Sie sind auf Seite 1von 239

Title

Author(s)

On fatigue failure prediction with damage mechanics: theory and


application

Wei, Yong.;

Citation

Issued Date

URL

Rights

1993

http://hdl.handle.net/10722/34268

The author retains all proprietary rights, (such as patent rights)


and the right to use in future works.

O N F A T I G U E F A I L U R E PREDICTION WITH D A M A G E MECHANICS:


THEORY AND APPLICATION

WEI YONG
Ph. D. THESIS

The University of Hong Kong


1993

DECLARATION

I hereby declare that the work reported in this thesis was done solely by me and that it
has not been submitted to the University of Hong Kong or any other institutions in
application for admission to degree diploma, or other qualification.

WEI Yong

July 1993

ACKNOWLEDGEMENT

The author wishes to thank my supervisors, Prof. C.L. Chow and Dr. B J . Duggan, for
their unfailing support and advice in the preparation of this thesis.

Thanks to the technical staff of Materials Technology Laboratory and Solids Mechanics
Laboratory of the University for their assistance.

Also, the support and advice of

colleagues and friends are gratefully acknowledged.

The author would like to thank the financial support from the Croucher Foundation.

Abstract of thesis entitled:

ON FATIGUE FAILURE PREDICTION WITH DAMAGE MECHANICS:


THEORY AND APPLICATION

submitted by

WEI Yong
B.Sc. (University of Science and Technology of China)
M.Eng. (University of Science and Technology of China)

for the degree of Doctor of Philosophy


at the University of Hong Kong
inM y 1993

ABSTRACT

Fatigue failure has been the subject of extensive studies due to its practical importance in
engineering applications. The prediction of fatigue life for most engineering structures
conventionally uses two major methods: one is centred on the S-N curve for components
without macro-cracks, and the other is the fracture mechanics methodology for
components with macro-cracks (e.g. Paris's Law). Up to the present, these two important
methods of analysis have been employed for most fatigue design analyses with a somewhat
artificial distinction with respect to crack initiation and propagation. However, such a
distinction is not based on sound reasoning or physical behaviour and is therefore
considered arbitrary. Recently, an unified approach to fatigue analysis has emerged based
on the concept of local approach. The objective of this investigation is to establish a
comprehensive theory capable of predicting the high and medium cycle fatigue failure of
engineering structures with the theory of damage mechanics. The investigation may be
subdivided into five areas:

1. A fatigue damage model has been proposed based on the thermodynamics theory of
irreversible processes with internal state variables. Depending upon the level of stress
developed within a particular material element, the damage accumulation under a loading
process is subdivided into two parts: fatigue damage and plastic damage.

With the

introduction of new damage effect tensor and damage dissipative potential function,
the damage evolution equations are formulated based on the hypothesis that overall
damage is induced by the linear summation of fatigue damage and plastic damage. When
the overall damage reaches critical value, the material element is postulated to have been
folly damaged or ruptured. The difference between fatigue damage accumulation under

tension and compression is also considered.


2. Finite element implementation of the proposed damage model is made as an acid test
to check the validity of the model. This was done by comparing the predicted results with
experimental measurements in several two dimensional components. Three associated
major problems were resolved in this computational effort: i) the treatment of asymmetric
constitutive equations resulting from the damage consideration, ii) the numerical
simulation of crack propagation, and iii) the simulation of cyclic fatigue loading. The
crack closure effect during the fatigue process was also considered in the numerical
analysis.

3. The fatigue life prediction for one dimensional problems.


A series of tests for A L 2024-T3 Alclad specimens under uniaxial cyclic loadings were
carried out to determine a number of fatigue damage variables which were considered as
intrinsic material properties required for predicting fatigue failure in complex structures.
The S-N curves at different mean stresses or stress ratios are predicted, showing excellent
agreement with the experimental results.

4. Fatigue failure prediction for two dimensional smooth specimens.


The damage model is employed to examine the fatigue failure of standard tensile specimen
under uniaxial cyclic loading. An important phenomenon observed from the test and
verified by theoretical analysis was that fatigue cracks initiate at the shoulder of the
specimen rather than in the parallel section of the specimen as in the case of ductile
fracture under uniaxial tensile loading. The model is further extended to predict the
fatigue life in plates containing an edge notch or an inclined notch under uniaxial cyclic
loading and good agreement with the experimental results has been achieved.

An

important observation from the numerical results is the ability of the model to identify the
phenomenon of stress redistribution due to material degradation during the fatigue loading
process.

5. Crack propagation under constant amplitude loadings or two-step loading.


The numerical simulation of crack propagation in a plate with a centre crack was
performed. The crack length versus cyclic loading curves were obtained both theoretically
and experimentally, and satisfactory results were obtained not only for the constant
amplitude loadings but also for the two-step loading prescribed in the investigation.

CONTENTS

DECLARATION
ACKNOWLEDGEMENTS
ABSTRACT
NOMENCLATURE

1. INTRODUCTION

2. LITERATURE REVIEW
2.1 FATIGUE OF SMOOTH SPECIMENS
2.1.1 S-N Curves
2.1.1.1 Fatigue life under completely reversed loadings
2.1.1.2 Effect of mean stress on fatigue strength
2.1.1.3 Fatigue strength under combined loadings
*

2.1.2 Fatigue Life Prediction


2.1.2.1 Basquin law
2.1.2.2 Coffin-Manson law
2.1.2.3 Fatigue life estimations under multiaxial loadings
2.1.3 Critical Plane Approaches
2.1.3.1 F-plane approach
2.1.3.2 Socie et al model

2.1.3.3 Dang Van et al model


2.1.4 Energy Criteria

17

2.1.4.1 Plastic energy criteria

17

2.1.4.2 Total energy criterion

19

2.1.4.3 AJ-based approach

20

2.1.5 Cumulative Damage Analysis for Variable Amplitude Loading


2.2 FATIGUE OF NOTCHED SPECIMENS

21
23

2.2.1 Cyclic Stress-Strain Curve

23

2.2.2 Local Stress/Strain Analysis

24

2.2.3 Fatigue analysis

27

2.3 FATIGUE OF CRACKED SPECIMENS

28

2.3.1 Long Cracks

29

2.3.2 Short Cracks

32

2.4 FATIGUE FAILURE MECHANISMS

35

2.4.1 Persistent Slip Bands

35

2.4.2 Dislocation Structures

36

2.4.3 Micro-Cracks Growth

38

2.5 APPLICATIONS OF D A M A G E MECHANICS

40

2.5.1 Introduction

40

2.5.2 Damage Variables

41

2.5.3 Thermodynamic Formulation

44

2.5.4 Fatigue Damage Models

49

2.5.5 Life Prediction in Structures

52

DAMAGE MODEL

55

3.1 INTRODUCTION

55

3.2 D A M A G E VARIABLE

57

3.3 CONSHTUTWE EQUATIONS

62

3.31 Elasticity Coupled with Damage

62

3.3.2 Plasticity Coupled with Damage

64

3.4 DAMAGE EVOLUTION

68

3.4.1 Basic Modelling

68

3.4.2 Damage Energy Release Rate Y

70

3.4.3 Plastic Damage

72

3.4.4 Fatigue Damage

75

3.4.5 Damage Accumulation

79

3.5 DETCRMINATION OF MATERIAL PROPERTIES

81

3.5.1 Damage Components d And /t

81

3.5.2 Damage Evolution Coefficient 7

82

3.5.3 Increment of Strain Hardening Threshold R(p)

83

3.5.4 Increment of Plastic Damage Threshold B(wp) and Critical Value


of Total Overall Damage

84

3.5.5 Fatigue Damage Surface

86

3.5.6 Fatigue Constants K0 and a

86

3.5.7 Summary

89

NUMERICAL ALGORITHM

90

4.1 INTRODUCTION

90

4.2 FINITE ELEMENT FORMULATION

90

4.3 SIMULATION OF CYCLIC LOADING

97

4.4 SIMULATION OF CRACK GROWTH

98

5. FATIGUE L D E PREDICTION OF SMOOTH SPECIMENS

100

5.1 INTRODUCTION

100

5.2 DETERMINATION OF MATERIAL CONSTANTS

100

5.2.1 Uniaxial Tensile Tests

100

5.2.2 Uniaxial Fatigue Tests

104

5.2.3 Summary

107

5.3 D A M A G E ACCUMULATION EQUATIONS

108

5.4 RESULTS

111

5.4.1 Experiments

111

5.4.2 Fatigue Life at Constant Stress Range

112

5.4.3 Fatigue Life at Constant Mean Stress

113

5.4.4 Effect of Rolling Direction

114

6. FATIGUE FAILURE PREDICTION OF NOTCHED SPECIMENS

115

6.1 INTRODUCTION

115

6.2 D A M A G E ACCUMULATION

116

6.3 STANDARD TENSILE SPECIMEN

117

6.4 SPECIMEN WITHA N INCLINED NOTCH

120

6.5 SPECIMEN WITHA N EDGE NOTCH

123

6.6 S U M M A R Y

126
iv

7. FATIGUE FAILURE PREDICTION OF CRACKED SPECIMENS

128

7.1 INTRODUCTION

128

7.2 EXPERIMENTS

129

7.3 NUMERICAL ANALYSIS

130

7.4 FATIGUE CRACK GROWTH


UNDER CONSTANT AMPLITUDE LOADINGS

133

7.5 FATIGUE CRACK GROWTH


UNDER TWO-STCP LOADINGS
7.6 S U M M A R Y

134
134

8. CONCLUSIONS

136

9. FUTURE WORK

139

REFERENCES

140

TABLES
EEGUKES

NOMENCLATURE

Crack length

B0

Initial plastic damage threshold

B, B(Wp)

Increment of plastic damage threshold

Transformation matrix

ACTOD

Crack tip opening displacement range

Elastic tensor of undamaged material

Effective elastic tensor of damaged material

Cep

Instantaneous tangent modulus tensor

Cep

Effective instantaneous tangent modulus tensor

Symmetric form of Cep

Characteristic dimension of a barrier. Differential symbol,


Independent component of damage tensor

df

Independent component of fatigue damage tensor

dp
da/dN

Independent component of plastic damage tensor


>
Crack propagation rate

Damage tensor

Df

Fatigue damage tensor

Dp

Plastic damage tensor

Damage rate tensor

Df

Fatigue damage rate tensor


Plastic damage rate tensor

Increment of total damage in a loading step


Damage components of D
Independent damage components
Young's modulus of undamaged material
Young's modulus of damaged material
Yield surface
Fatigue damage surface
Plastic damage surface
Shear elastic modulus, strain energy release rate
Critical strain energy release rate
Maximum strain energy release rate in a cycle
Strain energy release rate range
Plastic characteristic tensor of undamaged material
Effective plastic characteristic tensor
Rank four identity tensor
Positive hydrostatic stress
Second invariant of deviatoric stress
>
Critical value of J-integral
Maximum value of J-integral in a cycle
J-integral range
Finite element stiffness
Fatigue constant
Maximum stress intensity factor in a cycle
Measured value of fatigue strength reduction factor

Theoretical value of fatigue strength reduction factor


Hookian stress concentration factor
Stress intensity factor for mode I crack
Fracture toughness for mode I
Stress intensity factor at crack tip opening
Stress concentration factor
Threshold stress intensity factor
Strain concentration factor
Stress intensity factor range
Effective stress intensity factor range
Damage effect tensor
Number of cycles at a given stress level
Normal vector to a plane
Number of loading cycles
Number of cycles to failure
Expected life at a given stress level, number of cycles in a loading step
Overall macro-plastic strain
Overall plastic strain rate
Applied load
A set of plastic variable
Stress ratio o^/o
Initial strain hardening threshold
Increment of strain hardening threshold
Current cross-section area

Total overall damage


Critical value of w
Overall fatigue damage
Overall fatigue damage rate
Overall plastic damage
Overall plastic damage rate
Increment of total overall damage in a loading step
Elastic energy
Plastic energy
Total plastic energy up to failure
Total active energy up to failure
Plastic strain energy per cycle
Total strain energy per cycle
Active energy per cycle
Damage energy release rate, plastic damage energy release
Fatigue damage energy release rate
Components of damage energy release rate
Independent components of Y
Indqjendent components of Y
Independent components of Y f
Defined by equation (3.55)
Critical value of .Yfd on fatigue damage surface
Maximum value of Yfd
Defined by equation (3.42)

Maximum value of Ypd


Minimum value of Ypd
Damage efficiency factor
Damage evolution coefficient
Maximum shear strain amplitude
Local plastic strain range in the PSBs
Difference symbol
Principal strains
Elastic strain along a loading direction
Plastic strain along a loading direction
Elastic strain amplitude
Strain amplitude normal to the plane of the maximum shear strain amplitude
Plastic strain amplitude
Total strain amplitude
Von Mises equivalent strain amplitude
Von Mises equivalent plastic strain amplitude
Static fracture ductility
*

True strain tensor


Effective strain tensor
Elastic strain tensor
Effective elastic strain tensor
Plastic strain tensor
Effective plastic strain tensor
Plastic strain rate tensor

Lagrange multipliers
Independent component of damage tensor
Independent component of fatigue damage tensor

Mr

Independent component of plastic damage tensor


Poissons ratio of undamaged material
Poisson's ratio of damaged material
Overall micro-plastic strain
Triaxiality parameter, mass density
Uniaxial true stress
^13 ^25

Principal stresses

,act? ^2,act?
^12,801

Active stress components

^aO

Stress amplitude with zero mean stress


Maximum surface bending stress

O
b

Fatigue limit or endurance limit in pure bending


Bending stress amplitude
Von Mises equivalent stress amplitude
Hydrostatic pressure
Mean hydrostatic pressure
Maximum hydrostatic pressure
Fatigue limit or endurance limit under uniaxial loading

^10

Fatigue limit or endurance limit with zero mean stress


Maximum stress
stress

Minimum stress
Mean stress normal to the plane of the maximum shear strain amplitude
Tensile strength
Yield stress at 0.02% plastic strain
Defined by equation (3.20)
True stress tensor
Effective stress tensor
Stress tensor corresponding to
Positive part of a stress tensor
Negative part of a stress tensor
Active stress tensor
Critical value of oract on fatigue damage surface
Stress tensor on plastic damage surface
Stress tensor on fatigue damage surface
Maximum stress tensor
Minimum stress tensor
Stress range
Positive part of stress range in a cycle
Negative part of stress range in a cycle
Effective stress range
Shear stress
Shear stress amplitude
Fatigue limit or endurance limit in pure torsion
Octahedral shear stress amplitude

Mean octahedral shear stress


Shear stress range
Dissipation potential
Damage dissipative potential
Fatigue damage dissipative potential
Plastic dissipative potential
Plastic damage dissipative potential
Free energy
Free energy of undamaged material
Elastic part of free energy
Plastic part of free energy
= 0 for

= 1 ; for i=j
= 0 for i or j. = 4,5,6
= 1 for i and j = 1,2,3
Transpose of (...)
Inverse of (...)

CHAPTER ONE

1. INTRODUCTION

Fatigue failure has been the subject of extensive studies due to its practical importance in
engineering applications. Up till now, a large amount of experimental data has been
accumulated under simple laboratory testing condition. For many years, researchers have
attempted to develop a theory that will translate laboratory results to the design of
complex engineering structures.

However, a satisfactory theory has not yet been

universally approved. Consequently, many costly breakdowns of industrial structures still


occur today.

The prediction of fatigue life for most engineering structures conventionally uses two
major methods: one is centred on the S-N curve for components without macro-cracks,
and the other is the fracture mechanics methodology for components with macro-cracks
(e.g. Paris's Law). Up to the present, these two important methods of analysis have been
>
employed for most fatigue design analyses with a somewhat artificial distinction with
respect to crack initiation and propagation. However, such distinction is not based on
sound reasoning or mechanisms of material behaviour and is therefore considered
arbitrary. Consequently, there often exist problems in the transition between two stages
of fatigue failure during a loading process due to limited applicability of each method.

In the past decade, a new theory known as damage mechanics has been developed to a

stage ready for a wide range of engineering application. The theory, which introduces
a set of macroscopic internal state variables to describe the averaging effect of damage
caused by the initiation and growth of micro-cracks, is ideally suited for characterizing
material behaviour under the fatigue loading by the concept of local approach. The
primary objective of this investigation is to establish a comprehensive theory capable of
predicting high and medium cycle fatigue failure of engineering structures using material
parameters which can be determined using simple uniaxial tests on standard laboratory
specimens.

A literature review is given in chapter 2 based on the author's understanding of the


material behaviour under the fatigue loading. A fatigue damage model is proposed in
chapter 3 and the numerical algorithm follows in chapter 4.

Chapter 56 and 7

demonstrate the application of the proposed model to fatigue failure prediction of smooth
specimens, notched specimens and cracked specimens. Finally, the conclusions on this
investigation are summarize in chapter 8 and a brief description of future work is given
in chapter 9.

CHAPTER TWO

2. LITERATURE REVIEW

The reasons for fatigue failure of materials have engaged the engineer and metallurgist for
a long time because most of the failures of engineering components occur at stresses
below their tensile strength, sometimes even below their yield stress with very little
apparent plastic deformation. Our fundamental understanding of fatigue is increasing
nowadays and this can be very helpful to fatigue life prediction of engineering components
or structures. Detailed reviews of the fatigue behaviour can be found in the publications
of Frost, Marsh and Pook (1974); Fine (1980); Miller (1987); Tanaka (1989); Hertzberg
(1989), Lemaitre and Chaboche (1990) and Liu (1991). A brief review is given in this
chapter based on the author's understanding of the material behaviour under fatigue
loadings.

The discussions include (i) fatigue life estimations of smooth specimens,

notched specimens and cracked specimens; (ii) fatigue failure mechanisms; and (iii)
applications of damage mechanics.

2.1 FATIGUE OF SMOOTH SPECIMENS

Over the years, laboratory tests have been carried out in bending, torsion, direct axial or
combined cyclic loading on smooth specimens in order to describe the basic behaviour of
materials. Most tests were conducted under conditions of constant load or moment range
constant deflection or strain range, or other constant parameters for combined loadings.

For constant cyclic stress range loadings, standard definitions regarding key stress
variables are shown in Fig. 2.1 and defined by (Hertzberg, 1989)

stress

range

stress amplitude

msaii stzsss

stress

ratio

amean

R=

^max

(2.4)

Discussions in this section are restricted to fatigue life estimations of smooth specimens.

2.1.1 S-N Curves

2.1.1.1 Fatigue life under completely reversed loadings

It is generally accepted that the fatigue behaviour of materials under constant stress
amplitude loadings can be represented by the Wohler diagram (or S-N curve) in log-log
coordinates shown in Fig. 2.2 (Kocanda, 1978). Most tests are conducted under zero
mean stress conditions (R=-l). Several important facts may be seen from S-N curves.
First, fatigue life increased with decreasing cyclic stress amplitude. Second, below a
certain level of stress amplitude, which is defined as fatigue limit, fatigue failure of yield
point materials (most steels) did not occur, but it is not the case for non-yield point
materials (nonferrous metals) which have not apparent fatigue limits. Third, S-N curve

may be subdivided into two zones, namely low-cycle fatigue L C F and high-cycle fatigue
HCF. The zone L C F is bounded by a number of cycles not yet definitely established, but
most often up to 104 cycle. Usually, 107 cycles is sufficient to determine the fatigue limit
in yield point materials.

However, 5xl07 or more cycles is needed to establish an

effective endurance limit for non-yield point materials because there is no apparent "knee"
in their S-N curves (Frost, Marsh and Pook, 1974).

2.1.1.2 Effect of mean stress on fatigue strength

In fact, most loading conditions are not completely reverse. Mean stress can represent
an important role in the process of damage accumulation. It then becomes necessary to
investigate the effect of mean stress on fatigue strength.

A review of the historical

development has been made by Frost, Marsh and Pook (1974). For most materials,
fatigue life decreased with increasing amean for a given a a .

In general, empirical

relationship between fatigue strength and mean stress for a given fatigue life can be
expressed in the form
rr

= 1 _

{ C\

(2.5)

where aa is the fatigue strength when a mean stress is present, <110 is the fatigue strength
when

= 0 and ais the tensile strength. When m equals 1 or 2one obtains two

empirical design relationships, the modified Goodman relationship and the Gerber
relationship, which are generally accepted as representing the experimental results. Both
curves include two limit points, the tensile strength of the material at and the fatigue

strength at zero mean stress ffa0. The experimental tests on different materials showed that
90 per cent of the data lay above Goodman line, falling mainly between Goodman and
Gerber lines, then Goodman relationship represents a more conservative design criteria
for mean stress effects. However, some low- and medium-strength aluminium alloys give
values lying below the Goodman line, and there is no effect of mean stress on the fatigue
endurance limit when the minimum stress is negative for a few materials, such as 2014-T6
and 2024-T4 aluminium alloys (Frost, Marsh and Pook, 1974).

2.1.1.3 Fatigue strength under combined loadings

Under a combined loading, three criteria, namely maximum principal stress, maximum
shear stress and maximum shear-strain energy (or von Mises), are most commonly used
to predict the fatigue strength at zero mean stress from the corresponding uniaxial fatigue
strength. For a cylindrical bar subjected to combined bending and torsional loads, they
can be written as
Maximum principal stress criterion

[ ob+(al+4z2)

] = (3b0

(2.6)

Maximum shear stress criterion


(ai+4T2) 2 = Ob0

Von Mises criterion

(2.7)

(ai+3T2)

(2.8)

where ab is the maximum surface bending stress, r is the maximum surface shear stress
and abo is the fatigue limit in pure bending. Verification of these criteria may be obtained
by measuring the value of the ratio of the torsional fatigue limit to the rotating bending
fatigue limit (Frost, Marsh and Pook, 1974).

The experimental data fall between

maximum shear stress criterion (the ratio is 0.5) and von Mises criterion (the ratio is
0.577) for most metals. However, this ratio for various cast irons usually lies between
0.9 and 1.0, which is more close to the maximum principal stress criterion. It was found
that the experimental results for numerous ferrous alloys under alternating bending,
alternating torsion, and different combinations of alternating in-phase bending and
torsional stresses could be represented by an ellipse quadrant having as end points the
fatigue limits in pure torsion and pure bending (Gough, 1949; Gough, Pollard and
Clenshaw, 1951). The equation of the ellipse quadrant is

21

where ( and % are the fatigue limits in pure bending and torsion respectively,

and

r , are the stress amplitudes due to bending and torsion respectively at the fatigue limit
under a combined stress loading. It is obvious from equation (2.9) that this criterion
corresponds to the maximum shear stress criterion if
criterion if

equals 2 and to the von Mises

equals V 3 . Others (Findley, 1956; Rotvel, 1970 and Matake, 1977)

suggested the following general equation for different specimens instead of the ellipse arc,

i.e.

"Cio

ob0

(2.10)

It is equivalent to equation (2.9) when 0^)/ equals 2. In applying these two criteria, only
tests under uniaxial cyclic loading and torsion cyclic loading alone are required.

Another equation describing the fatigue strength under combined alternating bending and
twisting (McDiarmid, 1974) is

which allows for the effect of the normal stress amplitude acting on the plane of maximum
range of shear stress to be taken into account when calculating the maximum allowable
range of shear stress, where r a , a b , are the shear stress amplitude and the normal stress
amplitude on the plane of maximum range of shear stress. It was extended later to
>
include the effects of mean stress and stress concentration on fatigue under combined
bending and twisting (McDiarmid1985). Based on the experimental results that the
effect of mean normal stress cr1 on the plane of maximum range of shear stress is
usually much greater than the effect of mean shear stress on the same plane, only the
effect of amtm requires to be considered. Therefore, the allowable amplitude of shear
stress on this plane can be predicted from an equation of the form

where at is the tensile strength and Kt is the theoretical stress concentration factor.

Sines (1955) proposed the following criterion for fatigue failure with superimposed static
stresses on the bases of the observations that (1) the combined stress fatigue limits were
in good agreement with the von Mises criterion, (2) a mean torsional stress only had a
minor effect on fatigue strength if the material did not yield, and (3) the fatigue limit in
uniaxial or torsional stress were affected by a mean stress and the relationship was almost
linear i f the material did not yield
^0ct,a

+ a(J

where

S.mean

=P

(2.13)

is an octahedral shear stress amplitude, crHmejai is a mean hydrostatic pressure

and a , /3 are material constants which could be evaluated from uniaxial tests carried out
at zero mean stress and a pulsating tension cyclic loading. When the maximum stress
over a cycle exceeds the yield strength by around 30 to 50 percent, the linear behaviour
no longer holds.

A modified equation was proposed in order to account for the

macroscopic inelastic effects (Sines and Ohgi, 1981)


^oct.a

+ (ta

H,mean

oct.meaa

(2 . 14)

where t is a mean octahedral shear stress and a!m are experimental parameters.
Crossland (1956) also suggested a criterion close to equation (2.14) but in which the mean

hydrostatic pressure orH>m was substituted by the maximum hydrostatic pressure (rHimax.
The material determines which of these two criteria is more realistic.

These two criteria may be extended to the nonproportional loading condition (lemaitre and
Chaboche, 1990). For this case, the definition of the octahedral shear stress amplitude
should be written as

where J2 denotes the second invariant of deviatoric stress. However, there always exist
difficulties in the definition of mean stress or maximum stress for non-proportional loading
conditions.

2.1.2 Fatigue Life Prediction

2.1.2.1 Basquin law

Since the publication of Wohlers experimental results of uniaxial fatigue tests, many
investigators have attempted to express the shape of the S/N curve in mathematical form.
These suggested equations represent the relations between the number of cycles to failure
N f , the value of the maximum stress (7(or the stress range act) and the mean stress
(or the stress ratio R). Basquin's law is a typical example for high cycle fatigue (Ashby
and Jones, 1980)

10


where

(2.16)

is the stress amplitude with zero mean stress, Nf is the corresponding number

of cycles to failure, and C, m are constants which are determined from experimental data.
This law did not consider the effect of fatigue limit (or endurance limit). Other criteria,
known as Stromeyer's law, Yoshikawa's law and Weibull's law, have been proposed to
include the effect of this parameter on fatigue strength (Frost, Marsh and Pook, 1974).
However, these formulae based on the linear relationship between stress amplitude and
fatigue life in log-log coordinates are valid only for high cycle fatigue where neither the
maximum nor minimum stresses exceed the yield stress. The limitation in applications
is obvious because they are empirical formulae established on the simple uniaxial tests for
some special cases. They mostly fail in the case of low cycle fatigue.

2.1.2.2 Coffin-Manson law

For ductile materials at high stress levels, small changes in stress level will result in large
changes in fatigue life.

This has led to tests being carried out in which the strain

amplitude rather tljan the stress amplitude is maintained constant. Usually, total strain
may be subdivided into elastic strain component and plastic strain component. The plastic
strain amplitude dominates, of course, in the low-cycle range. However, as the fatigue
life increases, the elastic strain amplitude becomes more important. Due to the linear
relationship between the elastic strain and the true stress, it is easy to establish an
empirical law to represent the effect of elastic strain amplitude on fatigue life.

The

relationship between the plastic strain amplitude and the fatigue life is best described by
Coffin-Manson law (Coffin, 1954; Manson, 1954)
11

Nfm

(2.17):

where m, C are material constants. For most polycrystalline single-phase metals with Nf
< 105, Coffin found that the data obtained by reversed direct constant plastic strain
amplitude tests conformed to this relationship if m and C are 0.5 and
static fracture ductility. However, Manson preferred 0.6 and
accepted that these two parameters are not universal constants (Frost, Marsh and Pook,
1974).

Although cyclic plastic strain is responsible for low-cycle fatigue damage and thus more
likely to lead to a correlation of the data, it is the fact that the mechanical quantity
obtained directly from components in service is not a plastic strain, but a total strain
instead. On the other hand, if a designer is interested in lives in the region of 104 or 105
cycles, it would be impractical to carry out tests at the constant plastic strain amplitude
because the plastic strain part is very small. Therefore, the total strain-based life curve
is significant in fatigue-resistant design in engineering.

By adding the elastic strain

amplitude sae into equation (2.17), it can be written with the universal slopes as (Manson,
1965)
t

e + s p

= sa

(2.18)

o ^t * r-0.12+ Ey 0.6^^-0.6
=3.5N
Nf
f

where

is the total strain amplitude, crt is the tensile strength, and E is the Young's
12

modulus. This relationship is interesting because after a large number of tests on twentynine widely different metallic alloys in reversed direct loading over the range 10 - 106
cycles the data has been found to conform to this universal slopes equation. It has been
very useful over the years to estimate fatigue life when only tensile properties of the
material are known. Later, a modified equation was proposed to improve its accuracy
(Muralidharan and Manson, 1988)
\0.832
,-0.09

1.17

(2.19)
-0_53
0.155

r-0.56

0.02668/

which was derived using extensive data of fifty-seven materials including steels, aluminum
and titanium alloys. It would be expected that the total strain life curve would approach
the plastic strain life curve in the low-cycle range and approach the elastic strain life curve
in the high cycle range.

The trends are verified by many tests reviewed by Kocanda

(1978) and Hertzberg (1989).

2.1.2.3 Fatigue life estimations under multiaxial loadings

An approach that has been extensively studied for multiaxial low cycle fatigue is the
reduction of the three dimensional strain state to an equivalent scalar parameter which may
be used to replace the strain parameter in a uniaxial loading case. The octahedral shear
strain criterion has been widely used in tension-torsion or biaxial bending low cycle
fatigue (Shewchuk, Zamrik and Marin, 1968; Kikukawa, OhjiKotam and Yokoi, 1972;
13

Ellison and Andrews, 1973)but with limited success (Brown and Miller, 1982). Some
researchers proposed a modification of the Coffin-Manson equation (Libertiny, 1967;
Pascoe and De Villiers, 1967)

F=

(2.20)

c = g(C0 , M)

where

m = h(m0

is the equivalent strain amplitude for which different expressions have been

chosen according to the testing results, C0 and 11 are the values of C and m for the
uniaxial case, M is a stress- or strain-dependent parameter. Such theories axe no longer
widely used, since they require extensive experimental studies for the determination of the
empirical parameters (Lefebvre, 1989).

2.1.3 Critical Plane Approaches

The review in the last section shows that no single criterion has been proposed that will
correlate all experimental data on multiaxial fatigue, particularly in the high-strain regime.
Recently, the method known as the critical plane approach has become the basis of several
efforts in multiaxial fatigue. This approach deals directly with the multiaxial aspects of
the failure based on a critical plane that is a function of stress state, strain path,
temperature, and other as yet undefined parameters, such that it may be used to predict
the multiaxial fatigue behaviour under non-proportional loading conditions.
approaches are as follows.

14

Several

2.1.3.1 F-plane approach

It was suggested that a plot of the maximum shear strain amplitude against the tensile
strain amplitude normal to the plane of maximum shear would illustrate the controlling
processes in fatigue crack growth at each state of strain (Brown and Miller, 1973). Based
on the metallurgical evidence, they proposed that the maximum shear strain governed
plastic deformation, hence crack initiation and the cracks were assisted in propagation by
the normal strain on the maximum shear plane. The proposed equation on constant life
contours was written in terms of the principal strains as
(2.21)

where > s^. This criterion can be written in another form for a given fatigue life
(Kandil, Brown and Miller, 1979; 1982)
(2.22)

+ C e" = constant

where %are the maximum shear strain amplitude and the strain amplitude normal to
the plane of maximum shear strain amplitude and C is determined experimentally. This
approach is capable of correlating the experimental data for completely reversed biaxial
loading.

2.1.3.2 Socie et al model

In order to consider the effect of mean stress, a three parameters description of the shear
15

damage process has been proposed (Socie, Wail and Dittmer, 1985; Socie, Kurath and
Koch, 1989)

Ya +

ta
is

t0

where an0 is the mean stress normal to the plane of the maximum shear strain amplitude,
S and T are the material constants, E and G are the tensile and shear elastic modulii, 7 f 0
is the torsional fatigue shear ductility coefficient, Tf0 is the torsional fatigue shear strength
coefficient, and cb are torsional fatigue ductility and strength exponents respectively.

It is important to note that the above two models are based on the assumption that shear
damage is dominant, i.e. crack nucleation and growth on planes of maximum shear strain
amplitude. Therefore, the models are not appropriate for materials which may nucleate
in shear, but spend the majority of life in mode I tensile crack growth (Socie, 1987;
Morrow and Kurath, 1989).

The appropriate fatigue failure criterion is material

dependent.

2.1.3.3 Dang Van et al model

For high cycle fatigue, Dang Van et al proposed a criterion based on some physical
understanding, namely as cracks usually occur in intragranular slip bands, the local shear
stress acting on these planes is an important parameter and the hydrostatic stress

also

an important parameter with regard to the opening of cracks (Dang Van, Cailletaud,
Flavenot, Douaron and Lieurade, 1989; Dang Van, Griveau and Message, 1989)i.e.
IX

Max{n) Max[ t)

(2.24)

where T(t) is the shear stress on a plane with the normal vector -n, U

is the mean

shear stress on this plane during the cycle,ffH(t) is the hydrostatic pressure and ab are
the material constants. Usually,

is defined by the centre of the smallest circle

containing the path of r(t) in the plane. The radius of this cycle is taken to be the shear
amplitude (Lemaitre and Chaboche, 1990)

The essential difference between equation (2.24) and that of the Sines and Crossland
criteria mentioned in Section 2.1.1.3 is that the quantities appearing in the Dang Van
criterion, shear stress amplitude and hydrostatic pressure, are defined at the same instant.
This criterion also gives a critical plane, so that the direction of crack initiation can be
determined.

Due to three maximization in equations (2.24) and (^2.25)two time

parameters t, to at each plane and one parameter of the direction of the normal vector n
this criterion is generally very complex.

2.1.4 Energy Criteria

Another class of criteria which has received significant attention are energy-based
v.

approaches for multiaxial fatigue. Several models are as follows.

2.1.4.1 Plastic energy criteria


17

The concept that the resistance of a material to fatigue failure can be quantified by the
plastic work required to cause failure or the cyclic strain energy density has been adopted
by several investigators. The following relation is postulated to relate the plastic work per
cycle aW11 and fatigue life to crack initiation Nf based on the assumption that most of the
damage is accumulated during the stable cyclic response

Nf =

where F is a monotonically decreasing function of a W . A power law relation appears


to fit the uniaxial test data with sufficient accuracy over Nf = 50 cycles to 106 cycles for
many materials (Garud, 1981). Garud (1981) proposed a new "hardening rule" which he
incorporated into material constitutive relations to compute the plastic work per cycle
under complex loading with moving principal stress axes. Fatigue life prediction using
this approach requires only the-uniaxial cyclic stress-strain curve and the uniaxial fatigue
test results from smooth specimens.

Another criterion has been developed for high-strain biaxial fatigue with the plastic work
required for failure during the cycling process (Lefebvre, Neale and Ellyin, 1981)

= KN
where the equivalent plastic strain amplitude

and the equivalent stress amplitude

are based on the von Mises yield function of plasticity, K and c are parameters which
depend on the stress ratio and mechanical properties of the material.

IX

Despite plastic energy criteria having been found to be very useful predictors for
multiaxial fatigue life in experimental tests, they are difficult to integrate into practical
procedures involving structural analysis because the small values of plastic work per cycle
is not easy to determine with much confidence.

Furthermore, the plastic work Wf

required to cause failure is not a material constant at all life levels (Leese, 1988).
According to the experimental observations that a major part of this energy is dissipated
into heat, and the remaining mechanical energy causes dislocation movements along slip
lines and volumetric changes, and eventual material damage, a parameter W* which
represents the active energy stored in the material element has been proposed for the
uniaxial case (Lemaitre and Chaboche, 1990), i.e.
AW* = AWp -2ay Atp
(2.28)

W*

where aW* is the active energy per cycle, ay is the yield stress and Wf is the total plastic
work up to failure.

It is interesting that the total active energy Wf* is approximately

constant at all live levels and may be considered intrinsic to the material 316 steel.

2.1.4.2 Total energy criterion

A failure criterion for the multiaxial fatigue was proposed relating the total strain energy
per cycle aW1 to the number of cycles to failure Nf together with an imposed triaxiality
parameter

19

ATT' = ( ap+b ) Nf m +

p =

(2.29)

2e2

(er^)

where
evaluated. aW 1 includes the plastic work and the positive part of the elastic strain energy
and is suggested to be an unifying parameter which can describe both the initiation and
subsequent propagation of fatigue crack. It has the general form of
AW

= AW ( J2, /

where J2 is the second invariant of the deviatoric stress and If3* is the maximum value
of positive hydrostatic stress. Therefore, this criterion is hydrostatic stress sensitive.

2.1.4.3 AJ-based approach

An energy-like approach, which is motivated by the concept of AJ-integral, was proposed


by McDowell and Beraxd (1992) for correlation of biaxial fatigue.

Two microcrack

propagation laws were first introduced with the concept of elastic-plastic AJ-integral for
predominantly low cycle and high cycle fatigue regimes. Then the relationship between
maximum engineering shear strain range and fatigue life was obtained by integrating two
microcrack propagation laws. This approach has demonstrated the capability of containing
the most desirable features of plastic work type criteria and critical plane approaches.
Such an approach may be used to characterize fatigue behaviour in the transition between
short crack and long crack.

20

2.1.5 Cumulative Damage Analysis for Variable Amplitude Loading

The review discussed thus far did not concern variable amplitude loading.

However,

many structures are subjected to a range of load fluctuations and mean levels in actual
field service conditions. Therefore, it is important to predict, based on constant amplitude
test data, the fatigue lives of components or structures subjected to variable amplitude
loadings. Many attempts have been made to deal with this project since Palmgren first
suggested what is known as a linear damage rule (Frost, Marsh and Pook, 1974). The
same rule was later independently proposed by Miner (1945) based on the assumption that
damage of a specimen at a given stress level can be considered to accumulate linearly with
the number of cycles and failure occurs when the accumulated damage reaches a critical
value. For a multi-level test, the linear damage rule (LDR) is

(2.31)

where n
is the number of cycles at a given stress level, Nis the expected life at the same
stress level. Although this rule is commonly used in practical applications, it does not
conform to the long established fact that the life of a specimen is found to depend on the
order in which the stresses are applied. A wide range of values for the sum of cycle
ratios was obtained, but most values lay between 0.3 and 3.0. The sum of the observed
cycle ratios is usually less than unity when high stress loadings are applied first and low
stress loadings are subsequently applied until failure occurs. Conversely, in low-high
stress tests on plain specimens this value is often found to be greater than unity.

21

The inaccuracy of LDR led to a number of alternative methods being proposed to predict
the fatigue life under complex loading condition (Manson and Halford, 1986).

An

approach called damage curve analysis (DCA) has been used to reflect the interaction of
loading at different stress levels by the concepts of cumulative fatigue damages.

In a

general case when K loadings are applied before failure occurs, the equation for DCA is
(Manson and Halford, 1981)

.0.4

(Ay//j)0
(2.32)
("jtW 4

'r-i

where the subscripts 1, 2, 3, ..., K-l, K are the sequence numbers of the loadings
applied.

It is interesting that the only constant in equation (2.32) is the exponent 0.4

which can be taken as an universal value for most materials.

In two-level loadings, the

application of equation (2.32) is limited for low values of iij/N,. In order to improve the
life prediction in this troublesome range, a new approach called the double-damage curve
approach (DDCA) has been proposed by adding a term

that would have a large

significance at low values of n/N!but only relatively small effect for larger values
(Manson and Halford, 1986).

Another approach which has been widely used is the double-linear damage rule (DLDR).
Based on the concept that fatigue damage is at least a two-stage process, phase I and phase
II, and the two processes developed at different rates for different life levels, two damage
22

rules were established respectively for determining the two phases of life (Manson and
Halford, 1981). Instead of using the Nf curve to compute the cycle ratios in conventional
LDR analysis, the Nj curve is used to calculate the sum of cycle ratios in the conventional
manner until the summation reaches unity for which phase I is now complete. Then for
continued loadings, summation of cycle ratios is made using the Nn life curves. When
the phase n sum reaches unity, failure is presumed to occur. No material constants other
than those normally used in a conventional linear damage rule analysis are required for
application of this approach. The DLDR approach is intended as a simplification of the
DCA concept, and prediction by both methods are generally reasonably consistent.
Therefore, the DLDR may be usually regarded as preferable.

2.2 FATIGUE OF NOTCHED SPECIMENS

Since components generally have various notches, fatigue crack initiation at the notch root
is of great significance in engineering. Unlike smooth specimen, stress/strain distribution
in notched specimen is not uniform. Therefore, local stress/strain plays an important role
to fatigue crack initiation. In many applications, local stress/strain can be calculated only
from FEM analyses. In this case it is important to establish a accurate cyclic stress-strain
relationship. Before discussing numerical methods of estimating local stress/strain, we
first review previous work on material response under cyclic loadings.

2.2.1 Cyclic Stress-Strain Curve

By monitoring stress and strain during a cyclic loading test, the response of material can
23

be identified (Hertzberg, 1989). If only elastic deformation is involved under the applied
loads, the cyclic stress-strain curve is shown in Fig. 2.3a. Conversely, the cyclic stressstrain curve shown in Fig. 2.3b that reflects both elastic and plastic deformation is
produced. The area enclosed by a loop represents the plastic deformation work done on
the material. For most materials, the shape of hysteresis loop changes with continued
cycling until cyclic stability is reached. According to the change, the material is said to
cyclic hardening or softening.

The cyclic hardening and softening response of metal has been studied intensively
(Neumann and Tonnessen, 1987; Hatanaka, 1990). Many experimental studies shown that
cyclic deformation behaviour is very complex in detail, but with reasonable accuracy the
cyclic stress-strain curve may be written as an unique power law. However, it is not the
case for nonproportional loading where the principal axes of strain rotate during cyclic
plastic flow (McDowell and Socie, 1985). For this case, the cyclic stress-strain curve
based on equivalent stress and strain is not unique but depends on the degree of
nonproportionality of loading. A path-dependent integral was proposed by McDowell and
Socie (1985) to correlate the experiment results for several complex nonproportional paths
under axial-torsional loadings. Only uniaxial and 90 out-of-phase sinusoidal test results
are required by this integral representation.

2.2.2 Local Stress/Strain Analysis

To predict fatigue life of crack initiation (millimetre sized cracks) for notched specimen,
the so-called notch analysis concepts are often used to account for the local stresses/strains
24

(Tipton and Nelson, 1985). When elastic analysis is conducted, the maximum principal
stress at a notch is used to enter a smooth specimen S-N curve to predict fatigue damage.
Usually, the maximum principal stress is estimated with the geometric elastic stress
concentration Kt or the fatigue strength reduction factor K f which equals the ratio of the
plain fatigue to the corresponding notched fatigue limit,

The experimental results on the

relationship between Kt and K f may be summarized as follows (Frost, Marsh and Pook,
1974)

1. For low values of IQ, K f may equal to Kt, but in general is somewhat less.
2. For high values of Kj, K f is often very much smaller than Kt.
3. Different geometries producing the same Kt value may give different Kf values.
4. For a given material and certain notch geometries, there appears to be a particular
value of Kt at which K f reaches a maximum value. Higher values of K, result in no further
increase in Kf.

Because K f factors cannot be known in advance of testing, attempts have been made to
formulate relationships between Kf and Kt for minimizing the need to test a large number
of notched specimens of different materials. A possible equation has been proposed by
the assumption that the discrepancy between these two factors results from the influence
of notch stress biaxiality when the notch strain is elastic (Leis and Topper, 1977)

25

(2.33)

Aa2

where Kis the theoretical value of Kf, ac^ and A(X2 are the ranges of maximum principal
and transverse stress respectively. This approach produced good estimates of long-life
fatigue strength for blunt notches. When an elastic-plastic analysis is performed, Hookian
stress concentration factor KH in Neuber's rule is widely used for local stress-strain
analysis and low cycle fatigue life estimation. It may be expressed as (Neuber, 1961)
4

(2.34)

= Kt K t

where Kj is the notch stress concentration factor and K<. is the notch strain concentration
factor.

Another powerful approach for local stress/strain analysis is the finite element method
(FEM) which can perform detailed evaluations of the elastic-plastic stress/strain fields at
all fatigue critical locations of specimens, especially for some areas where the stress/strain
situation becomes so complicated that simple evaluation methods are no longer available.
In the case of the flat plate, the FEM results axe very close to the Neuber results.
However, the Neuber-predicted strains are much larger and the stress ranges smaller than
those predicted by FEM for the notched bars. This suggests that the Neuber approach is
applicable for plane stress analyses but requires further modification for notch analyses
where multiaxial stresses and strains develop (Domas and Antolovich, 1985).
26

To summarize, two classes of methods axe often used to account for the local
stresses/strain: notch analysis concepts, which are simple to apply to various specific
design problems but unsuitable in complex loading conditions, and FEM based methods,
which usually give more accurate results but need large computers (Nowack, Ott, Foth,
Peeken and Seeger, 1988).

Both approaches exhibit advantages but also certain

limitations.

2.2.3 Fatigue analysis

Based on the results of local stress/strain analysis, some criteria developed for smooth
specimens can be used to estimate fatigue life. Examples are: (1) maximum shear and
octahedral shear stress criteria, (2) maximum shear and octahedral shear strain, (3) cyclic
plastic work or total strain energy, and (4) other correlation parameters that incorporate
both local stress and local strain. The ability of these criteria to predict fatigue life is
highly dependent on knowledge of local stress/strain distribution around a notch.

So long as the stresses and strains remain predominantly elastic, conventional fatigue life
calculation methods may be applied (Nowack, Ott, Foth, Peeken and Seeger, 1988). The
criteria based on the maximum shear and octahedral shear stress/strain are still suitable
for elastic-plastic cases if the local stresses and strains remain proportional. The energy
criteria usually provided reasonably good predictions for both in-phase and out-of-phase
loadings. However, a drawback to the approach for use in long life fatigue evaluation is
that plastic energy is generally so small that slight variations in its computation can lead
to large differences in predicted life (Tipton and Nelson, 1985).
27

When the local

stresses/strains become non-proportional, life prediction approaches are still in an early


stage of development.

Along with the introduction of fatigue criteria mentioned above another concept which can
not be avoided is that of a "critical" notch root length/volume. Real materials cannot be
sub-divided indefinitely without affecting their mechanical properties. Neuber suggested
that there may be an elementary structural unit of length A, measured inwards from the
notch root on the plane of minimum cross-sectional area, over which the average cyclic
stress must exceed the plain fatigue limit of the material before cracking can occur.
Others preferred the cyclic stress/strain at that point with a distance h below the notch
root, instead of the average value, to predict the fatigue failure (Frost, Marsh and Pook,
1974).

Either A or h is assumed to be material constant.

This concept has been

successfully used to explain the effect of specimen diameter on the fatigue limit which
sometimes occurs in the rotating bending fatigue tests, and to predict the rotating bending
fatigue limit from the reversed direct stress fatigue limit. Domas and Antolovich (1985)
used a critical notch root volume to define the average notch energy density in the fatigue
failure criterion and found that plausiblefirst choice for this parameter was the entire
plastic volume surrounding the notch root.

In some studies, the critical size was

correlated to microstructural features, such as dislocation cell sizes in solid solution alloys
and precipitate spacing in precipitation hardening alloys (Chanani, Antolovich and
Gerberich, 1972; Antolovich, Saxena and Chanani, 1975). However, it is always difficult
to choose an appropriate characteristic size/volume when it is needed.

28

2.3 FATIGUE OF CRACKED SPECIMENS

2.3.1 Long Cracks

Since Paris and Erdogan (1963) first obtained the fourth-power law between the crack
propagation rate da/dN and the stress intensity factor range
K

= c (A 4

(2.35)

the cyclic stress intensity factor (SIF) range AK has been widely recognized as the useful
fracture mechanics parameter for analysing long crack growth (Frost, Pook and Denton,
1971; Paris, 1981; Paris and Hermann, 1981; Chand and Garg, 1985; Tanaka, 1989; Liu,
1991). Many workers in the following three decades have shown that the exponent of the
power relation is not a universal constant, but varies between two to seven for different
materials. Furthermore, the power law is correct only in the intermediate rate regime
from 10"9 to 10"6 m/cycle. Many formulae have been proposed for describing the fatigue
crack growth in low AK regimes, where there is a threshold stress intensity range, and
in high AK regimes, where unstablefracture starts at the maximum stress intensity factor
which equals the fracture toughness. These two special cases can be expressed as

-- f

- 0
(2.36)

where KIe is the fracture toughness for mode I and


factor.

is the threshold stress intensity

Among the hundreds of mathematical models proposed during the 1960s and
29

1970s, Forman's equation is a typical example which includes the effect of stress ratio R
(Forman, Kearney and Engle, 1967)
C (AiO m
{\-R)KIC-^K

(2.37)

Apart from these two empirical laws (Eqns. 2.35 and 2.37), some researchers proposed
crack propagation laws on the bases of theoretical considerations, such as deformations
ahead of crack tip (Raju, 1972; Schwalbe, 1974 and Duggan, 1977) or crack tip geometry
(Frost and Dixon, 1967 and Pook and Frost1973). A large number of unknown material
constants are needed to be determined experimentally.

Unfortunately, no universal

equations have been found.

For ductile materials, Elber (1971) noted that fatigue crack surfaces interfere with each
other through closure even for tension-tension cyclic loading. This is because the plastic
zone at the crack tip induces a field of residual stresses which tends to delay the opening
of the crack under increasing load. The effect of the mean stress on the crack growth rate
in low and intermediate AK regimes may be attributed to crack closure. So the effect of
stress ratio R diminishes in the Elber equation

(2.38)

when K is replaced by an effective stress intensity factor range


computed as

minus the SIF value at crack tip opening K ^ .

30

which was

The concept of crack closure can be applied to the fatigue crack growth under variable
amplitude loading. Once the value of is

determined experimentally or theoretically,

the subsequent crack growth rate can be predictedfrom the da/dN - aK^ relation obtained
in the test under constant amplitude loading (Tanaka, 1989). Another prediction method
for the cases of variable amplitude loading is based on Wheeler's model by estimating the
plastic zone size and including the load interaction effects (Lemaitre and Chaboche, 1990).

The stress intensity factor range K can be used as a loading parameter to describe the
fatigue crack growth only under small scale yielding (SSY) conditions. The rate of fatigue
crack growth under elastic-plastic and gross plasticity conditions is higher than that
predicted from the da/dN - AK relation (Tanaka, 1989). J-integral range AJ or strain
energy release rate range AG have been proposed to replace AK by many researchers
(Lamba, 1975; Dowling and Begley1976; Wuthrich, 1982; Tanaka, 1983; and Chow and
Lu1991). The fatigue crack propagation laws proposed by Chow et al with AG (Woo
and Chow, 1984; Chow and Woo, 1985; Chow, Woo and Chung) or AJ (Chow and Lu,
1990) are respectively
( A G ! ) m
G - G .

(2.39)

( AJ ) m

(2.40)

and

where Gc is the critical value of energy release rate, Jc is the critical value of J-integral,
and

are the corresponding maximum values. Under SSY conditions, AJ reduces


31

to the range of elastic energy release rate AG.

Other parameters, such as crack tip

opening displacement range A C T O D and plastic zone size PZS, have also been proposed
to replace AK.

Among these parameters for large scale yielding cases, the J-integral

range seems to be most successful (Tanaka, 1989; Chow and Lu, 1991).

Based on the local approach concept, a method was proposed by some researchers (Sih
and Moyer, 1983; Moyer and Sih, 1984; Sih and Chao, 1989) in which the crack growth
process is simulated by predicting a series of crack growth steps which are assumed to
occur when the material elements ahead of the crack tip accumulate a critical amount of
stored strain energy density. According to this approach, the fatigue crack growth can
be predicted by FEM analysis with some material constants measured by simple uniaxial
tests.

However, it is difficult to compute the accumulated strain energy density.

Furthermore, the fatigue life predicted by this theory is too conservative because of the
hypothesis that the stored strain energy density during each cycle is equal to the amount
accumulated during the first cycle of the current increment of crack growth, specially for
tension-tension fatigue loading. In addition, the numerical results are strongly dependent
on the finite element size.

2.3.2 Short Cracks

When the crack length is short, the applicability of the conventional fracture mechanics
approach is questionable (Leis, Hopper and Ahmad, 1986; Nowack and Marissen, 1987;
Skelton, 1988; Mcevily, 1989). Experimental results have shown that short crack growth
rates are higher than those predicted by the formulations established for long cracks. This
32

is due to the breakdown of the similitude concept which is the basis of these conventional
formulations. Three principal reasons have been proposed for the high growth rates of
short cracks: microstructural effects, premature crack closure, and macro-plasticity effects
(Tanaka, 1987).

When the crack length is of the order of the material microstructure dimensions, such as
the grain size, the fatigue crack growth is mainly influenced by microstructure. The crack
growth rate will be different due to microstructural inhomogeneity even if the same
loading parameter is applied to the crack.

These short cracks are named as

microstructural short cracks (MSG). The cracks grow first along the slip direction with
several kinks, then often decelerate when they hit the grain/phase boundaries or when they
make sharp bends.

The interaction of a propagating crack with grain and phase

boundaries has been studied by several investigators (Tanaka, 1989).


opening displacement range

ACTOD

The crack tip

was chosen as a controlling parameter to derive

irregular growth of MSG by assuming the crack growth rate as a power function of
ACTOD.

Generally, the configuration of the crack tip is complicated, and the

measurement of

ACTOD

is not practical except for some special cases. Therefore, the

reversible plastic zone size was taken as the controlling factor instead of

ACTOD

by

assuming that the crack growth rate is proportional to the reversible plastic zone size
(Nishitani and Kawagoishi1992). Another equation was proposed (Miller, 1987)
= C

(Ay)"1 (d-a)

(2-41)

where Ay is the shear strain range, d is the characteristic dimension of the barrier, a is
33

the crack length, and C, m are material constants.

This exhibits a decreasing crack

growth rate as the crack length increases in the range 0 < a < d for a constant stress or
strain range and that when a = d, the crack growth rate is zero.

However, the

mechanical equation for MSG growth is not well established due to the complicated
behaviour of fatigue cracks in this region.

It is argued that statistical investigation is

necessary to include the interaction of cracks with the material microstructure.

When the crack length is larger than several times the grain size, but less than the smallest
length of long crack, the crack is named as physical short crack (PSC) or mechanical short
cracks.

It is very difficult to define these three regions because the boundaries are

dependent on the dominant barrier and material properties. In this region, the plastic zone
size is usually large compared with the PSC length, and the crack closure is not fully
developed because of the short crack wake. The PSC growth rate is strongly and more
dependent on stress level rather than material microstructure, so that an appropriate choice
of fracture mechanics such as AK^ or A J will make crack growth prediction possible
(Tanaka, 1989).

Some investigators argued that it was better to provide directly the

appropriate crack growth rate for a given stress/strain range and crack length combination.
Two examples are given: one is based on the shear strain range a7 (Miller, 1987)

=C
dN

(Ayr

(2.42)

a - B

and the other is based on the stress range ao* (Nisitani, 1981)

34

C (Aa) B

(2.43)

where C, B and m are material constants.

2.4 FATIGUE FAILURE MECHANISMS

Fatigue failure is a very complex process.

It may be divided into several stages: (1)

cyclic plastic deformation prior to fatigue crack initiation, (2) microcracks initiation and
growth, (3) macrocracks form and propagate tofinal failure. In spite of the fact that the
principal micromechanisms on these stages are well known, there is still a lack of
quantitative description on the first two stages (Fine, 1980; Ceroid and Meier, 1987;
Neumann and Tonnessen, 1987).

2.4.1 Persistent Slip Bands

It has long been generally recognised that fatigue cracks in pure metals are associated with
persistent slip bands (PSBs).

When a fatigued specimen was lightly electropolished,

although many of the striations on the specimen surface disappeared, some, which are
called PSBs,

remained (Thompson,

Wadsworth

and Louat,

1956).

Continued

electropolishing eventually remove them, but PSBs form again in the same places when
the specimen is reloaded. For single crystals the PSB can be found throughout the whole
volume, so persistent slip is a bulk phenomenon in single crystals.

It was suggested by

some researchers that the PSBs are a surface phenomenon in polycrystals and their only
35

role is to aid the crack nucleation.

However, Rasmussen and Pedersen (1980) have

revealed the existence of PSBs in the interior grains in polycrystalline copper crystals.
It can be argued that it is easier to form PSBs in the grains on surface than in the interior
grains due to the effect of micro stress/strain concentration.

The PSBs usually pass

through twin boundaries, but they rarely cross grain boundary. Each grain tends to
contain only one set of parallel PSBs. If a grain contains more than one set of PSBs
there is a tendency for each set to occupy its own area.

The local plastic strain range in the PSBs

was found to be the controlling parameter

which correlates best with the life via a Coffm-Manson type relationship (Hunsche and
Neumann, 1986)

Nf = C (AY

In common experiments with single crystals

is independent of the applied plastic

strain range 7 within the plateau of the cyclic stress-strain curve. Therefore, there is
almost no dependence of specimen lives to the applied plastic strain amplitude in the range
from 0.1 to 1 % for copper single crystals.

2.4.2 Dislocation Structures

There is an extensive body of literature on dislocation structures in the fatigued materials


(Winter, 1974; Winter, Pedersen and Rasmussen, 1981; Tabata, Fujita, Hiraoka and
Onishi, 1983; Ackermann, Kubin, Lepinoux and Mughrabi, 1984; Jin, 1989).

The

dislocation arrangements in the cyclical saturated crystals are dependent on the applied
36

plastic strain amplitude. In particular, the well known ladder structure of the PSBs and
vein structure of the matrix are formed in copper single crystal when the shear strain
amplitude is low. PSBs almost always appear as narrow regions of wall structure parallel
to a {1 1 1} slip plane, and this ladder structure has a lower flow stress than the
surrounding matrix. This configuration is ideally suited to allow the PSBs to suffer plastic
deformation independent of the elastic matrix. The PSB pattern in the polycrystal suggests
a similar tendency for individual grains to avoid persistent slip on more than one slip
system and there is no obvious difference in dislocation structures comparing with that in
single crystal. So the Sachs model may be used to establish the relationship between the
polycrystal fatigue limit and single crystal plateau, and close agreement with experiment
has been obtained (Rasmussen and Pedersen, 1980).

It has been shown experimentally that there is no relationship between the slip system
active in a particular grain and its orientation with respect to the tensile axis (Winter,
Pedersen and Rasmussen, 1981). This suggests that the stress states in a given grain are
dominated more by its neighbours than by the applied stress states due to large local
variations in the stress field.

In addition to the PSBs, two other types of dislocation structure, misoriented cells and
labyrinth, are found at higher strain amplitudes. They are probably products of multiple
slip.

Some investigations have shown that these two structures are not active and their

function is to exclude PSBs from sizeable regions of the specimen (Winter, Pedersen and
Rasmusen, 1981).

37

2.4.3 Micro-Cracks Growth

Fatigue micro-cracks usually initiate at sites that are related to strain localization. They
may begin along slip bands, in grain boundaries or at constituent particles, and the mode
of fatigue crack initiation depends on which occurs most easily. Actually, it is difficult
to demonstrate and to study the initiation of micro-cracks due to fatigue because the
fatigue crack must be a certain length before it can be observed.

The value mainly

depends on the resolving power of the test system used, or researchers' definition.
Experimental evidence shows that fatigue cracks when first initiated are extremely small,
less than 0.5 /xm (Lin, Fine and Mura, 1986). Therefore, the origin of the fatigue cracks
which formed at grain boundaries or constituent particles is often obscure. For example,
some investigators proposed that such cracks near particles in high strength aluminum
alloys were due to debonding between the constituent particle and matrix.

However,

Kung and Fine (1979) argued that no direct evidence for such a mechanism was found in
their research. They suggested that the cracks seemed to form where slip bands in the
matrix impinged on constituent particles. In some cases it could be shown that the initial
cracks were parallel to slip planes. It was confirmed by many investigations that cracked
constituent particles in high strength aluminum alloys either formed during processing or
during fatigue loading do not initiate fatigue cracks in the matrix. This is because the
stress required to extend the crack into the matrix is larger than that to initiate the crack
in the particles (Fine, 1980).

A mathematical model for fatigue crack initiation proposed by Lin, Fine and Mura (1986)
is based on accumulation of dislocations in slip bands.
38

In the model, a systematic

accumulation of dislocation dipoles along slip band matrix boundaries from fatigue cycling
was derived using the theory of continuous distribution of dislocations.

Cracking was

assumed to occur when the accumulated dislocation dipoles reach the number to give the
critical displacement to cause fracture in a perfect metal.

It should be noted that they

introduced a damage accumulation efficiency factor and assumed that it was equal to the
increase in dislocation density which was subject to dynamic recovery especially during
the reverse cycle.

In Kung and Fine's investigation on aluminium alloys, several parallel cracks formed in
the interior of a grain and spread to the nearest grain boundary (Kung and Fine, 1979).
At first these cracks stop at the grain boundaries and did not grow further. Examination
of many such cracks has shown that they are parallel to slip plane traces. Similar results
in pure iron have shown that an early crack penetrates into grains along one or more
dominating slip systems having a common slip direction in respective grains, regardless
of whether it nucleated at a grain boundary or along a slip band (Tanaka, 1987).
Accordingly, the propagation of early micro-cracks occurs in the shear mode.

When

micro-cracks extend to more than one grain, they have a tendency to orient themselves
perpendicular to the direction of maximum principal stress. There are usually two cases:
one is that the crack becomes discontinuous and the microcracks penetrate into successive
grains individually, and the other is that crack propagates to an adjacent grain having a
preferential slip system whose orientation is close to that in the first grain.

When the

crack size becomes significant with a well-defined direction, it grows in a preferential


way,

partially

unloading other microcracks and generating a high stress concentration at

its tip (Lemaitre and Chaboche, 1990).


39

2.5 APPLICATIONS OF DAMAGE MECHANICS

2.5.1 Introduction

Since Kachanov (1958) published thefirst paper on creep failure of metals under uniaxial
loads, damage mechanics has been developed to a stage ready for a wide range of
engineering applications (Chaboche,

1981; Krajcinovic,

1984; Lemaitre,

1986a;

Chaboche, 1986; Murakami, 1987b; Hult, 1987). There are several important differences
between fracture mechanics and damage mechanics.

Conventional fracture mechanics

deals with the failure problems in structures containing major cracks which are assumed
to be embedded in defect free materials.

This method is usually based on a global

analysis by correlating the critical state or the crack growth rate to some global fracture
parameters, such as K, J, etc. It sometimes has limited applicability to very sophisticated
problems of damage and fracture (Lemaitre, 1986a).

On the other hand, damage

mechanics is a local approach to failure prediction for structures with or without macrocracks where the material itself is damaged due to the presence of micro-defects. The
process of the initiation and propagation of macro-cracks caused by the progression of the
local damage is analyzed by numerical methods with a unified approach.

Most damage mechanics models are based on a macro-mechanics approach by ignoring


the fine detailed information of micro-defects and employing some internal variables to
characterize, at the macro-scale, the material behaviour including material stiffness
degradation, damage-induced anisotropy, crack initiation and propagation, etc.
40

There

were also some micromechanical damage models proposed in last decade (Krajcinovic and
Fanella, 1986; Hult, 1987; ICrajcinovic and Sumarac, 1989; Ju, 1991). B u t ^ pointed
out by Krajcinovic (1991), micromechanical models are computationally inefficient in
practical applications despite their great advantages in modelling physical reality with a
minimum of ambiguity and arbitrariness. Further research on micromechanical models
is obviously necessary. The following review is focused on the damage mechanics models
based on a macro-mechanics approach.

2.5.2 Damage Variables

Since the deterioration of engineering materials as a result of the nucleation and growth
of distributed micro-defects is very complex, it is necessary to introduce some averaging
macro-variables in order to establish practically usable approaches.

It is assumed that

there exists a representative element which is large enough to include a sufficient number
of micro-defects so that a discontinuous state on the micro-scale may be represented by
averaging continuous macro-variables, but is small enough so that it can be taken as a
mathematical point to define mechanical variables, such as stress, strain, etc. The scale
of this material element is given by Lemaitre (1987) as follows:

0.05 to 0.5 mm for metals,


0.1 to 1.0 mm for polymers,
1.0 to 10.0 mm for woods,
10.0 to 100.0 mm for concrete.

The success of a damage model to a large extent depends on the definition of damage
variables which is related to the concept of effective stress. The effective stress tensor
:can be expressed in a generalized form as (Lemaitre and Chaboche, 1990)

where o
r is the true stress tensor and M(D) is the damage effect tensor. Most of the early
work was based on a scalar damage variable D which may be interpreted as the net area
reduction due to the development of micro-defects. For isotropic damage, equation (2.45)
becomes

In order to extend this damage model to the case of anisotropic damage, two modified
models were proposed in the principal stress coordinate system.

One proposed by

Kachanov in 1974 on the basis of introducing three scalar variables D!, D2 and D3 to
characterize the net area reductions of the cross planes normal to the direction of three
principal stresses q and
aj1 =

02 =

a,
3 =

1^3

(2.47)

Another, based on the assumption that only the majcimum principal stress is affected by
damage evaluation (Hayburst and Leckie, 1973) gives

42

Y-D

= a

3 =.a3

(2-48)

These two models are improper for the cases of rotating principal stress directions. Other
investigators introduced a vector (Krajcinovic and Fonseka, 1981) or a second rank tensor
(Murakami and Ohno, 1981Cordebois and Sidoroff, 1982) as damage variable to
describe the anisotropic damage behaviour. While it is always true that a vector or tensor
variable can store more information than a scalar variable and more reasonable from the
physical point of view, the fact is that the great inconvenience of these variables is often
linked to the difficulty of the identification of the material parameters and the numerical
analysis. In order to limit the number of degrees of freedom as far as possible, Chaboche
(1982) proposed a model for the case of periodic array of penny shaped cracks based on
a fourth rank damage tensor. The damage effect tensor is introduced as follows

l-A

1 - v 1-Z),

M{D)

(2.49)

1 - v !->,

l-DK
l-D,

where direction 1 is perpendicular to the plane of the cracks and v is poisson's ratio. This
model is quite elegant because it can describe general states of orthotxopic damage by two
damage parameters D] and D5. However, it is very difficult, even for this simple damage
43

variableto carry out the numerical analysis because the formulation of the corresponding
damage energy release rate is very complex in general cases.

Due to the different mechanisms involved in the development of micro-defects in different


materials under various loading conditions, it is difficult to propose a best method for
damage measurements. Lemaitre and Dufailly (1987) examined eight methods to measure
damage defined as the effective surface density of micro-cracks and cavities. They are:
(1) direct measurements such as the observation of micrographic pictures and the
measurement of the variations of density; (2) non-direct measurements including
destructive methods such as the measurement of the variations of the elastic modulus, of
the ultrasonic waves propagation, of the cyclic plasticity or creep responses, and nondestructive methods such as the measurement of the variation of micro-hardness and of
the electrical potential. The comparative results are shown in table 2.1. Unfortunately,
no method is suitable for the case of high cycle fatigue.

2.5.3 Thermodynamic Formulation

In order to establish a proper averaging process to transmit the information of material


damage from microscale to macroscale, it is assumed that the damage states in a material
element can be represented by certain macroscopic damage variables which are precisely
the internal state variables in the sense of thermodynamic theories. Therefore, it is easy
to develop a systematic approach to derive the damage evolution equations and the
constitutive equations of damaged materials in the framework of continuum mechanics
(Lemaitre, 1987).
44

Usually, damage variables are introduced through the following assumptions:

1. Effective stress and the hypothesis of strain equivalence (Lemaitre and Chaboche,
1978): any strain constitutive equation of a damaged material is written exactly as for an
undamaged material except that the stress is replaced by the effective stress.
2. Effective strain and the hypothesis of stress equivalence (Simo and Ju, 1987): the stress
associated with a damaged state under the applied strain is equivalent to the stress
associated with its undamaged state under the effective strain.

These two approaches lead to dual but not equivalent formulations, neither physically nor
computationally (Simo and Ju, 1987). However, it was proved that both these approaches
lead asymmetry of the stiffness matrix when anisotropic damage is considered.

To

overcome this inconsistency, other hypotheses are proposed:

3. Effective stress and the hypothesis of elastic energy equivalence (Cordebois and
Sidoroff, 1982): the complementary elastic energy for a damaged material is the same in
form as that of an undamaged material except that the stress is replaced by the effective
stress in the stress-based energy formulation.
4. Effective stress and the hypothesis of energy equivalence (Chow and Lu, 1989): the
damaged material state can be replaced by an undamaged material state which is
characterized by effective stress or effective strain in the sense that the total energy
involved in the two processes, reversible or irreversible, should be equal.

When the inelastic strain is neglected, the above two hypotheses are equivalent. For the
45

case of anisotropic damage, the constitutive equations of damaged materials derived from
these two hypothesis satisfy the symmetry requirements of strain energy representations.

It is postulated that there exists a thermodynamic potential which is a convex function


of all state variables. One may choose as a homogenized free energy function
T

= T ( e 5 , >, g )

(2.50)

where
e is the elastic strain tensor, D is a damage variable and q denotes a suitable set
of plastic variables. Therefore, the thermodynamic forces conjugate to state variables can
be derived as (Lemaitre and Chaboche, 1990)
a = p

(2.51)

r = p

(2.52)

des

and

where a is the stress tensor, Y is the damage energy release rate, and p is the mass
density. Based on the following assumption: (1) no coupling between plastic strain and
elasticity; (2) a small coupling between micro-plasticity and elasticity; (3) no direct
coupling between damage and plasticity, Lemaitre (1987) proposed that the free energy
could be chosen as
T ( c 6 , U , g ) = 7 a ( C e , D)

q )

46

{2.53)

where

% are the elasticplastic part of free energy respectively. HoweverJu

(1989) argued that it was physically incorrect to neglect the effect of damage on the plastic
part of free energy. For the case of isotropic damage, his form can be written as
Wi ze9 D9 q ) = ( I - D ) Y 0 (E tf , q )

where

is the total potential energy of an undamaged material.

damage energy release rate Y equals 0not the elastic part of

(2-54)

Accordingly, the
0

as proposed by

Lemaitre (1984). However, it is not: clear which one is better because both forms are
introduced from a locally averaged free energy function which itself is based on a macromechanics approach. This point requires further work.

In accordance with the thermodynamic method, a dissipation potential $ is postulated to


exist as a convex function of all the flux variables. It can be written in a general form
as
$ = ( a, Y, ce, D, q )

(2-55)

So the plastic strain rate p and damage evolution law can be derived as (Lemaitre and
Chaboche, 1990)

where X is the Lagrange multiplier. For simplification, it is assumed that the dissipations
due to the deformation process and the damage process are independent of each other
(Lemaitre and Chaboche, 1990). Therefore, the function $ can be decomposed into

^ = %(o,q)

where

^ ^D( F, c tf, D

$D are the plastic potential and damage potential respectively.

In the

conventional theory of plasticity, the yieldsurface is usually taken as the mathematical


form of plastic potential. In a similar way,the form of damage potential is postulated to
be the same as that of the damage surface.

The concept of damage surface in the stress- or strain-space has been widely used by
many investigators (Dragon and Mroz, 1979; Krajcinovic and Fonseka, 1981; Cordebois
and Sidoroff, 1982; Chow and Wang, 1987).

Several attempts were made to define

damage surfaces using acoustic emission (Holcomb and Costin, 1986). In this approach,
the onset of acoustic emission during loading is taken to indicate the initiation of damage.
By following a number of loading paths, the locus of onset stress can be mapped and
identified as a damage surface. However, this method can only be used to measure the
damage for brittle materials. In almost all damage models the function of damage surface
is a priori postulated, then several material parameters are determined by simple
experiments.

48

In some eases, the difference between the damage processes undertensile and compressive
-

..

...-.

loading should be considered. A method was proposed to divide the stress tensor into a
positive part a + and a negative part a. (Ladeveze and Lemaitre, 1984). In the principal
stress system, they are

a.

<>

<a2>

<a->

(2.59)

where the McAuley bracket is defined by < x > = x if x > 0 and < x > = 0 if x <
0. Two scalar variables, Dt and Dc, are introduced to characterize the damage processes
corresponding to the positive part and the negative part of stress tensor respectively. This
model is very simple, but cannot describe the anisotropic elastic behaviour induced by
damage in an initially isotropic material. Other models have been proposed with two
second order damage tensors (Ramtani, 1990) or a fourth order damage tensor (Ju, 1989)
based on the strain decomposition. However, there is no acceptable theory which can
describe simultaneously the damage induced anisotropy and the different effects between
tensile and compressive loading on the damage process.

2.5.4 Fatigue Damage Models


49

Fatigue damage occurs when materials are subjected to periodic or variable loadings.
Under uniaxial cyclic loading condition a general form of the damage rate is expressed
in terms of cycles N as (Lemaitre and Chaboche, 1978)
r

dD = D"^'

gmax _

iP

dN

whereffmax and crmean are, respectively, the maximum and mean stress. The fatigue life NF
for constant cr^ and (TmeaQcan be calculated by integrating equation (2.60) between D =
0 and D = 1 which corresponds the fatigue failure

This model is also capable of predicting the fatigue life under multilevel fatigue loading
conditions. For a two-level test, it has been shown that

Fl

(2.62)

l-a 2

1-

where NF1 andN r are the fatigue lives for the two loading conditions, N! is the number
of cycles at the first level, N2 is the remaining life at the second level. This non-linear

cumulative fatigue damage model has been successfully applied to different steels under
i...

..

.... * ^ ^ . . . . . .

, i.

.. i .

- * *- - .

. .

various loading conditionsincluding two-level tests and block-programs (Chaboche and


Lesne1988).

Several attempts have been made to generalize this model to multiaxial fatigue loadings
(Lemaitre and Chaboche, 1990). However, two fundamental problems are inevitably
confronted: (1) the definition of the multiaxial stress parameters associated with the
maximum and mean stress; (2) the applicability of the model to complex loading
conditions, including non-proportional loading for engineering structures.

Another model is based on a dissipation potential (Lemaitre, 1987)

= 1 ^
2
S
o (l-D)ao

(2.63)

where p and r are respectively the accumulated macro-plastic and micro-plastic strain, S0
and amay be functions of Y.

It was supposed that in low cyclic fatigue x may be

neglected in comparison with p, and conversely p may be neglected in high cyclic fatigue.
This model has been applied to predict the fatigue life under random loadings defined by
the probability of Rayleigh with the Monte Carlo numerical method (Lemaitre, 1986b).

The life prediction under fatigue loadings at elevated temperatures must consider the creep
damage. A simple approach was proposed by adding the two damage functions directly
(Chaboche, 1980)

51

dD

/c(

where fc and fF can be determined from pure tensile creep tests and pure fatigue tests at
high frequency.

On the basis of their experimental results, Hayburst, Leckie and

McDowell (1985) suggested that this model could be readily extended to nonproportional
loading.

Several other models, including both linear and nonlinear accumulation and

interaction, were reviewed by Chaboche (1981)Murakami (1987) and Lemaitre and


Chaboche (1990). Most of these models are restricted to the uniaxial states of stress, and
characterize the damage process by a single damage variable. Generalization to multiaxial
loading conditions is a very difficult problem.

2.5.5 Life Prediction in Structures

Damage mechanics based on a macromechanics approach is a new branch of solid


mechanics.

Therefore, the constitutive equations and the damage evolution laws of

damaged materials can be readily applied to various boundary value problems to calculate
the stress/strain fields by conventional analytical or numerical methods. Furthermore, it
can also be used to simulate the damage process of structures, including both the initiation
and propagation of macrocracks. A classical decoupling method includes four essential
steps (Chaboche, 1987):

1. determination of the constitutive equations;


2. computation of stress/strain fields in structures from the applied loading conditions;
3. determination of the damage evolution laws and failure criteria;
4. computation of the damage development in each material element to failure by the
52

stresses/strains computed in step 2.

In this method, the stress distribution in structures is assumed to be the same as that
without damage up to macro-crack initiation.

The modem approach is to take into

account the coupling between strain and damage.

In this way, the stress and damage

fields are calculated at the same time by coupled constitutive equations of damaged
material.

Of course, this method gives more reasonable results, but costs much more

computing time, especially for the case of high cycle fatigue loading. Some applications
are given by Lemaitre (1986a) and Chaboche (1987).

By the concept of local approach, a crack can be defined as aflat zone of high gradients
of rigidity and strength in which the critical conditions of damage have been reached
(Lemaitre, 1986a). This means that the macro cracks can be simulated by failed elements
in which therigidity and strength may be taken as zero. In order to determine the critical
state of the material element, the following criteria have been used (Lemaitre, 1986a):

1. a critical value of an equivalent stress at a fixed distance from the crack tip;
2. a critical mean value of an equivalent strain over a finite volume;
3. a critical value of the load on a node in finite element analysis;
4. a critical value of the damage D as a mean value on a finite mesh;
5. a critical value of the overall damage at a fixed distance from the crack tip (Chow and
Wang, 1988).

All these criteria concern a critical size which is interpreted as a material constant. It is
53

very difficult to determine its value. Unfortunately, there is usually no convergence of


numerical results if it is not used

Several techniques are proposed to remedy this

difficulty (Chaboche, 1987). However, this is still an open problem (Bazant, 1986).

54

3. DAMAGE MODEL

3.1 INTRODUCTION

A great deal of work has been devoted in the past to the material characterization of
fatigue crack initiation and propagation which provides some understanding and insight
into material behaviour during fatigue failure. The fatigue life predictions for specimens
or structures have been realized mainly by conventional phenomenological models, such
as the S-N curve, the Coffin-Manson law, the Paris law and the linear damage rule
(LDR), etc.

Unfortunately, most work is concerned with methods of testing, data

collection and correlation, and there are no systematic ways to elaborate them further or
to generalize them to more complicated problems. Even for the same material, many
experimental tests need to be carried out to determine the material constants for various
models which are used to describe different phases of fatigue behaviour, such as fatigue
crack initiation, short crack growth and long crack propagation.

Consequently, there

often exist problems in the transition between two phases of life due to the ambiguous and
artificial definition for each phase and the limited applicability of each model.

satisfactory theory is required for the safe design of complex engineering structures
against fatigue damage.

55

Damage mechanics theory provides a more systematic approach to fatigue life prediction.
Since Kachanov published his paper on creep failure in 1958the theory of damage
mechanics has been well developed.

This area of solid mechanics centred on micro

defects development provides a better understanding of the mechanisms of fracture in


structures by means of macroscopic damage variables which represent the deterioration
of material elements. Many damage models have been successfully applied to such failure
problems as brittle fracture, ductile fracture and creep fracture (Lemaitre, 1986a;
Murakami, 1990). However, there are only a few models based on damage mechanics
theory, to author's knowledge, proposed to deal with fatigue problems. This is because
fatigue damage is more local than other types of damage, thus much more difficult to
describe with the macroscopic variables.

A damage model known as NLCD (nonlinear continuous damage) proposed by Lemaitre


and Chaboche (1978) attempted to describe the governing features of fatigue damage with
some degree of success under various uniaxial loading conditions.

As the stress

parameters maximum stress and mean stress were employed in this kind of model for the
case of uniaYial fatigue loading, it is difficult to generalize these models to complex
fatigue loading conditions. Therefore, it is necessary to develop a comprehensive model
capable of characterizing fatigue damage behaviour of engineering materials under
multiaxial fatigue loading conditions.

This chapter presents a fatigue damage model for the case of multiaxial high- and
medium-cycle fatigue loading based on the damage mechanics theory. A new damage
effect tensor M ( d ) is proposed to describe the effect of damage on material elements at

56

the macro-scale. Depending upon the level of applied stress in a material element, the
damage accumulation under cyclic loading is postulated to undergo two processes, the
accumulation of fatigue damage and plastic damage.

Two corresponding damage

evolution laws are introduced using the thermodynamics of irreversible processes by two
damage surfaces which are used to identify respectively fatigue damage and plastic
damage.

Finally, an experimental method is demonstrated to determine the material

constants required in the proposed damage model.

3.2 DAMAGE VARIABLE

During a loading processthe micro-structure of a material element changes due to


nucleation and growth of micro-cracks or defects. It is difficult to describe this change
in detail by a mathematical model in view of the complexity of the irreversible physical
microcracking processes, grain size and surface effects, especially in the case of fatigue
loading. In order to establish a damage model which can be used to predict fatigue failure
in engineering structures, it is necessary to introduce an averaging macro-variable D to
characterize the gradual deterioration of the material under continuous load application.
The averaging process is assumed to be carried out on a material element which is small
enough to be taken as a "mathematical point" to define mechanical variables, such as
stress, strain, damage, etc., but large enough to be considered that the elementary
mechanisms of the material are well represented by the continuous variables based on a
macro-mechanics theory (Lemaitre, 1986a). Following the damage mechanics theory, an
effective stress which wasfirst introduced by Kachanov (1958) is defined as

57

where a is the true stress tensor and the damage effect variable M(D) is a fourth rank
tensor. If the following notations are chosen,
a T

= Ca i a 2 3 o

] = [

a22

CT33 a 2 3

a31 a12 ]
(3.2)

CT

" [

2 e3

[ e i l S22 S33 ^ 2 3 2S 3 i 2S 12 ]

M(D) can be expressed as a 6 x 6 matrix. In fact, one could define M(D) by the concept
of net area reduction, including the microcrack area, the microstress concentration and the
interaction among microcracks, due to the development of micro-defects (Lemaitre and
Chaboche, 1990)
M ( D ) = ( I - D )-1

(3-3)

where I is the rank four identity tensor. In general, D is also a fourth rank tensor which
has 36 independent components due to the symmetry of the stress tensor. It is impractical
to introduce so many degrees of freedom in a damage model due to the difficulties in
physical measurement of the necessary material constants and in numerical analysis.
Some assumptions need to be made to simplify the problem on the basis of the physical
understanding.

Based on the microscopic studies on the high cycle fatigue failure in smooth specimens,
some experimental results reviewed in last chapter are: (1) most micro-cracks initiate in
58

PSBs and grow along the slip planes; (2) the development of early micro-cracks occurs
in the shear mode independently in various grains; (3) there is no relationship between the
slip system active in a particular grain and its orientation with respect to the direction of
applied stress; (4) when the size of micro-cracks becomes significant, they have a
tendency to orient themselves perpendicular to the direction of maximum principal stress
due to the local stress concentration. Accordingly, the direction of early micro-crack in
a given grain is dominated more by its neighbours than by the applied stress direction due
to large local variations in the stress field. For the grains on the surface of material, the
micro-cracks usually occur first in some preferential grains in which the slip plane traces
are parallel to the direction with 45inclination to the direction of the largest principal
stress. For simplification the discussion here is confined to two dimensional isotropic
material elements.

The distribution of micro-cracks can been simplified as two

perpendicular systems of micro-cracks with


maximum principal stress.

4
5
inclination to the direction of the local

The number of micro-cracks in the two systems is

approximately the same from statistical considerations. The micro-cracks along the two
directions exhibit similar behaviour with smaller variations in the material between the two
directions, so the degree of anisotropy will be much less (Lemaitre and Chaboche, 1990).
Based on the above analyses, original isotropic materials may be assumed to remain
isotropic during the fatigue process. The assumption is considered reasonable until the
very late phase of damage when linking up of micro-cracks has begun and the material
element is well on its way to failure.

For the case of isotropic damage, a scalar variable d is often used to describe the damage
process. The damage effect M and the effective stress ff simply reduce to
59

M(D) =

1-d

I
(3.4)

o =

1-d

where d = 0 is for undamaged state and

S [0 1] corresponds to a critical state of

damaged state for local rupture. This is the typical form frequently chosen to describe
the isotropic damage phenomenon. However, it does not include the experimental fact
that the values of Poisson's ratio of damaged materials vary with the increase of applied
stresses- (Cordebois and Sidoroff, 1982; Chow and Wang, 1987). This phenomenon is
often considered as anisotropic damage, and the experimental results are used to determine
the damage variables in anisotropic models.

It may be argued that original isotropic

materials remain isotropic after damage although the values of elastic modulus and
Poisson's ratio may change during loading processes.

In order to establish the

mathematical model based on this isotropic assumption, a fourth rank damage tensor d
is proposed with two independent components D! and D2. It can be expressed as

Djj = D1 5^- + D2 ( PJJ- djj )

no summation for i ,

(3.5)

where i, j = 1,2,...,6 and

(3.6)
0

Then the damage tensor can be written in 6x6 matrix form as


60

The damage effect tensor is derived by the definition of M(D) (Eqn. 3.3) as
Jif(P)

(I

1-c

- D)

1 X( |i 0
H 1 H 0
p 1 0
0 0 0 1-|.
0 0 0
0 0 0

(3.8)

0
0

where d and x
f are two new independent components which have the relationships with Dj
and D2 as

( 1 - ) (1--0,-) - 22
(3.9)

U,=

D2
1-^-

Obviously, this new damage effect tensor reduces to the typical form (Lemaitre and
Chaboche, 1990) M = 1/(1-d) when^ = 0 which corresponds to D2 = 0 and d = D
So the damage component /x (or

can be taken as a parameter which represents the

effect of damage on the value of Poisson's ratio.

The split of the effective total sixain tensor s into the effective elastic strain ee and the
effective plastic strain p parts is assumed for damaged materials as
e = e + c15

(3.10)

Similarly, definitions of these effective parameters are given as follows


e = hf' 1
p

d'z

(3.11)

= 2^'

Latter discussions will show that the elastic matrix deduced for damaged material is in the
form of isotropic material although the proposed damage effect tensor is a fourth rank
tensor. This special form is capable of accounting for the changes both in the elastic
modulus and in the Poisson's ratio. Due to the isotropic property of damaged material,
it is practical to apply this damage variable on fatigue life predictions of structures under
complex conditions.

3.3 CONSTITUTIVE EQUATIONS

3.3.1 Elasticity Coupled with Damage

According to the hypothesis of strain energy equivalence due to Chow and Lu (1989)the
elastic energy for a damaged material is the same as that of the undamaged material,
except that the stress/strain tensor is replaced by the corresponding effective parameter
in the stress/strain-based form. It then can be written in the stress-based form as
62

where C is the elastic tensor of the undamaged material and can be written in a matrix
form as
1 -v v
-v 1 "V
-v V 1
0

2 (1+v)

2 (1+V)

2 (1+v

(3.13)

E and v are the values of Young's modulus and Poisson's ratio for an virgin or
undamaged material. Therefore, the elastic law is derived by the thermodynamic theory
in the true stress-true strain space as
c e = dWe(afD)
da

c-i . a

(3.14)

or in the effective stress-effective strain space as


=

where C is the effective elastic tensor for damaged material and expressed as

63

-v

->

0 2(1+7)

(3.16)
0

2(1+v)

2(1+?)

E and
have the relationships with the damage components as
_

l-d)2
l - 4 v i i + 2 ( 1 - v ) [i 2
(3.17)
v - 2 (1-v) n-(l-3v) n2
1 - 4 V J 1 + 2 ( 1 - v ) p.2

It is worth noting that equation (3.17) is derived from the hypothesis of energy
equivalence and is different from the conventional relationship of E = E(l-d) (Lemaitre
and Chaboche, 1990).

The conventional form is obtained from the hypothesis of

stress/strain equivalence which has been found to yield an anomaly in the establishment
of constitutive equations of anisotropic damaged material (Cordebois and Sidoroff, 1982;
Chow and Wang, 1987).

3.3.2 Plasticity Coupled with Damage

For damaged materials, following Von Mises theory, the yield surface may be written as

Fp(a,D,R)

and the thermodynamic potential function may be defined as the same form of yield
surface

= op2 -

[i?a + i?(p) ]

(3

'19)

where

1 aT : H
2

(3.20)

H is the plastic characteristic tensor of undamaged material which is a positive semidefmite tensor, R0 is the initial strain hardening threshold, p is the overall plastic strain,
and R(p) is the increment of strain hardening threshold. For an isotropic material, the
plastic characteristic tensor is (Hill, 1950)
2
-1
-1
0
0
0

-1
2
-1
0
0
0

-1
-1
2
0
0
0

0 0 0
0 0 0
0 0 0
6 0 0
0 6 0
0 0 6

(3.21)

Then the effective plastic characteristic tensor H is derivedfrom equation (3.20) as

65

(1 - ti)2
(1-d)2

(3.22}

Similar to the conventional theory of plasticity (Hill, 1950)the constitutive plastic


equations incorporating material damage are derived in the true stress-true strain space as

dzp
dt

da

(3.23)

2o 1 / 2

or in the effective stress-effective strain space as

dEp
dt

da

(3.24)

20 1 / 2

and
dF

dt

(3-25)

d(-R)

whereis a Lagrange multiplier which can be determined by means of the yield surface.
Fp = 0.

From equation (3.18)


, ( a ,

D, R )
(3.26)

da

dFp
ID

dFn

dD

Then the Lagrange multiplieris deduced from equation (3.25) as

66

dFp
da

dF

'dR
if

F=0 and

if

:d > 0

da

(3.27)

F <

o r F,, = 0 and

dF

da

a+

(dF^

From the definition of yield surface (Eqn. 3.18),

dR

dFD
do

-1
i
a
20 1 / 2

(3.28)

dF^
dD

40172

T h e r e f o r e , t h e L a g r a n g e m u l t i p l i e r X p c a n b e e x p r e s s e da s

'.H-.b +
2a

1/2

:aiD > 0

i f F=Q and ar:H

i f

or Fp=Q and aT:"H: a +

From equation (3.23),


67

F <

SS

\D ^

(3.29)

zP T-.H- x'.z p

(3.30)
4CT P

So the overall plastic strain rate can be expressed in an alternative form as


p =2 [ 1 c'r

( 3

.31)

For an isotropic material, H shown in equation (3.21) is singular matrix. A generalized


inverse form of H may be chosen as
-1
2
-1
0
0
0

2
-1
-1
0
0
0

-1
-1
2
0 1 . 5 0
0 0 1 . 5 0
0 1 . 5
0 0

(3.32)

Then the generalized inverse of H can be deduced as


1

= -

(3.33)
(l-n)2

3.4 DAMAGE EVOLUTION

3.4.1 Basic Modelling

Depending upon the level of the applied stresses several cases of damage should be

considered under uniaxial loading condition:

1. If the applied stress is less than the fatigue endurance limit, there is no significant
irreversible plastic deformation and hence no damage accumulation.
2. If the applied stress is larger than the fatigue endurance limit but less than the yield
stress of material, damage accumulates due to micro-plasticity and is defined as fatigue
damage.
3. If the applied stress is larger than the yield stress, the resulting damage is defined as
plastic damage.

The above considerations enable the damage accumulation under any uniaxial loading
condition to be divided into two processes: fatigue damage and plastic damage. In order
to extend this idea to more complex loading conditions, it is postulated that there exists
a fatigue damage surface Ffd=0 corresponding to the fatigue endurance limit and a plastic
damage surface Fpd=0 corresponding to the start of plastic damage. These surfaces are
used to identify different damage processes.

Therefore, there may be three cases of

damage accumulation under any complex loading condition:

l.If the stress state is inside the fatigue damage surface, there is no damage accumulation.
2.If the stress state is between two damage surfaces, the resulting damage is defined as
fatigue damage.
3.If the stress state is outside the plastic damage surface, the resulting damage is defined
as plastic damage.

69

It is postulated that the total damage is the summation of fatigue damage and plastic
damage. When total damage in a material element reaches a critical state, this element
is said to be ruptured. Therefore, unlike the conventional fracture mechanics concepts,
the entire process of damage accumulation followed by initiation and propagation of
macro-cracks leading to final rupture of structures can be described by this model. The
following sections describe a mathematical model to simulate the whole process.

3.4.2 Damage Energy Release Rate Y

It is Postulated that a specific homogenized free energy can be written as (Lemaitre and
Chaboche, 1990)

where %,

3.34)

are the elastic, plastic part of free energy respectively, D is thefourth rank

damage tensorand q is a suitable set of plastic variables.

Then the thermodynamic

conjugate force of the damage variable Dwhich is called as damage energy release rate,
can be deduced from equation (3.12) as

dD

dD

dD
(3.35)

- \

' (

where p is the mass density, WB is the elastic energy, Y is a fourth rank tensor
corresponding to D. Due to the isotropic assumption for damaged materials, Y has a
70

similar form as that of d (Eqn. 3.5) and can been expressed as

if i=j
(3.36)

if i*j
for i and j =
for

where YD1, YD2 are two independent components of the fourth tensor Y. They can be
expressed as

dW'
'dE

(3.37)

dWl
~dK

It is observed that the damage processes are characterized by two independent scalar
parameters

(D^ D2) based

correspondence between (D

on the isotropic assumption.

Due

to a one-to-one

and (d, fx), shown in equation (3.9)one may choose

(dn) as two new damage parameters instead of (D!Dto characterize the damage
process in order to simplify the derivation. Therefore, the new damage energy release
rate Y is
= Yd 5ij

+ y,(

and two corresponding components are derived from equation (3.12) as

71

(3.39)

dW'

where A"1 is expressed in the matrix form as


^1 -^2 ^-2
A2
A2
^2 -^2
l-d)

2 (-

2\i

2 {A^-A^
2
0

(3.40)

(1 + li) (1-v) -2nv

3.4.3 Plastic Damage

It is supposed that there exists a plastic damage surface defining the start of plastic
damage under any loading condition and may be formulated as (Cordebois and Sidoroff,
1982)
Fpd(

Y,B

2
) - Yp 2d - i B^B(wp)

(3.41)

] =0

where Ypd is defined as

72

(3.42)

Yd, Y , are two independent components of Y, Bq is the initial plastic damage threshold,
B is the increment of plastic damage threshold, wp is the overall plastic damage, and 7 is
the damage evolution coefficient.

An essential step in developing damage evolution

equations is to formulate an appropriate damage dissipative potential.

Similar to the

conventional theory of plasticity (Hill, 1950), the plastic damage dissipative potential may
be chosen as the same form as the plastic damage surface
$pd( ^ 3 ) =

Fpd( Y, B )
(3.43)
1

=Y/d

- [ B0 ^ B(wp)]

In a similar way as that leading to the damage-affected plastic constitutive equations, the
plastic damage evolution laws are derived as
i)

= -X

Pd

d pd

Qy
(3.44)

where Dp is the plastic damage tensor, Xpd is the Lagrange multiplier which can be
determined by the plastic damage surface. From equation (3.41),

73


(3.45)

dYd

dY^

11

dB

dwp

Then the Lagrange multiplier is derived as

Pd

yY^
1

+ +

and two independent components of Dp are expressed as


,

_A p d (7 d n^)
2
(3.47)

, = -

Pd 0y

2rn,

_i

From equation (3.47),

l-v2

p -

2p p

"

4 (l-y 2 )

-2y(Yd+yYil) (yY^)

_
y

+(

3.48)

Then the overall plastic damage rate may be presented alternatively as

^ = ^ = 2[

(-2 +

74

.49)

3.4.4 Fatigue Damage

The fatigue damage represents a microscopic change of material behaviour at the scale of
one or a few grains. At this scale the material is neither homogeneous nor isotropic, and
the damage remains very localized.

In general, this difficult but challenging practical

problem of localized damage has not yet received a rigorous mathematical treatment. The
attempt is made in this section to describe a fatigue damage process by an internal variable
d f , which is called as fatigue damage tensor, and fatigue damage evolution laws.
Although these concepts lack a sound physical basis, they offer some possible methods
dealing with fatigue failure of structures.

For most engineering materials, the contribution of tensile stress to damage accumulation
is different from that of compressive stress due to the existence of microcracks and
microvoids (Krajcinovic and Fonseka, 1981; Ladeveze, 1983). The different responses
of these microcracks may be associated to the closure of some microcracks. Therefore,
a parameter, which is named as effective stress range, is introduced to take into account
this difference. In general, the stress range in a cycle may be decomposed into two parts,
positive part + and negative part tf as follows:
A<y+

=
(3.50)

AcT = a 0 - a m i n
where

and <r0 are the stress tensors corresponding to respectively the maximum and

minimum value of Ypd (Eqn. 3.42) over a particular cycle, and


the instant half a cycle after (or before) appears.
condition, there are two cases:
75

is the stress tensor at

Under uniaxial cyclic loading

1.For tension-compression loading, the minimum stress is negative. The minimum value
of Ypd is zero so that <0r equHls 0. This means that the positive part is the maximum stress
Omax and the negative part, -o-^, is the absolute value of the minimum stress.
2.For tension-tension loading, the minimum stress is positive.

So the value of Ypd

increases monotonically as the loading increases and the value of

0 is equal to the

minimum stress. Therefore, the positive part is the total stress range and the negative part
is zero.

For a complex loading condition, such as near the crack tip, the minimum value of Ypd
in a cycle is used to distinguish these two parts. Specially, if

in a material element

increases monotonically as the applied stress rises from the minimum stress to the
maximum stress, <T0 corresponding to the

equals the minimum local stress and the

negative part is zero. In order to consider the different damage behaviour between the
positive part and the negative part, a damage efficiency factor, a, is introduced and the
effective stress range is defined as:
L<3 et = Acr+ + a cT

(3.51)

a, which can be determined experimentally, varies from 0 to 1 for most engineering


materials.

It is obvious from equation (3.51) that there are two special cases

corresponding to 0 and 1 for a. There is no contribution of the negative part on the


damage accumulation when a equals 0and the effect of the negative part is the same as
that of the positive part if a is 1.

Based on the above analysis, a new stress parameter


is introduced as
76

which is called as active stress,

0~00 + C
O (OQ""Om^n) /
a(CJ-Omin)

(3.52)

i f ^

CT

[ o' min^

This active stress accounts the difference between the fatigue damage accumulation under
tension and compression, and can be used to calculate the fatigue damage instead of true
stress a.

It is postulated that there exists a fatigue damage surface which represents the fatigue
endurance limit
Ffd(

f d i I

(3.53)

) = 0

or
(3.54)
a

actfl

'

0 '

which can be determined by experiment, where Y^,, is the critical value of Yfd on the
fatigue damage surface,

is the minimum value of

over a cycle, ^

is the

critical value of on the fatigue damage surface, and


=

( Yl + 2yYdifY}lif

--2

\
f )

(3 55)

The fatigue damage evolution laws are postulated to have similar forms as that of the
plastic damage evolution laws (Eqns. 3.46 and 3.47)

77

24
(3.56)

d, f^d, f +y Yd,

'FD -

(3 .57)

2Yf dK(w)
where df and

are two independent components of fatigue damage variable Df, wf is the

overall fatigue damage, w is the total overall damage which is the summation of wf and
wp, and the fatigue damage energy release rate Yf is defined as
Y{

a,

D )

= Y

aact,

D )

(3.58)

Therefore, the two components of Yf can be written as


0 , 0 ) =

YD{

OACE,

D )
(3.59)

,f(

= r u ( v act f D )

The function K(w) in equation (3.57) may be expressed to a first approximation as

K{w)

= KQ

( 1

(3.60)

where Ko and ware material constants.

Similarly, the overall fatigue damage rate may be presented alternatively as


78

=fd

= 2

2 (2Y<5 f li/ + (if)


. 2 ll-Y

(3.61)

3.4.5 Damage Accumulation

Several cases can now be considered under fatigue loading conditions as follows:

1. Ffd( Yfd", Ypd11 ) < 0.

In this case the material is operating under the fatigue

endurance limit, and there is no damage accumulation.

2. Ffd(Y ^ ,

) > 0 and Fpd( Y , , ^ ) < 0.

This case corresponds to fatigue

damage. The total damage rate in a cycle is


A ?
AN

AN

= dD
f
f

(3.62)

Jalc

and the total overall damage rate in a cycle is


Aw
A

Ai^

where

AiV

Jo,.

wffdt

is the critical stress tensor which can be determined by the fatigue damage

surface Ffd( Yfd(<rlc),

) = 0. Then the two components of total damage rate in a

cycle can be written as

79

(3.64)

At

AN

dt

Jalc

Integrating equations (3.56) and (3.61) allows the accumulated damage in a cycle to be
calculated.

3. Fpd( Yp^) = 0. The total damage in this case is postulated to be the linear summation
of fatigue damage and plastic damage.
AD _ Al>f
AN
AiV

, ADD
A.N
(3.65)

=P

Jo

Aw _
AN

lc

d Df

J<

'd

Aw
AN

l^N

(3.66)

=
J

dw

^1C

where stress tensor ad can be determined by the plastic damage surface

YpdOJ )

0. Then the two components of total damage rate in a cycle can be written as
d

r
,

AIA
II

(3.67)

Integrating equations (3.47) and (3.56) allows these two components to be calculated.
80

3.5 DETERMINATION OF MATERIAL PROPERTIES

This section gives a general method to determine the material constants or curves which
are required in this model. In principle, all material constants used to predict fatigue life
of structures can be measured by uniaxial tensile tests or uniaxial constant amplitude
fatigue tests. The details follow.

3.5.1 Damage Components d And x


f

From the formulation of effective elastic tensor (Eqn. 3.16), Young's modulus E and
Poisson's ratio u of damaged materials are

L - 4 V U + 2 ( 1 - V ) \I2
(3.68)
_

Y- 2 ( 1 - V ) N . - ( L ~ 3 V ) \ I 2
2

l-4v\i+2 (1-V) N

Then the damage components d and fi are derived froni


_ N
[ 2 ( 1 - V ) V + L - 3 V ] N 2 + 2 ( L - V - 2 V V ) ^ + V_V = 0

(1-D)

[ 1

4V[JL + 2 ( 1 - V ) H 2 ]

The positive solutions are taken for these two equations because d and

(3.69)

(3.70)

are always

positive. Therefore, the damage components can be determined using the measured values
of elastic modulus and Poisson's ratio of the damaged materials under uniaxial tensile
81

tests. In this case, the plastic damage components dp and ixv may be taken the same as d
and n because fatigue damage components df and m are relatively very small.

3.5.2 Damage Evolution Coefficient y

Under uniaxial tension, two components of damage energy release rate Y shown in
equation (3.39) can be reduced as
Y
d

= - 1-4VIJ.+2 (1-V) H2 ct2


1-d) 3
(3.71)

y
1

F(l-d)2

and the damage components evolution laws shown in equation (3.47) become

j.

(3.72)

"

27

where a is the tensile stress. From equation (3.72)


dd=

"

+Y ( l - d )

yf(\x)

=1-4VH+2(1-V)^2

(3.73)

When the values of d and are small at the beginning, the approximate solution of
equation (3.73) is
82

L - v \2
\ 1-v/

l-v-2v 2
2 (1-v)2

=c

(-"
(3.74)

a=

+ Y

yf(ii) + l - d

where c is a constant which can be determined by substituting initial condition, d = 0 and


fi = 0, into the solution as
1-2VY

2(1-v)

(l-Y2) e

y(y-2V)

2(y2-l)

(3.75)

Alternatively, y can be determined directly from equation (3.73)

^(M) -

(3.76)

- ( " I S
where Ad is the increment of d and Afx is the increment of fi. This equation is valid
during whole loading process.

Therefore, the damage evolution coefficient 7 can be

determined by the calculated values of damage components d and


3.5.1.

3.5.3 Increment of Strain Hardening Threshold R(p)

Under uniaxial tension, the yield surface of damaged material (Eqn. 3.18) becomes
4 2 a - [ i?0 +
1- a

Rip)

(3.77)

l-t^o
l-dn

83

where ay is the initial yield stress, and d0, /i0 are the values of the corresponding damage
components on the initial yield surface. Usually, they can be assumed to be zero for the
case of uniaxial tension, so Rq is the same as a r The value of R is accordingly obtained

Izit

-R(p)

(3.78)

1-d

The overall plastic strain can be derived from equation (3.31) as


1 _ d

1-H

dsf
(3.79)

where Sj is the total strain along the loading direction, E is the elastic modulus
corresponding to a. Therefore, the curve of R(p) can be obtained with the measured
variables a,

E and calculated variables d and \i.

3.5.4 Increment of Plastic Damage Threshold B(wp) and Critical Value of Total Overall
Damage wc

Under uniaxial tension, the plastic damage surface can be written from equation (3.41) as

(al+ai^ya^)

- B0 - B = 0

where
84

(3.80)

1-4VH+2 (1-v) p 2
(1-d)3
(3.81)

2\i
(1-d)2

(3.82)

(1+4V2-4YV)

crd

to be the same as

the yield stress (rr

This implies that the plastic damage starts wh


ee
n applied stress
th

exceeds the yield stress. Of course, it is not the case for some special
materials. So the
value of B can be calculated by the following equation

(ai+a2 +2Ya1a2)

(3.83)

The total overall damage can be derived from equation (3.49) as

dw = / dw
(3.84)
2

2(l-y )

((cfdJ2

- 2y dd0 d\iD + (d\xj

Therefore, the curve of B(w) can be determined by equations (3.83) and (3.84) with the
calculated variables d, jx and y. The critical value of total overall damage wc which
corresponds to the critical state of complete specimen fracture is also determined by
equation (3.84).

85

3.5.5 Fatigue Damage Surface

In general, the fatigue damage surface is


Ffd(

Yfdtl,

Under uniaxial tension-tension cyclic loading, from equations (3.42) and (3.55)
=

=I

(1+4V2-4YV)
(3,86)

Y S

= I

( 1 + 4 V

- 4 Y V )

where ^ is the fatigue endurance limit and

is the minimum stress in a cycle. The

fatigue damage surface represented by Yfd>1 - Yp^ under multiaxial cyclic loading
conditions is postulated to be the same as that under the uniaxial loading conditions.
Therefore, the fatigue damage surface can be determined by the fatigue endurance limit
-minimum stress

curve measured in tension-tension cyclic tests.

3.5.6 Fatigue Constants K0 and a

Under uniaxial tension-tension cyclic loading, the damage rate in a cycle for
can be written, from equation (3.64), as
=
M
AN

=-

AN

"

(3.87)

86

A\i _
'AN

Y A ,
AN

AYfdK0(l-^.)
(3.88)
+

Y ^d,

and the total overall damage rate in a cycle is by equation (3.63)

AN

A.DF

2 (l-y )

AN

AN

)
(3.89)

2
2Y-

where

AiV" AiV

is determined by fatigue damage surface Ffd( Yfd(o"lc)Y^o-^ ) = 0. Two

components of Yf can be deduced from equation (3.59) as


,{o-o min;
2

1 - + 2 (1-v) M-

1-d)

( _a . )/
wCT vmin
(3.90)

(a2 n ( 1 - v ) _2v v/ u
)
,N 22
Uroin7
E{l-d)

Integrating equation (3.89) for the number of cycles N between 0 and Nf leads to
d

(3.91)

Aw

87

where wc is the critical value of w and Nf is the number of cycles to failure. Therefore,
Kq can be determined by this equation from the measured value Nf corresponding to cyclic
loading (a-min, a ^ .

Under uniaxial tension-compression or compression-compression cyclic loading, equations


(3.87) and (3.88) are valid for the case of cr^ < crd and

> -crd, where crd is the stress

on the plastic damage surface, except that the integrating process should be completed
from 0 to Acrcff. The value of Ao-eff can be calculated from equation (3.51)
= aroax -

Ydf and
T,

crmin

(3.92)

are calculated by following equations


_

l - 4 v n + 2 ( l - v ) ja2 2
~

'

(3.93)
T,

_ . 2(1 ( 1 - v ) - 2 v 1

EU-dV

and the active stress a1)act is deduced from equation (3.52)


a (
1, act

,ijf

a <

, i f

o ^

(3.94)

Similarly, the material constant a can be also determined by equation (3.91) with the
measured value of Nf under tension-compression cyclic loading

88

cr^).

3.5.7 Summary

In order to determine the basic material properties required in proposed fatigue damage
model, uniaxial tests are performed:

1. Some uniaxial tensile tests should be carried out to measure true stress-true strain
curve, elastic modulus-true stress curve and Poisson's ratio-true stress curve. Based on
these experimental results, one can calculate the damage components d and [i, damage
evolution coefficient
Bq and B(w), and the critical value of total overall damage wc.

2. The uniaxial tension-tension cyclic loading test for

< ay is needed to determine

the material fatigue constant K0 by measuring the number of cycles to failure.

3. The uniaxial tension-compression cyclic loading test for

<

O
y and

necessary for the determination of another material fatigue constant a if this damage
model is applied to the compressive as well as tensile loading conditions.

4. The effect of minimum stress on the fatigue endurance limit under uniaxial tensiontension cyclic loading conditions should be studied.
determine the fatigue damage surface Ffd(Y

89

f ( U

The testing results are used to

) = 0.

CHAPTER FOUR

4. NUMERICAL ALGORITHM

4.1 INTRODUCTION

Due to damage accumulation, the localrigidity in a material element decreases during the
loading process. Generally, the stress distribution in a structure is not uniform due to
geometrical discontinuities, such that the material properties from one location to another
in the structure may be different. This will cause stress/strain redistribution. The changes
should be included in the numerical analysis in order to achieve an accurate assessment
of material behaviours around notches or cracks.

This chapter is intended to formulate a finite element analysis of the proposed damage
model described in the last chapter.

In addition to the implementation effort, two main

problems are also addressed and discussed: 1. The simulation of fatigue loading; 2. The
simulation of crack initiation and propagation.

4.2 FINITE ELEMENT FORMULATION

Although the constitutive equations of damaged material are more complex than those of
undamaged material, the conventional finite element method can be conveniently modified
to include the effect of deterioration of material properties.
90

The only modification

required is to derive a new stiffness matrix which incorporates the effects of damage in
the constitutive equations of elasticity and plasticity and the damage evolution.

From equation (3.15)the elastic law of damaged material can be expressed in the
effective stress-effective strain space as

By the assumption of effective strain split (Eqn. 3.10)


dB = C i dze

= C

Multiplying the above by (3Fp/3o)T yields

From equation (3.18),

Substituting equations (3.24)(3.25) and (4.4) into (4.3), it can be shown that

91

Thus the relationship between


(4.5) as follows

where C is the instantaneous tangent modulus tensor in the effective stress-effective


strain space and has the formula

By the definition of yield surface (Eqn. 3.18),

(4.8)

^ ,
71/2

Substituting it into equation (4.7), Cep can be derived as


(C:H:M:a)

4a

(C:S:M:a)T

( 4 > 9 )

Finite element implementation of the proposed damage model is made by a modified

NFAP program which is expressed in terms of the true stress-true strain space.
Therefore, the relationship between
da = C ep

is required, where Cep is the effective instantaneous tangent modulus tensor in true stresstrue strain space which can be derived by the following method.

From the definitions of effective stress and strain in equations (3.1) and (3.11)
(4.11)

da
dip

(4.12)

de

where
(4.13)

By the plastic damage evolution law (Eqn. 3.44), dM can be expressed as


(4.14)

dn

"WJ

and from the plastic damage surface expressed in equation (3.41)

Then, with the formulae

equation (4.15) becomes

(4.17)

Combining equations (4.6), (4.11), (4.12), (4.14) and (4.17), the effective instantaneous
tangent modulus tensor is derived as
gap

j f . - X

where
I

dY

(4.19)

dY

'-.If'

Some of the terms in equations (4.17) and (4.19) can be expressed as


94

. dFpd

dY

ar .

dM + dFpd dM
dYd dd
dY^ d\i

dF
Pd

T^7i[
^pd

( W V

pd

ay

4.

(iVA) (Y^+r,)

dY _

1
2Y^2

dY, dY,,
dd

dY,
( V y V - ^

1-d

1-d

M1
M2 m1 M2

(1-d)

M2
0 0 o
0 0 0
0

o
0

(yyd +Ya)

+( Y W

1
1-d

a2=
dd 2

0
1
1
0
0
0

1
0
1
0
0
0

1
1
0
0
0
0

0
0
0
-1
0
0

0
0
0
0
-1
0

0
0
0
0
0
-1

dM
1
ad " 1 - d
(4.25)

= __i_

Usually the matrix form of

dM

d2M .

is a 6x6 asymmetric matrix for which most general

purpose finite element programs may encounter computational difficulties in convergence.


Consequently, the symmetric form
C

= (

(4.26)

is taken as a tangent modulus matrix when the finite element stiffness matrix is computed
JT = f

BT \ Cs

\ B

(4.27)

where B is the transformation matrix- It should be emphasized that Cs is only used for
evaluating the element stiffness matrix in an iteration procedure, as the relationship
between the incremental true stress and strain still follows equation (4.8). Consequently,
the accuracy of results obtained by this method is the same as that by the conventional
FEM analysis.

However, the solution convergence rate may be very slow in the

numerical integration of the differential stress-strain equations when total damage is large.
This is because the difference between Cs and C increases as total damage accumulates.
A further study is needed to reduce computing time.
96

4.3 SIMULATION OF CYCLIC LOADING

In order to achieve the most accurate prediction of fatigue failure in structures, it would
be desirable to compute damage accumulation for each loading cycle. This numerical
method would however lead to prohibitively long computing time and cost, and would not
therefore be considered practical for routine applications, especially under high cycle
loading conditions. In order to save computing time, a simplified approach is adopted in
this investigation for which fatigue loading process is simulated with several load steps,
rather than cycle by cycle.

A fatigue loading process is divided into several loading steps, and the number of cycles
for each step may be different.

In the first cycle of a loading step, the elasto-plastic

deformation occurs first in a material element when an external load is applied and local
stress exceeds the yield stress.

This is followed by elastic deformation under the

unloading process and the subsequent cycles.

The loading process is based on the

assumption that stress/strain distribution at each step remains approximately constant and
there is no obvious reversed plastic deformation during the fatigue loading process (as
shown in Fig. 4.1).

Therefore, in a loading step a relatively large material damage,

including both fatigue damage and plastic damage, is induced at the first cycle which is
followed by smaller fatigue damage accumulation for a considerable number of subsequent
loading cycles. It is postulated that the fatigue damage accumulation in a cycle is constant
at a particular loading step and there is no plastic damage accumulation except in the first
cycle of the step.

The increment of damage under loading step can be accordingly

written as
97

(4.28)

where i is the serial number of a loading step, Nis the number of cycles in the ith step,
AD is the increment of total damage, Aw is the increment of total overall damage,
(AD/AN)! is the total damage rate in the first cycle of the ith load step which can be
calculated by equation (3.65), (AD/AN); is the total damage rate in the remaining cycles
of the load step which is determined by equation (3.62)and (Aw/AN)ii, (Aw/AN) are the
*

corresponding total overall damage rates in the ith cycle which can be calculated by
equations (3.66) and (3.63) respectively. For a simple case in which there is no plastic
damage accumulation in a loading step, equation (4.28) becomes

^ (if),

(4.29)

Due to the damage accumulation in this step, the stress/strain distribution in structures will
change in the following loading step, and the new values of Awi+1 and AD i+i should be
re-calculated.

The process described above is repeated until a material element is

ruptured. According to this method, the FEM analysis is only required in the first cycle
for each loading step. The effect of the number of cycles chosen for a loading step Nj on
the numerical results is examined in the following chapters.

4.4 SIMULATION OF CRACK GROWTH

Based on the finite element analysis proposed above, the total overall damage at each
integration point can be calculated. When the total overall damage at the integration point
reaches its critical value we, this material point is said to be fully damaged resulting
complete rapture.

In thefinite element implementation of the proposed damage model,

crack initiation and propagation in structures can be simulated with the following
computational steps. The first step is to calculate the damage accumulation in all elements
of the structure and then to identify those integral points where the damage reaches a
critical value. The second step is to release these integration points so that the rigidity
and strength of these points are reduced to zero. The third step is to calculate the new
distribution of damage accumulation in the structure subjected to its subsequent load
increment and then identify additional integration points which have failed.

The

aggregation of ruptured points is interpreted as a part of a macro-crack. The dependence


of numerical results on the finite element mesh discretization will be discussed in detail
in the following chapters.

In order to consider the crack closure effect, the capability of a ruptured integral point to
sustain the stress is recovered to the value it had at its critical state if the strain component
perpendicular to the crack direction at that point is found to be negative. Therefore, the
whole process of fatigue crack propagationfrom several material grain sizes to the final
rupture of a structure can be simulated by the proposed method with material parameters
which can be determined by uniaxial tests on smooth specimens.

99

5. FATIGUE LIFE PREDICTION OF SMOOTH SPECIMEN

5.1 INTRODUCTION

The model is first applied to predict the fatigue lives of smooth specimens under uniaxial
cyclic loading conditions. The material chosen for the present investigation was an asreceived Al alloy 2024-T3 Alclad plate with thickness 3.175 mm. Two main problems
are addressed in this chapter. One is the determination of material constants required in
the proposed damage model for the material used. The other is the prediction of S-N
curves of the material under different load levels.

5.2 DETERMINATION OF MATERIAL CONSTANTS

5.2.1 Uniaxial Tensile Tests

The uniaxial tensile tests on the material used in this investigation are identical to those
used by Wang (1987) to measure the true stress-true strain curve, the values of Young's
modulus E and Poisson's ratio
small number of tests were performed on the present batch of material to establish the
applicability of Wang's data, and the results were identical. Therefore Wang's data is
used in this part of the thesis. The dimensions of tensile specimens are shown in Fig.

IX

5 1. Light marks of 1 mm spacing were machined within the gauge length of 60 ram for
the measurement of large strains. The experiments were conducted on an Instron-IOT
universal testing machine.

When the initial deformation was small, the strain was

assumed to be uniform within the gauge length. Hence, only the current gauge length L
needed to be measured and the true strain could be calculated from

A s the uniaxial loading increased, the plastic deformation localized into a necking area.
In this case, the uniform assumption could no longer be used, instead each distance
between two neighbouring markswas measured and the corresponding true strain was

The corresponding true stress can be calculated by measuring the current cross-section
area S at the necking position
p

where P was the lied load. ^

(5.3)

relationship between true stress and true strain is

shown i n Fig. 5.2 for the damaged specimens along parallel or transverse to the rolling
direction. The experimental results indicated that only slight anisotropy could be
(Wang, 1987).

Due to localized damage in the specimen, three precautions had to be taken for the
101

measurements of E and 7 when necking occurred. First, small strain gauges were chosen
to ensure the accuracy of the strain measurements. Second, a strain gauge should be
replaced by interrupting the test when the strain accumulated in the gauge exceeded its
maximum limit.

Third, the values of E and

stresses/strains in the central part of unloading path in order to obtain a better accuracy
and minimize backlash at the beginning of either loading or unloading. The measured
results of E and u are shown in Fig. 5.3 and Fig. 5.4 respectively. These two figures
reveal the gradual material degradation behaviour with the increase of applied stress.
There was no significant difference between the specimens machined along the rolling
direction and those transverse to the rolling direction. A l l data is tabulated in Table 5.1
according to Wang's measurements (Wang, 1987).

The mechanical properties of

undamaged material were measured as follows:

E = 74300 MPa

= 0.33

cTy = 330 M P a

The damaged material parameters mentioned in section 3.5 can not be directly calculated
from above measured results due to the data scatter. The measured results are smoothed
out using a curve-fitting technique as shown in Fig. 5.1, 5.2 and 5.3 for the determination
of damaged material properties. The data used to calculate the damage parameters are
shown in Table 5.2.

From equations (3.69) and (3.70)the damage components dand/z were calculated by the
following equations

102

-l_v_2vv) +

l - y - 2 v v ) 2 + ( v -v ) [2 ( 1 - v v + 1 - 3 v ]
2 ( 1 - v ) v + 1 - 3v

[ 1 - 4V|i + 2 (1-v) n 2 ]

(54)

(5.5)

The calculated results are tabulated in Table 5.3. Therefore, the evolution curves of
damage components with true stresses/strains are depicted in Fig. 5.5 to Fig. 5.8. The
damage evolution coefficient 7 can be determined with equation (3.76) from the calculated
values of damage components d and
relationship between damage evolution coefficient 7 and true strain e is shown in Fig. 5.9.
It can be observed from the figure that 7 may be approximately taken as a constant of 0.7
in the loading process.

After the measurement of true stress and true strain, the strain hardening threshold can
be determined with the calculated values of damage components by a general method
described in section 3.5.3 for an isotropic material. Under uniaxial tension, the initial
threshold

is equal to crr

The increment of strain hardening threshold R follows

equation (3.78), and the overall plastic strain p is determined with equation (3.79). The
calculated results are shown in Table 5.5 and Fig. 5.10. Therefore, the function dR/dp
required in the constitutive plastic equations can be derived from the curve of R(p) shown
in Fig. 5.10. The yield surface of A l alloy 2024-T3 determined with equation (3.18) is

Fp{a,

(5.6)

D, R )

103

Similarly, another function dB/dWp required in plastic damage evolution equations can be
derived from the curve of B(wp).

Under uniaxial tension, the initial plastic damage

threshold of A l alloy 2024-T3 Alclad is given by equation (3.82)


330z

BQ =

74300

=0.741

The increment of plastic damage threshold B and overall plastic damage wp are calculated
with equations (3.83) and (3.84) respectively based on the calculated values of damage
components. The results are tabulated in Table 5.6. The value of critical total overall
damage wc, which corresponds to the state when the specimen fractures completely, is
obtained as 0.245 from the table. The relationship between B and wp is illustrated in Fig.
5.11 which is used to determine the function dB/dwp. Therefore, the plastic damage
surface shown in equation (3.41) can be written as

Fpd(

Yr B ) =

5.2.2 Uniaxial Fatigue Tests

The effect of mean stress on uniaxial direct fatigue endurance limit, defined as the fatigue
strength corresponding to 5 x l 0 7 cycles, is examined by a series of simple fatigue tests.
The dimensions of fatigue test specimens are shown in Fig. 5.12. A l l tests were
conducted on an AMSLER high frequency vibrophore machine. The loading is in tensioncompression or tension-tension with the loading direction normal to the specimen's rolling

104

direction at about 100 Hz. The results of tests are given in Table 5.7. As shown in Fig.
5.13, the fatigue endurance limit increases as the minimum stress decreases. There is a
sharp change when the minimum stress is close to zero. This implies that there exist the
different effects on fatigue damage between tension and compression. Under uniaxial
tension-tension cyclic loading, the critical value of Y fd and the minimum value of
determined by equation (3.86) can be calculated for the case of

< ay by following

formulae
(20,)4

^ d . l = 0-2558

(5.9)
=0.2558

If

>

d
u
e to the effect of plastic damage equation (5.9) is replaced by the equation

derived from equation (3.93)

Yfcii = ^ (3^+32^273^2) - t
1

pd

.-1.32u+1.34u2
(1-d)3

! )

1 0 )

1.34H-0.66
(1-d)2

Therefore, the relationship between Yfdjl and Ypd" is obtained by the measured results in
Table 5.7, and the calculated results are summarized in Table 5.8. It is postulated that
the value of Yfd corresponding to any value of Ypd""" can be determined by the following
method of interpolation
105

y
Y

4 _ /y-

fd.l

\ 4 j. (

^1*1

fd, i) ^

,,niin7

/T,min\ 1

(is n )7 + 1 " ( ^ ) 1

(5.11)

for

where (Yp^), ( Y ^ ) ^ ! are the measured values of Yp/^ and

(Y fd)1 )i,

(Yfdil)i+1 are the

corresponding values of Yfdji. If there is no plastic damage during a fatigue loading


process, equation (5.11) becomes

,
( a _i)i
f

o r

aj)i+i (a
( av . )
_ (a
kw . )
min':i+l
min' i

amin

min 1

i
( 5 12 )

( a min)i ^ 0 mn <

It has a similar form as that of the Goodman linear law within the minimum stress range
[ ( a )

except that the mean stress is replaced by the minimum stress. If the

number of tests i is more than 2the fatigue damage surface can be expressed as a multilinear relationship between (Yfdjl)I/4 and (Ypd0110)174 shown in Fig. 5.14.

From the general method proposed in section 3.5.6the material constant a can be
determined using finite fatigue life under tension-compression cyclic loading. According
to the measured results of two fatigue loadings, -111 139 MPa and -125 125 MPa,
shown in Table 5.11the average value of a is 0.48.

Alternatively? a simple method is given as follows. Under uniaxial tension-compression


fatigue loadings,

is always equal to 0. From equations (3.55) and (3.93), the

critical value of Yfd corresponding to the fatigue damage surface is

106

(^max _

Yfdtl

(5.13)

As shown in Table 5.7 and Table 5.8, the fatigue endurance limit corresponding to Y p ^
= 0 is 68 MPa. Thereforewe have

^max "

aa

m i n =2

x 6 8

136

= " ~
0

- min

(5-14)

Equation (5.14) can be used to determine the material constant a if the fatigue endurance
limit is measured under tension-compression cyclic loading. According to the measured
results of two fatigue loadings, -100 ~ 100 MPa and -47 -- 111 MPa, shown in Table 5.7,
the average value of a for Al alloy 2024-73 Alclad is 0.45 which is close to the value of
0.48 determined by the method proposed in section 3.5.6. In this investigation the value
of a was chosen as 0.45 for Al 2024-T3 Alclad.

In order to determine the material constant Kq, a total of 8 specimens were tested under
two different tension-tension fatigue loadings. The specimen failure is defined as the state
when a 2 to 3 mm crack appears.

The measured results are given in Table 5.9.

Therefore, the average value of Kq for material Al alloy 2024-13 Alclad is 423500 MPa
by equation (3.91) in conjunction with equations (3.87)(3.88) and (3.89).

5.2.3 Summary

In summary, the measured material constants for Al alloy 2024-T3 Alclad are

107

=74300 MPa

^=0.33

^ = ^ = 3 3 0 MPa

K O = 4 2 3 5 0 0M P a

wO=0.245

B 0 = 0 . 7 4 1M P a

7=0.7
=0.45

The function of R(p), increment of strain hardening threshold, is shown in Table 5.5 and
Fig 5 10. The function of B(wp)increment of plastic damage threshold, is shown in
Table 5 . 6 and Fig. 5.11.

For simplicity, these two functions can be approximately

expressed in multi-linear forms for numerical analysis such as

:a1(p1-0)

+a

(p-pi-i)

(5.15)

where

a.

:"

x ^ P < Pi

and
B = PDW+

(5.16)
11

where (^ =w

:
,.
p,i-wP,i-i

p. i-i

After the determination of the material properties, the proposed damage model can be
used to predict the fatigue behaviour of components or structures under cyclic loadings.

5.3 DAMAGE ACCUMULATION EQUATIONS

Under uniaxial loading condition, the plastic damage energy release rate components (Eqn.
3.39) are

E
l-4vu+2(l-v)tA
(1-d)3

-
2

(5.17)
= i H -(

"

v )

(1-d)

108

~2v-

and the fatigue damage energy release rate components (Eqn. 3.59) are

r.d, j

'act

(5.18)
min

'act

a (a

min

if crmin < 0 and a s


if amin < 0 and a < 0

Hence the plastic damage evolution equations (3.47) and (3.49) become
7) ( a 1

y a , ^ + y a

^ )
( 5 . 1 9 )

2 {al+a2+2ya x a 2 )

(ygi^ga)
( 5 . 2 0 )

2{ai*al^a1a2)-^

wp

2 (L-Y )

( - 2ydp\ip + \s.2p)

( 5 . 2 1 )

and the fatigue damage evolution equations (3.56) and (3.61) have similar forms
{ a x ^ a 2 ) {axydif^axyvif+ya2ydif+a2y^i
2 (

D-2+ 2 Y3 ^ 2 )1 _

( 5 . 2 2 )

Y
( +a2) (a1
( 5 . 2 3 )

2 (L-Y )

(c^f - 2 y ^ # f

af)

The stress a d can be determined by the plastic damage surface (Eqn. 3.41)
109

( 5 . 2 4 )

(a^ai^ya^)

B0

(5.25)

= 0

and the stress o"el csii be determined by the fatigue damage surface (Eqn. 5.11) as

'act, 1

2E^
( Yfd, l)l+1

(Yfd:

(5.26)

mm 4

(^ n)L
for

(FX). ^ FPT < (C

lc

jmin

^ 1c

'act, 1

iin\

i f

a (oIC - amiI1) ,

Therefore, the total damage can be calculated by a step-by-step method proposed m


section 4.3 and expressed as
v

=
.

v i

d = 5. a d ,

^ =E i

( 5

'

2 8 )

where Aw, Adand AMi are the increments of damage during a loading step. They can
be calculated from equation (4.28). For a simple case in which there is no plastic damage
accumulation, we have

110

^lc

dt

(5.29)

Afij = Ni
Ja

lc

where Nj is the number of cycles at a step.

5.4 RESULTS

5.4.1 Experiments

The dimensions of specimen used in uniaxial cyclic loading tests are shown in Fig. 5.12.
The material is an as-received A l alloy 2024-T3 Alclad plate with thickness 3,175 mm.
The intercept length of the grain is about 0.08 mm, and only slight differences is observed
between along the rolling direction and transverse to the rolling direction. Fatigue tests
were carried out on an AMSLER high frequency vibrophore machine at a cycle frequency
of 100 H z and the loading direction was normal to the specimen's rolling direction. The
fatigue life was defined as the number of cycles corresponding to the critical state when
a 2 to 3 mm crack appeared in the specimen.

The testing machine was set to stop

automatically when the specimen reached its critical state. About 130 specimens were
tested under different cyclic stress levels, including both tension-tension and tensioncompression loadings with at least three specimens tested at each stress level for most of
the cases. The test results are shown in Table 5.10 to Table 5.15. The first three tables
show test results of fatigue lives under the constant cyclic stress ranges. Three various

111

rangesnamely 200 MPa, 250 MPa and 340 MPa, are chosen. The later three tables
display test results of fatigue lives under the constant mean stresses. The mean stresses
chosen in tests are 60 MPa, 135 MPa and 245 MPa. The range of fatigue life is 10^ to
107 cycles. It is observed that the data scatter is less than 20%except in some cases
when the fatigue life exceeded 106 cycles.

5.4.2 Fatigue Life at Constant Stress Range

The accumulated damage can be calculated during fatigue loading process with equation
(5.28). When the total overall damage w reaches its critical value 0.245the specimen
is said to have reached its critical state and the corresponding number of cycles is defined
as the predicted fatigue life. For three different constant cyclic stress ranges of 200 MPa,
250 MPa and 340 MPa, the effects of cyclic maximum stress on the fatigue life were
examined by both theoretical and experimental methods. The experimental results are
summarized in Table 5.10, 5.11 and 5.12, and compared with the predicted results shown
in Fig. 5.15, 5.16 and 5.17 respectively. The solid line represents the calculated results
which are in good agreement with the experimental results. As the cyclic maximum stress
increases, the corresponding fatigue life decreases sharply. It is observed from Fig. 5.15
that there are two discontinuities on the predicted curve slope: one corresponds to the
yield stress which is associated with the onset of plastic damage, and the other is
associated with the change of cyclic minimum stress from tension to compression. As the
stress range increases from 200 MPa to 250 MPa, these two discontinuous points move
gradually close shown in Fig. 5.16. When the stress range reaches 340 MPa which is
large enough to include two discontinuous points at the same time, only one discontinuity
112

can be observed because these two points become close together shown in Fig. 5.17.
is suggested that these discontinuities are associated with the change of damage mechanism
from one phase to another. The higher discontinuous point corresponds to the transition
between macro-plasticity and micro-piasticityand the lower discontinuous point is
associated with the transition from tension to compression. The test results shown in Fig.
5.155.16 and 5.17 provide evidence of some degree of success. Due to the scatter of
fatigue data, more tests are desirable and a statistical method should be introduced in
order to give greater confidence.

Of particular interest is that discontinuities associated

with plastic instability in the S-N curves have been reported by some investigators for
many materialsincluding mild steels and aluminium alloys (Tabeshfar and Williams,
1974).

Further studies are required to find the mechanisms associated with the

discontinuities.

Fig. 5.18 demonstrates the relationships between the cyclic maximum stress and fatigue
life at two different cyclic stress ranges, 200 MPa and 340 MPa respectively. It is
obvious that the fatigue life decreases with the increase of cyclic stress range for a certain
value of cyclic maximum stress.

5.4.3 Fatigue Life at Constant Mean Stress

The relationships between the stress range and fatigue life at constant mean stresses are
now examined. The fatigue lives corresponding to various cyclic stress ranges can be
calculated from equation (5.28) and the comparison between the predicted and the
experimental results is shown in Fig. 5.195.20 and 5.21 for three different mean

113

It

stresses, 60 MPa., 135 M P a and 245 MPa. Good agreement is obtained. The fatigue life
increases as the cyclic stress range decreases for a certain value of mean stress. The
comparison of S - N curves between two different mean stresses, 60 M P a and 245 MPa,
is shown in Fig. 5.22. It is observed from the figure that the difference between two
predicted curves decreases as the fatigue life increases. This means that the effect of
mean stress on fatigue life is relatively small for the case of high cycle fatigue.

It is obvious that Fig. 5.15 to Fig. 5.22 demonstrate the validity of the proposed damage
model for the fatigue life prediction of smooth specimens over the range 1(^-107 cycles.
For the range of 10-lCf cycle, other factors, such as plastic instability and creep damage
may cause complication in the analysis. Modifications to the proposed model are required
to account for these important factors.

These are outside the scope of the present

investigation.

5.4.4 Effect of Rolling Direction

The fatigue tests with the loading direction parallel to the specimen's rolling direction
were also conducted on an AMSLER high frequency vibrophore machine at 100 Hz in
order to check the effect of specimen's direction on fatigue life. Five specimens are
tested for two different cyclic stress levels, 0-250 MPa and 50-300 MPa. The test results
are tabulated in Table 5.16.

Comparing the corresponding data of two specimen's

directions shown in Table 5.10 and Table 5.16, the discrepancy is within 10%.
Therefore, the material used in this investigation can be taken as an isotropic material.

114

chapter six

6. f a t i g u e f a i l u r e p r e d i c t i o n o f n o t c h e d s p e c i m e n s

6.1 i n t r o d u c t i o n

To predict life to crack initiation (millimetre sized cracks) of notched specimens, two
approaches are often used to account for the local stresses/strains: the so-called notch
analysis concepts such as Neuber rule, and the FEM-based analysis. Based on these local
stresses/strains analyses, several criteria can be used to estimate fatigue life, such as: (1)
maximum principal stress, maximum shear stress or octahedral shear stress, (2) maximum
shear or octahedral shear strain, (3) cyclic plastic work, (4) other correlation parameters
that incorporate both local stress and local strain. Although these criteria have been
successfully used in some fields, an essential problem in establishing an active notch root
size/volume still remains unresolved.

There always exist difficulties in choosing a

possible characteristic size/volume which is needed in these approaches.

An approach to notch fatigue life prediction based on the proposed damage model is
described in this chapter. The introduction of a "critical" size is avoided by this method.
According to the fatigue damage model proposed in chapter three, fatigue failure in a
component can be dealt with by a concept of the "material element". The mechanical
properties of a material element change during the fatigue loading process because the
micro-cracks initiate and grow in this element. The overall damage in a material element

115

can be calculated based on the damage evolution equations by F E M method. When the
overall damage reaches a critical value, which is a material constant, the material element
is said to be ruptured. The aggregation of these ruptured elements forms a macro-crack.
By this means, it is possible to simulate the whole process of fatigue crack propagation
from a crack with the length of several times of grain size to the final failure. Therefore,
the fatigue life corresponding to a crack with a certain length can be predicted for a
notched specimen.

Three different notched specimens were chosen for the present investigation for both
experiment and theory. The material used was an as-received A l alloy 2024-T3 Alclad
plate with thickness 3.175 mm. Special emphasis was placed on the discussion of the
effect of the finite element discretization on the fatigue life prediction.

6.2 D A M A G E A C C U M U L A T I O N

The specimens are assumed to be in a state of plane stress during loading process. For
the case of plane stress,
o

Then, the plastic damage energy release rate components (Eqn. 3.39) become

116

[ o i - 2^0^2
(6.2)

[al + 2b2a

where

a2,

b2 are the functions of damage components and can be expressed as

1-4VH+2 ( 1 - v ) n2
(1-d)3

=
2

2n(1-v) -2v
(l-cf)2~
( 6 . 3 )

v~2p. ( 1 - v ) - [ i ( l - 3 v )
1 - 4 V | J I + 2 ( 1 - v ) \i2

(1-v)-2v

The fatigue damage energy release rate components (Eqn. 3.59) are

^drf

(6.4)
f

l,acc

-^2 a i,acC Clr 2,act

where
equation (3.52). The fatigue damage evolution equations are the same as equations (5.19)
to (5.24). Therefore, the damage accumulation in specimens can be calculated by the
method described in chapter four.

6.3 STANDARD TENSILE SPECIMEN

A n important difference between the behaviour of statically-loaded specimen and a


cyclically-loaded specimen of the same geometry such as the standard tensile specimen
117

shown in Fig. 5.1 is examined. As an example, a numerical analysis was performed with
a 2 by 2 Gaussian numerical integration scheme on the standard tensile specimen for both
the tensile loading and the cyclic loading. Only a quarter of the specimen needed to be
analyzed due to the symmetry of the geometry and the loading conditions. The finite
element network used is shown in Fig. 6.1 and Fig. 6.2. The dimension of the smallest
element normal to the loading direction shown in Fig. 6.2 is 0.48 mm which was used to
simulate the fatigue crack propagation described later.

A total of 86 B-nodes

isoparametric plane stress elements with 292 nodes were chosen. A step by step method
described in section 4.3 was used to simulate the cyclic loading. The number of cycles
at each step N was 2000 cycles. The material constants are given in section 5.2. Stress
and damage parameters were calculated at the Gaussian points by a modified version of
NFAP (Nonlinear Finite element Analysis Program).

For the uniaxial tensile loading, the location of maximum equivalent stress and initial yield
stress appears first at the shoulder if the applied stress is less than the yield stress, as
shown in Fig. 6.3. As the applied stress increases, the location of maximum equivalent
stress moves gradually to the parallel section of the specimen (Fig. 6.4). Consequently
ductile fracture occurs in general in the parallel section of the specimen.

However, this is not the case for the fatigue failure under high- and medium-cycle
loadings. As the cyclic loading proceeds, the fatal fatigue crack first appears at the
shoulder (Fig. 6.5) and then propagates along the direction normal to the applied loading
direction (Fig. 6.6) which is in satisfactory agreement with the experimental observations
(Fig. 6.7). It is observed, from Fig. 6.5 and 6.6 that the fatigue failure happens quickly
8

IX
1

after a short crack (about 0.1 mm) appears. TMs means that the number of cycles spent
on macro-crack propagation to final fracture is very small for this type of specimen. Only
about 2% the loading cycles are spent on macroscopic fatigue crack propagation during
the loading process.

A series of both numerical analyses and experiments were carried out to determine the
number of cycles to fatigue failure in the specimen. The fatigue life is defined as the
critical state when a 3 mm crack appears in the specimen. All tests are conducted on
an AMSLER high frequency vibrophore machine.

The machine was set to stop

automatically when the specimen reached its critical state. The applied loading was in
tension-tension with the loading direction normal to the specimen's rolling direction at a
frequency of about 100 Hz. A total of 17 specimens were tested for five different loading
ranges with same maximum cyclic stress of 290 MPa and the results are summarized in
Table 6.1. The table includes the predicted results not only from the two-dimensional
model described here but also from one-dimensional model described in chapter five using
the assumption that the fatigue failure happened at the parallel part of the specimen where
the stress/strain state was uniform. The values of applied stress are the corresponding
engineering stresses at the parallel part of the specimen. The discrepancy between the test
results and the calculated results based on the two-dimensional model is within 10% for
the first four stress ranges and about 24% for the last stress range.

However, the

difference between the test results and the calculated results based on the one-dimensional
model ranges from 40% to 240%depending on the cyclic stress range. It is obvious that
the two-dimensional model is much better than the one-dimensional one.

119

6.4 SPECIMEN W I T HA N INCLINED NOTCH

In this section, the fatigue failure of a plate with an inclined notch (shown in Fig. 6.8) is
analyzed. The material is an as-received A l alloy 2024-T3 Alclad plate with thickness
3.175 mm. A l l experimental tests under the cyclic loading were performed on an
A M S L E R high frequency vibrophore machine. The loading was applied in the direction
normal to the specimen's rolling direction at a frequency of about 100 Hz. Fatigue crack
initiation and growth were identified using a travelling microscope.

The material

constants required in the fatigue damage model were those determined in section 5.2 and
all numerical analyses were performed using a modified version of NFAP.

First, the effect of N chosen at each step on the convergence of numerical results was
examined with an applied cyclic stress of 50+40 MPa.

A total of 142 8-nodes

isoparametic plane stress elements with 446 nodes were chosen, as shown in Fig. 6.9 and
Fig. 6.10 and the element area integrations were performed with 2 by 2 Gaussian
numerical integration scheme. The dimension of elements used to simulate macro-cracks
(shown in Fig. 6.10b) was 0.5 mm along the crack propagation direction. Three different
values of N, 1000, 2500 and 5000 cycles, were chosen to simulate the cyclic loading.
Based on the numerical analysis presented in chapter four, the fatigue lives of the
specimen corresponding to a crack of length 3 mm were 108000, 110000 and 115000
cycles respectively. These results are so close that the effect of N on the numerical result
can b e neglected. However, divergence of numerical results occurs if N is larger than
a value for which two integral points are in the critical state at the same step. This is
because releasing two or more integral points in a step results in instability in the FEM

120

analysis and so care has to be taken in choosing an appropriate N to simulate the cyclic
loading.

It is important to check whether the number of integral points in an element affects the
fatigue life prediction because the aggregation of ruptured integral points is used to
simulate the macro-crack growth. Three numerical integration schemes, 2 by 2, 3 by 3
and 4 by 4 Gaussian points, were chosen to perform element area integrations for the
finite element network shown in Fig. 6.9 and Fig. 6.10. The number of cycles N at each
step was 1000 cycles. The numerical results show that the fatigue lives for crack with
an initial length 3 mm are 108000, 109000 and 110000 respectively corresponding to these
three different schemes. The difference is very small. The effect of finite element size
on the estimated fatigue life of the specimen was also examined. The fatigue lives were
110000, 97500 and 100000 cycles respectively for three meshes with the dimension 0.5
1.0 and 1.5 mm along the crack propagation direction. The biggest discrepancy is about
13%. These analyses imply that the numerical results converge for the different numbers
of integral points used to simulate crack. If a large number of integral points were
chosen to simulate a crack by increasing Gaussian points in an element or alternatively by
decreasing the element size in the direction of crack propagation, the distance between two
neighbouring points became small and less loading cycles for a crack passing these two
integral points were needed. Conversely, more loading cycles are required due to the
increment of the distance between two nearby points i f a small number of integral points
are chosen by decreasing Gaussian points in an element or increasing the finite element
size in the direction of crack propagation. Therefore, there is the negligible net effect of
the finite element size and the number of Gaussian integral points on the numerical
IX

results. The reason for this result is that the stress gradient near the notch is not very
large. When the stress gradient is high, the discrepancies increase. This problem is
discussed in chapter seven.

Fig. 6.11 is the equivalent stress distribution near the notch under an applied stress of 90
M P a in a tensile test. There are two small yield fields in which the equivalent stress is
over 330 MPa. As the number of cycles increases, the stress amplitudes decrease in the
local elements near the notch during the fatigue loading process.

The reduction of

amplitude reaches 15-20% in the material element when it is in the critical state, i.e. at
the failure threshold.

This is attributed to the fact that damage is developed in the

material under the cyclic loading, causing a reduction in the elastic modulus and inducing
stress redistribution in the specimen. Fig. 6.12 is the equivalent stress distribution near
the notch under applied stress 90

at 50000 cycles with the cyclic loading 50+40

MPa. It is obvious by comparing Fig. 6.11 and Fig. 6.12 that the maximum equivalent
stress decreases after the fatigue loading because the capability of sustaining stress
decreases in the material elements near the notch boundary, but the stress distributions are
same for those elements which are far from the notch boundary. This is because the
overall damage is very high near the notch boundary, and decreases sharply from this
position, as shown in Fig. 6.13. From the numerical results, it is clear that the fatigue
crack initiates in the element close to the notch root, where the overall damage first
reaches its critical value (shown in Fig. 6.13) and grows normal to the loading direction
shown in Fig. 6.14 as the cyclic loading continues. The predicted crack track is very
close to that observed in experiments (Fig. 6.15). It is observed from Fig. 6.14 that the
stress distribution around the crack is different from that of notch (shown in Fig. 6.12).
I
x

Howeverthe damage still concentrates on the elements close to the crack, as shown in
Fig. 6.16.

Two stress ranges, 50 2 5 MPa and 50 4 0 MPa, were applied to test fatigue crack
initiation and propagation in the specimen. Five specimens for each loading were used.
It was not so easy to identify the crack length when the crack was very small because the
crack always initiated at the comer of the notch boundary and there were many microcracks or arrested cracks in the same position. For this case, the stress state around the
crack was very complex. However, when the crack grew to size which equalled the
thickness of the plate, it was found that the crack lengths measured from both surfaces of
the specimen were nearly identical. Therefore, the plane stress assumption can only be
used in the case for which the crack length was larger than several millimetres.
Accordingly, the predicted and experimental results for fatigue life corresponding to
cracks with a certain length, 2 mm or 3 mm, could be compared, which is done in Tables
6.2 and 6.3.

For a 3 mm crack, the discrepancy between the test results and the

calculated results is 1% for cyclic loading 50 4 0 MPa and 24% for the cyclic loading
5 0 + 2 5 MPa.

For a 2 mm crack, the corresponding discrepancy is 5% and 33%

respectively. It is obvious that the discrepancy for a 3 mm crack in Table 6.3 is much
less than that for a 2 mm crack in Table 6.2. This is because the estimated life is based
on the assumption of plane stress, so the error involved in this assumption for a short
crack is larger than that for a long crack.

6.5 SPECIMEN WITH AN EDGE NOTCH

123

The fatigue failure of a specimen with an edge notch is discussed in this section. The
dimensions of the specimen are shown in Fig. 6.17. Only one half of the plate was
analyzed due to the symmetry of geometry and loading. Fig. 6.18 and Fig. 6.19 aie the
finite element networks for the numerical analysis and four different networks near the
notch region were chosen to examine the effect of element size on the numerical results,
as shown in Fig. 6.19. In Fig. 19a there are 118 8-nodes isoparametric plane stress
elements with 397 nodes; in Fig. 19b there are 108 8-nodes isoparametric plane stress
elements with 367 nodes; in Fig. 19c there are 91 8-nodes isoparametric plane stress
elements with 314 nodes and in Fig. 19d there are 84 8-nodes isoparametric plane stress
elements with 291 nodes. The corresponding dimensions of the minimum elements along
the crack growth direction are 0.24 mm, 0.48 mm, 0.72 mm and 0.96 mm respectively.
The element area integrations were performed with 2 by 2 Gaussian numerical integration
scheme. The number of cycles in step was 1000 cycles. Under the cyclic loading
5030 MPa, the calculated fatigue lives corresponding to a 2 mm crack are 102,000
cycles, 104,000 cycles, 105,000 cycles and 106,000 cycles respectively for the four
different element sizes. These results are very close. Therefore, the effect of element
size on the numerical results can be neglected if the dimensions of the finite element is
about several times the material's grain size.

The convergence of the numerical results for several values of N


was also examined. Six
values of Ni, namely 500, 1000, 2000, 2500, 5000 and 10000 cycles, were chosen. The
corresponding calculated fatigue life for a 2 mm crack is 104000, 104000, 108000,
125000, 130000 and 140000 cycles respectively. The discrepancy is within 5% when the
value of Nj changes from 500 cycles to 2000 cycles. The results start diverging when the

124

value o f

excccds 2500 cycles. Quite obviously, it is better to choose a smaller N^, but

a smaller value of Nj costs computing time. Therefore, careful choice of Nj is essential.

From the above analyses, some parameters can be determined for the numerical
simulation. A total of 108 8-nodes isoparametric plane stress elements with 367 nodes,
as shown i n Fig. 6.16 and 6.17were chosen for FEM analysis. An integration scheme
with 3 by 3 Gaussian points was used to perform element area integration. The dimension
of the elements used to simulate the fatigue crack propagation is 0.5 mm along the crack
growth direction and the number of cycles in a step N was chosen to be 1000 cycles.
Under the cyclic loading 50+30 MPa, a small crack (about 0.1 mm) is predicted to
appear at the root of notch at 57000 loading cycles, as shown in Fig. 6.20. It was
observed that the plastic deformation zone is very small, as is the damage zone shown in
Fig. 6.21. As the cyclic loading continues, the simulated crack grows gradually normal
to the loading direction. The predicted crack track is shown in Fig. 6.22 and 6.23 which
is very close to the result observed in experiments (Fig. 6.24). It is clear fromfigures
6.22 and 6.23 that both the plastic deformation and total damage are still limited to the
area near the crack. The plastic deformation zone increases quickly as the applied cyclic
mean stress increases. This change is shown in Fig. 6.25 and Fig. 6.26 for the cyclic
loading 10030 MPa. The corresponding damage zone does not change much at first
under the higher stress, as shown by comparing Fig. 6.21 and Fig. 6.27. Howeveras
the fatigue proceeds, the differences become obvious, c.f. Fig. 6.23 and Fig. 6.28.

The fatigue lives of the specimen corresponding to a 3 nun crack were obtained for four
different applied cyclic loadings by both analysis and experiment. The test method was

125

the same as that described in section 6.4. The applied stresses were 5020 MPa, 5030
MPa, 50+37.85 MPa and 10030 MPa. Five specimens were tested for each cyclic
loading. The comparison between the predicted and experimental results are summarized
in Table 6.4. The discrepancy between thetest results and the calculated results is about
15% for the cases of 5020 MPa and 5030 MPa. As the applied cyclic maximum
stress increases, the discrepancy increases from 23% for the cyclic loading 5037.85
MPa to 42% for the cyclic loading 100 30 MPa.

Compared with the data scatter

observed in the experimental tests, the differences between the predicted fatigue life and
the measured ones shown in Table 6.4 is reasonable except for the case of cyclic loading
10030 MPa.
It is suggested that the error is mainly caused by the assumption that total damage is the
linear summation of fatigue damage and plastic damage and there is no effect of fatigue
damage accumulation to the plastic damage threshold. The contribution of the plastic
damage to the fatigue life is over-estimated in the proposed damage model. Nonlinear
interaction should be considered when the plastic deformation is large. A possible way
to do this is to consider the effect of fatigue damage on the plastic damage threshold
B(wp).

If the plastic damage threshold is supposed to increase with fatigue damage

accumulationthe estimated plastic damage accumulation will decrease during a loading


processHowever, it is not so easy to identify the plastic damage threshold.

6.6 SUMMARY

The proposed damage model has been applied to the fatigue failure of three kinds of
notched specimens under different tension-tension cyclic loadings and good agreement

126

between the estimated ^nd the experimental results have been obtsincd. Some of the main
results are summarized as follows:

1. The model has been successfully used to predict not only the fatigue life of notched
specimens, but also the fatigue crack track in the specimen. An important phenomenon
is observed from the test and verified by theoretical analysis that the fatigue crack initiates
at the shoulder rather than at the parallel section of the standard tensile specimen, as in
the case of ductile fracture under uniaxial tensile loading.

2. The estimated results are independent of the distance between two neighbouring
integration points i f the finite element dimension along the crack growth direction is about
several times the material grain size. Due to convergence, the choice of a reasonable
value for the number of cycles in a step Nj is possible in order to simulate the cyclic
loading.

3. Compared with the data scatter observed in the fatigue tests, the discrepancy between
the calculated results and the measured results is reasonable for most applied cyclic
loadings concerned in this chapter. Some modifications are required to take into account
the nonlinear interaction between fatigue damage and plastic damage for the cases when
large plastic deformation appears in a material element during the fatigue loading process.

127

chapter seven

7. f a t i g u e f a i l u r e p r e d i c t i o n o f c r a c k e d s p e c i m e n s

7.1 i n t r o d u c t i o n

Since the inception of the Paris law for fatigue crack propagation (FCP) based on the
fracture mechanics, this global approach has become popular among practising engineers
in the prediction of fatigue failure in engineering structures. Generally, the fatigue crack
propagation rate is governed by a general loading parameter, such as the applied stress,
the stress intensity factor or the J-integral. Thus, a generalized law is

where P is a general loading parameter. By integrating Eqn.(7.1) from the initial to a


certain crack length, the corresponding number of cycles is determined. Therefore, it is
very important to establish a function f(P) under complex loading conditions. A large
number of F C P models have been proposed empirically and theoretically to describe the
crack behaviour in their respective phases and these have been reviewed in section 2.3.
Due to the complex nature of fatigue damage, the process of fatigue crack growth has
been artificially divided into several phases, such as the behaviour of micro-cracks, short
macro-cracks and long cracks. Consequently, many experimental tests need to be earned
out to determine the material constants for each FCP formulation. Furthermore, there
128

often exist problems in the transition between two phases of the fatigue crack growth due
to the limited applicability of each formulation.

This chapter is concerned with the application of the proposed fatigue damage model to
fatigue crack growth in a plate containing a centre crack. Problems with the simulation
of cyclic loading and the dependence of the numerical results on the finite element mesh
discretization are discussed in detail and a simplified procedure is proposed for numerical
analysis of fatigue crack propagation.

7.2 E X P E R I M E N T S

The material was an as-received A l alloy 2024-T3 Alclad plate of thickness 3.175 mm.
The dimensions of the specimen were shown in Fig. 7.1. The experimental tests are
carried out on an A M S L E R high frequency vibrophore machine. The cyclic loading was
in tension-tension with the loading direction normal to the specimen's rolling direction at
a frequency of about 100 Hz. The fatigue crack growth was followed using a travelling
microscope. The material constants required in the fatigue damage model were measured
by a series of uniaxial tensile tests and uniaxial fatigue tests on smooth specimens and are
presented i n section 5.2.

A 20 mm centre crack was first machined in the specimen. Then the pre-loading of 458
MPa was applied to the specimen in order to make a sharp crack. The preparation of the
cracked specimen was finished when a 28 mm centre crack appeared. Four different
levels of constant amplitude cyclic loadings, 4524 MPa, 4516.5 MPa, 4 5 1 1 . 5 MPa
1A

and 90 11.5 MPawere applied. A total of 18 specimens were tested. The test results
of the fatigue crack propagation are summarized in Tables 7.1 to 7.4. The total increment
of the crack length was about 30 mm. The crack growth rate increased about 500% from
the start to the finish during the tests. It was observed from these tables that the data
scatter is about 30% at the beginning and changes slightly as the fatigue loading continues.
When the crack length was close to its critical value, the scatter increased sharply due to
unstable crack propagation.

In order to examine the fatigue crack behaviour under block cyclic loadings, a two-step
fatigue loading was applied to the specimen. A cyclic loading of 4 5 + 1 6 . 5 MPa was first
employed from 0 to 60000 cycles, then this loading was replaced by 45 11.5 MPa until
the increment of crack length was about 30 mm. Two specimens were tested and fatigue
crack growth results are tabulated in Table 7.5. The data scatter is about 10%. As
expected, there is a discontinuity in the crack growth rate at 60000 loading cycles. The
fatigue crack growth rate decreases sharply when the applied loading changed from
4516.5 Mpa to 4511.5 Mpa.

7.3 NUMERICAL ANALYSIS

The first problem to be examined was the simulation of cyclic loading. It is impracticable
to calculate the damage accumulation cycle by cycle, especially under high cycle fatigue.
So a step-wise procedure was proposed in chapter four to save computing time. In this
way, the whole loading process can be divided into several steps. In order to check the
effect of the number of cycles in a step N on the convergence of the numerical results,

130

a FEM analysis was performed. The applied loading was 45 16.5 Mpa. The specimen
was assumed to be in a state of plane stress in the numerical analysis. Only a quarter of
the specimen needed to be analyzed due to the symmetry of the geometry and the loading
condition. A total of 90 8-nodes isoparametric plane stress element with 306 nodes were
chosen. The element area integrations were performed with 2 by 2 Gaussian numerical
integration scheme. The finite element mesh near the crack tip is shown in Fig. 7.2b.
The dimensions of the elements used to simulate the macro-crack are 0.08 mm x 0.32
mm. Three different values of N, 1000 cycles, 500 cycles and 250 cycles, were chosen
and the corresponding results on the fatigue crack growth simulation are shown in Fig.
7.3. No sign of convergence can be seen so that more steps should be taken in order to
attain a more accurate result. However, if the damage accumulation is first calculated by
a larger Nuntil the overall damage at the integration point near to the crack tip is close
to the critical value and then a relatively smaller N i 0 =50 cycles is chosen for the release
of rigidity and strength at this point, more accurate results in terms of convergence were
obtained (Fig. 7.4). There is almost no difference for Nj < 500 cycles. It is thus clear
that the numerical divergence is mainly caused by the release of ruptured integration point
due to the sharp change of stress distribution near the crack tip in the simulation. It is
known from the numerical results that the number of cycles is about 1500-2000 cycle
when a crack propagates from one integration point to the ensuing point at the beginning
of the crack growth, so it is enough to divide this loading process into 3 steps for this
second method.

There is only about 10% discrepancy even if one step is chosen.

Comparing the curves in Fig. 7.3 and Fig. 7.4 for 1^=250 cycle, the difference is about
10%.

Therefore, for the first method when the number of cycles in a step remains

constant during the whole loading process, the loading process from the rupture of an
IX

IX

integration point to the following one should be divided into at least 6 steps to achieve a
satisfactory result. Therefore, the number of cycles in a step Nshould be less than 250
cycles for the first method and 500 cycles for the second under the cyclic loading
4516.5 MPa in order to limit the discrepancy within 10%.

If the value of N is

increased to 500 cycles for the first method, the computing time is reduced by about 40%
but the discrepancy increases to 25 %. Although the second method is more efficient than
the first, it is difficult to realize in the numerical program for some complex cases because
some criteria are required to control the change between the Nand N;o. Furthermore,
more memory space in the computer is necessary for the second method.

Another problem is the dependence of numerical results on the finite element size. Four
different sizes of finite element mesh (Fig. 7.2a) are used in the direction of crack
propagation. They are 0.16 mm, 0.24 mm, 0.32 mm and 0.48 mm respectively with the
same dimension 0.08 mm along the loading direction. The cyclic loading was simulated
by NSOO cycles and^ = 5 0 cycles according to the procedure described above. It can
be seen in Fig. 7.5 that there is no marked difference except for the case of 0.48 mm.
The maximum difference in fatigue life between the first three cases is about 22%. The
effect of the mesh discretization, on the fatigue crack growth was also examined. The
calculated results show that the difference in simulated fatigue crack growth between two
kinds of mesh in Fig. 7.2 is about 10%.

It is obvious from the above discussions that reasonable numerical results can be obtained
in comparison with the data scatter normally observed in fatigue testing if the simulation
parameters are carefully chosen. In order to check the accuracy of the predicted results,

132

the numerical analyses were performed for fatigue crack propagation under different cyclic
loadings. A total of 117 elements with 405 nodes were chosen. The mesh discretization
near the crack tip is shown in Fig. 7.2b with several 0.08 mm x 0.32 mm elements
which are used to simulate the crack growth. In order to save computing time, the cyclic
loading simulation was carried out by a third method which is a modified version of the
first. In this, Nis a constant only for several steps, not for whole loading process. The
value of Nis dependent on the applied loadings and the crack growth rate and is chosen
with the condition that the discrepancy resulting from the cyclic loading method is limited
within 10%. This third method allows significant computer time saving, and is used to
compare the model with experiment.

7.4 FATIGUE CRACK GROWTH UNDER CONSTANT AMPLITUDE LOADINGS

Both calculated and measured results on the fatigue crack growth are shown in Fig. 7.6
and Fig. 7.7. Four different levels of constant amplitude cyclic loadings are concerned.
They are 45 2 4 MPa, 4516.5 MPa, 4511.5 MPa and 9011.5 MPa respectively.
It is shown in Fig. 7.6 that the calculated results are in good agreement with the measured
results at the beginning of the crack growth under the cyclic loadings 45+16.5 MPa and
45 + 11.5 MPa, and the discrepancy increases slightly as the fatigue loading continues.
The maximum difference between calculated and average number of cycles corresponding
to crack of a certain length is about 15%. The discrepancy increases to about 25%
when the applied cyclic stress amplitude increases to 24 MPa for the same mean stress of
45 MPa, as shown in Fig. 7.7. Compared with the data scatter observed in the tests, the
discrepancy is reasonable.

It is obvious that the estimated results of fatigue crack

133

propagation are acceptable for the first three loadings.

7.5 FATIGUE CRACK GROWTH UNDER TWO-STEP LOADINGS

Fatigue crack propagation under the two-step loading was examined. The cyclic loading
45 16.5 MPa was first applied to the specimen from 0 to 60000 cycles, then the loading
was replaced by 45 11.5 MPa.

The results obtained both experimentally and

theoretically are displayed in Fig. 7.8. The dashed line is the predicted result for the case
of constant amplitude loading 45 16.5 MPa during the whole loading process, and the
solid line represents the calculated result under the two-step loading. When applied cyclic
amplitude changes from 16.5 MPa to 11.5 MPa after 60000 cycles, the local cyclic stress
range in an element near the crack tip also decreases. The damage accumulation rate in
a cycle becomes smaller in the element. Thus, more loading cycles is required for the
crack propagation from one integration point to another. Therefore, the crack growth rate
becomes smaller as the applied cyclic amplitude decreases, as shown in Fig. 7.8. It is
known from the results that the discrepancy of the fatigue life corresponding to the crack
with a certain length is about 20%, which is quite reasonable for the fatigue loading.
Therefore, the proposed damage model may be used to predict the fatigue crack
propagation under block loading which is often used to simulate variable amplitude
loading.

7.6 SUMMARY

The application of the fatigue damage model to fatigue crack growth in a 2024-T3 Alclad

134

plate containing a centre crack under both constant amplitude loading and two-step loading
has been presented.

The analyses of fatigue crack propagation were carried out both

experimentally and numerically and good agreement has been achieved for most cases.
Two main problems are addressed and resolved in this chapter:

1. The simulation of cyclic loading was performed in a step-wise procedure, not cycle by
cycle, in order to reduce the computing time. The effect of number of cycles chosen for
each step on the accuracy of the numerical results was examined and a simple criterion
proposed to gain better convergence in the numerical solutions.

2. The dependence of the numerical results on the finite element mesh size and the mesh
discretization were discussed. Satisfactory results have been obtained in comparison with
the data scatter normally observed in the fatigue tests if the mesh size is not too large.

135

CHAPTER EIGHT

8. CONCLUSIONS

1. A damage model based on the damage mechanics theory has been proposed for highand medium-cycle fatigue problems without extensive plastic deformation. Depending
upon the level of stress developed within a material element, the damage accumulation
under a loading process is subdivided into two parts, fatigue damage and plastic damage,
which are identified by the fatigue damage surface and the plastic damage surface. With
the introduction of a new damage effect tensor, the damage evolution equations are
formulated based on the hypothesis that overall damage is induced by the linear
summation of fatigue damage and plastic damage. When the total overall damage reaches
a critical value, the material element is postulated to be ruptured. The difference between
fatigue damage accumulation under tension and compression is also considered.

2. A finite element implementation of the proposed damage model has been made. Three
major problems associated with this computational effort are solved. These are:
(i) the asymmetric effective instantaneous tangent modulus matrix was replaced by a
symmetric form to evaluate the element stiffness matrix in an iteration procedure. In this
way, the conventional FEM program can be readily modified to include the effect of
damage.

However, the solution convergence rate may be slow in the numerical

integration of the differential stress-strain equations when total damage is large.


(ii) in order to save computing time, simplified approach was proposed such that
i
x

fatigue loading was simulated with several steps, rather than cycle by cycle.

The

discrepancy resulting from this approximate method on the nunierical results was
examined and a simple criterion proposed to gain better convergency in the numerical
solution.
(iii) The crack propagation was simulated by releasing a series of integral points when,
they reached the critical state. The crack closure effect was considered in the numerical
analysis.

Satisfactory results have been obtained in comparison with the data scatter

normally observed from the fatigue tests if the mesh size used to simulate the crack
growth is not too large.

3. More than 150 smooth specimens of Al 2024-T3 Alclad were tested under uniaxial
tension-tension or tension-compression cyclic loadings. The S-N curves at different mean
stresses were predicted, and these showed showing excellent agreement with the
experimental results.

4. The model has been successfully used to predict not only the fatigue life of notched
specimens, but also the fatigue crack track in the specimen. The damage model was first
employed to examine fatigue failure of a standard tensile specimen under uniaxial cyclic
loading. An important phenomenon was observedfrom thetest and verified by theoretical
analysis that fatigue crack initiates at the shoulder of the specimen rather than in the
parallel section of the specimen, as in the case of ductile fracture under uniaxial tensile
loading. The model was further extended to predict the fatigue life in plates containing
an edge notch or an inclined notch under uniaxial cyclic loading and good agreement with
the experimental results has been achieved. An important observationfrom the numerical
IX

7
3

results is the ability of the model to identity the phenomenon of stress redistribution due
to material degradation during its fatigue process.

5. The numerical simulation of crack propagation in a plate with a centre crack was
performed. The crack length versus cyclic loading curves were obtained both theoretically
and experimentally, and satisfactory results are obtained not only for the constant
amplitude loadings but also for the two-step loading.

138

CHAPTER NINE

9. FUTURE WORK

1. Some modifications to the proposed damage model are required to consider the
nonlinear interaction between fatigue damage and plastic damage when large plastic
deformation appears in a material element during the fatigue loading process.

2. The application of the model is limited to the case of no obvious reverse plastic
deformation.

Further research is needed to consider the effect of reverse plastic

deformation on the damage accumulation in a material element and to establish cyclic


stress-strain relationship under complex loadings.

3.

An efficient numerical method for calculation of damage accumulation for large

numbers of cycles is required in order to apply the proposed damage model to the fatigue
failure prediction of engineering structures under complex loadings.

139

REFERENCES

Ackermann, F.Kubin, L. P., Lepinoux, J. and Mughrabi, H., 1984The dependence


of dislocation microstructure on plastic strain amplitude in cyclically strained copper single
crystals, Acta Metall., 32, pp.715.

Antolovich, S.D., Saxena, A. and Chanani, G.R., 1975A model for fatigue crack
propagation, Eng. Fracture Mech., 7pp.649.

Ashby, M. E. and Jones, D. R. H., 1980Engineering Materials --- an introduction to


their properties and applications, PERGAMON PRESS.

Bazant, Z. P., 1986, Mechanics of distributed cracking, Appl. Mech. Rev., 39, pp.675.

Brown, M. W. and Miller, K. J., 1973A theory for fatigue failure under multiaxial
stress-strain conditions, Proc. Instn. Mech. Engrs 745pp.745.

Brown, M. W. and Miller, K. J., 1979Biaxial cyclic deformation behaviour of steels,


Fatigue Engng Mater. Struct, pp.93.

Brown, M. W. and Miller, K. L , 1982, Two decades of progress in the assessment of


multiaxial low-cycle fatigue life, ASTM STP 770pp.482.
1A

Chaboche, J. L.1980,

Lifetime predictions and cumulative damage under high

temperature conditions, Int. Symp. on Low Cycle Fatigue and Life Prediction, ASTM
STP 770, 1982.

Chaboche, J. L., 1981, Continuous damage mechanicsa tool to describe phenomena


before crack initiation, Nucl. Eng. and Design, 64, pp.233.

Chaboche, J. L.1982,

The concept of effective stress applied to elasticity and

viscoplasticity in the presence of anisotropic damage,

Mechanical Behaviour of

Anisotropic Solids, pp.737.

Chaboche, J. L.1986, Continuum damage mechanics: present state and future trend,
Seminaire International sur I'Approcfae Locale de la Rupture, Moret-sur-Loing.

Chaboche, J. L.1987Continuum damage mechanics and its application to structural


lifetime predictions, La Rech. Aerospatiale, 4, pp.38.

Chaboche, J. L. and Lesne, P. M.1988A non-linear continuous fatigue damage model,


Fatigue Fract. Engng. Mater. Struct., 11pp.1.

Chanani, G. R. Antolovich, S. D. and Gerberich, W. W.1972,

Fatigue crack

propagation in metastable austenitic steels, Metall. Trans., 326pp.612.

Chand, S. and Garg, S. B. L., 1985

Crack propagation under constant amplitude


1
4
1

loading, Eng. Fract. Mech., 21, pp.1.

Chow, C. L. and Wang, J 1987An anisotropic theory for continuum damage


mechanics for ductile fracture,

Eng. Fract. Mech., 27pp.547.

Chow, CL. and Wang, J., 1988A finite element analysis of continuum damage
mechanics for ductile fracture, Int. J. of Fract., 38pp.83.

ChowC. L. and Woo, C. W.1985 A unified formulation of fatigue crack propagation


in aluminum alloys and PMMA, Eng. Fract. Mech., 21pp.589.

Chow, C. L.Woo, C. W. and Chung, K. T.1986, Fatigue crack propagation in mild


steel, Eng. Fract. Mech., 24, pp.233.

Chow, C. L. and Lu, T. J., 1989 On evolution laws of anisotropic damage, Eng. Fract.
Mech., 34, pp.679.

Chow, C. L. and Lu, T. J., 1990, An unified approach to fatigue crack propagation in
metals and polymers, J. Mater. Sci. Lett., 9pp.1427.

Chow, C. L. and Lu, T. L , 1991, Cyclic J-integral in relation to fatigue crack initiation
and propagation, Eng. Fract. Mech., 39pp.1.

Coffin, L. R1954, A study of the effects of cyclic thermal stresses on a ductile metal,
42
1

Transactions of ASME, 76pp.931.

Coffin, L. F., 1988Some perspectives on future directions in low cycle fatigue, Low
Cycle Fatigue, ASTM STP 942, pp.5.

Cordebois, JP. and Sidoroff, F., 1982, Anisotropic damage in elasticity and plasticity,
Numero special, J. De Mecanique Theorique Et Appliquee, pp.45.

Crossland, B., 1956,

Effect of large hydrostatic pressures on the torsional fatigue

strength of an alloy steel, Proceedings on International Conference on the Fatigue of


Metals, Institution of Mechanical Engineering, London, pp.138.

Dang Van, K., Griveau, B. and Message, 0 . , 1989, On a new multiaxial fatigue limit
criterion: theory and application, Biaxial and Multiaxial Fatigue, EGF 3Mechanical
Engineering Publications, London, pp.479.

Dang Van, K.Cailletand, G., Flavenot, F., Le Douaron, A. and Lieurade, H. P., 1989,
Criterion for high cycle fatigue under multiaxial loading, Biaxial and Multiaxial Fatigue,
EGF 3, Mechanical Engineering Publications, London, pp.459.

Domas, P. A. and Antolovich, S. D.1985, An integrated local energy density approach


to notch low cycle fatigue life prediction, Eng. Fract. Mech., 21pp.187.

Dowling, N . E. and Begley, J. A., 1976Fatigue crack growth during gross plasticity and

143

the J-integral, Mechanics of Crack Growth, ASTM STP 590pp.82.

Dragon, A. and Mroz, Z., 1979, A continuum model for plastic-brittle behaviour of rock
and concrete, Int. J. Eng. Sci., 17, pp.121.

Duggan, T. V.1977A theory for fatigue crack propagation, Eng. Fract. Mech., 9,
pp.735.

Elber, W., 1971,

The significance of crack closure, Damage Tolerance in Aircraft

Structures, ASTM STP 486pp.230.

Ellison, E. G. and Andrews, J. M. H.1973, Biaxial cyclic high strain fatigue of


aluminium alloy RR58, J. Strain Analysis, 8 pp.209.

Ellyin, F., 1989, Cyclic strain energy density as a criterion for multiaxial fatigue failure,
Biaxial and Multiaxial Fatigue, EGF 3pp.571.

Findley, W. N.1956, Fatigue of metals under combinations of stress, Trans. ASME, 56


pp.1337.

Fine, M. E 1980Fatigue resistance of metals, Met. Trans. A, 11A, pp.365.

Forman, R. G., Kearney, V. E. and Engle, R. M.1967, Numerical analysis of crack


propagation in cyclic loaded structures, J. of Basic Engineering, 89, pp.459.

144

Frost, N. E. and Dixon, J. R., 1967, A theory of fatigue crack growth, Int. J- Fracture
Mech., 3, pp. 301.

Frost, N. E., Pook, L. P. and Denton, K.1971, A fracture mechanics analysis of


fatigue crack growth data for various materials, Eng. Fract. Mech., 3pp.109.

Frost, N . E.Marsh, K. J. and Pook, L. P., 1974Metal Fatigue, Oxford University


Press.

GaruciY. S., 1981A new approach to the evaluation of fatigue under multiaxial
loadings, J. Eng. Mat. Tech., 103pp.118.

Gough, H. J., 1949, Engineering steels under combined cyclic and static stress,
Proceedings of the Institution of Mechanical Engineers, 60, pp.417.

Gough, H. JPollard, H. V. and Clenshaw, W. J., 1951Some experiments on the


resistance of metals to fatigue under combined stresses, Aero. Research Council, R and
M 2522 HMSO, London.

Gerold, V. and Meier, B.1987, Deformation mechanisms and crack initiation in fatigue
Fatigue'87, pp.1517.

K., 1990, Cyclic stress-strain response and low-cycle fatigue life in metallic
materials, JSME Int. J., 33, pp.13.
145

Hayhurst, D. R. and Leckie, F. A., 1973, The effect of creep constitutive and damage
relationships upon the rupture time of a solid circular torsion bar, J. Mech. and Phys.
Solids, 21pp.431.

Hayhurst, D. R.Leckie, F. A. and McDowell, D. L., Damage growth under


nonproportional loading, Multiaxial Fatigue, ASTM STP 853pp.688.

Hill, R.1950, The Mathematical Theory of Plasticity, The Clarendon Press.

HolcombD. J. and Costin, L. S.1986Detecting damage surfaces in brittle materials


using acoustic emissions, J. Appl. Mech., 53, pp.536.

Hertzberg, R. W.1989Deformation and Fracture Mechanics of Engineering Materials,


John Wiley & Sons, Inc..

Hopper, C. D. and Miller, K. J., 1977Fatigue crack propagation in biaxial stress fields,
J. of Strain Analysis, 12pp.23.

Hult, J., 1987Introduction and general overview,

Continuum Damage Mechanics:

Theory and Applications, CISM No.295., pp.1.

Hunsche, A. and Neumann, P., 1986Quantitative measurement of persistent slip band


profiles and crack initiation, Acta Metall., 34pp.207.

146

Jin, N. Y.1989,

Formation of dislocation structures during cyclic deformation of

F.C.C. crystals-I. Formation of PSBs in crystals oriented for single-slip, Acta Metall.,
37pp.2055.

Ju, J. W 1989On energy-based coupled elastoplastic damage theories: constitutive


modelling and computational aspects, Int. J. Solids Struct., 25pp.803.

Ju, J. W., 1991, On two-dimensional self-consistent micromechanical damage models for


brittle solids, Int. J. Solids Struct., 27pp.227.

Kachanov, L. M.1958,

On the time to failure during creep, Izv. An SSSROTN,

pp.26.

Kandil,

F. A., Brown, M. W. and Miller, K. L, 1982,

Biaxial low-cycle fatigue

fracture of 316 stainless steel at elevated temperatures, The Metals Society, London 280,
pp.203.

Kikukawa, M., Ohji, K. Kotani, S. and Yokoi, T 1972, A comparison of axial and
reversed-torsional strain cycling low cycle fatigue strength of several structural metals,
Bull. JSME, 15pp.889.

Kocanda, S.1978Fatigue Failure of Metals, Sijthoff & Noordhoff International


Publishers.

Krajcinovic, D.1984Continuum damage mechanics, Appl. Mech. Reviews, 37, pp.1.

Krajcinovic, D. and Fanella, D.1986A micromechanical damage model for concrete,


Eng. Fract. Mech., 25, pp.585.

Krajcinovic, D. and Sumarac, D., 1989A mesomechanical model for brittle deformation
processes part 1 and 2, Appl. Mech., 56, pp.51.

Krajcinovic, D.Basista, M. and Sumarac, D., 1991Micromechanically inspired


phenomenological damage model, J. Appl. Mech., 58pp.305.

Kung, C. Y. and Fine, M. E. 1979, Fatigue crack initiation and microcrack growth in
2024-T4 and 2124-T4 aluminum alloys, Metall. Trans. A, 10pp.603.

Ladeveze, P.1983On an anisotropic damage theory, Rapport Interne N 0 34, Laboratoire


de Mecanique et Technologie, ENSET/Universite PARIS 6/CNRS.

Lamba, H. S., 1975, The J-integral applied to cyclic loading, Eng. Fract. Mech., 3,
pp.693.

Leese, G. E.1988Engineering significance of recent multiaxial research, Low Cycle


Fatigue, ASTM STP 942 pp.861.

Lefebvre, D., Neale, K. W. and EllyinF., 1981A criterion for low-cycle fatigue

148

failure under biaxial states of stress, J. Eng. Mat. Tech., 103pp.1.

Lefebvre, D. F., 1989Hydrostatic pressure effect on life prediction in biaxial low-cycle


fatigue, Biaxial and Multiaxial Fatigue, EGF3, pp.511.

Leis, B. N . and Topper, T, 1977, Long-life notch strength reduction due to local biaxial
state of stress, J. Eng. Mat. Tech., pp.215.

Leis, B. N . , Hopper, A. T.Ahmad, J., Broek, D. and Kanninen, M. F.1986Critical


review of the fatigue growth of short cracks, Eng. Fract. Mech., 23pp.883.

Lemaitre, J. and Chaboche, J. L., 1978Aspects phenomenologiques de la rupture par


endamagement, J. Mec. Appl., 2pp.317.

Lemaitre, J 1984How to use damage mechanics, Nucl. Eng. and Design, 80pp.233.

Lemaitre, J., 1986a, Local approach of fracture, Eng. Fract. Mech., 25, pp.523.

Lemaitre, J., 1986b,

Plasticity and damage under random loading,

In Proc. Tenth

National Congress of Applied Mechanics, pp.125.

Lemaitre, J. and Dufailly, L , 1987Damage measurements, Eng. Fract. Mech., 28


pp.643.

149

LcmsitreJ.1987,

Fonxiulation End identification of damage kinetic constitutive

equations, Continuum Damage mechanics:

Theory and Applications, CISM No.295,

pp.37.

Lemaitre, J. and Chaboche, J, L.1990Mechanics of Solid Materials, Cambridge


University Press.

Libertiny, G. I., 1967Short life fatigue under combined stress, J. Strain Analysis, 2
pp.91.

Lin, M.-R.Fine, M. E. and Mura, T1986Fatigue crack initiation on slip bands:


theory and experiment, Acta Metall.,
34, pp.619.

Liu, H. W., 1991A review of fatigue crack growth analyses, Theor. Appl. Fract.
Mech., 16pp.91.

Matake, T.1977, An explanation of fatigue limit under combined stresses, Bulletin of


the JSME, 20pp.257.

Manson, S. S.1954Behaviour of materials under conditions of thermal stress, NACA


Report 1170Lewis Flight Propulsion

Laboratory, Cleveland.

Manson, S. S. and Halford, G. R.1981Practical implementation of the double linear

150

damage rule and damage curve approach for treating cumulative fatigue damage, Int. J.
Fract., 17, pp.169.

Manson, S. S. and Halford, G. R., 1981Practical implementation of the double linear


damage rule and damage curve approach for treating cumulative fatigue damage, Int. J.
Fract.17pp.169.

Manson, S. S. and Halford, G. R., 1986Re-examination of cumulative fatigue damage


analysis- an engineering perspective, Eng. Fract. Mech., 25pp.539.

Manson, S. S., 1988,

Future directions for low cycle fatigue, Low Cycle Fatigue,

ASTM STP 942, pp.15.

McDiarmid, D. L.1974A new analysis of fatigue under combined bending and


twisting, Aeronaut. J., 78pp.325.

Mcdiarmid, D. L., 1985, The effects of mean stress and stress concentration on fatigue
under combined bending and twisting, Fat. Fract. Eng. Mat. Struct., 8pp.1.

McDowell, D. L. and Socie, D. F.1985Transient and stable deformation behaviour


under cyclic nonproportional loading, Multiaxial Fatigue, ASTM STP 853, pp.64.

McDowell, D. L. and Berard, J. Y., 1992, A AJ-based approach to biaxial fatigue,


Fatigue Fract. Engng Mater. Struct., 15, pp.719.

151

McevilyA. J., 1989On the growth of small/short fatigue cracks, JSME Int. J. Series
I32pp.181.

Miller, K. J., 1987The behaviour of short fatigue cracks and their initiationpart I. a
review of two recent books, part E. a general summary, Fat. Fract. Eng. Mat. Struct.,
10pp.75.

Miner, M. A., 1945Cumulative damage in fatigue, J. Appl. Mech., 67pp.159.

Morrow, D. L. and Kurath, P., 1989, Proportional biaxial-tension low cycle fatigue of
Inconel 718, Biaxial and Multiaxial Fatigue, EGF3, pp.551.

Moyer, E. T. and Sih, G. C.1984Fatigue analysis of an edge crack specimen:


hysteresis strain energy density, Eng. Fract. Mech., 19, pp.643.

Mqnson, S. S., 1965, A complex subject - some simple approximations, Experimental


Mechanics, 5, pp.193.

Murakami, S. and Ohno, N., 1981A continuum theory of creep and creep damage,
Proceeding of 3rd IUTAM Symp. on Creep in Structures, Springer, Berlin, pp.422.

Murakami, S
continuum damage mechanics, Continuum Damage Mechanics: Theory and Applications,
CISM No.295, pp.91-133.

152

Murakami, S.1987b, Progress of continuum damage mechanics, JSME Int. J., 30


pp.701-710.

Murakami, S., 1990, A continuum mechanics theory of anisotropic damage Yielding,


Damage, and Failure of Anisotropic Solids,

EGF5, pp.465-482.

Muralidharan, U. and Manson, S. S.1988, A modified universal slopes equation for


estimation of fatigue characteristic of metals, J. Eng. Mat. Tech., 110, pp.55.

Neumann, P. and Tonnessen, A., 1987Cyclic deformation and crack initiation, Fatigue
'87, pp.3-22.

Nisitani, H.1981Unifying treatment of fatigue crack growth in small, large and nonpropagating cracks, ASME AMD, 47, pp.151.

Nishitani, H. and Kawagoishi, N.1992Fatigue crack growth laws in small and large
cracks and their physical background, JSME Int. J. Series I35, pp.1-11.

Neuber, H., 1961Theory of stress concentration for shear-strained prismatical bodies


with arbitrary nonlinear stress-strain law, J. Appl. Mech., 28pp.544.

Nowack, H. and Marissen, R. 1987Fatigue crack propagation of short and long cracks:
phys. basis, prediction methods and engineering significance, Fatigue 87, pp.207.

153

Nowack, H., Ott, W., Foth, J., Peeken, H. and Seeger, T.1988, Some contributions to
the further development of low cycle fatigue life analysis concepts for notched components
under variable amplitude loading, Low Cycle Fatigue, ASME STP 942, pp.987.

Paris, P. C. and Erdogan, F.1963A critical analysis of crack propagation law, J. of


Basic Engineering, 85 pp.528.

Paris, P. C.1981, Twenty years of reflection on questions involving fatigue crack


growth, part I: historical observations and perspectives, Fatigue Thresholds, pp.3.

Paris, P. C. and Hermann, L., 1981Twenty years of reflection on questions involving


fatigue crack growth, part II: some observations of crack closure, Fatigue Thresholds,
pp.11.

Pook, L. P. and Frost, N. E.1973, A fatigue crack growth theory, Int. J. Fracture, 9
pp.53.

Pascoe, K. J. and DE Villiers, J.WR, 1967, Low-cycle fatigue of steels under biaxial
straining, J. Strain Analysis, 2pp.117.

RajuK. N . , 1972, An energy balance criterion for crack growth under fatigue loading
from considerations of energy of plastic deformation, Int. J. Fracture, 8, pp.1.

Rasmussen, K. V. and Pedersen, O. B., 1980Fatigue of copper polycrystal at low


Ii

4
5

plastic strain amplitudes, Acta Metall., 28, pp.1467.

Rotvel, F., 1970, Biaxial fatigue tests with zero mean stress using tubular specimens, Int.
J. Mech. Sci., 12, pp.597.

Schwalbe, K. H.1974Comparison of several fatigue crack propagation laws with


experimental results, Engng Fracture Mech., 6pp.325.

Shewchuk, J Zamrik, S. Y. and Marin, J 1968Low cycle fatigue of 7075-T651


aluminium alloy m biaxial bending, Expl. Mech., 8, pp.504.

Sih, G. C. and Moyer, E. T., 1983Path dependent nature of fatigue crack growth,
Eng. Fract. Mech., 17, pp.269.

SihG. C. and Chao, C. K.1989Fatigue failure initiation analysis of wing/fuselage


bolt assembly, Theor. Appl. Fract. Mech., 11pp. 109.

sim

o , j . c . and Ju, J. W.1987, Strain- and stress-based continuum damage models -

I. formulation; EL computational aspects, Int. J. Solids Struct., 23pp.82L

SinesG., 1955Failure of material under combined repeated stresses with superimposed


static stress, national Advisory

Committee for Aeronautics Technology Note 3495,

pp.69.

155

SkeltonR. P., 1988, Application of small specimen crack growth data to engineering
components at high temperature: a review Low Cycle Fatigue, ASTM STP 942, pp.209.

Socie, D . R , Wail, L. E. and Dittmer, D. F., 1985Biaxial fatigue of Inconel 718


including mean stress effects, Multiaxial Fatigue, ASTM STP 853pp.463.

Socie, D . , 1987Multiaxial fatigue damage models, J. Eng. Mat. Tech., 109, pp.293.

Socie,

D . F.Kurath, P . and Koch, J., 1989, A multiaxial fatigue damage parameter,

Biaxial and Multiaxial Fatigue, EGF 3, pp.535.

Tabata, T.Fujita, H., Hiraoka, M. and Onishi, K. 1983Dislocation behaviour and the
formation of persistent slip bands in fatigued copper single crystal observed by highvoltage electron microscopy, Philosophical Magazine A, 47, pp.841.

Tabeshfar, K. and Williams, T. R. G., 1974Discontinuities in the S/N curves of


rimming and stabilized mild steels, Metal Science, 8pp.291.

Tanaka, T.1987, Effect of grain orientation on the origination and the initial growth of
fatigue cracks in pure iron, Fatigue 87, pp.23.

Tanaka, K 1983The cyclic J-integral as a criterion for fatigue crack growth, Int. J.
Fracture, 22, pp.91.

156

Tanaka, K.,

1989 Mechanics and micrornechanics of fatigue crack propagation,

Fracture Mechanics: Perspectives and Directions, ASTM STP 1020 pp.151.

Thompson, N.Wadsworth, N. J. and Louat, N.1956Philosophical Magazine A, 1,


pp.113.

Tipton, S. M. and Nelson, D. V., 1985, Fatigue life predictions for a notched shaft in
combined bending and torsion, Multiaxial Fatigue, ASTM STP 853pp.514.

Wang, J., 1987, Development of an anisotropic damage mechanics model in ductile


fracture, Ph.D. Thesis, University of Hong Kong.

Winter, A. T., 1974, A model for the fatigue of copper at low plastic strain amplitudes,
Philosophical Magazine A, 30 pp.719.

Winter, A. T., Pedersen, 0 . B. and Rasmussen, K. V., 1981, Dislocation microstructure


in fatigued copper polycrystal, Acta Metall., 29, pp.735.

Woo, C. W. and Chow, C. L.1984, Fatigue crack propagation in aluminum and


PMMA, Int. J. of Fract, 26pp.R37.

Wuthrich, C., The extension of the J-integral concept to fatigue cracks, Int. J. Fracture,
20, pp.R35.

157

Table 2.1 Quality chart of methods of damage measurement


(Lemaitre and Dufailly, 1987)

Damage

Brittle

Micrography

Density

Ductile

* ,

* *

Elastic Modulus

Ultrasonic waves

Cycle stress

Creep

Low cycle

High cycle

fatigue

fatigue

* *

amplitude
Tertiary creep

Micro-hardness
Electrical

* *

* *

resistance

where
*** means "very good"
-means "do not try"

** means "good"

* means "try to see"

Table 5.1 Measured results for Al alloy 2024-T3 Alclad


(Wang, 1987)

C5c5c;c5c5c5o.c5o.o.cjo.o.cicjo.o.o.o,o.c5c5c5o.o.cicjo.o.o.o.c5o.c5o.o.o.o.o.o.o.

4
0 0 04
3
1 24 7 4 6 2 42
32
84
0684976504090610195)7)050
32
.2

30
32
3834 7
ii2 1 2 2 3 22^< 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3<3 3 3 3 3 3 3 3 3 2 3 3 2 2
33333333

.^
8r
.^
62
..
4l
ol
^L
.o
2.
48
0.
^8
.o
7c
a5
^0
c6
J0
5^
27
4.
39
3.
18
r.
^
46
24
3
c
.0
60
.l
6r
.J
43
.4
46
.0
53
.0
44
45
.2
49
c.
^3
3.
.l
3.
.o
r.
^9
c.
^o
r*
4r
lJ
..
Lo
L.
L5
o..
<).o.^.98

38
31
48
66
15
14
96
58
65
67
88
32
41
14
69
42
27
55
0S
02
60
44
83
7&
41
59
47
46
26
93
81
30
14
71
11
05
0G
W
54^
7
70

7777777666666666666^>V66^66^66666666666666655

E, GPa

0.0063
0.0137
0.0147
0.0172
0.0247
0.0292
0.0348
0.0386
0.0545
0.0550
0.0563
0.0584
0.0597
0.0638
0.0758
0.0872
0.0927
0.0959
0.1004
0.1089
0.1159

0.1222

0.1301
0.1398
0.1415
0.1536
0.1561
0.1624
0.1823
0.1917
0.1989

0.2070
0.2151
0.2231
0.2390
0.2469
0.2546
0.2624
0.2927
0.3221

585500323255367498990932454564757889PNJ2;5)2,)4'
3
3334444444444444444545555555555555556666

2.0.0.1.5.5.1.8.0.9.8.6.0.6.5.9.8.8.3.9.8.3.3.1.2.0.5.2.5.7.1.3.6.2.5.5.9.)4.)9.8.^6.

8 7 7 5 5 6 3 8 2 7 6 5 5 5 8 74 2 2 0 3 5 6 9 7 ) 9 9 8 7 7 4 0 2 ) 6 7 ) 6 ) 9 t 8 ^ 2 i l 5 2
414376472220832991278548703 1.3700^.3.7.84.42.8

True stress a.
True strain e

0.2021

Table 5.2 The data used to calculate the damage parameters

True strain e

True stress a, MPa

E, GPa

0.00444

330

74300

0.3300

0.02

390

71500

0.3287

0.04

430

69500

0.3271

0.06

455

67800

0.3255

0.08

475

66500

0.3239

0.10

492

65400

0.3224

0.12

515

64500

0.3208

0.14

532

63700

0.3193

0.16

547

62800

0.3177

0.18

562

62100

0.3162

0.20

575

61500

0.3147

0.22

587

61000

0.3131

0.24

600

60400

0.3115

0.26

612

60000

0.3100

0.28

625

59500

0.3085

0.30

635

59200

0.3069

0.32

647

58800

0.3054

Table 5.3 The calculated damage components d and fi

True strain e

True stress a, MPa

lo-1

0.00444

330

0.000

0.000

0.02

390

0.020

0.014

0.04

430

0.034

0.032

0.06

455

0.048

0.049

0.08

475

0.058

0.067

0.10

492

0.067

0.083

0.12

515

0.074

0.100

0.14

532

0.081

0.116

0.16

547

0.088

0.133

0.18

562

0.094

0.149

0.20

575

0.100

0.164

0.22

587

0.104

0.181

0.24

600

0.109

0.197

0.26

612

0.113

0.213

0.28

625

0.117

0.228

0.30

635

0.121

0.244

0.32

647

0.124

0.260

Table 5.4 The calculated damage evolution coefficient 7

True strain e

True stress cr, MPa

0.012

366

0.69

0.03

410

0.70

0.05

445

0.70

0.07

465

0.72

0.09

483

0.72

0.11

504

0.74

0.13

522

0.74

0.15

539

0.73

0.17

555

0.74

0.19

570

0.75

0.21

582

0.78

0.23

594

0.76

0.25

606

0.80

0.27

618

0.77

0.29

630

0.85

0.31

641

0.79

Table 5.5 The strain hardening threshold

p
0.005
0.015
0.024
0.033
0.043
0.052
0.062
0.071
0.080
0.090
0.099
0.108
0.117
0.126
0.135
0.145

RMPa

34.8
68.5
92.6
116
134
147
160
172
183
194
208
222
230
243
253
263

P
0.154
0.163
0.172
0.181
0.190
0.199
0.208
0.217
0.226
0.235
0.243
0.252
0.261
0.270
0.279
0.287

R, MPa
274
283
294
301
310
317
327
335
344
352
361
372
380
387
396
405

Table 5.6 The plastic damage threshold

03142533.43.51.6373..84.9520..32.41.53.

052.050^2.2.0^2.2.2.3.3.3.111

rxj
161717181819S2020.21.21.22.23.23.24.24

179:368.521673:803.901.10202841:53.61.73.82.93

0.10.0.0.0.1.111111111

0.027
0.043
0.061
0.072
0.084
0.094
0.104
0.110
0.120
0.124
0.130
0.137
0.140
0.148
0.154
0.160

o.o.o.CJCJo.cio.o.o.CJo.cio.o.o.

MPe

66

Table 5.7 Fatigue endurance limit for different mean stresses

Mean stress
OmeanMPa

Endurance limit
o-j, MPa

Maximum stress
o-max, MPa

100
79
68
61.5
59
47.5

0
32
68
85
135
347

Minimum stress
ffmin, MPa

100
111
136
146.5
194
394.5

-100
-47
0
23.5
76
299.5

Table 5.8 The relationship between YfU and


Yfdl
(MPa)2
1.59 XlO"2
1.06 x 10-2
0.90 x 10-2
0.44X10-2

V
I p
dnin
(MPa)2

(Yfd)1)1/4
(MPa)I/2

0
1.41 x 10-5
1.55 x 10-3
4.34 X10-1

0.355
0.321
0.308
0.258

(MPa) '
0
0,061
0.198
0.811

Table 5.9 The fatigue test results for smooth specimen (1)
Minimum stress
o-min, MPa

Maximum stress
"maxj MPa

12

Fatigue life
Nf, 104 cycles

50

300

13.06
13.00
12.00
11.57
11.26

10

260

17.51
14.45
14.43

Table 5.10 The fatigue test results for smooth specimen (2)
(stress range

Max. stress
MPa

Mean stress
ffmMPa

Fatigue life
Nf, 104 cycle

Average life
104 cycle

430

330

CO
-H" O O
*H t-H T<

10.95

200

400

300

22.23
19.60
19.43
13.52

18.70

145

345

245

20.15
19.95
17.70

19.27

100

300

200

28.17
26.82
23.31
22.20

25.13

65

265

165

32.16
28.97
27.04

29.39

35

235

135

41.92
32.63
31.21
27.39

33.29

205

105

40.11

-40

160

60

63.91
58.38
45.05
39.89
37.64

48.97

-70

130

30

82.72
82.25
77.86

80.94

-85

115

15

227.26
101.90
82.87

137.34

-94

106

>1500
608.73
111.74

230

LO U> 4^OO VO JO
Ln

O O

Min. stress
(TMPa

On 00

Table 5.11 The fatigue test results for smooth specimen (3)
(stress range Acr = 250 MPa, load frequency 100 Hz)

Min. stress

Max. stress

Mean stress

aMPa

cr M P a

^mean> MPa

203

453

328

6.01
5.24
5.18

5.48

190

440

315

6.33

6.33

120

370

245

9.34
8.98
8.18

8.83

50

300

175

13.06
13.00
12.00
11.57
11.26

12.18

10

260

135

17.51
14.45
14.43

15.46

-11

239

114

18.32
16.30
15.70

16.77

-65

185

60

21.11
20.27
19.45
19.18

20.00

-111

139

14

32.73
31.62

32.18

-125

125

40.54
39.68
36.25
34.59
34.03

37.02

Fatigue life
Nf, 104 cycle

Average life
104 cycle

Table 5.12 The fatigue test results for smooth specimen (4)
(stress range Ao- = 340 MPa, load frequency 100 Hz)

Fatigue life
Nf, 104 cycle

Average life
104 cycle

245

3.34
3.07
3.01

3.14

370

200

4.80
4.60
4.44

4.61

-5

335

165

6.83
6.34
6.14

6.44

-35

305

135

8.39
6.79
5.30

6.83

-70

270

100

9.28
8.68
7.97

8.64

-110

230

60

(N
CO (N T-H
o
t-H oH o
t-H

185

15

17.41
16.09
13.39
11.89

Min. stress
OmkMPa

Max. stress

75

415

30

0MPa

Mean stress
O-mean, MPa

10.20

-155

14.70

Table 5.13 The fatigue test results for smooth specimen (5)
(mean stress crmean = 60 MPa, load frequency 100 Hz)

Max. stress
^maxMPa

Stress range
Aa, MPa

Fatigue life
Nf, 104 cycle

Average life
104 cycle

-110

230

340

10.37
10.12
10.12

10.20

-90

210

300

15.08
14.29
13.29

14.22

-65

185

250

21.11
20.27
19.45
19.18

20.00

-52

172

224

31.15
28.62
25.67
21.14

26.65

-40

160

200

63.91
58.38
45.05
39.89
37.64

48.97

Min. stress
MPa

\l/

/fv \J/

e
c
l
o

Im00

00

^en

o
t
t

Stress range
A<t, MPa

mfre

-23

s f
r d
fo
f
0a
s 1

305

s
u
p
rcM

-35

St}5
e3 11
t
e =
u
s
II

Max. stress

f
a
t
i
j

e
h s

I
. s
4

k m

Min. stress
cr, MPa

Fatigue life
Nf, 104 cycle

Average life
Nf, 1C? cycle

340

8.39
6.79
5.30

6.83

293

316

10.80
8.27
6.70

8.59

-7

277

284

12.80
12.07
12.00

12.29

10

260

250

17.51
14.45
14.43

15.46

25

245

220

27.65
22.86
22.46

24.32

35

235

200

41.92
32.63
31.21
27.39

33.29

40

230

190

43.01
35.49
34.62
25.64

34.69

48

222

174

60.00
58.00
46.86

54.95

56

214

158

80.65
77.86
77.36

78.62

6 4

206

142

129.83
123.72
105.50

119.68

69

201

132

168.12
146.57
129.85

148.18

7 4

196

122

973.35

"rnaxs M P a

258.85
214.89

482.36

Table 5.15 The fatigue test results for smooth specimen (7)
(mean stress
Min. stress

Max. stress

aMPa

crMPa

Stress range
Aa, MPa

75

415

340

3.34
3.07
3.01

3.14

120

370

250

9.34
8.98
8.18

8.83

25

245

220

27.65
22.86
22.46

24.32

Fatigue life
Nf, 104 cycle

Average life
104 cycle

Table 5.16 The fatigue test results for smooth specimen (8)
(stress range Acr = 250 MPa, load frequency 100 Hz, L-direction)

Fatigue life
Nf, 104 cycle

Average life
104 cycle

Min. stress

Max. stress

Mean stress

o^MPa

o,MPa

^mcan> MPS

250

125

16.21
14.25
12.98

14.48

50

300

175

11.48
10.70

11.09

Table 6.1 Comparison between predicted and experimental


results for fatigue life of standard tensile specimen

Applied
Stress
MPa

Experimental Result, 103

Calculated result,

103 cycles

two-dimensional

one-dimensional

testing

average

76-290

162.1
132.0
129.5

141.2

127.9

200.6

101-290

250.9
172.8
159.8
142.4

181.5

176.5

282.2

141-290

397.5
369.6
369.4
360.4

374.2

354.0

616.4

164-290

695.3
656.9
610.6
439.6

600.6

626.2

1386.6

172-290

853.1
508.9

681.0

844.1

2311.2

Table 6.2 Comparison between predicted and experimental


results for fatigue life corresponding to a 2 mm crack
in the specimen with an inclined notch

Experimental result

103 cycles

Calculated
result
103 cycles

Applied
stress
MPa

50+25

365

380

395

435

435

402

270

50+40

94

100

108

108

110

102

97

testing value

average
value

Table 6.3 Comparison between predicted and experimental


results for fatigue life corresponding to a 3 mm crack
in specimen with an inclined notch

Experimental result

103 cycles

Applied
stress
MPa

5025

385

392

408

448

455

418

316

5040

98

104

104

113

115

107

108

testing value

average
value

Calculated
result
103 cycles

Table 6.4 Comparison between predicted and experimental


results for fatigue life corresponding to a 3 mm crack
in specimen with an edge notch

Experimental result

103 cycles

Calculated
result
103 cycles

Applied
stress
MPa

5020

402

493

545

610

611

532

450

5030

127

143

156

171

176

155

130

50 37.85

81

81

98

120

129

102

78

10030

76

127

130

137

140

122

71

testing value

average
value

Table 7.1 Test results for fatigue crack propagation


under the cyclic loading 4524 MPa

test 1

test 2

test 3

test 4

test 5

N
103

a, mm

N
103

a, mm

N
103

a, mm

N
103

a, mm

N
103

a, mm

10
15
20
25
30
35
40
45
50

1.88
2.98
4.15
5.45
6.90
8.33
10.13
12.68
16.93

7.5
12.5
17.5
22.5
27.5
32.5
37.5
42.5
47.5

1.48
2.63
3.63
4.88
5.93
7.40
8.95
10.95
13.80

10
20
30
40
45
50

1.70
3.93
6.48
9.73
11.58
14.40

10
20
30
40
45
50

1.88
3.95
6.33
8.98
10.75
13.43

10
20
30
40
45

1.75
3.75
5.65
7.73
8.75

* N is the number of loading cycles,


a is half the increment of the crack length.

Table 7.2 Test results for fatigue crack propagation


under the cyclic loading 45 11.5 MPa

test 5

test 4

test 3

test 2

test 1
N
103

a, mm

N
103

a, mm

N
103

a, mm

N
103

a, mm

31
60
90
120
150
180
200
220
240
260
280
300
310

0.75
1.60
2.50
3.50
4.85
6.08
7.20
8.35
9.58
11.00
12.78
14.75
15.90

45
75
105
135
165
195
230
250
270
290
310
330

0.80
1.50
2.23
3.05
4.15
5.48
7.25
8.35
9.78
11.20
13.05
14.95

40
70
100
130
160
190
220
240

1.30
2.40
3.65
5.10
6.98
9.23
11.75
14.33

30
60
90
120
150
180
210
240
270

0.65
1.48
2.50
3.65
5.03
6.68
8.78
10.55
13.75

* N is the number of loading cycles,


a is half the increment of the crack length.

N
103
30
60
90
120
150
180
210
240
270

a, mm
0.63
1.48
2.60
3.90
5.30
7.10
9.15
11.83
15.23

Table 7.3 Test results for fatigue crack propagation


under the cyclic loading 45 16.5 MPa

test 2

test 1

test 3

N
103

a, mm

N
103

a, mm

N
103

8
13
17
21
25
29
33
37
41
45
49
53
57
61
65
69
73
77
81
85
89
93
97
101
105

0.63
1.10
1.48
1.87
2.33
2.65
3.10
3.43
3.73
4.25
4.61
5.18
5.70
6.15
6.68
7.20
7.85
8.53
9.23
9.95
10.73
11.38
12.45
13.60
14.73

10
15
20
25
30
35
40
45
50
55
60
65
70
75
80
85
90
95
100
105
110

0.60
1.18
1.55
1.98
2.45
2.98
3.45
4.13
4.60
5.18
5.65
6.50
7.00
7.75
8.55
9.38
10.50
11.35
12.43
13.48
15.00

15
30
45
60
75
101
115

* N is the number of loading cycles,


a is half the increment of the crack length.

test 4

test 5

a, mm

N
103

a, mm

N
103

a, mm

0.95
2.43
3.88
5.75
7.73
10.58
14.05

10
20
30
40
50
60

0.58
1.45
2.30
3.28
4.18
5.38

15
25
35
45
60

1.05
1.88
2.78
3.75
5.40

Table 7.4 Test results for fatigue crack propagation


under the cyclic loading 9011.5 MPa

test 1
3

N, 10

test 2
a, mm

20
40
60
80
100
120
140
160
180
200

0.40
1.13
1.85
2.78
3.95
4.88
6.28
7.88
9.73
12.28

N, 10

20
40
60
80
100
120
140
160
180

test 3
a, mm

N, 103

0.55
1.38
2.23
3.25
4.25
5.45
7.13
8.85
11.28

a, mm

20
40
60
80
100
120
140
160
180
200

0.48
0.95
1.53
2.25
3.18
4.13
5.43
6.78
8.50
10.73

* N is the number of loading cycles,


a is half the increment of the crack length.

Table 7.5 Test results for fatigue crack propagation


under the two-step cyclic loading
(45 16.5 MPa for 0-60000 cycles, 45 11.5 MPa for 60000-

test 2

test 1
N, 103

a, mm

10
20
30
40
50
60
70
80
90
100

0.58
1.45
2.30
3.28
4.18
5.38
5.70
6.23
6.60
7.18

cycles)

N103

110
120
130
140
150
160
170
180
190
200

a, mm
7.73
8.30
8.93
9.63
10.30
11.10
11.90
12.80
13.68
14.85

* N is the number of loading cycles,


a is half the increment of the crack length.

N, 103
15
25
35
45
60
75
85
95
105
115

a, mm

N, 103

a, mm

1.05
1.88
2,78
3.75
5.40
6.10
6.65
7.20
7.70
8.33

125
135
145
155
165
175
185

8.98
9.70
10.43
11.23
12.13
13.18
14.25

m a x i m u m stress
streks a m p l i t u d e

m e a n stress
time
m i n i m u m stress

N o m e n c l a t u r e t o describe t e s t p a r a m e t e r s
i n v o l v e d i n c y c l i c stress t e s t i n g

f o r y i e l d point m a t e r i a l s

for non-yield point materials

LCF

HCF

n u m b e r of cycles to f a i l u r e

Fig. 2.2

The Wohler d i a g r a m

i n l o g - l o g coordinates

(6)

F i g . 2.3

H y s t e r e s i s loops f o r c y c l i c l o a d i n g

i n (a) i d e a l l y e l a s t i c m a t e r i a l a n d
(b) m a t e r i a l u n d e r g o i n g e l a s t i c a n d
p l a s t i c d e f o r m a t i o n (Hertzberg, 1989)

mi n!:

IliFH

iii!
i ii

ii
ii:!
i;i
:::: ::::
1.:::

i M P l

ii

i
i
:
l
r
-

L
*

::: :rg- :.ru.

flBBBflBBBBBH|CSSSSSSSnSSSSSS3SSBSSMSS jSlSSffEwB
fffffpniiiiwtwiigggiggM^BSBgggi gaagffl

H
:

m}?;n!;

strain
Fig. 4.1

S t r e s s - s t r a i n response u n d e r c y c l i c loading

Unitm m

200
25

15

2 5

5 0

Fig. 5 . 1

Specimen

used i n uniaxial tensile tests

cds

0.00

b.io

0.20

0.30

t r u e strain
F i g

. 5.2

The relationship between t r u e stress


and true strain

80000n

70000

60000

50000

)0

40000

30000
0.00

5.3

o.io

0.20

0.30

true strain

T h e m e a s u r e d r e s u l t s of elastic m o d u l u s

0.35 q .

0.30

0.25

0.15 -

0.10

4-r

0.00

.4

oio

O.io ."o^O

true strain

T h e measured r e s u l t s of P o i s s o n ' s r a t i o

0.08

0.04

0.00

200

400

500

800

true stress, MPa


5.5

E v o l u t i o n of d a m a g e c o m p o n e n t d
w i t h t r u e stress

-JU9uodxJUOa

0.025

p-t-juauoduloo
9

0.020

0.015

0.010

0.005

0.0000 , 0 0

0.10

0.20

0-30

0.40

true strain
F i g . 5.6

E v o l u t i o n of d a m a g e c o m p o n e n t
with true strain

0.025

3.

c
C
o!
O
h
o
o

0.020

CD

dc
0
T
3

0.015

0.010

0.005

0.000

200

400

600

800

true stressMPa
5.7

E v o l u t i o n of d a m a g e c o m p o n e n t M
with t r u e stress

0.025

ri

0.020

0.015

0.010

0.005 -

0.000

0.00

0.10

0.20

0.30

0.40

true strain
5.8

E v o l u t i o n of d a m a g e c o m p o n e n t fx
with t r u e strain

0.6 ~

0.4

J
0.0 1

iI 1

0.00

0:10

1 1

0.20

0.30

11

0.40

true strain
5.9

The relationship between 7 a n d s

500

400

300

200

100-

0.00

0.05

0.10

0.15

0.20

0.25

0.30

overall p l a s t i c s t r a i n

5.10

The r e l a t i o n s h i p b e t w e e n R a n d

4.00

3.00

2.00

1.00

0.00

0.00

0.05

0.10

0.15

0,20

0.25

overall plastic damage ivp

F i g . 5.11

The r e l a t i o n s h i p between B a n d

Wp

180

Fig. 5.12

Specimen used i n u n i a x i a l fatigue tests

^Im-flFI
90XI1gJnTZ!XI9

" . " I

-200

-100

100

m i n i m u m stress,

Fig. 5 . 1 3

I I

Juu

MPa

T h e e f f e c t of m i n i m u m s t r e s s

o n t h e fatigue endurance

limit

0.40

0.30

0.20

0.10

0.00
0.00

0.20

0.40

0.60

0.80

(MPa)

F i g . 5.14

F a t i g u e damage s u r f a c e

00

600

B p h S

500

400.

discontinuity

m n s p c T e m

300.

200 t

discontinuity

100

10 5

10

o6

10

cvoles t o f a i l u r e
_ c a l c u l a t e d r e s u l t
experimental r e s u l t

Fig-

5.15

M a x i m u m stress vs f a t i g u e l i f e

a t c o n s t a n t stress r a n g e 200 M P a

400 t

discontinuity
100^

cycles t o f a i l u r e
calculated result
" e x p e r i m e n t a l result

F i g . 5.16
a

M a x i m u m stress vs f a t i g u e l i f e

t c o n s t a n t stress range 2 5 0 MPa

^
mnm-f-lXTSm

400

discontinuity

300

200

100

10

io

i i i ' ' i

10

c y c l e s to f a i l u r e

Fig. 5 . 1 7

++

calculated result
experimental result

Maximum stress vs fatigue life

a t c o n s t a n t s t r e s s r a n g e 340 MPa

400-

300

200

100.

rrr
10

10

10

c y c l e s to f a i l u r e
c a l c u l a t e d r e s u l t f o r 200 M P a
experimental result
^ f e i m l i t a l ' r ^ s u i r f o r 340 M P a

F i g . 5.18

M a x i m u m stress vs f a t i g u e l i f e

a t c o n s t a n t s t r e s s ranges 200 a n d 340 M P a

cycles to f a i l u r e
calculated result
e x p e r i m e n t a l result

5.19

Stress range vs f a t i g u e

t c o n s t a n t m e a n s t r e s s 60 M P a

600-]

500

ojdM

400

300

200

100

cycles to f a i l u r e
calculated result
e x p e r i m e n t a l result

F i g . 5.20

Stress r a n g e vs f a t i g u e l i f e

a t c o n s t a n t m e a n stress 135 M P a

cycles to f a i l u r e

++++

5.21

calculated result
experimental result
S t r e s s range vs f a t i g u e l i f e

c o n s t a n t m e a n stress 245 M P a

500 n

rtdM
"cutLOucdJ S S Q J - P S

10

'0 5

cycles to f a i l u r e
c a l c u l a t e d r e s u l t f o r 60 M P a

* # *

experimental result

catculated result f o r 245 M P a


e x p e r i m e n t a l r e s u l t f o r 245 M P a

F i g . 5.22

Stress range vs f a t i g u e l i f e

a t c o n s t a n t m e a n stresses 60 a n d 2 4 5 M P a

7.5

Unit: m m

30

see Fig. 6.2 f o r d e t a i l s

70

25

12.5
F i n i t e e l e m e n t n e t w o r k f o r ttxe t e n s i l e s p e c i m e n

Unit:

8.23

Fig. 6.2

Fine network at the shoulder

F i g . 6.3

Contours of e q u i v a l e n t s t r e s s

a t t h e applied stress 300 M P a

Unit: MPa

260

24
F i g . 6.4

Contours of e q u i v a l e n t s t r e s s

a t t h e applied stress 450 M P a

0.12
0.18
0.22

crack initiation

Fig. 6.5

Contours of overall damage at 346,000

cycles under the loading 141290 MPa

fatigue crack

24

Fig. 6,6

Contours of overall d a m a g e a t 354,000

cycles u n d e r t h e loading 141-290 MPa

see Fig. 6.9 for details

F i eo . 6.9

Finite element network

for the specimen with an inclined n o t c h

Unit: m m
24

24

see f i g u r e (b) f o r details


(a)

(b)

Fig. 6.10

Network at notch

Unit: MPa

33'

24

120

30

iff

B.H

C o n t o u r s of e q u i v a l e n t s t r e s s

. u n d e r t h e applied s t r e s s 9 0 MPa

Unit: MPa

crack initiation
24

Fig. 6.12 Contours of e.uivalents t ^


s t r e s s 90 MPa at 50,000 cycles with the loading

12 m m

crack initiation

Fig. 6.13

Contours of overall damage under the applied

stress 90 MPa at 50,000 cycles with the loading 50 40 MPa

Unit: MPa

crack

12 m m

contours of equivalent stress under the applied


J e s s go MPa at 95,000 cycles with the loading 50 40 MPa
fi 1 4

Unit: m m

uono?.=p
pdol

F i g . 6.17

S p e c i m e n w i t h a n edge n o t c h

Unitm m
100

80

\
s e e Fig. 6.17 f o r d e t a i l s

Fig. 6.18

Finite element network

f o r t h e s p e c i m e n with a n edge n o t c h

(a)

6.19

(b)

Network at n o t c h region

Unit: m m
mm

150

180

270
300
330
c r a c k initiation.

Fig. 6.20

Contours of equivalent stress

a t 57,000 cycles under the loading 50 30 MPa


mm

0.07
0.10
0.16,
crack initiatioa

Fig. 6.21 Contours of overall damage


a t 57,000 cycles under the loading 50 30 MPa

Unit: MPa

180
mm

\300
fatigue c r a c k

Fig. 6.22

330

Contours of equivalent stress

a t 120,000 cycles under the loading 50 30 MPa

8 mm

5 mm
O.Oi

0.07

0.04

0,13
0.16
fatigue c r a c k

Fig. 6.23 Contours of overall damage


a t 120,000 cycles under the loading 50 30 MPa

mm

o.ia

0.1

o.ia
crack initiation

Fig. 6.27

Contours of overall damage

at 38,000 cycles under the loading 100 30 MPa

mm

fatigue crack

Fig. 6.28

0.16/
0.19/
Contours of overall d a m a g e

a t 70,000 cycles u n d e r t h e loading 100 30 MPa

Unit: m m

initial crack

Pre-cracked specimen

F i g . 7.2

Two k i n d s of m e s h n e a r t h e c r a c k t i p

30000
1000 cycles

25000
500 cycles

s9o>.a

20000

250 cycles

jo

15000

10000

5000

2.0

0.0

crack length increment, m m

Fig. 7.3

The effect of the number of cycles for each step (I)

J9quinu

20000

step
500 cycles
250 cycles
125 cycles

15000
s9a>^o

jo
10000

5000

0.0

1.0

2.0

crack length increment, m m


Fig. 7.4

The effect of the number of cycles for each step (II)

25000

0.24

20000

0.32

s9oico

15000

jo

0.43

10000

5000

0.0

0.5
crack length increment, m m

Fig. 7.5

The effect of the f i n i t e e l e m e n t m e s h size

pp'fiH

^i aa

p
m
mmm

.55. 5

::

161.6.
1 1 i1

i5*55

for
for
for
for

x 1000

V
.

Fig. 7.6

result
result
result
result

300

measured
measured
predicted
predicted

number of

100

Comparison of fatigue c r a c k grwoth (1)

^00

ah

20

bu

80

100

number of cvcles,

measured
predicted
measured
predicted

Fig. 7.7

r e su l t
result
r e su l t
r esu l t

for
for
for
for

45 16.5 MPa
4 5 j
6.5 M P a
45 24 MPa
4 5 i :24 MPa

C o m p a r i s o n of f a t i g u e c r a c k g r o w t h (2)

120

I I n i 1 1 IT]1 I I > i 1 I I i | i 1 1 i I 1 1 i I | i Ij i ( Ii i > i > . i i i I [ 1 [

50
100
150
200
250
n u m b e r of cycles,

x 1000

m e a s u r e d r e s u l t f o r 4 5 16.5 MPa
p r e d i c t e d r e s u l t f o r 45 16.5 MPa
* " * m e a s u r e d result f o r t w o - s t e p loading
p r e d i c t e d r e s u l t f o r twostep loading
xxxxx

Fig. 7.8

C o m p a r i s o n of fatigue c r a c k growth (4)

Das könnte Ihnen auch gefallen