔK 0 ⎜ ⎟ /⎜ ⎟ (4-10)
⎝ a + a0 ⎠ ⎝ (1 − A0 )(1 − R ) ⎠

where R is the stress ratio, f is the Newman closure function, A0 is a constant (Eq.
4.9) used in f , ΔK0 is the threshold stress intensity factor range at R = 0, Cth is an
empirical constant, a is the crack length, and a0 is an intrinsic crack length. Values of
Cth for positive and negative values of R , and ΔK0 are stored as constants in the
NASGRO materials files. The intrinsic crack size a0 has been assigned a fixed value
of 0.0015 in. (0.0381 mm).

In practice, some materials exhibit a very small stress ratio effect. In these special
cases, a curve-fitting option that allows the crack opening function to be bypassed is
used. The parameters for this bypass option are α = 5.845, Smax/σ0 = 1.0. These
values are selected in order that f in Eq. (4.9) would be equal to zero for negative
stress ratios and would be equal to R (Kop = Kmin) for 0 ≤ R < 1. Thus, for positive
stress ratios, the crack growth relationship (Eq. (4.7)) reduces to:

p
⎛ ΔK th ⎞
C ΔK ⎜ 1 −
n

da ⎝ ΔK ⎠
= q
(4-11)
dN ⎛ K ⎞
⎜ 1 − max ⎟
⎝ Kc ⎠

where the entire ΔK range contributes to crack propagation. For negative stress ratios,
the reduced expression is:

p
⎛ ΔK th ⎞
CK n
⎜1 − ⎟
da max
⎝ ΔK ⎠
= q
(4-12)
dN ⎛ K ⎞
⎜ 1 − max ⎟
⎝ Kc ⎠

Version 4.0 (and subsequent versions) of NASGRO have been developed and are
distributed under the terms of Space Act Agreement between NASA and Southwest
Research Institute. These versions of NASGRO will require purchasing a license.

12
M. Mirzaei, Fracture Mechanics

AFGROW
AFGROW, which is a software tool for analysis of fatigue crack initiation and growth
in metallic structures, has emerged from a number of computer programs developed at
the Air Force Wright Aeronautical Laboratories (AFWAL/FIBEC). AFGROW
implements five different material models (Forman Equation, Walker Equation,
Tabular lookup, Harter-T Method and NASGRO Equation) to determine crack growth
per loading cyclic. Other user options include five load interaction (retardation)
models (Closure, FASTRAN, Hsu, Wheeler, and Generalized Willenborg), a strain-
life based fatigue crack initiation model, and the ability to consider the effect of the
bonded repair. AFGROW is a very powerful tool for fatigue crack growth studies of
aircraft and non aerospace structures and may be used as a learning and research tool
in fracture mechanics.
It can be downloaded at: http://www.siresearch.info/projects/afgrow

Demonstration
In this section, we will briefly review the simulation of fatigue crack growth in a
detonation tube using AFGROW and FRANC3D codes. More details are reported in:
Mirzaei, M., Salavatian, M, Biglari, H., "Simulation of Fatigue Crack Growth in a
Detonation Tube," Proceedings of the ASME PVP-2006 /11th International
Conference on Pressure Vessel technology, ICPVT-11, 23-27 July 2006, Vancouver,
Canada.

In this work, we studied the cyclic growth of a pre-existing crack in a pressure vessel
subjected to repeated internal gaseous explosions. Such situation may occur in
pipeline systems, pressurized aircraft fuselages, rocket casings, pulse detonation
engines (PDE), and many other engineering applications. In practice, the internal
detonation loading of cylindrical tubes involves loads that propagate at high speeds,
causing the formation of flexural waves in the tube wall. These waves can result in
high strains, which may exceed the equivalent static strains (caused by the same
nominal loading pressure) by up to a factor of 4. Moreover, every single blast
initiates a spectrum of vibrational strains which includes both original and reflected
waves as shown in Fig.14.

Figure 14. (a): experimental result for detonation velocity of 1478.8 m/s.
(b): prediction of the main signal by an analytical model. (c): prediction of the
main signal and the reflected waves by an analytical model.

Two different methods were used for fatigue life calculations. In the first method, the
stress analysis results of an analytical model were combined with the crack growth

13
M. Mirzaei, Fracture Mechanics

simulation capabilities of the AFGROW software. In the second method, fatigue life
calculations were performed by a cycle-by-cycle integration of the desired growth rate
expression along the crack path using FRANC3D. The growth rate expression was the
Paris model modified to incorporate fatigue crack closure. The fatigue lives were
predicted using three different design approaches described below:

A. Design through the consideration of the maximum pressure multiplied by an


amplification factor to simulate the dynamic effects,
B. Design through the consideration of the vibrational spectrum due to traveling
load,
C. Design through the consideration of the vibrational spectrum due to the
traveling load, plus the reflected waves.

Figure 15. Loading spectrums.

Figure 16 shows crack growth curves for the detonation velocity of 1478.8 m/s,
obtained based on the above three design approaches. It is seen that for each design
approach there is a good agreement between the two different methods of crack
growth modeling, although in general, the FRANC3D simulations predict shorter
lives. Nevertheless, the striking fact is that the predicted lives obtained based on the
design approach C (the full loading spectrum) are almost twenty times less than those
obtained from design A (the simplistic assumption of a single loading cycle multiplied
by a dynamic amplification factor). The difference between the two approaches B
and C is also significant, indicating the importance of the reflected waves in reducing
the fatigue life. The above simulations indicate that realistic fatigue life predictions
for tubes subjected to internal detonation require the consideration of the entire

14
M. Mirzaei, Fracture Mechanics

spectrum. Failure to do so can result in life estimations which are orders of


magnitude more than real lives. Hence, in such situations, the usual practice of design
based on the simplistic assumption of a single loading cycle modified by a dynamic
load factor can be quite non-conservative and very dangerous.

FRANC3D-Design A FRANC3D-Design B FRANC3D-Design C


AFGROW-Design A AFGROW-Design B AFGROW-Design C

0.025
Crack Length (mm)

0.02

0.015

0.01

0.005
1.E+02 1.E+03 1.E+04 1.E+05

Number of Cycles

Figure 16. crack growth curves for detonation velocity of 1478.8 m/s.

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 of their melting point in
Kelvin scale. However, for polymers substantial 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

15
M. Mirzaei, Fracture Mechanics

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

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.

Elastic
Tertiary

Steady State
Primary

Figure 19. Creep zones ahead of crack tip

It should be mentioned that for short-term applications at high 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

16
M. Mirzaei, Fracture Mechanics

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.

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 stress and
temperature gradients it is likely that creep 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 creep zone relative to the specimen dimensions.

Elastic Elastic Elastic

Creep Zone
Creep Zone Creep Zone

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


conditions.

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 ⎟ (4-13)
Γ⎝
∂x ⎠

in which
εkl
w = ∫ σ ij d ε ij (4-14)
0

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. This is the so-called Small Scale condition (SSC), as opposed to the

17
M. Mirzaei, Fracture Mechanics

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 path-
independent integral C* is usually used. The transition time for SSC condition to turn
to SS condition can be estimated by :

1 + 2 β n K 2 (t1 )(1 − ν 2 )
t1 = (4-15)
n +1 C * (t1 ) E

in which β is dependent on the waveform of loading and is defined by:

K ( t ) = K 1t β (4-16)

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

Demonstration
As a demonstration of the material covered above, we will briefly review the stress
analysis and life assessment of a first-stage air-cooled blade made of the superalloy
IN738LC. More details are reported in: Mirzaei, M., Karimi, R., "Stress Analysis and
Life Assessment of a Gas Turbine Blade," Proceedings of the 10th International
Congress of Fracture, (ICF10) USA. 2001.

18
M. Mirzaei, Fracture Mechanics

The most sever cyclic duty for an industrial turbine blade is the peak-load generation
commonly experienced in the utility combined-cycle plants, where the blades are
subjected to frequent startups and shutdowns and also large numbers of working
hours. In this study, three-dimensional finite element thermal and stress analyses of
the blade were carried out for the steady-state full-load operation. The results of these
analyses were used for determination of the regions where the combination of high
temperature and high tensile stress was sufficient for significant creep-fatigue crack
growth. Accordingly, a critical point at the leading edge of the airfoil, near the root,
was selected for crack modeling. With the assumption of occurrence of small-scale
creep and thermal-fatigue during each start-stop cycle, the pertinent crack tip
parameters were calculated using the energy domain integral method. An incremental
crack growth scheme was considered and the total life for the growth of a 0.5mm
surface crack to a 5mm through-thickness crack was calculated.

(a) (b)

Figure 22: a) Finite Element mesh of the entire blade. b) Contours of maximum
principal stress due to the mechanical loadings (stresses are in Pascal).

The thermal and mechanical stress analyses were carried out using the general-
purpose finite element package LUSAS. Figure 22a depicts the finite element mesh
of the entire blade. Figure 22b shows the distribution of the maximum principal stress
component, due to the centrifugal and pressure loadings, in the airfoil. It is clear that
the maximum tensile stress occurs at the suction side of the airfoil near the root.
However, if the mechanical and thermal stresses are combined, the location of the
maximum tensile stress will shift to the corner of an internal rib, which is in fact the
coolest point according to the thermal analysis results depicted in Figure 23a. The
results also indicate that the leading edge is the hottest region of the blade. Hence, the
crack modeling was performed in a critical region of the leading edge near the root, in
spite of relatively lower tensile stresses in this region.

As previously mentioned, the two major parameters used for correlating creep crack
growth data are the stress intensity factor K and the integral C*. 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.

19
M. Mirzaei, Fracture Mechanics

Since the IN738LC is a high-strength creep-resistant material and the start-stop cycle-
time of the peak load turbines are usually less than 10 hours, the assumptions of SSC
condition and validity of applying K as the correlation parameter were considered in
this study.

(a) (b)

Figure 23: a) Temperature distribution in a section of the airfoil near the root (Cº).
b) Modified finite element mesh at the same section for crack modeling.

Figure 23b depicts the section of the airfoil at the crack location, where the mesh was
refined for crack modeling. Several procedures are available for numerical evaluation
of stress intensity factors. In this study we used the energy domain integral method for
calculation of the J-Integral and the SIF. The general formulation of the energy
domain integral method in the absence of plastic strains and crack surface tractions
can be written as:

⎧ ⎡ ∂u j ⎤ ∂q ⎡ ∂θ ∂u j ⎤ ⎫
J = ∫ ⎨ ⎢σ ij − Wδ 1i ⎥ + ⎢ασ ii − Fi q ⎬dA (4-17)
A* ⎩ ⎣
∂x1 ⎦ ∂xi ⎣ ∂x1 ∂x1 ⎥⎦ ⎭
ε kl
in which W is the stress work given by: W = ∫ σ ij dε ij
0
In the above equations A* is the integration area, σij and εij are the Cauchy stress and
strain tensors, uj is the displacement vector, δ is the Kronecker delta, q is an arbitrary
function that is equal to unity on Γ0 and zero on Г1 (the inner and the outer integration
paths respectively), α is the coefficient of thermal expansion, θ is the temperature, and
Fi are body forces. The discretized form of the above expression was evaluated for
different crack lengths. Accordingly, the K values were calculated using the
following expression:

J =
(1 − ν ) K
2
2
(4-18)
I
E

Finally, using a regression analysis, the following expression was obtained for the
variation of K with the crack length at the desired region:

20
M. Mirzaei, Fracture Mechanics

K I = 58.4 × a 0.206 (4-19)

The total crack growth per working cycle, da/dN, was then expressed as:
tC
da da
= + ∫ a ( K )dt (4-20)
dN dN Fatigue 0

in which, tC is the holding time during each start-stop cycle. Using the pertinent
experimental results for IN738LC, the numerical integration of the above equation
between the two crack lengths of 0.5mm and 5mm, for a holding time of 8 hours,
resulted in a total life of 11368 hours and 1421 start-stop cycles. The obtained results
were consistent with the observations reported by Bernstein and Allen, who
performed detailed failure analysis of cracked first-stage blades, for General Electric
MS1001E industrial gas turbine, made of IN738LC. For an 11000-hour blade with
874 start-stop cycles, they observed a 0.5mm crack at the leading edge of the blade.
They observed extensive leading edge cracking for a blade that had experienced 1800
start-stop cycles and 24000 hours of operation, with cracks penetrated through to the
leading edge cooling hole.

These notes have been prepared as a student aid and should not be considered as a book. Little
originality is claimed for these notes other than selection, organization, and presentation of the
material.

The following references have been used for preparation of the lecture notes and are
recommended for further study in this course.

References:
Anderson, “Fracture Mechanics Fundamentals and Applications.”
Meguid, “Engineering Fracture mechanics”
Kanninen, "Advanced Fracture Mechanics"
Broek, “Elementary Engineering Fracture Mechanics”
Dowling, "Mechanical Behavior of Materials"
“Structural Integrity Lecture Notes”, by Prof. Gray

21

Das könnte Ihnen auch gefallen