Anda di halaman 1dari 173

A Dissertation

Entitled

Fatigue Life Prediction and Modeling of Elastomeric Components

by
Touhid Zarrin-Ghalami

Submitted to the Graduate Faculty as partial fulfillment of the


requirements for the Doctor of Philosophy Degree in Engineering

Dr. Ali Fatemi, Committee Chair

Dr. Efstratios Nikolaidis, Committee Member

Dr. Mehdi Pourazady, Committee Member

Dr. Vijay Goel, Committee Member

Dr. Yong Gan, Committee Member

Dr. Patricia R. Komuniecki, Dean


College of Graduate Studies

The University of Toledo


May 2013

i
Copyright 2013, Touhid Zarrin-Ghalami

This document is copyrighted material. Under copyright law, no parts of this document
may be reproduced without the expressed permission of the author.

ii
An Abstract of

Fatigue Life Prediction and Modeling of Elastomeric Components

by

Touhid Zarrin-Ghalami

Submitted to the Graduate Faculty as partial fulfillment of the


requirements for the Doctor of Philosophy Degree in Engineering

The University of Toledo


May 2013

This study investigates constitutive behavior, material properties and fatigue

damage under constant and variable amplitude uniaxial and multiaxial loading

conditions, with the goal of developing CAE analytical techniques for durability and life

prediction of elastomeric components. Such techniques involve various topics including

material monotonic and cyclic deformation behaviors, proper knowledge of stress/strain

histories, fatigue damage quantification parameters, efficient event identification

methods, and damage accumulation rules.

Elastomeric components are widely used in many applications, including

automobiles due to their good damping and energy absorption characteristics. The type of

loading normally encountered by these components in service is variable amplitude

cyclic loading. Therefore, fatigue failure is a major consideration in their design and

availability of an effective technique to predict fatigue life under complex loading is very

valuable to the design procedure. In this work a fatigue life prediction methodology for

rubber components is developed which is then verified by means of analysis and testing

iii
of an automobile cradle mount made of filled natural rubber. The methodology was

validated with component testing under different loading conditions including constant

and variable amplitude in-phase and out-of-phase axial-torsion experiments. The analysis

conducted includes constitutive behavior representation of the material, finite element

analysis of the component, and a fatigue damage parameter for life predictions. In

addition, capabilities of Rainflow cycle counting procedure and Miner’s linear

cumulative damage rule are evaluated.

Fatigue characterization typically includes both crack nucleation and crack

growth. Therefore, relevant material deformation and fatigue properties are obtained from

experiments conducted under stress states of simple tension and planar tension. For

component life predictions, both fatigue crack initiation approach as well as fatigue crack

growth approach based on fracture mechanics are presented. Crack initiation life

prediction was performed using different damage criteria. The optimum method for crack

initiation life prediction for complex multiaxial variable amplitude loading was found to

be a critical plane approach based on maximum normal strain plane and damage

quantification by cracking energy density on that plane. The fracture mechanics approach

was used for total fatigue life prediction of the component based on specimen crack

growth data and FE simulation results. Total fatigue life prediction results showed good

agreement with experiments for all of the loading conditions considered.

iv
This dissertation is dedicated to my dear parents, Jila and Siawash.

v
Acknowledgements

I would like to sincerely thank my dear advisor, Prof. Ali Fatemi, who has

supported me throughout this long route with his knowledge, patience and useful hints.

This work could not be completed without his guidance, encouragement and advice

during long meetings, almost daily, and lots of emails. Dr. Yong Gan, Dr. Mehdi

Pourazady, Dr. Efstratios Nikolaidis, and Dr. Vijay Goel are highly appreciated for

serving on my Ph.D. committee.

I would also like to thank Chrysler Group LLC and specifically Dr. Yung-Li Lee

for funding the project and Paulstra CRC for providing the components and FE model.

The time and effort of Mr. John Jaegly, Mr. Randall Reihing and Mr. Tim Grivanos from the

machine shop of the MIME Department for helping in making different fixtures for uniaxial

and multiaxial testing of the components is highly acknowledged as well. I would also like to

thank my colleagues at the fatigue and fracture research laboratory of the University of

Toledo for their help, support and efforts during this period of hard working.

I would also like to thank Ms. Debbie Kraftchick and Ms. Emily Lewandowski in the

MIME department office for their help in providing the requirements of this study.

Finally, I would like to thank my dear parents and sisters for all their support and

encouragement throughout my education from the very beginning to doctorate degree.

vi
Table of Contents
Abstract .............................................................................................................................. iii

Acknowledgements ............................................................................................................ vi

Table of Contents .............................................................................................................. vii

List of Tables…… ............................................................................................................. xi

List of Figures ................................................................................................................... xii

List of Abbreviations ...................................................................................................... xvii

List of Symbols ................................................................................................................ xix

1 Introduction…… ....................................................................................................1

1.1 Preview…. .........................................................................................................1

1.2 Motivation for the study and objectives.............................................................3

1.3 Outline…............................................................................................................5

2 Literature Survey….. .............................................................................................8

2.1 Introduction ........................................................................................................8

2.2 Rubber deformation mechanisms and behavior .................................................8

2.3 Fatigue crack initiation and growth approaches ..............................................10

2.4 Multiaxial fatigue behavior and models ..........................................................14

2.5 Component fatigue testing and life prediction including FE simulations........20

2.6 Damage accumulation and cycle counting in variable amplitude

loading……......................................................................................................23

vii
3 Material Characterization and Fatigue Behavior …. .......................................27

3.1 Introduction ......................................................................................................27

3.2 Material and specimens…................................................................................27

3.3 Testing equipment ............................................................................................28

3.4 Specimen test methods and procedures ...........................................................28

3.4.1 Deformation behavior characterization.............................................28

3.4.2 Constant amplitude crack initiation testing.......................................29

3.4.3 Crack growth testing .........................................................................30

3.4.4 Variable amplitude crack initiation testing .......................................33

3.5 Specimen experimental results and observations ............................................34

3.5.1 Monotonic and cyclic deformation behavior results .........................34

3.5.2 Constant amplitude crack initiation tests results ...............................36

3.5.3 Crack growth tests results .................................................................38

3.5.4 Relationship between crack growth and crack initiation

approaches........................................................................................40

3.5.5 Variable amplitude crack initiation tests analysis methodology.......42

3.5.6 Variable amplitude crack initiation tests results and analysis ..........43

3.6 Conclusions ......................................................................................................45

4 Component Finite Element Model and Results …............................................65

4.1 Introduction ......................................................................................................65

4.2 Hyperelastic material deformation characterization ........................................65

4.3 Uniaxial FE model definition and specifications .............................................69

viii
4.4 Stiffness test results comparison with predictions ...........................................71

4.5 Uniaxial constant amplitude FE simulation results .........................................72

4.6 Conclusions ......................................................................................................73

5 Component Uniaxial Fatigue Behavior …. ........................................................80

5.1 Introduction ......................................................................................................80

5.2 Component life predictions ..............................................................................81

5.2.1 Crack initiation life predictions ........................................................81

5.2.2 Crack growth-based life predictions .................................................82

5.2.3 Variable amplitude loading life predictions ......................................84

5.3 Constant amplitude fatigue experimental program and validation of life

predictions .......................................................................................................86

5.3.1 Component fatigue tests ....................................................................86

5.3.2 Damage development........................................................................88

5.3.3 Comparison of predicted and experimental fatigue lives .................88

5.4 Variable amplitude fatigue experimental program and validation of life

predictions .......................................................................................................89

5.4.1 Component fatigue tests ....................................................................89

5.4.2 Damage development........................................................................91

5.4.3 Comparison of predicted and experimental fatigue lives .................91

5.5 Conclusions ......................................................................................................92

6 Component Multiaxial Fatigue Behavior …. ..................................................105

6.1 Introduction ....................................................................................................105

ix
6.2 Life prediction methodology for general random and multiaxial loading

of components ................................................................................................106

6.3 Applications to multiaxial constant amplitude loading .................................110

6.3.1 Experimental program and results ..................................................110

6.3.2 Life predictions and validation of methodology .............................112

6.4 Applications to variable amplitude multiaxial loading ..................................114

6.5 Discussion of results ......................................................................................116

6.6 Conclusions ....................................................................................................118

7 Summary and Possible Future Research… .....................................................136

7.1 Summary….. ..................................................................................................136

7.1.1 Constitutive behavior ......................................................................137

7.1.2 Fatigue material relations obtained from specimen experiments ...138

7.1.3 Uniaxial constant and variable amplitude component fatigue

behavior and life predictions ..........................................................139

7.1.4 Multiaxial constant and variable amplitude component fatigue

behavior and life predictions ..........................................................140

7.2 Possible future research .................................................................................143

References…..……….. ...................................................................................................145

x
List of Tables

3.1 Constant amplitude fatigue crack growth test conditions…………………. 48

3.2 Crack nucleation test conditions and results………………………………. 48

3.3 Simple tension specimen fatigue test conditions and results with the

random loading history used...……………………………………….......... 49

4.1 Summary of parameters used for numerical model………………………. 74

5.1 Summary of component fatigue test results……………………………… 94

5.2 Predicted nucleation and total fatigue lives of the rubber mount

component and comparison with experimental fatigue lives……….......... 95

5.3 Component fatigue test results and predictions under uniaxial variable

amplitude loading with R~0………………………………......................... 96

6.1 Component experimental conditions and test results.......………………... 120

6.2 Summary of component crack initiation and total life experiments and

predictions………………………………………………………............... 122

xi
List of Figures
3-1 Specimen geometry and dimensions for (a) simple tension, and (b) planar

tension. Specimen thickness is 1 mm for both geometries.........................................50

3-2 Axial servo-hydraulic Instron frame used for specimen testing..................................51

3-3 Simple tension and planar tension specimens with the corresponding

stretch and stress states [36]........................................................................................52

3-4 Crack initiation test setup showing five specimens and grips.....................................53

3-5 Maximum strain energy density versus maximum engineering strain

obtained from uncracked planar tension specimen.....................................................53

3-6 (a) Mullins effect showing initial transient softening in planar tension specimen

at 133% maximum strain, and (b) stabilization of stress with applied cycles

in displacement-controlled incremental step cyclic deformation tests of

planar tension specimen at different maximum strain levels.....................................54

3-7 (a) Loading history used for random loading tests of simple tension

specimens in displacement control and of cradle mounts in load control,

and (b) Rainflow cycle count for max/min values of ±100 with relative

xii
damage distribution from each cycle [63]..................................................................55

3-8 Superimposed plot of monotonic and stable cyclic curves for simple

tension and planar tension specimens.........................................................................56

3-9 Stable cyclic stress-strain loops from incremental step tests at different peak

strain levels from (a) simple tension test, and (b) planar tension test.........................57

3-10 Crack nucleation life as a function of peak strain in simple tension tests.................58

3-11 Equivalent fatigue life as a function of equivalent maximum strain by using

Mars-Fatemi R-ratio model........................................................................................59

3-12 Crack length versus cycle linear fits from crack growth test at (a)RT = 0,

(b)RT = 0,(c)RT = 0.05, and (d) RT = 0.10.................................................................60

3-13 Fatigue crack growth rate data comparisons at different R-ratios.............................61

3-14 Fatigue crack growth rate data correlations at different R-ratios based on

Mars-Fatemi R-ratio model.......................................................................................62

3-15 Comparison of fatigue lives obtained from crack initiation and crack

growth approaches....................................................................................................63

3-16 Experimental versus predicted fatigue lives in blocks to failure for simple

tension specimen tests subjected to variable load history.........................................64

4-1 Vehicle cradle mount used as illustrative example (a), and rubber mount FE

xiii
model where due to symmetry half of the model is shown (b)...................................75

4-2 Component mid-life hysteresis loops under axial loading condition for (a)

R ~ 0 and, (b) R = 0.2.................................................................................................76

4-3 Stiffness comparison between FEA simulation and monotonic test of the

mount (a) and cyclic tests of the mount (b)...............................................................77

4-4 Strain (a) and stress (b) histories at the critical element location obtained

from FE simulation during one sinusoidal loading cycle between 750 N

and 3750 N.................................................................................................................78

4-5 Maximum principal strain distribution under R =0.2 and load amplitude of

1,500N.........................................................................................................................79

5-1 Maximum R = 0 equivalent engineering strain versus crack depth for load

amplitude of 1500 N and R = 0.2 from FEA simulation.............................................97

5-2 Component failure definition based on stiffness degradation in fatigue tests.............98

5-3 Displacement amplitude versus cycles in fatigue tests................................................99

5-4 Experimental failure locations and crack growth direction in component tests........100

5-5 Experimental versus predicted fatigue life based on crack initiation approach

(a) and based on crack growth approach (b).............................................................101

5-6 Component displacement amplitude response as a function of applied blocks

xiv
of loading under variable amplitude fatigue loading................................................102

5-7 Crack length versus applied load blocks for a test at R ~ 0 and with a load

range of 3,600 N......................................................................................................102

5-8 Maximum principal strain contour results for R~0 and loading range of

3,000 N obtained from FE simulation.......................................................................103

5-9 Component fatigue testing and predictions for variable amplitude random

loading at R~0, both for crack initiation and total component lives.........................104

6-1 Vehicle cradle mount FE model................................................................................124

6-2 Definition of failure illustrated by variation of axial and rotational

displacements in constant amplitude in-phase axial-torsion tests of the cradle

mount........................................................................................................................124

6-3 Experimental response for constant amplitude axial-torsion loading of

components for (a) IP and (b) OP tests.....................................................................125

6-4 Axial displacement versus rotation angle of the component for different

cycles throughout the CA and A-T tests of (a) IP and (b) 90 OP...........................126

6-5 Axial displacement versus rotation angle of the component for mid-life cycles

for CA and A-T IP and 90° OP tests.........................................................................127

xv
6-6 Evolution of crack length and depth for constant amplitude axial-torsion

(a) in-phase test with Nf= 31,290, and (b)out-of-phase test with Nf= 21,454...........128

6-7 Crack initiation to total fatigue life ratio for all types of loadings............................129

6-8 Maximum principal strain location for constant amplitude in-phase loading

simulation.................................................................................................................130

6-9 Experimental versus predicted component initiation life for all loading

conditions based on (a) maximum normal strain, (b) critical CED, and

(c) CED on MNS plane.............................................................................................131

6-10 Experimental versus predicted component total fatigue life for all loading

conditions................................................................................................................133

6.11 Variable amplitude in-phase and out-of-phase axial-torsion loading history

used for component testing.....................................................................................134

6.12 Normal strain on maximum normal strain (MNS) plane history of variable

amplitude and axial-torsion loading for (a) in-phase, and (b) out-of-phase

loading....................................................................................................................135

xvi
List of Abbreviations
A....................Axial
ASTM............American Society of Testing and Materials
A-T................Axial-Torsion

CA.................Constant Amplitude
CAE...............Computer Aided Engineering
CED...............Cracking Energy Density

DC.................Displacement-Controlled

EPDM............Ethylene Propylene Rubber

FCG...............Fatigue Crack Growth


FE..................Finite Element
FEA...............Finite Element Analysis

IP...................In-Phase

LC................. Load-Controlled
LDR...............Linear Damage Rule

MNS..............Maximum Normal Strain

NR.................Natural Rubber

OP..................Out-of-Phase

xvii
SAE...............Society of Automotive Engineers
SBR...............Styrene-Butadiene Rubber
SED...............Strain Energy Density
SWT..............Smith-Watson-Topper

VA.................Variable Amplitude

xviii
List of Symbols
a……………..crack length
a0………….... initial crack length or flaw size
af………......... final crack length
Bf…………… blocks to failure
B……………. fatigue crack growth coefficient, number of blocks
C……………. initial Young’s modulus, Green deformation tensor
cij.................... material constants for Rivlin strain energy density function
dA…………... change in new crack surface area
Du………….. change in stored mechanical energy
dWc ………….increment in cracking energy density

d ………….. unit displacement vector
E …………… modulus of elasticity, Green-Lagrange strain tensor
emax…………. maximum engineering strain
F……………. deformation gradient tensor, fatigue crack growth exponent
Fa…………… load amplitude
F(R)………… power-law exponent for Mars-Fatemi model
f..…………... test frequency
G…………… shear modulus
G(R)………... power-law exponent for Mars-Fatemi model
h…………... specimen height
I …………… unit matrix
I1 , I 2 , I 3 …… the first, second and third invariants of the stress tensor
J el ………… elastic volume ratio
k……………. strain dependent material parameter
L,L0………… length, initial length of specimen
Ma………...... torque amplitude
N…………… number of cycles
NE………..... nominal strain
N f …………. cycles to failure
N̂ ………….. unit vector normal to the plane in space
ni…………… principal stretch direction
Pa…………... load amplitude
R…………… minimum to maximum ratio
r, da/dN…..... crack growth rate
rC ………….. critical crack growth rate

xix
RM ………... torque R- ratio
R P ………... load R- ratio
R ……….... displacement R- ratio
R …………. strain R-ratio
R ………… rotation R- ratio
RT ………… energy release rate R-ratio
RW ………… strain energy density R-ratio

R …………....unit vector in the undeformed configuration
r …………... unit vector in the current configuration
S…………… stress
Smax………… maximum engineering stress
~ nd
S ................... 2 Piola-Kirchhoff stress
T…………… tearing energy, energy release rate, Cauchy stress tensor
TC………….. critical energy release rate

T …………... traction vector
U…………… strain energy density per unit of reference volume
U vol …………volumetric part of U
U dev ………...deviatoric part of U
V……………..left stretch tensor
W…………….strain energy density
a ……………displacement amplitude
 …………...strain
 1 ……………maximum principal strain
 C ………….. critical strain for crack initiation
 ………….. .stretch ratio
 i ,  i ………. material constants for Ogden strain energy density function
 …………..strain range
V ………….change in volume
……………. mass density
a…………….rotation amplitude

xx
Chapter 1

Introduction

1.1 Preview

The definition of fatigue as currently stated by ASTM [1] is “the process of

progressive localized permanent structural change occurring in a material subjected to

conditions which produce fluctuating stresses and strains at some point or points and

which may culminate in cracks or complete fracture after a sufficient number of

fluctuations.”

Fatigue failure is a typical service failure mode resulting from alternating and

repeated loads over a period of time. Although elastomeric components are commonly

used in many industries and their field of application has broadened in recent years, there

is very limited research related to fatigue failure aspects of elastomeric components, as

compared to metallic components. As a result, fatigue life evaluation and prediction of

rubber components are important topics to investigate to assure their safety and

reliability [2].

Elastomers are highly nonlinear elastic materials with large deformations. These

characteristics along with high energy absorption ability and dynamic damping capacity

as well as low cost of manufacturing rubber parts make this material ideal for many

1
diverse engineering applications. This includes the automotive industry (e.g. car tires,

bumpers, engine mount), piping (e.g. gaskets, bearings, hoses, seals) and as vibration

isolators in many other applications.

However, the behavior of rubber materials is complex due to high deformability,

quasi incompressibility, stress softening, and time-dependent effects. Therefore, a correct

description of the behavior of elastomers must include both geometric and material

non-linearities as well as incompressibility [3]. Furthermore, properties of rubber can

change drastically depending on material composition, manufacturing conditions (mixing

and curing), loading condition (strain state and strain rate) and other service conditions

(temperature and period of usage, for example) [4].

Availability of an effective technique to predict fatigue life under actual complex

loading is very valuable to the design procedure. Such a technique or model relates

fatigue properties obtained from constant amplitude fatigue characterization tests to

variable amplitude (VA) service loading conditions. Most researchers studying fatigue

behavior of elastomers have mainly focused on constant amplitude loading conditions,

rather than variable amplitude loading applications.

In order to be able to predict life of elastomeric components, several inputs are

necessary. These include the component geometry, relevant material properties, and the

loading history. For analysis, two basic steps are often required. First, stress and/or strain

analysis is used, typically by finite element method, to obtain stress/or strain histories at

critical location(s) from the applied load history. Then, fatigue life prediction analysis is

performed based on a fatigue damage quantification parameter and damage accumulation

model using the critical location stress and/or strain history. Linear elastic and small

2
strain assumptions, which are typically used for metals, do not apply to elastomers. As a

result, continuum mechanics parameters such as deformation gradient, stretch ratio, and

Green-Lagrange strain components are used in rubber mechanics analysis.

Numerical FE simulation softwares have been broadly used for design and they

have become more reliable in recent years. However, both analysis and testing are still

required for fatigue design. The process of producing inexpensive and high quality

products requires reducing the amount of testing by more accurate analysis.

1.2 Motivation for the study and objectives

The motivation for this study was a practical need to predict fatigue life for

industrial components made of elastomeric materials. The main input of the study is the

material properties obtained from material testing. Different loading histories from

simple to complex are evaluated, including variable amplitude non-proportional

multiaxial loading, which is a realistic loading condition in many industrial applications.

Crack nucleation life approach and total fatigue life approach based on crack growth

properties are used for life predictions, and illustrated by using a passenger vehicle cradle

mount.

At the critical location(s) of most components and structures, multi-axial states of

stress and strain are typically present during each loading cycle. Such components and

structures are also usually subjected to variable amplitude or random loading histories.

Therefore, the study of multiaxial fatigue and deformation under variable amplitude

loading is very important. The extreme values for the strain and/or stress components do

not usually coincide with each other under complex loading conditions. This indicates the

complex nature of the strain history in applications such as vibration isolators.

3
The typical life prediction methodologies for uniaxial loading have been well

developed and used for many years. These methodologies for the more complex case of

multiaxial variable amplitude loading are not yet well established, particularly when the

loads are non-proportional. Very limited papers can be found in the literature related to

fatigue life prediction of elastomeric components under variable amplitude multiaxial

loading. Therefore, the objective of this study is to develop a robust methodology for

fatigue life prediction of such complex loading conditions and validate it by using an

automobile cradle mount. More specifically, the following are the objectives of this

research:

1) To develop CAE analytical techniques for durability and life prediction of

elastomeric components.

2) To validate the CAE analytical techniques with cradle mount component testing.

3) To be able to significantly reduce component bench tests to save time and cost,

thereby facilitating optimization.

4) To evaluate the recent promising and other commonly used damage parameters

for fatigue life assessment of elastomeric industrial components under multi-axial

loads.

5) To investigate the effect of dwell periods and loading rate on behavior of rubbers

and its industrial applications.

6) To evaluate the capability of commonly used cycle counting procedures (rainflow

cycle counting method) and damage accumulation rules (linear damage rule) for

life predictions of elastomers under different loading conditions.

4
1.3 Outline

In Chapter 2, a brief literature review is presented. This chapter reviews different

works on material properties and constitutive behavior models, crack initiation and

growth approaches, multi-axial fatigue models, FE simulations, damage accumulation

and application to variable amplitude loading.

An understanding of stress-strain behavior, particularly under cyclic loading, is

necessary for the study of fatigue. Material fatigue behavior characterization for crack

initiation and crack growth, and constitutive behavior characterization from simple

tension and planar tension tests are presented in Chapter 3. Test procedures, testing

apparatus, test methods used, and obtained data are presented and discussed in this

chapter.

Chapter 4 presents the finite element modeling aspects of the component. FE

model details, preprocessor modules description, material properties used in FE analysis

and analysis type, step and loading considerations are discussed in this chapter. Mesh

sensitivity analysis and the effect of load control versus displacement control input on the

results are other issues which are discussed in this chapter. Stiffness comparison between

FE simulation and experiment as well as simulation results for uniaxial constant

amplitude loading of the component are also presented.

Chapter 5 presents the component uniaxial constant and variable amplitude

loading experiments and life predictions. FE simulations, fatigue crack initiation as well

as total fatigue life prediction methodologies are also discussed. Damage evolution in the

component is also described in this chapter.

5
Component multiaxial fatigue behavior and life predictions are presented in

Chapter 6. First, the life prediction methodology for multiaxial and random loading is

presented. This methodology discusses both crack initiation and growth approaches.

Then, the capability of these damage quantification parameters in fatigue life prediction

for different axial-torsion loadings of both constant and variable amplitude is discussed.

This comparison is done with bench component testing and by using the simulation

results and the methodology presented. The results of the prediction methods are also

discussed in this chapter. Cracking behavior and damage evolution are also presented in

Chapter 6. Effects of control mode on fatigue life, hysteresis loops, crack initiation

location and crack initiation life are also discussed in this chapter.

Finally, a summary of the dissertation along with the important conclusions are

presented in Chapter 7. Some suggestions for further studies in multiaxial fatigue of

elastomers are also presented in this chapter. These suggestions include further studies to

develop or improve a multiaxial fatigue life prediction results by finer meshing at the

critical locations of the component and utilizing a Fracture Mechanics FE model of the

component for simulations.

Based on the contents of this dissertation, five papers have been published.

Material deformation and fatigue behavior of the natural rubber used in manufacturing

the cradle mount used are characterized in [5] based on what is presented and discussed

in Chapters 2 and 3. The component FEA and fatigue life analysis and predictions for

uniaxial constant amplitude and variable amplitude loadings are presented in [6] and [7],

respectively. The materials in these papers are extracted from Chapters 2, 4 and 5.

A robust methodology for fatigue life prediction of elastomeric components under

6
complex loading conditions which is developed in Chapter 6 is presented in [8] for

general random and multiaxial loading. Then applications to constant amplitude

axial-torsion in-phase and out-of-phase loading as well as variable amplitude loading are

demonstrated using experimental results from the cradle mount. The information

presented in this paper is mainly chosen from the contents of Chapter 6. An overview of

the CAE durability analytical techniques for predicting crack initiation life of the

automotive cradle mount and their validation under constant and variable amplitude

uniaxial and multiaxial loadings is also presented in [9].

7
Chapter 2

Literature Survey

2.1 Introduction

In this chapter analytical and numerical approaches that are currently available for

predicting fatigue life of elastomers are reviewed. Deformation mechanisms of

elastomers are first discussed. Different studies on two main approaches of crack

initiation and crack growth are then reviewed. Multiaxial fatigue models and different

damage quantification parameters related to complex loading conditions constitute the

next part of this literature survey.

Numerical methods and fatigue life predictions of complex geometries such as

those used in industrial applications are also covered in this review. Recent works on

damage accumulation and application in variable amplitude loading of elastomers is the

last topic which is reviewed in this chapter.

2.2 Rubber deformation mechanisms and behavior

Rubber can have a wide range of mechanical properties by changing the

compound formulation and manufacturing process. Strain crystallization has been shown

8
to have a beneficial effect on fatigue life at moderate or high strain levels. The addition of

carbon black to rubber compound could strengthen the material against fatigue failure

and drastically change its mechanical properties. Antidegradants are added to rubber

compounds to avoid the deleterious effects of oxygen and ozone. Vulcanization is used in

thermosetting elastomers to create covalent bonds or crosslinks between adjacent

polymer chains. Crosslink density determines the physical properties of rubber, with

higher crosslink density resulting in increased stiffness and reduced hysteresis.

Compound stiffness has a direct effect on energy release rate. Filled rubbers show more

dissipative mechanical responses at both high strains because of network chain breakage

and strain crystallization, and at small or moderate strains under alternating loading.

Hysteresis is an important aspect when evaluating fatigue properties in terms of

deformation, crack nucleation, or crack growth.

In natural rubber material, hysteresis behavior appears at high strains above 200

to 250% and is usually associated with the phase transformation process called

crystallization. The initial softening, which is called the Mullin’s effect, depends

non-linearly on the maximum principal strain reached during a cycle [10].

Mars and Fatemi [11] investigated the monotonic and cyclic behaviors of filled

natural rubber. They conducted pure axial, pure torsion as well as proportional and

non-proportional axial-torsion experiments by using short thin-walled cylindrical

specimens. They observed initial softening within the first 10 cycles of fatigue loading

compared to monotonic tests and also the stabilized stress-strain response after a short

number of cycles. This is associated with irreversible breakage of different types of bonds

in the elastomer-filler composite. They also described the deviations from non-linear

9
elasticity under cyclic loading for filled natural rubber. Their results also showed the

effect of phase angle on the shape and size of the hysteresis loops with the 90° out of

phase loading histories leading to the maximum hysteresis. Initial overload was found to

have a major effect on the subsequent evolution of the stress amplitude response. By

applying the initial overload, the logarithmic decrease trend of stress amplitude drop with

cycles totally disappeared and the stress response remained constant. Mars and Fatemi

[12] also described the inability of current hyperelastic models to capture the Mullin’s

effect and proposed a new model relating flaw growth to the softening of cyclic stress-

strain response to explain the behavior of the specimen during late portion of fatigue life.

2.3 Fatigue crack initiation and growth approaches

Mars and Fatemi [13] conducted a literature study on different approaches

associated with fatigue analysis for rubber. Fatigue failure process is divided into two

distinct phases, the crack nucleation phase in which crack nucleates in regions that were

initially crack free, and crack growth phase when the nucleated cracks grow to the point

of failure or rupture. They identified maximum principal strain (or stretch) and strain

energy density as two broadly used damage quantification parameters for the nucleation

phase. Although it is commonly observed that cracks initiate on a plane normal to the

maximum principal tensile strain direction, the strain energy density criterion applied as a

scalar criterion can’t account for this preferred orientation. For the fatigue crack growth

stage, a fracture mechanics analysis is used based on energy release rate. Mars and

Fatemi also discussed the relationship between crack nucleation and crack growth

approaches by using an integrated power-law model, which is also proposed in [13].

10
Stevenson [14] conducted a study of fatigue crack growth of rubber in

compression with a cylindrical specimen. He concluded that crack growth in compression

occurs at an approximately constant rate and is restricted to the outer regions of the test

specimen with high local shear strains. Legorju-jago and Bathias [15] showed that a mean

stress in tension improves the fatigue behavior by crystallization of the stretched bonds in

pure tension cycles. On the other hand, a minimum stress in compression seriously

damages the material.

Abraham et al. [16] investigated the effect of minimum stress and stress

amplitude on fatigue life of non-strain crystallizing elastomers such as ethylene

propylene (EPDM) and styrene-butadiene (SBR) rubbers. They used cylindrical dumbbell

specimens and concluded that increasing minimum stress with constant stress amplitude

can increase the service life by a factor of 10 despite the increase in maximum stresses.

To account for Rε ratio (minimum to maximum strain ratio) effect on either crack

nucleation or crack growth life, Mars and Fatemi [17] proposed a phenomenological

model. The ability of the model to represent data for different R ratios was shown to be

reasonable based on their own experimental data, as well as data of Lindley et al. [17].

Harbour et al. [18] also utilized this model to represent their non-zero fatigue crack

growth rate data.

Wang et al. [19] proposed a continuum damage mechanics model for fatigue

damage behavior of elastomers. They used the Ogden model to construct the constitutive

relation for carbon-filled natural rubber. They introduced an equation for fatigue life as a

function of applied nominal strain amplitude. They concluded their proposed relation

11
could describe experimental data very well, though restricted to R ratio of zero and fixed

frequency values [19].

A crack initiation life diagram is the Haigh diagram which is a contour plot of

strain amplitude versus mean strain in which lines of equal fatigue life are plotted [20].

The Cadwell diagram is another format which also summarizes the dependence of the

crack nucleation life on the changing limits of constant amplitude cycle. In this diagram

base 10 logarithm of the fatigue life is plotted versus minimum strain with the contours of

strain amplitude kept constant [21].

Two main mechanisms, decohision and cavitation, are responsible for fatigue

damage nucleation which are both independent of the type of loading and dependent on

the nature of the inclusion [22]. The decohision process could be easily seen on the

fracture surface, because the surface of the inclusion is free of rubber after decohision. It

is found to be predominant at rigid inclusions such as SiO2 and CaCO3. The interface

between the inclusion and rubber matrix is the weakest location, therefore, fatigue

damage initiates at the interface which causes the decohision between the matrix and

inclusion. Cavitation is the process of sudden void initiation under stress state [22].

Mars and Fatemi [23] observed a direct relationship between pre-existing flaws in

the rubber material and the crack initiation process by assuming that the pre-existing

flaws in the materials were small cracks and crack initiation occurs when one of these

small cracks becomes large.

Several factors including environmental conditions, formulation of the rubber,

mechanical load history, and dissipative effects of the stress-strain behavior affect the

12
fatigue life of rubber. The synergistic effects of all abovementioned factors define the

fatigue life. Mars and Fatemi [24] investigated a detailed survey of these factors.

Kim and Jeong [25] investigated natural rubber compound with three different

filler types to examine the effect of carbon black on fatigue life. They concluded that for

larger carbon black agglomerates, separation of fillers from rubber matrix could relatively

easily occur resulting in shorter fatigue life, compared to smaller size fillers.

Santangelo and Roland [26] verified that double network of natural rubber (NR) has a

higher modulus than single network of equal crosslink density. Fatigue life of double

networks NR was also found to be as much as a factor of 10 higher than conventionally

crosslinked NR, although tensile strength was slightly lower.

Environmental effects can significantly affect fatigue life of elastomers. High

temperature has a detrimental effect both on crack nucleation life and crack growth rate.

This effect is intensified in amorphous rubber. Ozone could shorten the crack growth life

due to reaction with carbon bonds at the crack tip. Presence of oxygen increases the

fatigue crack growth rate. Mars and Fatemi discussed the aforementioned effects

including temperature, ozone, oxygen, and electrical charges in detail in [24].

The crack growth approach is based on the growth of pre-existing flaws or cracks

using fracture mechanics. This approach for rubber was developed in the 1950s and

1960s. The strain energy release rate introduced by Griffith for brittle fracture and

extended to rubber by Rivlin and Thomas [27] is generally used as the governing fatigue

damage parameter for describing fatigue crack growth rates.

13
2.4 Multiaxial fatigue behavior and models

Rubber structures in service usually encounter time-varying loads in several

directions necessitating a proper multiaxial fatigue criterion for relating stress and strain

histories to fatigue life. An important issue that should be considered in multiaxial fatigue

is in-phase versus out-of-phase loading. When the different loading channels reach to

their peak values at the same time, the loading is called in-phase, while when the time of

reaching maximum or minimum value for different load channels are different, the

loading is called out-of-phase.

Life prediction methodologies for uniaxial loading have been relatively well

developed. These methodologies for the more complex case of multiaxial variable

amplitude loading are not yet well established, particularly when the loads are

non-proportional. Very limited papers can be found in the literature related to fatigue life

prediction of elastomeric components under variable amplitude multiaxial loading.

Multiaxial fatigue life prediction methodologies for rubber can be classified into

four main approaches. These consist of the equivalent strain approaches, energy

approaches, equivalent stress approaches, and the more recent critical plane approaches.

Critical plane models can be used for both proportional and non-proportional loading

conditions and are based on the physical process of the damage process. Critical plane

approaches could be strain-based (such as the maximum normal strain), stress-based

(such as the maximum principal Cauchy stress), or energy-based (such as cracking

energy density). Energy-based critical plane approaches, which use both stress and strain,

can reflect the constitutive behavior of the material, while stress or strain-based critical

plane approaches do not require the constitutive behavior of the material.

14
In general, a good multiaxial fatigue model should be robust, sensitive to load

phasing and mean stress, and applicable to variable amplitude loading. Another important

characteristic is the ability to consider crack closure effects.

Crack initiation is normally related to the continuum mechanics quantities

(e.g. stress or strain) which are macroscopic. The most widely used parameters for crack

initiation are the maximum principal strain, maximum principal Cauchy stress and strain

energy density (SED). Approaches such as the cracking energy density (CED) or based

on configurational mechanics (Eshelby stress tensor) are more complex and recent

approaches dealing with multiaxial fatigue analysis of elastomers.

The maximum principal strain criterion was introduced by Cadwell et al. [21] for

unfilled vulcanized natural rubber in 1940 and still remains one of the most commonly

used criteria for rubber. For incompressible materials, this criterion always gives a

positive value for maximum principal strain. Octahedral shear strain is also another

strain-based predictor which also always gives a positive value for rubber-like

materials. In addition, since rubber is incompressible, the hydrostatic stress is

independent of strain tensor.

Ayoub et al. [28] established a relationship between the fatigue life of

styrene-butadiene rubber and the stretch amplitude. They conducted multiaxial constant

and variable amplitude tests on cylindrical (axisymmetric) specimens. The specimen

curvature radius was large enough to minimize the stress triaxiality effect. Their results

showed a good agreement between predicted and experimental fatigue lives. The

predictions were based on continuum damage mechanics improved by incorporating CED

15
criterion. The developed model is both a fatigue damage criterion and an accumulative

damage rule.

Rivlin and Thomas first applied strain energy density (SED) criterion to rubber

material under static loading. This criterion has some drawbacks including inability to

differentiate between simple tension and compression, inability to account for crack

closure and the fact that all of the stored energy in the material would not release due to

crack growth [29]. Strain energy density usually uses a hyperelastic formulation which is

defined in terms of strains. If strain energy density and maximum principal strain would

be applied as scalar criteria, the fact that cracks are usually observed to orient in specific

orientation could not be predicted.

Mars and Fatemi [30] designed a novel specimen for investigating the mechanical

behavior of elastomers under multiaxial loading conditions. By utilizing this specimen

and based on their experimental observations [23] they concluded that the maximum

principal strain as a fatigue damage parameter gave the best prediction of fatigue life,

while the traditionally used strain energy density gave the worst correlation of

experimental data.

Mars and Fatemi [31] discussed the observations of crack initiation and small

flaw growth in filled natural rubber under multiaxial loading conditions. They used their

designed ring specimen with axial, torsion and both proportional and non-proportional

axial-torsion loadings. They suggested that crack nucleation in rubber starts from existing

flaws in the virgin material such as voids, surface cavities and non-rubber particles.

Fatigue crack initiation and growth were observed to occur on preferred failure planes.

For axial, torsion and in-phase axial-torsion loading the cracking plane was transverse to

16
the maximum principal strain direction. For more complex loading they still observed

preferred nucleation planes but their relation to the principal strain directions was

sometimes different. Crack closure affected the nucleation plane and more closure was

observed in cyclic torsion under static compression. This also happened in fully reversed

axial-torsion experiments due to friction during the torsion-compression part of the

loading.

Andre et al. [32] performed torsion experiments on axisymmetrical notched

sample geometry specimens and concluded that the maximum principal Cauchy stress

could be related to multiaxial fatigue damage mechanisms. They proposed that the

maximum principal Cauchy stress could be used as a fatigue damage quantification

parameter and observed that crack orientation is perpendicular to its direction.

Brunac et al. [33] extended Haigh’s diagram to arbitrary 3D loadings by considering the

smallest sphere containing the closed path of the positive part of the Cauchy stress tensor.

They concluded that this approach results in successfully predicting fatigue life.

Saintier at al. [22] investigated the micro-mechanisms leading to crack initiation

under multiaxial fatigue loading conditions and suggested that cracks initiate from

inclusions or large carbon black agglomerates. They observed decohision and cavitation

as damage mechanisms and suggest that cyclic loading does not produce a new damage

mechanism in rubber, in contrast to metallic materials. They observed that the maximum

first principal stress reached during the cycle defined the orientation of the crack in all of

the fatigue loading conditions they considered. In another study, Saintier et al. [10]

proposed a critical plane approach for fatigue crack initiation based on the

micro-mechanisms of crack initiation. In their study, under non-proportional multiaxial

17
loading, dependent on the material, fatigue crack growth orientation was found to be on

the plane of maximum shear stress amplitude (shear cracking), maximum normal stress

plane (tensile cracking), or even mixed-mode shear and tensile cracking.

Mars and Fatemi [34] concluded that scalar equivalence criteria were not capable

of predicting the fatigue initiation life in natural rubber. They suggested using cracking

energy density which is the portion of the strain energy density that is available to cause

crack growth on a particular plane. The application of this method involves knowledge of

the constitutive behavior of the material. Zine et al. [35] applied the cracking energy

density criterion in a FE code and found good agreement between numerical and

analytical results for common strain states. Their experiments results also showed the

efficiency of this criterion to explain fatigue life of elastomers under multiaxial loading

conditions.

Harbour et al. [36] used the multiaxial ring specimen in [30] and performed both

constant and variable amplitude axial-torsion experiments. They used two rubber

materials, one which strain crystallizes (natural rubber) and one which does not (SBR).

They concluded that both cracking energy density and normal strain approaches were

able to predict the dominant crack orientations for some of the test signals in each

material (i.e. axial and multilevel axial tests), but not successful for some other loadings

such as for fully reversed torsion experiments. In another study Harbour et al. [37] used

the maximum normal strain to find the critical plane and the cracking energy density on

that plane to determine fatigue life. Their results showed that this criterion produced

similar fatigue life results compared to other approaches such as cracking energy density,

strain energy density and maximum principal strain. Because maximum normal strain is

18
independent of constitutive behavior, the critical plane can be identified more easily than

for CED. They [37] also studied the effect of variable amplitude multiaxial loading and

concluded that Miner’s linear damage rule gave reasonable predictions of their

experimental results. Verron et al. [38] proposed a multiaxial criterion for crack

nucleation based on the local properties of the Eshelby second-order tensor [39] and

defined in terms of continuum mechanics parameter.

Wang et al. [29] evaluated fatigue life prediction approaches by using

experimental results from proportional and non-proportional loading paths applied to

small axisymmetric diabolo specimens made of vulcanized NR. They conclude maximum

principal strain and octahedral shear strain provide good predictions of the fatigue life.

They also found that compared to traditional criterion of strain energy density, the

cracking energy density model gives better life prediction results.

Mars [20] investigated the duty cycle on each material plane along with its

corresponding damage to transform the multiaxial loading into the localized flaws. After

identifying the damaging events, the original duty cycle is simplified and reconstituted to

a new duty cycle including the number of the most damaging events of the original duty

cycle. The new shortened duty cycle maintains those features of the original duty cycle

corresponding to the original mode of failure and shortens the time scale of the test. For

multiaxially loaded rubber parts, Flamm et al. [40] proposed discretizing the continuous

signal by a level crossing cycle counting method for each loading channel. Then they

constituted the stress amplitude history and performed Rainflow cycle counting on these

alternating points.

19
2.5 Component fatigue testing and life prediction including FE

simulations

In order to be able to evaluate or predict fatigue life of the component, it is

necessary to obtain the stress and strain history for critical locations of the component.

Due to hyper-elastic material behavior and complex component geometry, measurements

of mechanical quantities such as stress, strain, and strain energy density often necessitate

the use of numerical methods such as finite element (FE). The finite element method has

been a very useful tool for simulating the actual service performance of components at

their design and development stage. This method can significantly help to reduce

trial-and-error cycles, prototyping and testing efforts, and time to market and total

cost [41].

Selection of a proper failure criterion is another key issue in analyzing fatigue life

of rubber components. Many studies discuss capabilities of different criteria in

determining fatigue life and correlation of the predicted lives with experimental lives.

Mars and Fatemi [13] provide a literature survey of fatigue analysis approaches for

rubber, as well as a review of the many factors that affect fatigue life of rubber [24].

Determination and realistic representation of material deformation and fatigue behavior is

another important input to accurate FE analysis and life prediction of rubber components.

Rubber mounts are often used as vibration isolators in vehicles. They provide

damping for high-amplitude as well as low frequency road induced vibrations, and in the

case of engine mount, low-amplitude high frequency engine induced vibrations.

Li et al. [42] predicted fatigue life of a rubber mount by obtaining material properties and

utilizing the maximum principal strain at the critical region as a fatigue damage

20
parameter determined from FEA. To validate their predictions, they performed constant

amplitude fatigue tests in load control and in directions perpendicular to the axial

direction.

In another study of an engine rubber mount by Kim et al. [43] the maximum

Green-Lagrange strain and the maximum energy density were used as fatigue damage

parameters for life predictions. Load versus Green-Lagrange strain relation was

determined by FEA of the engine mount and the material fatigue behavior was derived

based on fatigue tests of dumbbell-shaped specimens. Based on their load-controlled

constant amplitude fatigue tests of rubber mount they concluded that Green-Lagrange

strain is a better fatigue damage parameter than strain energy density for component life

predictions.

Woo et al. [2] performed fatigue life predictions of natural rubber components by

using the Green-Lagrange strain at critical location determined from FEA. Fatigue tests

under displacement control and for different values of mean displacement and amplitude

were performed. They found their predictions to agree fairly well with the experimental

fatigue lives.

Fatigue failure analyses of rubber springs, which are widely used as anti-vibration

components, were conducted by Luo and Wu [44]. They showed that a nonlinear

quasi-static FE simulation capable of modeling geometric and material nonlinearities can

effectively be used for product design and failure analysis. Luo et al. [45-47] analyzed

fatigue life of anti-vibration rubber springs used in a rail vehicle suspension system with

the aid of FEA to obtain the stress contour and using the S-N curve obtained under

uniaxial loading for fatigue life predictions. They used the Mooney-Rivlin hyper-elastic

21
material model and proposed a three-dimensional effective stress criterion describing an

ellipsoidal failure envelope for fatigue failure based on principal stress values.

Verification tests of this approach carried out on two types of rubber springs in the

longitudinal direction under uniaxial loading indicated good agreement with the

predictions both for crack location and for crack initiation orientation [47].

Zhao et al. [48] analyzed a rubber mount using FEA. They determined the main

reason for fatigue cracking was the stress concentration at the interfaces between the

rubber and metal layers, which was also observed in their constant amplitude mount

bench testing. They then modified the geometry of the mount by dig grooving in the

rubber layers and changing the shape of the inner bushing. This resulted in a reduction of

maximum strain due to lower stress concentration and, therefore, improved fatigue life.

Takeuchi at al. [49] devised a procedure for endurance fatigue testing using

optimum test piece geometry for fatigue testing. The geometry was designed by FEA

with the aim of producing the same maximum tensile strain experienced by the rubber

component in bench endurance tests. The experimental fatigue life of the material was

shown to have a close correlation with component experimental results. Lee and

Kim [50] attempted to minimize both the weight and maximum stress of engine rubber

mount to maximize the fatigue life cycle subjected to constraints on the static stiffness of

the mount.

Sundararaman et al. [51] evaluated fatigue crack growth of V-ribbed belts by

using fracture mechanics and FE. They used a power-law relation to characterize the

fatigue crack growth rate based on J-integral as well as sub-modeling capability of the

ABAQUS software, which eases the evaluation of large models by concentrating on the

22
critical location in the local sub-model. In their study the J-integral analysis was

performed by using hyper-elastic strain energy density function of Ogden.

2.6 Damage accumulation and cycle counting in variable amplitude

loading

In most component and structures variable amplitude loading is typical in service.

In order to analyze variable amplitude loadings, a cycle counting method is used to

disintegrate the loading events to constant amplitude loads. Then by using a cumulative

damage rule, the amount of damage caused by each individual cycle is defined and the

damage is accumulated to predict failure or to evaluate remaining life.

Steinweger et al. [52] developed a test time reduction method for rubber parts by

using Rainflow filtering to reduce the length of the load block by removing

non-damaging cycles. In multi axial loading it is important to preserve the phasing angle

between load channels and their model accounts for this aspect. They suggest

conventional methods used for steel parts are not applicable to rubber parts due to their

nonlinearity at higher load amplitude and damping properties at higher frequencies.

Ayoub et al. [28] mentioned that the damage rules used for rubber are generally

based on experimental data which were obtained for specific loading cases and materials.

Because of its simplicity, Miner’s rule [53] is widely used for variable amplitude loading

applications. This rule is applicable to loading blocks where the sequence of loading is

not important. Klenke and Beste [54] applied the linear damage rule (LDR) to predict

rubber fatigue life of a rubber-metal mount. Their results show that the Miner’s linear

damage rule is a reliable tool for fatigue life predictions.

23
Sun et al. [55] studied the effect of step loading sequence on residual strength for

different natural rubber and styrene butadiene rubber (SBR) compounds and reported that

Miner’s rule is not applicable to multilevel loadings. They used two strain levels to

investigate the effect of loading sequence. The step-up sequence showed a higher crack

growth rate and larger cracks and lower tearing energy, compared to the step-down

sequence. The results also indicated that this effect is larger for longer blocks of load and

higher strain levels.

Flamm et al. [56] studied the effects of very high loads on fatigue life of

elastomeric materials made of natural rubber. By using a spectrum of two different

amplitudes, they concluded that Miner’s LDR rule is applicable and valid for damage

accumulation when the loading block contains few excessively high loads. They used

Lagrange strain as the damage parameter and concluded that the application of Miner’s

rule gives damage ratios of close to one. They also found that the combination of small

number of cycles of very high load amplitude and large number of low load amplitude

cycles could give reasonable values for damage sum close to optimum value

of 1.

In another study by Flamm et al. [57], it was shown that in many cases LDR is

suitable for use as a cumulative damage rule in rubber parts. They conclude that tests

with different sequences, mixed signals and signals resulting in a rotation of maximum

principal stress of almost 60° had little influence on the fatigue lifetime. For multiaxially

loaded rubber parts, Flamm et al. [40] propose discretizing the continuous signal by a

level crossing cycle counting method for each loading channel. Then they constituted the

24
stress amplitude history and performed Rainflow cycle counting on these alternating

points.

Harbour et al. [37] studied the effect of VA multiaxial loading on rubber

specimens made of two types of rubber materials, natural rubber and SBR. They

concluded that the VA loading prediction results by using Miner’s linear damage rule

gave reasonable predictions compared to the experiments for NR, while the predicted

results did not agree well for SBR, although the predicted values were still within a factor

of three of the experimental fatigue lives.

In another study, Harbour et al. [18] developed a linear crack growth model

equivalent to Miner’s linear damage rule. This model equates the crack growth rate for

variable amplitude loading to the sum of the constant amplitude crack growth rates for

each individual cycle. They studied the effects of R ratio, load level, load sequence and

dwell period on crack growth rates of planar tension specimens. They used the

Mars-Fatemi model [17] to accurately account for the effect of R ratio. They found that

changing R-ratio, load level and load sequence did not significantly affect crack growth

rates for repeated block test signals. Adding short dwell periods at the near zero stress

level between loading blocks produced faster crack growth rates in both NR and

especially SBR. This was explained in terms of the time-dependent recovery in the rubber

microstructure at the crack tip which could increase the localized stress state at that point.

Roland and Sobieski [58] studied some aspects of variable amplitude behaviour

with annealing periods in a strained state for rubber material. They used pre-cycles and

then annealed the specimens either with or without annealing strains for different designs

of natural rubber and synthetic variant rubber to evaluate the effect on fatigue life. They

25
concluded that annealing of polyisoprene-based elastomers can either improve or worsen

the deterioration properties based on the deformation type of the rubber during annealing

period. These changes were more severe under conventional fatigue loads.

Kim et al. [59] studied fatigue life prediction methodology of automotive rubber

components under variable amplitude loading. They performed displacement-controlled

fatigue testing and used maximum Green-Lagrange strain as a damage parameter. The

SAE transmission load history [60] was used in their study. Displacement-controlled

testing of dumbbell specimens provided the relationship between displacement and

fatigue life. They then used FE analysis results by utilizing Ogden strain energy density

function to relate the maximum Green-Lagrange strain at the critical component location

to displacement, and subsequently to fatigue life. Racetrack cycle counting method was

used to reduce the complex load history. The results showed that predicted fatigue lives

based on Miner’s rule were within a factor of two, compared to the experimental lives.

26
Chapter 3

Material Characterization and Fatigue

Behavior

3.1 Introduction

This chapter presents the experimental program for characterizing material

properties and fatigue behavior using simple geometry specimens under constant

amplitude loading. The material used and the method of making specimens as well as

testing equipments used for testing are described. Different test procedures for obtaining

constitutive behavior response and fatigue initiation life and crack growth properties are

explained in detail. Applicability of Miner’s cumulative damage rule is also investigated

by using variable amplitude loading. The experimental results obtained and prediction

methodology used for variable amplitude crack initiation experiments are discussed.

3.2 Material and specimens

A filled natural rubber material with 21% carbon black and 9.5% plasticizer was

used for specimen experiments. The compound used for specimen tests was identical to

that used for the cradle mount manufacture.

27
The stress states of simple tension and planar tension are often used to

characterize the deformation behavior. Simple tension specimens were cut by using

specimen cutting dies according to ASTM standard D4482-99 [61]. The test specimen

used in this study had an hourglass shape specimen with 1 mm thickness and dimensions

shown in Figure 3.1(a). Specimen preparation conditions are described in the

aforementioned ASTM standard. For planar tension deformation and crack growth tests,

a wide rectangular specimen of 1 mm thickness with dimensions shown in Figure 3.1(b)

was used. For crack growth specimen, a pre-crack is cut at mid-length, as shown in

Figure 3.1(b), but for cyclic deformation test no pre-crack was cut.

3.3 Testing equipment

An Instron closed-loop servo-hydraulic axial load frame, as shown in Figure 3.2,

in conjunction with a digital controller was used to conduct the experiments. The capacity

of the load cell was 5 kN.

3.4 Specimen test methods and procedures

3.4.1 Deformation behavior characterization

Monotonic and cyclic deformation curves properties are typically needed for FE

modeling and stress-strain relations for fatigue analysis. Phenomena associated with

elastomers, such as Mullin’s effect, can also be evaluated.

Extension ratio or stretch ratio (  ) is commonly used for finite strain analysis and

is defined as the ratio of the extended length (L) to the original length (L0):

28
L
 (3.1)
L0

The relationship between engineering strain (ε) and stretch ratio is given by:

   1 (3.2)

As the rubber is an incompressible material, there is no change in volume under

deformation, V  0 . This results in the determinant of the deformation gradient to be

one, which in turn results in the following relation for principal stretch ratios:

1 2 3  1 (3.3)

Figure 3.3 shows the stretch and stress states for each specimen condition under

uniaxial loading. For simple tension condition, uniaxial state of stress is present, with a

multiaxial stretch state with the longitudinal stretch value of  and two equal transverse

stretch values of -1/2 to satisfy the incompressibility condition. Planar tension specimen

is under plane stress condition with longitudinal and transverse stresses. The stretch in the

width dimension is 1, indicating there is no strain in this direction. A stretch of  in

loading direction and 1 exist in the out of plane direction. It should also be mentioned

that due to the incompressibility condition, Poisson’s ratio is 0.5 and the initial modulus

of elasticity (E) is approximately equal to three times the shear modulus (G), E = 3G.

3.4.2 Constant amplitude crack initiation testing

Cracks often nucleate from pre-existing flaws in the compound (i.e. such as

undispersed carbon black agglomerates), due to processing defects (such as

contaminations, voids, and molding flaws), or at stress concentrations. In the context of

mechanical design and fatigue analysis, crack initiation life is referred to the life involved

in growing a crack from the pre-existing flaw to a small macro-crack typically on the

29
order of 1 mm. Strain is commonly used for crack initiation life analysis as an ideal

fatigue damage quantification parameter, due to ease of determining displacements in

elastomers.

The specimen geometry used in crack initiation tests is the geometry shown in

Figure 3.1(a). This type of specimen is designed in such a way that when a crack grows

to a length on the order of 1 mm, there would be little remaining life to fracture.

Therefore, these tests characterize fatigue life for a crack on the order of 1 mm, although

the criterion for defining fatigue nucleation life in the test is the number of cycles in

which the specimen ruptures completely according to ASTM standard D 4482 [61].

Multiple specimens can be used for each test. The test setup with five specimens used in

this study is shown in Figure 3.4.

Tests were performed in displacement control and three strain R ratios of 0.02, 0.1

and 0.2 were chosen for testing. As the specimen cannot support compression load, these

Rε ratios represent tension-tension loading condition. For each Rε ratio, three levels of

maximum strain were tested. The testing frequency used for all conditions was 1 Hz,

except those with peak strain values higher than 2.75 which utilized testing frequencies of

0.5 Hz.

3.4.3 Crack growth testing

Crack growth testing of elastomers typically uses a pre-cracked planar tension

specimen, as shown in Figure 3.1(b) with a pre-crack length of 25 mm. While the original

application of fracture mechanics approach to rubber was to predict static strength, in the

late 1950s, Thomas extended the approach to analyze the growth of cracks under cyclic

loads in natural rubber [62]. Since rubber is a nonlinear elastic material, energy release

30
rate is often used as the crack driving parameter, rather than the stress intensity factor

which is typically used for linear elastic nominal material behavior, such as in metals.

This parameter relates the global specimen loading to the local stress and strain fields at

the crack tip, and is also applicable to finite strain materials, such as elastomers.

Energy release rate is the change in the stored mechanical energy dU, per unit

change in new crack surface area dA, also often called the tearing energy T, and given as:

dU
T  (3.4)
dA

Under cyclic loading, the maximum energy release rate achieved during a cycle is related

to the crack growth rate.

Because of a simple relationship between energy release rate and strain energy

density and the fact that energy release rate is independent of crack length in planar

tension specimen, this specimen is an ideal choice for fatigue crack growth experiments.

In the planar tension specimen, shown in Figure 3.1(b), energy release rate depends only

on the strain energy density (W) remote from the crack and specimen edges, and the

specimen height (h):

T  Wh (3.5)

This relation makes the planar tension specimen geometry quite practical for use in a

fatigue crack growth characterization of elastomers. In this specimen, a state of plane

stress under uniaxial tension is present, as shown in Figure 3.3.

The strain energy density (W) is the area under the loading stress-strain curve for

a stable cycle for each peak strain level. The energy release rate and crack growth rate are

independent of crack length in the planar tension specimen. Stress-strain hysteresis loops

from an uncracked test specimen can be used to obtain the test signal parameters for the

31
fatigue crack growth rate. Numerical integration methods (such as Simpson’s rule), as

explained in the ASTM standard D4482-99 [61], can be used to calculate the area under

the stress-strain curve from the experimental data.

For a constant displacement or strain, strain energy density is constant and so the

crack growth rate would be constant with repeated cycles of the same amplitude. For

non-fully relaxed conditions (R> 0), the peak strain energy density ( Wmax ) is the sum of

the area under the loading stress-strain curve (WL) plus the strain energy density at the

minimum strain level ( Wmin ):

Wmax .  WL  Wmin . (3.6)

Control parameters for crack growth test with planar tension specimen are the

amplitude and mean values of the applied displacement. Figure 3.5 presents the

maximum strain energy density as a function of maximum engineering strain obtained

from uncracked test specimens. Tests were performed with both fully relaxing conditions

( RW  0 ), as well as at three RW ratios of 0, 0.05, and 0.10, while changing peak strain

values from 40% to 65%, as listed in Table 3.1.

To pre-condition the material to avoid transient Mullin’s effect, stable

deformation curves were defined at the 128th cycle of planar tension uncracked specimen,

where the peak stress was 10% lower than the first cycle stress (see Figure 3.6(b)). For

the cracked specimen 110% of maximum strain was used as pre-conditioning load for

500 cycles at each strain level used to minimize the transient deformation response as

well as to produce a natural crack tip.

A single test specimen can produce results for multiple crack growth tests as long

as the crack length and remaining specimen length are sufficiently longer than the

32
specimen height of 16.83 mm. Due to the transient softening, however, it is necessary to

conduct tests on the same specimen in ascending order of maximum strain level to make

sure that the higher levels of load would not affect the lower load level test results. The

desired range of fatigue crack growth rates were between 10 6 and 10 3 mm/cycle. Using

a cycling frequency of 5 Hz, it takes about 5 hours to grow the crack 0.1 mm at a crack

growth rate of 10 6 mm/cycle.

A traveling microscope which can track the growth of the crack tip along the

crack line can be used to measure the crack growth. Crack growth data based on a

minimum interval of approximately 0.1 mm of crack growth was used for measurements.

After obtaining several data points for each test condition, the R ratio was changed by

increasing the minimum displacement. After all R ratio test data were obtained, the value

of peak strain was then increased to obtain crack growth rate at the next strain level. If the

crack grew irregularly or at an inclined angle, re-cutting was performed by a razor blade,

followed by some initial cycles to initiate a natural crack tip.

3.4.4 Variable amplitude crack initiation testing

The test specimen geometry used is the same as that used for constant amplitude

loading to characterize the material fatigue behavior with a thickness of 1 mm and

dimensions shown in Figure 3.1(a).

In order to investigate the effect of cumulative damage from variable amplitude

loading on fatigue crack initiation life, a series of tests on the simple tension specimen

geometry were performed with three peak load levels and three mean load ratios. This

helps to examine the applicability of linear damage rule under simple tension stress state

without the complication of multiaxial stress effects due to stress concentration at the

33
critical location of the component. A random load history [63] depicted in Figure 3.7(a)

was used. This load history consists of 16 individual events, and damage from each event

calculated based on Rainflow cycle counting method is tabulated in Figure 3.7(b). The

mean and amplitude values listed in the table are calculated using maximum and

minimum values of ±100. The desired maximum and minimum strain or stress values in

the block are reached by scaling the history. This random loading history was used for

both specimen tests in displacement control and component tests in load control.

3.5 Specimen experimental results and observations

3.5.1 Monotonic and cyclic deformation behavior results

To determine the stress-strain behavior of the material, tests were conducted in

displacement control. The applied displacement was then converted into gauge section

strain and the measured load from the load cell, which had a maximum capacity of 5 kN,

was used for stress calculation. The monotonic deformation curve is shown in Figure 3.8,

indicating a markedly non-linear behavior.

A cyclic incremental test on simple tension specimen in displacement control with

the sinusoidal waveform was performed to obtain the cyclic stress-strain curve. Due to

Mullin’s effect, stabilized cycle’s data is used for each strain level (i.e. after about 20

applied cycles) and all tests are performed in ascending order of displacement amplitude.

Figure 3.9(a) shows the resultant hysteresis loops, where the area between loading and

unloading curves represents energy dissipation during the cycle.

Hyperelastic constitutive models are used to represent the nonlinear elastic

deformation curve. A Fictitious curve for simple tension condition is obtained by using a

34
least squared fit polynomial to the end points of stress and strain from each loop [64].

This curve is depicted in Figure 3.8 and the mathematical representation of maximum

engineering stress ( S max ) versus maximum engineering strain ( emax ) for simple tension

condition is given by:

S max  0.25 emax


2
 1.63 emax (3.7)

The superimposed plot of monotonic and cyclic simple tension curves illustrate that using

monotonic properties in a cyclic loading application can underestimate the level of strain

which can be present. It is, therefore, important to obtain and use cyclic material

properties in fatigue life analysis and applications.

Cyclic incremental step tests were also conducted with the planar tension

specimen. Figure 3.9(b) shows the hysteresis loops from the 128th cycle from each

incremental step. The 128th cycle is used as the stable cycle, subsequent to initial

softening due to Mullin’s effect. As can be seen, after the initial straining the material

does not return to zero strain at zero stress due to some degree of permanent deformation.

A curve for planar tension condition is also depicted in Figure 3.8. The nominal stress-

strain relationship for cyclic planar tension test is given by:

2
S max  0.84 emax  2.10 emax (3.8)

According to Figure 3.8, cyclic simple tension deformation curve is less stiff than

cyclic planar tension at strain levels below 0.8. This is consistent with other deformation

behavior characterization studies done by Sharma [65] and Mars-Fatemi [11]. However,

at higher strain values the simple tension curve is stiffer than planar tension specimen

curve. This could be mainly due to the higher amount of strain crystallization in simple

tension, which in turn causes higher stiffness. In simple tension specimen there is no

35
lateral constraint to cause specimen thinning and therefore the polymer network chains

could become sufficiently aligned. The choice of the curve to use in FE simulations

depends on the stress state at the critical location of the component being analyzed.

The initial softening phenomenon, also called the pre-stretch or preconditioning

effect, is widely referred to as the Mullin’s effect [66]. Subsequent loadings of equal or

lesser magnitude rapidly tend towards a steady state, nonlinear elastic response. This

effect is considered to be a consequence of the breakage of links inside the material and

both filler-matrix and chain interaction links are involved in the phenomenon [67]. The

Mullin’s effect is transient and is exhibited primarily by filled rubbers [68].

Figure 3.6(a) provides an illustration of the Mullin’s effect as the stress level

drops for each successive loop. The stress-strain loop stabilizes after 3 to 30 cycles of

loading for most elastomers. For a higher maximum strain, the initial softening is larger,

as can be seen from Figure 3.6(b). This Figure also shows that by increasing peak strain,

it takes a longer time for the material to show the stabilized response. Therefore, due to

the load history dependence associated with the Mullin’s effect, peak loading should be a

key consideration in fatigue analysis of rubbers, in addition to the load amplitude and the

mean load.

3.5.2 Constant amplitude crack initiation test results

Test conditions and result from these tests are tabulated in Table 3.2. The ratio of

maximum fatigue life to minimum fatigue life in each testing condition ( N f , max N f , min )

which is an indicator of data scatter is also shown in this Table. This ratio for all crack

initiation tests was in the range of 1.11 and 2.58, which is quite reasonable for

elastomeric material fatigue tests.

36
A mean value for each strain ratio and maximum strain was calculated and used in

curve fitting and data analysis. The reason for using a mean life value for each test

condition rather than direct fit of all the data is so that there is equal contribution of each

test condition in test data analysis. Figure 3.10 shows raw data of maximum strain versus

crack nucleation life for all tests at different R rates. It is observed that the power trend

lines shown fit the crack initiation data well. The best-fit lines use the least squares fit

method with fatigue life as the dependent variable. The fatigue life equations based on

the different strain R ratios (denoted by R ) are given by:

 max  28.9 ( N f ) 0.30 for R  0.02 (3.9)

 max  17.4 ( N f ) 0.25 for R  0.10 (3.10)

 max  9.41 ( N f ) 0.14 for R  0.20 (3.11)

From Figure 3.10 it can be seen that the effect of Rε= 0.2 cycles is to increase the

nucleation life. The life improvement is very significant at low strain, about an order of

magnitude, but less important at high strain, less than a factor of two. This is in contrast

with metals where a tensile mean load has detrimental effect on fatigue life. This life

improvement in natural rubber is mainly due to strain crystallization. By applying a

nonzero minimum strain, rubber does not come back from crystalline state to amorphous

rubbery state. This crystalline state in which the polymer chains are aligned highly in the

loading direction increase resistance to crack growth, therefore fatigue life would be

longer.

37
The influence of strain ratio can be estimated by an empirical relationship by the

Mars-Fatemi model [17] that relates fatigue failure at a given life for R  0 conditions to

fatigue failure at the same life for R  0 condition by the following relation:

G( R) G( R)
 max,0   max,R G (0)  c [1 G (0) ] (3.12)

G( R)  3.5  258R 3 (3.13)

where  max, R is the maximum engineering strain at a given R ratio,  max,0 is the equivalent

maximum engineering strain at R  0 , G(R) is the power law exponent, and  c  6.25 .

Figure 3.11 shows correlation of all crack nucleation data on a single plot for all strain

ratios. The equivalent R = 0 maximum strain (  max,0 ) is then given by:

3
 max,0   max,R (174R )  6.25( 74R )
3
(3.14)

Once  max, 0 is determined for a given Rε loading condition, it is then used in

Equations (3.9)-(3.11) to calculate the fatigue crack initiation life.

3.5.3 Crack growth tests results

Fatigue crack growth rate (da/dN) is obtained by fitting a linear relationship to the

crack length versus cycles data and determining the slope of the linear fit. Crack growth

data fits at different energy release rate ratios and peak strain levels are shown in

Figure 3.12. The test conditions used for the constant amplitude fatigue crack growth

experiments produces crack growth rates in the region of crack growth that can be

characterized by a power-law relation. The coefficient and exponent of this relation are

calculated from RT  0 data as:

38
da
 4  10 5 (Tmax ) 2.0 (3.15)
dN

Note that for planar tension specimen, due to the relation given by Equation (3.5),

RT = RW . In strain crystallized rubbers, such as the natural rubber used in this study,

increasing the minimum strain or energy release rate ratio has a significant beneficial

effect on fatigue life compared to RT  0 condition, as fatigue crack growth rate

decreases. This behavior is depicted in Figure 3.13 where the fits to data show that

increasing R ratio causes slower crack growth rate and, therefore, being beneficial to

fatigue life. The power law equations for the two energy release rate ratios higher than

zero are obtained as:

da
 1  10 5 (Tmax ) 2.73 for RT  0.05 (3.16)
dN

da
 3  10 6 (Tmax )1.99 for RT  0.10 (3.17)
dN

By using Mars-Fatemi model [17], it is possible to reduce fatigue data taken at

various levels of R ratio to a single equivalent R = 0 curve. This is achieved by plotting

the crack growth rates against the equivalent R = 0 maximum energy release rate,

Tmax, 0. The Mars-Fatemi model is given by:

da T
 rc ( max ) F ( R ) (3.18)
dN Tc

where Tc  10 kJ/m2 is based on Lindley’s estimate for fatigue crack growth of natural

rubber [17]. The critical crack growth rate for this condition is defined as

rc  0.004 mm/cycle based on R = 0 crack growth Equation. The equivalent R = 0

maximum energy release rate ( Tmax,0 ) is given by:

39
F ( R) T ( R)
(1 )
Tmax,0  Tmax,R F ( 0) Tc T ( 0) (3.19)

where:

F ( R)  F0 e F4 RT (3.20)

This form of F(R) is used when there are limited data points and it adds just one variable

to be calculated to the material properties obtained from R = 0 analysis. Based on

Lindley’s model and using energy release rate ratios of 0 and 0.10 constants, the values

of F0 , obtained from R = 0 power law fit data for this material, and F4 are calculated as 2

and 8.27, respectively. Figure 3.14 shows correlation of all fatigue crack growth (FCG)

data on a single plot for different RT ratios by the use of Mars-Fatemi model.

3.5.4 Relationship between crack growth and crack initiation approaches

In a single edge cut simple tension specimen geometry the energy release rate

depends on the gauge section strain energy density W, crack length a, and a stretch

dependent parameter k [69]:

T = 2kWa (3.21)

By combining the power-law fatigue crack growth rate relation (Equation (3.15)) and the

above equation and then integrating the resulting equation, the following relationship is

obtained [13]:

1 1 1 1
Nf  [ F 1  F 1 ] (3.22)
F  1 B(2kW ) a0
F
af

where B and F are the fatigue power-law coefficient and exponent for R = 0 condition,

respectively. If the initial flaw size a0 is much smaller than the critical flaw size, a f , the

life becomes nearly independent of the critical flaw size:

40
1 1 1
Nf  (3.23)
F  1 B(2kW ) a0F 1
F

Effective flaw sizes in the range of 0.02 mm to 0.06 mm were observed in a study

by Lake and Lindley, which covered different polymer types, and various fillers,

curatives and other compounding variables [70]. In order to compare the relationship

between crack initiation and crack growth approaches, Equation (3.23) was used to obtain

fatigue life based on the crack growth approach and to compare with the fatigue life

based on the crack initiation approach. By assuming a0  0.02 mm, utilizing B  4  10 5

and F = 2 as crack growth rate coefficient and exponent, respectively, in Equation (3.23),

and by assuming k  2 (in the stretch range studied), relatively good agreement

(i.e. about a factor of two in fatigue life) between the R = 0 crack initiation life and crack

growth life is obtained, as shown in Figure 3.15. Therefore, the crack growth approach

could be used as a total life approach, based on growth of pre-existing flaws to failure.

Correlation between crack nucleation life obtained from simple tension specimen and

crack growth life obtained from planar tension specimen was studied by Mars and

Fatemi [68] and they also found good agreement between the results. It should be

mentioned that a change of initial assumed crack length from 0.02 mm to 0.04 mm

resulted in very small change in the obtained results.

If N f , represents the fatigue life associated with a long final crack size (i.e. a

crack size much longer than the initial crack size of 0.02 mm), the following relation

based on Equation (3.22) is obtained [71]:

F 1
Nf [1 a0F 1  1 a Ff 1 ] a 
  1  0  (3.24)
N f , 1 a0F 1 a 
 f 

41
The failure crack nucleation size used for this study is a f  1 mm. Therefore, for the initial

flaw size in the range observed by Lake and Lindley, it is estimated that the observed

initiation life is in the range of 94% to 98% of the life if crack grew in an infinitely wide

specimen. Thus, the crack initiation results obtained are nearly geometry independent.

3.5.5 Variable amplitude crack initiation tests analysis methodology

In order to be able to predict the fatigue damage, there are several general

prerequisites. First, it is necessary to identify the fatigue critical location and map the

load history for this location by using a proper cycle counting method. A fatigue damage

parameter is then needed to compute damage for the loading cycles identified by the

cycle counting method. Finally, by using a proper cumulative damage criterion the

damage is integrated and the fatigue life is calculated for the critical location of the

specimen or component.

For complex geometries, such as industrial components, a numerical method such

as FEA is also necessary. The component life prediction procedure can utilize crack

initiation and/or crack growth approaches. Constant amplitude specimen test data are

used as the basis of fatigue life curve for crack nucleation, while specimen crack growth

data and the initial natural flaw size in the material are used for crack growth or total

fatigue life of the component.

The objective of all cycle counting methods is to be able to compare the effect of

variable amplitude loading to fatigue data and properties obtained from constant

amplitude load cycles. Many cycle counting methods have been proposed, with the most

popular method being the Rainflow method proposed by Matsuishi and Endo [72].

Miner’s LDR [53] is the simplest form of cumulative damage rule. However, load

42
sequence effect or interaction between cycles is not accounted for in this rule [37].The

mathematical representation of Miner’s linear damage rule is as follows:

N N1 N N
N i   2  i  1
Nf1 Nf 2 Nfi
(3.25)
fi

where Ni is the number of applied cycles at a given load level and Nfi is the constant

amplitude fatigue initiation life at the same load level.

Continuum mechanics parameters such as different definitions of stretch or strain,

stress, and energy are used as damage quantification parameters. Cauchy stress,

maximum Green-Lagrange strain and strain energy density are some examples. Due to

the ease of measuring strain, maximum principal strain is used for life predictions in this

work. This agrees with the experimental observation of fatigue cracks usually initiating

and growing perpendicular to the maximum principal strain direction [31].

The strain R ratio (Rε) effect during each loading cycle at the fatigue critical

location can be accounted for by calculating a maximum equivalent R = 0 engineering

strain (ε1,max,0) based on the Mars-Fatemi R ratio material model [17] as discussed by

Equation (3.14). By using the material strain-life, as shown in Figure 3.11, represented by

the equation below, crack initiation life is then predicted at each equivalent strain level:

N f  6.86 10 4 (1,max, 0 ) 3.3 (3.26)

Crack initiation life based on blocks of loading is then calculated based on the linear

damage rule, as expressed by Equation (3.25).

3.5.6 Variable amplitude crack initiation tests results and analysis

Table 3.3 tabulates the variable amplitude fatigue crack initiation experimental

results for the simple tension specimens. While the first crack to traverse the specimen

43
cross section determines the experimental fatigue life, many cracks were visible on the

specimen surface, similar to that observed in constant amplitude loading in [68]. As can

be seen from this table, scatter of fatigue lives for each test condition is within a factor of

about two. Table 3.3 also shows the life prediction results. Rainflow cycle counting

method in conjunction with the linear damage rule was used for life predictions. Each

cycle’s damage was calculated based on constant amplitude specimen behavior by using

Equation (3.14) to account for strain R-ratio and Equation (3.26) for fatigue life.

Experimental fatigue life versus predicted fatigue life is depicted in Figure 3.16.

Based on the comparison of predicted and experimental fatigue crack initiation lives, it is

observed that most of the predictions (73%) fall within a scatter band of two and nearly

all the remaining experimental data fall within a predicted factor of five. Considering a

scatter factor of two in duplicate experimental data, this can be considered reasonable

prediction of the crack initiation life based on LDR. The effects associated with a rest

period between blocks of cyclic loading on the material fatigue behavior was investigated

by 33 seconds hold periods (dwell period) between load blocks in a R ~ 0 strain ratio test

with strain range of 3.14, with the results included in Table 3.3. As can be seen from

comparison of fatigue lives between tests with or without rest period, no detrimental

effect on fatigue life is observed. In variable amplitude crack growth behavior of another

filled natural rubber in [18], dwell periods effect, although not significant, increased the

fatigue crack growth rate by a factor of two.

The effect of test frequency or loading rate on the variable amplitude fatigue

crack initiation life was investigated by using a loading rate three times faster than the

original baseline test. The results are also included in Table 3.3. As can be observed,

44
fatigue life results at higher and at lower rates are nearly identical, indicating no

significant effect of test frequency on the fatigue life for the range of frequency

investigated.

3.6 Conclusions

A key ingredient of fatigue analysis and life prediction of elastomeric components

is relevant material properties. These properties include deformation behavior under

different stress states, crack initiation life, and crack growth rate properties. Specific

experimental procedures and data analysis techniques were presented and discussed to

obtain each of these material characterizations.

Two simple specimen geometries can be used for deformation and fatigue

behavior characterizations of elastomers: simple tension and planar tension specimens. In

the simple tension specimen a uniaxial state of stress with a multiaxial stretch state is

present, while the planar tension specimen is under plane stress condition with

longitudinal and transverse stresses. As monotonic and cyclic deformation behaviors can

be vastly different, it is important to obtain and use cyclic deformation properties in

fatigue life analysis and applications. Cyclic incremental tests on simple tension and/or

planar tension specimen can be performed to obtain the stabilized cyclic stress-strain

curve. The choice of the curve to use in FE simulations depends on the stress state at the

critical location of the component being analyzed.

Due to the load history dependence associated with the Mullin’s effect, peak

stress (or strain) should be a key consideration in fatigue analysis of elastomers, in

addition to the stress (or strain) amplitude and the mean stress. Fatigue crack initiation

tests can be performed with the simple tension specimen geometry in displacement

45
control and with different strain ratios (Rε ratios) representing tension-tension loading

conditions. Best fit lines in log-log scale, where the fatigue life is treated as the dependent

variable, represent maximum strain versus crack nucleation life for each strain ratio. The

effect of Rε> 0 is to increase the nucleation life and is significant at low strains. This is in

contrast to metals and this effect in natural rubber can be attributed to strain

crystallization.

Energy release rate is often used as the crack driving parameter in characterizing

elastomers fatigue crack growth behavior. A pre-cracked planar tension specimen is

typically used for fatigue crack growth tests since energy release rate is independent of

crack length for this specimen geometry. Therefore, a single specimen can produce

results for multiple crack growth tests with different loading conditions and RT ratios. A

power-law relation can be used to describe crack growth rates in terms of the maximum

energy release rate. The crack growth approach could be used as a total life approach,

based on growth of pre-existing flaws to failure.

The Mars-Fatemi model can be used to correlate test results from different R ratio

conditions. This model can be used to obtain an equivalent maximum strain in crack

initiation tests, or an equivalent maximum energy release rate in crack growth tests, to

account for the effect of R ratio.

Most of the life predictions for specimen random loading history tests based on

Rainflow cycle counting method in conjunction with the linear damage rule were within

the factor of two scatter bands. The R-ratio equation for fatigue life based on constant

amplitude material data was found to be applicable to random loading. The effect

associated with a rest period (dwell period) between blocks of cyclic loading on the

46
material fatigue behavior was found to be insignificant. Also, loading rate in the range

investigated did not have a significant effect on fatigue life in specimen fatigue crack

initiation tests.

47
Table 3.1: Constant amplitude fatigue crack growth test conditions

da/dN Tmax Freq.


min max R RW 2
(mm/cycle) (kJ/m ) (Hz)
0 0.40 0 0 2.82E-04 2.63 5
0 0.45 0 0 3.71E-04 3.13 5
0 0.50 0 0 4.87E-04 3.66 4
0 0.55 0 0 8.40E-04 4.23 3
0 0.60 0 0 9.76E-04 4.84 2.5
0 0.65 0 0 1.51E-03 5.49 2
0 0.40 0 0 3.60E-04 2.63 5
0 0.45 0 0 4.70E-04 3.13 5
0.04 0.45 0.09 0.05 1.77E-04 3.13 5
0 0.50 0 0 6.30E-04 3.66 4
0.05 0.50 0.09 0.05 4.70E-04 3.66 4
0.09 0.50 0.17 0.10 4.43E-05 3.66 4
0 0.55 0 0 8.86E-04 4.23 3
0.05 0.55 0.10 0.05 6.14E-04 4.23 3
0.10 0.55 0.18 0.10 5.48E-05 4.23 3
0 0.60 0 0 1.05E-03 4.84 2.5
0.06 0.60 0.10 0.05 6.93E-04 4.84 2.5
0.11 0.60 0.18 0.10 6.85E-05 4.84 2.5
0 0.65 0 0 1.19E-03 5.49 2
0.07 0.65 0.10 0.05 9.69E-04 5.49 2
0.12 0.65 0.19 0.10 1.01E-04 5.49 2

Table 3.2: Crack nucleation test conditions and results

Freq Number of Mean Nf Standard


min max R a Nf,max / Nf,min
(Hz) Tests (cycles) Deviation
0.02 1.00 0.49 1 4 68,528 12,145 1.41
0.04 2.00 0.02 0.98 1 5 6,682 275 1.11
0.06 3.06 1.5 0.5 5 1,704 666 2.58
0.09 0.95 0.43 1 5 103,940 6,904 1.18
0.16 1.60 0.10 0.72 1 10 18,321 3,295 2.03
0.33 3.33 1.49 0.5 5 712 32 1.11
0.36 1.80 0.72 1 5 107,966 30,989 2.37
0.55 2.75 0.20 1.1 0.5 5 18,641 4,095 1.79
0.75 3.75 1.5 0.5 5 976 201 1.75

48
Table 3.3: Simple tension specimen fatigue test conditions and results with the random loading history used.

Load block Number of blocks to failure, Bf


Predicted
time
εmin εmax ∆ε Rε blocks to
duration
Std. failure
(sec) Test #1 Test #2 Test #3 Test #4 Test #5 Mean Bfmax/Bfmin
Dev.

0.03 3.18 3.14 0.01 33 288 299 340 441 522 378 101 1.81 384

0.03 3.18 3.14 0.01 66 342 421 470 511 522 453 74 1.53 384
*
0.03 3.18 3.14 0.01 11 266 384 391 412 426 376 64 1.6 384

0.03 1.86 1.83 0.01 33 1,941 1,945 2,023 2,064 2,385 2,072 183 1.23 2379
0.22 3.36 3.14 0.06 33 88 89 90 102 118 97 13 1.34 401
0.22 3.36 3.14 0.06 33 74 94 98 100 100 93 11 1.35 401
0.15 2.29 2.14 0.06 33 1,279 1,364 1,412 1,767 1,863 1,537 260 1.46 1605
0.78 3.92 3.14 0.2 33 309 461 484 497 717 494 146 2.32 865


This load block time consists of 33 seconds of loading time block and 33 seconds of dwell period between each block at near zero strain.
*
This load block time is one third of normal block duration for high rate effect evaluation.

49
(a)

(b)

Figure 3.1: Specimen geometry and dimensions for (a) simple tension, and (b) planar
tension. Specimen thickness is 1 mm for both geometries.

50
Figure 3.2: Axial servo-hydraulic Instron frame used for specimen testing.

51
Figure 3.3: Simple tension and planar tension specimens with the corresponding stretch
and stress states [36].

52
Figure 3.4: Crack initiation test setup showing five specimens and grips.

0.45
y = 0.439x 2 + 0.2111x
0.4
R2 = 0.9983
0.35
Maximum SED (MPa)

0.3
0.25
0.2
0.15

0.1
0.05
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
Maximum Engineering Strain

Figure 3.5: Maximum strain energy density versus maximum engineering strain
obtained from uncracked planar tension specimen.

53
(a)

(b)

Figure 3.6: (a) Mullins effect showing initial transient softening in planar tension
specimen at 133% maximum strain, and (b) stabilization of stress with
applied cycles in displacement-controlled incremental step cyclic
deformation tests of planar tension specimen at different maximum strain
levels.

54
(a)

cycle max min range mean % Damage


2-3-2' 16 -78 94 -31 2.39
4-5-4' 53 -79 132 -13 6.62
7-8-7' 6 -91 97 -43 2.61
9-10-9' 56 -91 147 -18 9.20
11-12-11' 45 -89 134 -22 6.96
14-15-14' 67 6 60 36 0.51
13-16-13' 75 -91 165 -8 13.04
18-19-18' 62 -50 112 6 4.10
22-23-22' 7 -53 60 -23 0.51
6-17-6' 93 -96 189 -2 19.52
20-24-20' 48 -95 143 -23 8.44
25-26-25' 33 -41 74 -4 1.14
27-28-27' 21 -40 61 -10 0.55
29-30-29' 17 -64 82 -23 1.53
1-31-1' 100 -99 199 0 22.71
32-33-32' 100 53 46 76 0.17
100.00
(b)

Figure 3.7: (a) Loading history used for random loading tests of simple tension
specimens in displacement control and of cradle mounts in load control, and
(b) Rainflow cycle count for max/min values of ±100 with relative damage
distribution from each cycle [63].

55
Figure 3.8: Superimposed plot of monotonic and stable cyclic curves for simple tension
and planar tension specimens.

56
2.5

Engineering Stress (MPa) 2

1.5

0.5

0
0 0.5 1 1.5 2 2.5

-0.5
Engineering Strain

(a)

1.6

1.4

1.2
Engineering Stress (MPa)

0.8

0.6

0.4

0.2

0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6
Engineering Strain

(b)

Figure 3.9: Stable cyclic stress-strain loops from incremental step tests at different peak
strain levels from (a) simple tension test, and (b) planar tension test.

57
10.0

Maximum Engineering Strain

1.0

Re=0.02
R = 0.02
ε

Rε = 0.10
Re=0.10

Rε = 0.20
Re=0.20

Power
(Re=0.02)
Power
(Re=0.10)
Power
(Re=0.20)
0.1
100 1000 10000 100000 1000000

Crack Nucleation Life (N f), cycles

Figure 3.10: Crack nucleation life as a function of peak strain in simple tension tests.

58
Figure 3.11: Equivalent fatigue life as a function of equivalent maximum strain by using
Mars-Fatemi R-ratio model.

59
(a) (b)

(c) (d)

Figure 3.12: Crack length versus cycle linear fits from crack growth test at (a) RT  0 ,
(b) RT  0 , (c) RT  0.05 , and (d) RT  0.10 .

60
1.00E-02

RRT
T== 00

RRT
T= 0.05
= 0.05

RRT = 0.10
T = 0.10
RRT
T=0
=0

Power (RT = 0)

Power (RT =
1.00E-03
Crack Growth Rate (mm/Cycle)

0.05)
Power (RT =
0.10)

1.00E-04

1.00E-05
1 10
2
Maximum Energy Release Rate, T max ,(kJ/m )

Figure 3.13: Fatigue crack growth rate data comparisons at different R-ratios.

61
1.00E-02

RT = 0.05
RT=0.05

RT = 0.10
RT=0.10

RT = 0
RT=0
RT =specimen1
RT=0, 0
-5 2
da/dN = 4×10 (Tmax,0 )
1.00E-03
Crack Growth Rate (mm/cycle)

1.00E-04

1.00E-05
0.1 1 10
2
Maximum Equivalent Energy Release Rate, T max,0 ,(kJ/m )

Figure 3.14: Fatigue crack growth rate data correlations at different R-ratios based on
Mars-Fatemi R-ratio model.

62
Figure 3.15: Comparison of fatigue lives obtained from crack initiation and crack growth
approaches.

63
Figure 3.16: Experimental versus predicted fatigue lives in blocks to failure for simple
tension specimen tests subjected to variable load history.

64
Chapter 4

Component Finite Element Model and

Results

4.1 Introduction

Because the measurement or analytical calculation of stress and strain for the

complex mount geometry is difficult, a numerical method such as finite element method

must be used to compute them. FE methods are commonly used tools for simulating the

actual service performance of the components at their design and development stage.

The nonlinear FE program ABAQUS was used to simulate the nonlinear and large

deformation behavior of the mount. This section discusses the relevant details and aspects

of the FE simulations including material deformation characterization, FE model

construction, and the obtained results from the FE model. Numerical results which are

obtained from FE simulations will then be used in fatigue life predictions.

4.2 Hyperelastic material deformation characterization

There are some common approaches in FE software to deal with hyper-elastic

materials. Rubber can be considered as a hyper-elastic material, showing highly nonlinear

65
elastic isotropic behavior with incompressibility. The relationship between stress and

strain in a hyper-elastic material is generally characterized by the strain energy potential.

Strain energy potential defines the strain energy stored at a point as a function of the

strain at that point assuming the material is isotropic and homogeneous.

There are two general approaches for isothermal rubber mechanical behavior, the

kinetic theory which is based on statistical thermodynamics considerations, and the

phenomenological approach which treats the material as a continuum regardless of its

micro-structural and molecular nature [3]. The kinetic energy approach is based on

statistical distribution of rubber molecular chains and assumes rubber elastic deformation

is originated from the decrease in entropy resulting from the increase in applied

deformation. For example, Arruda and Boyce proposed a three-dimensional model of

eight chain network [73]. The phenomenological approach assumes that the elastic

properties can be described in terms of a strain energy function W which is either defined

as a polynomial function of strain invariants, or in terms of principal stretch ratios.

Rivlin [74] proposed that the strain energy density function can be approximated

by a power series in terms of strain invariants, which for an incompressible material it

reduces to:


W c
i  j 1
ij ( I 1  3) i ( I 2  3) j (4.1)

where cij are material constants and I1 and I2 are the first and second strain invariants

given by:

1 1 1
I 1  12  22  , I 2  (12 ) 2   (4.2)
2
1
2
2 
2
1 22

66
and 1, and 2 are principal stretches or extension ratios. Note that I3 = (123)2 is the

square of the ratio between the volumes of a material element in the deformed and

undeformed states and for an incompressible material 123= 1 and, therefore, I 3  1 .

By using the first term of Equation 4.1, the well-known Neo-Hookean model is generated

and by using both terms the Mooney-Rivlin form is generated [75].

Ogden [76] presented the strain energy density function based on a polynomial of

principal stretches, which for incompressible materials is expressed as:

n
2 i
W  (1 i  2 i  3 i ) (4.3)
i 1  i
2

where μi and αi are constants to be determined from experimental data. It should be noted

that the aforementioned strain energy density functions assume the material behaves as a

perfectly elastic material. This is an idealization of filled rubber behavior since such

rubbers exhibit some degree of hysteresis, stress softening, and plastic deformation upon

unloading.

Of the several strain energy potential forms available in ABAQUS to model

incompressible isotropic elastomers, the Marlow form was used in this study. In this case

strain energy potential is constructed in a way to reproduce the test data exactly for a

given deformation state, for example simple tension, and have reasonable behavior in

other deformation states such as planar tension. In order to minimize undesirable

nonlinearity, sufficient data points should be specified in the low, intermediate, and high

strain ranges. The form of the Marlow strain energy potential which is based on

deviatoric part of strain energy for an incompressible material is given by [75]:

U  U dev ( I1 )  U vol ( J el ) (4.4)

67
where U is the strain energy per unit of reference volume with U dev as its deviatoric part

and U vol as its volumetric part and J el is elastic volume ratio. The deviatoric part of the

potential is defined by providing test data, while the volumetric part is not considered due

to incompressibility [75].

The Marlow model does not assume any explicit form [77], as this function is

calculated based on integrating the stress-strain curve up to a certain strain ε1 which is a

function of the first strain invariant, ε1 = 1(I1) – 1. It should also be noted that there is no

curve-fit involved. If simple tension, planar tension and equi-biaxial tension data are all

available, then the Ogden or Van-der-Waals [78] forms are more accurate in fitting the

experimental results [75].

Strain in this study is defined as the nominal strain. Nominal strain  can be

expressed in terms of the principal stretch ratios as [75]:

3
  V  I   (i  1) ni ni T (4.5)
i 1

where V  F F T is the left stretch tensor, F is the deformation gradient, and ni’s are the

principal stretch directions in the deformed shape. Principal values of nominal strain

show direct values of deformation and could be used in life predictions due to the fact

that these values are the ratios of change in length to length in the initial shape. These

strain values show the importance of deformation gradient in analyzing rubber-like

material continuum mechanics terms.

68
4.3 Uniaxial FE model definition and specifications

The FE model consists of half of the actual mount due to the symmetry of the

mount geometry and uniaxial loading applied in a symmetry plane. The mount and the

FE mesh are shown in Figure 4.1. To adequately model the half-mount geometry, a fine

mesh with 14,355 elements was used. Load and/or displacement are applied to the

reference point which is coupled with inner FE nodes of the mount at the interface with

metallic part by utilizing the rigid body constraint capability of the software. In load

control simulations half of the testing load was applied to the model due to the model

symmetry. To prevent rigid body motion the outer portion of the mount was fixed.

Proper element type, integration scheme and reasonable meshing strategy are all

important factors to consider for appropriate modeling and simulation of large

deformation rubber components. The fully hybrid formulation elements are

recommended for use with hyper-elastic materials [75]. Wang et al. [41] showed that

hybrid elements with full integration and lower-order interpolation show less distortion

than higher order reduced integration elements and are suitable for large deformation

simulation computations. ABAQUS/Standard has a special family of hybrid elements that

could be used to model the fully incompressible behavior. The element type used in this

work was the eight node linear brick (hexahedral) element.

Hybrid element locking is another key issue in the large deformation simulation

computations of 3-D models, which is caused by large element distortion. Element

locking can be lessened by proper meshing and element selections. Too fine a mesh

exhibits more sensitivity to element volumetric locking, especially in large strain areas.

The mesh verification showed no error in element meshing. This tool verifies elements

69
aspect ratios, and larger or smaller face corner angles to assure there would be no

inappropriate element distortion. Therefore, the mesh quality was quite good for

performing FE analysis.

For judging the quality of fit to experimental data a Drucker stability check is

performed in the software [75] to assure the tangential material stiffness is positive

definite. According to this criterion, the deformation of the body with

time-independent properties under isothermal conditions is stable provided that the work

done by infinitesimal increments of generalized forces associated with infinitesimal

increments of displacements is positive [79].

Quasi-static simulation was performed for constant amplitude axial loading. For

cyclic loading, the amplitude tool of the FE software was used, so that a loading cycle

could be exactly simulated. The stress–strain relationship of the natural rubber showed

significant softening during initial straining cycles but stabilized after almost 20 cycles.

This softening is associated with the Mullin’s effect. Therefore, a stabilized cyclic

stress–strain response was used for determining the cyclic material properties. The stable

loops at different strain levels are shown in Figure 3.9(a). Due to the application of stable

deformation curve, initial softening (Mullin’s effect) is inherently considered. Primary

Hyper-elastic curve is constructed through fictitious curve points. This curve is obtained

from peak strain and peak stress points of the different stabilized hysteresis loops [64], as

shown in Figure 3.8.

To evaluate mesh sensitivity, a refined mesh with 114,840 elements was utilized.

Although the results changed by 19% for strain at the critical location for 550 N load,

critical locations were the same for both analyses. The run time, memory usage and

70
output database file size for the refined mesh, however, increased 34, 11, and 8 times,

respectively. In addition, the possibility of element locking at large strain area and under

high load increases with the refined mesh. Therefore, it was decided to use the original

mesh from a practical application viewpoint.

In order to evaluate the displacement versus load input effect on simulation results

two sets of simulation are performed, under load control and under displacement control.

The displacement-controlled simulation used the displacements obtained from the

load-controlled simulation. As expected, load, displacement, maximum principal stress,

and strain values between the two analyses were identical.

FE simulations assume non-linear perfectly elastic constitutive behavior,

therefore, the same deformation curve for loading and unloading. This is not

unreasonable with regards to actual component behavior obtained from actual component

fatigue testing at midlife. The hysteresis loops shown in Figure 4.2 for two R ratios

(i.e. minimum to maximum load ratios) and several load levels all indicate small amount

of hysteresis during a stabilized cycle. Brief description of the model and the parameters

used are summarized in Table 4.1.

4.4 Stiffness test results comparison with predictions

A displacement-controlled monotonic compression test of the mount was

conducted to compare with the FEA results. For FE simulation of this test condition

monotonic stress-strain curve was used. Also, 30 pre-cycles were applied at the

intermediate displacement level to account for initial softening before the monotonic test.

Figure 4.3(a) shows stiffness comparison between the FEA and the test results indicating

reasonable agreement between results of less than 20%.

71
Incremental step displacement-controlled cyclic tests at 1 Hz and 3 Hz with

R ~ 0 were also performed on the cradle mount to compare with FEA results. In order to

avoid Mullin’s effects, 50 pre-cycles at intermediate displacement levels were applied

prior to each displacement amplitude increment. At each displacement level 260 cycles

were applied. The data from the last cycle are assumed to represent the stabilized

behavior for that test level. FE simulations were based on the stabilized cyclic material

behavior from specimen simple tension tests and SED function of Marlow, as described

earlier. Figure 4.3(b) shows the comparison between the simulation and test results. As

can be seen, the experimental cyclic curves agree well with FEA predictions.

4.5 Uniaxial constant amplitude FE simulation results

The numerical simulation results show that multiaxial states of stress and strain,

although proportional, are present at the critical location under axial load. Figure 4.4

shows the strain and stress history for R = 0.2 and load amplitude of 1,500 N during a

sinusoidal load cycle. As can be seen, one normal and one shear component of strain and

stress have the largest values compared to the other components. By examining the

principal strain and stress histories of the critical element during a cycle, the observed

states of stress and strain result in the mid and minimum principal stress and strain values

to be small.

Maximum principal strain distributions of an engine mount under R = 0.2 and

load amplitude of 1,500 N is shown in Figure 4.5. The maximum principal strain occurs

at the interface between the inner bushing and the bulk rubber material at a location of

abrupt geometry change. This region is the critical region where fatigue cracks are

72
expected to nucleate. The maximum principal strain at the critical location determined

from FEA will be used for evaluating the fatigue damage parameter of the material.

4.6 Conclusions

This chapter presented a generalized fatigue FE simulation analysis for rubber

components. Application of the approach was illustrated by analysis of a vehicle cradle

mount and verified by means of component stiffness tests. As it was discussed, quasi-

static simulations can be performed for cyclic loading where by using the stabilized

cyclic stress–strain curve the initial softening associated with the Mullin’s effect is

inherently accounted for. Due to large deformations, strain may be considered a natural

choice to characterize the fatigue behavior of rubber. Among the strain quantities,

maximum principal strain can be used for predictions since fatigue cracks in rubber

typically initiate and grow perpendicular to the maximum principal strain directions.

The R ratio can vary at different critical locations of the component even under

constant amplitude loading. As the R ratio can significantly influence fatigue life, its

effect can be taken into account by using a model such as the Mars-Fatemi R ratio model.

Simulation of component deformation and load-deflection curves under monotonic as

well as cyclic loadings agreed well with the component test results under the

corresponding loading condition.

73
Table 4.1: Summary of parameters used for numerical model.

Parameter Conditions
Material Behavior Elastic, Isotropic
Poisson’s Ratio ~0.5, Incompressible
Material Property Used Stabilized Cyclic Deformation Curve
Strain Energy Potential Model Marlow
Element Type Hybrid Continuum 8-Node Brick
Simulation Type Quasi-Static Implicit
Simulation Control Mode Uniaxial Load

74
(a)

(b)

Figure 4.1: Vehicle cradle mount used as illustrative example (a), and rubber mount
FE model where due to symmetry half of the model is shown (b).

75
(a)

(b)

Figure 4.2: Component mid-life hysteresis loops under axial loading condition
for (a) R ~ 0 and, (b) R = 0.2.

76
(a)

(b)

Figure 4.3: Stiffness comparison between FEA simulation and monotonic test of the
mount (a) and cyclic tests of the mount (b).

77
(a)

(b)

Figure 4.4: Strain (a) and stress (b) histories at the critical element location obtained
from FE simulation during one sinusoidal loading cycle between 750 N
and 3750 N.

78
Figure 4.5: Maximum principal strain distribution under R =0.2 and load amplitude of
1,500N.

79
Chapter 5

Component Uniaxial Fatigue Behavior

5.1 Introduction

The main objective of this chapter is developing and validating CAE durability

analytical techniques under constant and variable amplitude uniaxial loading. Application

of the approach is illustrated by analysis of a vehicle cradle mount. The approach is then

verified by component test results. The capabilities of Rainflow cycle counting procedure

and Miner’s linear damage rule are also evaluated with component tests. A filled natural

rubber material with 21% carbon black and 9.5% plasticizer was used for component

experiments.

In this chapter, first the component life prediction methodology and predictions

involving cycle counting, cumulative damage, and damage parameter are presented and

discussed, where both crack initiation and crack growth approaches are considered.

Finally, verification of the approach used is presented based on the conducted fatigue

tests of both constant amplitude and variable amplitude loadings. Validation of life

prediction methodology used is discussed with respect to both failure location and fatigue

life.

80
5.2 Component life predictions

In order to be able to predict the fatigue damage, there are several general

prerequisites. First, it is necessary to identify the fatigue critical location and map the

load history for this location by using a proper cycle counting method. A fatigue damage

parameter is then needed to compute damage for the loading cycles identified by the

cycle counting method. Finally, by using a proper cumulative damage criterion the

damage is integrated and the fatigue life is calculated for the critical location of the

specimen or component.

For complex geometries, such as industrial components, a numerical method such

as FEA is also necessary. The component life prediction procedure developed is based on

both crack initiation and crack growth approaches. Constant amplitude specimen test data

are used as the basis of fatigue life curve for crack nucleation, while specimen crack

growth data and the initial natural flaw size in the material are used for crack growth or

total fatigue life of the component. It should be noted, however, that the load levels used

for component testing as listed in Table 5.1 are higher than those applied to the

component in actual service condition. This was necessary in order to induce fatigue

failure and in a reasonable time duration (i.e. hours and days, rather than months or

years).

5.2.1 Crack initiation life predictions

There are generally two methods to characterize the fatigue behavior of the

material. One approach is based on nominal stress amplitude (S–N equation and curve).

Another approach is based on local strain amplitude (ε–N equation and curve). Because

strain is readily calculated from deformation and is relatively easy to measure due to

81
large deformations, the ε–N equation and curve may be considered a natural choice to

characterize the fatigue behavior of rubber. Among the strain quantities, maximum

principal strain is used for predictions. This is because fatigue cracks in rubber typically

initiate and grow perpendicular to the maximum principal strain directions [31].

Maximum principal strain at the critical location is obtained from FEA; thereby

local strain R ratio during the loading cycle can be obtained at this location. Maximum

equivalent R = 0 engineering strain is then calculated by using the Mars-Fatemi R ratio

material model [17] as discussed in chapter 3 and given by Equation (3.14). Predictions

for failure at critical location based on crack initiation approach are then made based on

the material strain-life curve shown in Figure 3.11 and represented by Equation (3.26).

Crack initiation lives of the rubber mount at different loads were then obtained by

substituting the maximum equivalent principal strains at the critical location into this

equation. Values of Rε, ε1,max, ε1,max,0, and predicted life Nf for different component tests

conducted are tabulated in Table 5.2.

5.2.2 Crack growth-based life predictions

Maximum principal strain is also used for crack growth-based fatigue life

predictions, where crack growth life is used as fatigue life. Maximum principal strain and

R ratio at 10 elements along the crack path adjacent to the critical location element was

obtained from FEA. Maximum equivalent R = 0 engineering strain could then be

calculated for each element by using the Mars-Fatemi R ratio material model based on

Equation (3.14).

Mars and Fatemi [80] used the energy release rate estimate for small edge crack

under simple tension condition which is given by [27]:

82
Tw  2kWa (5.1)

where W is the strain energy density, k is a factor dependent on maximum principal strain

[69], and a is the crack length. Strain energy release rate (pseudo-energy release rate

equation) is then obtained by observing the fact that energy density varies approximately

with the square of the maximum principal strain, given by [80]:

T ,max,0  C ( 1,max,0 ) 2 a (5.2)

where C is the material’s initial Young’s modulus obtained from monotonic stress-strain

curve and is 2.34 MPa for the material in this investigation and a is crack length. It

should be mentioned that the constant C defined as the initial Young’s modulus value is

not strictly consistent with the classical form of the strain energy release rate for simple

tension frequently used for rubber strength and fatigue analyses. However, Equation (5.2)

still represents the important characteristic of energy release rate varying linearly with

crack length. Recent work by Ait-Bachir et al. [81] validates the form of Equation (5.2)

by showing energy release rate to depend on crack length and one far-field property,

irrespective of the state of loading. Crack growth rate is then obtained based on specimen

crack growth rate data which is shown in Figure 3.14 and represented by:

da
 4  10 5 (Tmax, 0 ) 2  4  10 5 C 2 a 2 14,max, 0 (5.3)
dN

By using the maximum equivalent R = 0 engineering strain as a damage

parameter along the crack, which is in the depth direction for the component considered,

the relationship between strain and crack length (depth) was obtained. Integrating

Equation (5.3), fatigue life can then be obtained. Based on Lake and Lindley [82], the

natural flaw size for 8 different polymer types and with various fillers, curatives and

83
other compounding variables were reported to be around 0.02 to 0.05 mm. Mars and

Fatemi [68] obtained an initial crack size of 0.01 mm for natural rubber based on

agreement between crack initiation and crack growth properties. Based on this approach,

for the natural rubber material of the component used in this study the effective initial

flaw size was found to be 0.02 mm. This value was, therefore, used as an initial crack

length at the critical location in this study. Predictions based on integration results along

the critical dominating crack path are finally obtained from:

10
4566  1 1 
Nf  4    (5.4)
i 1  i ,1,max, 0  ai ai 1 

The variation of maximum equivalent R = 0 principal strain versus crack length (depth) is

shown in Figure 5.1. This figure shows the importance of initial flaw size in fatigue life

prediction, due to the drastic decrease of the value of 1,max, 0 away from this critical

element. Therefore, the critical element has the highest effect in total life prediction. The

total fatigue lives of the rubber mounts were predicted using this procedure and the

predicted lives for the four load levels and two R ratios considered are tabulated in

Table 5.2.

5.2.3 Variable amplitude loading life predictions

As discussed in section 3.5.5, many cycle counting methods have been proposed.

Miner’s LDR [53] is the simplest form of cumulative damage rule. Continuum mechanics

parameters are used as damage quantification parameters. Due to the ease of measuring

strain, maximum principal strain is used for life predictions in this work. This agrees with

the experimental observation of fatigue cracks usually initiating and growing

perpendicular to the maximum principal strain direction [31].

84
The strain R ratio (Rε) effect during each loading cycle at the fatigue critical

location can be accounted for by calculating a maximum equivalent R = 0 engineering

strain (ε1,max,0) based on the Mars-Fatemi R ratio material model (Equation (3.14)).

By using the material strain-life, as given in Equation (3.26) crack initiation life is

then predicted at each equivalent strain level. Crack initiation life based on blocks of

loading is then calculated based on the linear damage rule, as expressed by

Equation (3.25).

By utilizing maximum equivalent R = 0 engineering strain as a damage parameter

along the crack (depth direction for this component), the relationship between strain and

crack length (depth) was obtained. Predictions based on the damage parameter along the

critical dominating crack element for each individual event is obtained from:

4566  1 1
Nf  4    (5.5)
 1,max, 0  a0 a1 

where a0 and a1 are natural flaw size and critical element’s depth, respectively. It should

be noted that the unit of crack length in Equations (5.2) and (5.5) is mm. It is found that

the critical element has the highest effect in the total life prediction. Total fatigue lives of

the rubber mounts were predicted by using the linear damage rule based on blocks of

loading shown in Figure 3.7.

85
5.3 Constant amplitude fatigue experimental program and validation

of life predictions

5.3.1 Component fatigue tests

Component level fatigue tests were conducted in order to validate the accuracy of

the fatigue life predictions discussed. Fatigue tests were conducted at ambient

temperature under constant amplitude load-controlled conditions with a sinusoidal

waveform of 0.5 to 1 Hz using a servo-hydraulic fatigue testing load frame with the load

capacity of 50 kN. Displacement response of each test specimen was periodically

recorded. In order to define the fatigue crack nucleation life of the engine mount in a

consistent manner, cracking of several engine mounts was also monitored.

The load amplitudes chosen resulted in fatigue lives between about 19,000 cycles

and 161,500 cycles. With initial cycling, the maximum displacement increased due to the

Mullin’s effect. The load-displacement response of the mount then stabilized after several

hundred cycles. When the fatigue crack initiated and then grew to a critical size, the

displacement amplitude increased rapidly due to reduced stiffness and failure occurred.

Failure was defined as the number of cycles at which the displacement amplitude

increased drastically from its gradual linear change, as illustrated for a typical test in

Figure 5.2. The experimental fatigue lives at different load levels are tabulated in

Table 5.1.

As can be observed from Table 5.1, experimental life scatter is within a factor of

two between duplicate tests. This table also shows the first crack observations for the

tests which were stopped at regular intervals for monitoring. As can be observed, the

crack growth life is a significant portion of the total life in all tests. The experimental
86
results show that by increasing the mean load or R ratio, the total component life

increases. This is the same as what was observed for specimen testing of this material.

Displacement amplitude versus applied cycles for all of the fatigue experiments is

depicted in semi-logarithmic scale in Figure 5.3. As can be seen, near failure there is a

steep change in displacement amplitude due to reduced stiffness resulting from presence

of macro-crack(s).

In order to preliminary investigation of the effect of higher frequency on the

component life, one test with the loading rate of 3 Hz was performed to compare with the

test at 1 Hz. The fatigue life was similar to the lives at 1 Hz. Frequency effect has been

observed to be small for rubber compounds which strain crystallize under loading

(i.e. filled natural rubber) under isothermal conditions for the frequency range of 10 -3 to

50 Hz [83]. Young [84] also reported that while the strain rate could affect the crack

growth rate for natural rubber compounds at high energy release rates close to fracture,

for smaller values of energy release rate changing the strain rate by factor of 20 does not

significantly change the crack growth rate.

To evaluate the difference between load control (LC) and displacement control

(DC) on component fatigue life, one axial test in DC was also performed based on

midlife response displacement amplitude and R ratio of the corresponding LC test. The

crack grew more rapidly under load control than under displacement control, as also

observed in [85]. DC test total life was 61,450 cycles, compared to 69,683 and 107,411

cycles in LC tests (see Table 5.1)

87
5.3.2 Damage development

Crack growth life was a significant portion of total fatigue life (between 70% and

93%). For R ~ 0 tests at the two load levels the ratio of crack initiation life to total life

decreases compared to R = 0.2, but still exceeds 70% of total component life. For R = 0.2

ratio and load amplitudes of 1,100N and 1,500 N, this ratio increases to about 78% and

93%, respectively.

Crack length as well as its depth and their changes were measured at critical

(failure) location by periodic test interruptions and visual inspection. Crack length did not

have a significant effect on fatigue life, as even at a length of about 4 cm, the component

is still at only about half of its total life. Crack depth had a more dominant effect on life

than crack length. When crack depth was on the order of 3 cm, the component was near

failure. Stiffness drop also correlated better with crack depth than with crack length.

5.3.3 Comparison of predicted and experimental fatigue lives

Finite element contours at the maximum load are shown in Figure 4.5. The value

of highest maximum principal strain defines the crack initiation location. The same

critical element was observed in all of the FE simulations. Experiments failure locations

and crack growth directions are schematically depicted in the Figure 5.4. The

experiments show the same location as predicted by FE simulations.

Experimental and predicted initiation lives are tabulated in the Table 5.2. The

results show life predictions based on crack initiation are within about a factor of two of

experimental lives. The crack size used for both experimental and predicted nucleation

lives was 1 mm, as the specimen fatigue initiation life data equation was generated for a

crack length of about 1 mm.

88
Figure 5.5(a) shows crack initiation life prediction versus experimental life.

Initiation life for the tests not monitored for crack growth is calculated based on crack

initiation to total life ratio for each loading condition. It can be seen that most of the

predictions are within a factor of two and the predictions for R ~ 0 loading were closer to

the experimental results. This may be mainly due to the fact that the strain-life equation

was developed for R ~ 0 condition and an R ratio model was used to find the equivalent

R ~ 0 strain for the R = 0.2 condition.

Figure 5.5(b) shows predictions based on total life (crack growth approach) as

compared to experimental life. This figure illustrates that for R ~ 0 loading condition,

similar to the crack initiation approach, predictions are better than for R = 0.2 loading.

The predictions for total component life are within a factor of two and on the

conservative side for most of the loading conditions.

5.4 Variable amplitude fatigue experimental program and validation

of life predictions

5.4.1 Component fatigue tests

To evaluate variable amplitude loading damage accumulation and life prediction

methodology, the rubber component shown in Figure 4.1(a) was used. Quasi-static FE

simulation was performed for the variable amplitude loading shown in Figure 3.7, using

ABAQUS software, so that a loading block was exactly simulated. Because the critical

element location is the site for crack initiation, maximum principal strain and R ratio at

the integration point of this element (therefore, the averaged value in the element) were

obtained from FEA for each individual cycle.

89
Four component fatigue tests were conducted in order to investigate the

applicability of fatigue life prediction methodology discussed. Load R ratio of near zero

with two load ranges of 3,000 N and 3,600 N were used and each test was duplicated to

evaluate variability. The fatigue test results are presented in Table 5.3. A sinusoidal

waveform with 16.5 second duration per block was used and tests were conducted using a

servo-hydraulic fatigue testing machine with load capacity of 50 kN. Displacement

response of each test specimen was recorded based on the logarithmic intervals of 2. In

order to define the crack initiation life of the mount in a consistent manner, damage

evolution of two of the mounts was monitored by periodic visual inspection. For the tests

in which crack initiation life was not monitored, nucleation life was estimated based on

the ratio of the average initiation life to total life of the tests for which crack initiation

was monitored.

After application of the initial few blocks of load, the maximum displacement

increased because of Mullin’s effect. The response then stabilized after several blocks.

After fatigue crack initiation occurred, the displacement amplitude gradually increased

with macro-crack growth because of reduced stiffness. Once the crack grew to a critical

size, the displacement amplitude increased rapidly leading to failure. Failure was defined

as the number of blocks at which the displacement amplitude increased abruptly from its

gradual linear change, as shown in Figure 5.6 for a typical test.

90
5.4.2 Damage development

As can be seen in Table 5.3, experimental life scatter was within a factor of two

between duplicate tests. This table also tabulates the first crack observations for the tests

which were stopped at regular intervals for inspection. By increasing the load range, the

total life decreases, as expected. By increasing the load range, the ratio of crack initiation

life to total life decreases. This means that more life is spent in crack growth. Crack

growth life is a significant portion of the total fatigue life (about 80% to 90%). Crack

length and its growth were measured at critical (failure) location by periodic test

interruptions and visual inspection. Crack length did not have a significant effect on

fatigue life, while crack depth had a more dominant effect and this fact is illustrated in

Figure 5.7.

5.4.3 Comparison of predicted and experimental fatigue lives

Maximum principal strain history at the critical element was generated from FEA.

By using Rainflow cycle counting and Miner’s linear damage rule, life predictions were

performed for both crack initiation and total life approaches using the methodology

discussed in section 5.2. The predicted results are also tabulated in Table 5.3.

Figure 5.8 shows maximum principal strain contour for R ~ 0 and loading range

of 3,000 N at the highest peak load during a block of loading. The location of critical

element which is the indicator of crack initiation site is shown in red color and it matches

observed crack initiation location in experiments. Figure 5.9 shows experimental versus

predicted fatigue initiation as well as total lives. For the lower load range initiation

predictions are within a factor of 3, whereas for higher load range initiation predictions

are about a factor of 5 with the predictions being on the non-conservative side.

91
Total life predictions are within a factor of 2, thus total life gives more reasonable

predictions than crack initiation. The total life predictions are more meaningful since the

component fatigue life is dominated by macro-crack growth.

5.5 Conclusions

This chapter presented a generalized fatigue analysis and life prediction approach

for rubber components. Application of the approach was illustrated by analysis of a

vehicle cradle mount and verified by means of component tests. Based on the analysis

and test results presented, the following conclusions can be made:

Quasi-static simulations can be performed for cyclic loading where by using the

stabilized cyclic stress–strain curve the initial softening associated with the Mullin’s

effect is inherently accounted for.

Due to large deformations, strain may be considered a natural choice to

characterize the fatigue behavior of rubber. Among the strain quantities, maximum

principal strain can be used for predictions since fatigue cracks in rubber typically initiate

and grow perpendicular to the maximum principal strain directions.

The R ratio can vary at different critical locations of the component even under

constant amplitude loading. As the R ratio can significantly influence fatigue life, its

effect can be taken into account by using a model such as the Mars-Fatemi R ratio model.

The location of critical element as indicator of the crack initiation site based on

the life prediction methodology used was identical to the observed crack initiation

location in the cradle mount component experiments for both CA and VA loadings. The

predicted fatigue lives based on the crack initiation approach used were within about a

factor of two of the experimental lives for CA loadings. Comparison of predicted versus
92
experimental cradle mount fatigue lives show satisfactory life predictions based on the

linear cumulative damage rule and the maximum principal strain as a damage criterion.

Crack growth constitutes a significant portion of the component total fatigue life.

Life predictions for total component life based on a crack growth analysis approach were

within a factor of three of the experimental lives, with most of the predictions being on

the conservative side. The total life approach used resulted in more accurate fatigue life

predictions (within a factor of 2), as compared to the crack initiation approach.

93
Table 5.1: Summary of component fatigue test results.

Midlife Life to Failure (Cycles)


Control R Control Control Frequency
Response
Mode ratio Amplitude Mean (Hz)
Amplitude Initiation† Total
Load ~0 1025 N -1125 N 1 6.24 mm - >148,000
Load ~0 1100 N -1175 N 1 7.16 mm 20,455* 69,683
Load ~0 1100 N -1175 N 1 7.52 mm 28,000<N<35,000 107,411
Load ~0 1300 N -1400 N 1 7.88 mm - 36,359
Load ~0 1500 N -1600 N 1 10.1 mm 5,685* 19,069
Load ~0 1500 N -1600 N 3 8.85 mm 5,667* 19,100
Load ~0 1500 N -1600 N 1 10.4 mm 4,000<N<8,000 20,180
Load 0.2 1100 N -1650 N 1 7.39 mm 16,000<N<32,000 110,129
Load 0.2 1100 N -1650 N 0.5 - 1 8.04 mm 35,189* 161,447
Load 0.2 1500 N -2250 N 1 9.7 mm N<3,000 20,414
Load 0.2 1500 N -2250 N 0.75 9.44 mm 1,915* 26,034
Displacement 0.1 7.32 mm -9.03 mm 1 1344 N N<7,000 61,450
† st
1 observed crack (with the length of 1 mm) by periodic test interruptions and visual examination
*
Estimated initiation life calculated based on the ratio of initiation life to total life of the test with the same
loading which was stopped in regular intervals for crack growth monitoring

94
Table 5.2: Predicted nucleation and total fatigue lives of the rubber mount component and comparison with experimental
fatigue lives.

FEA results at critical Location Predicted Predicted


Load Mean Experimental Experimental ( N nuc) pred ( N total ) pred
Nucleation Total
RP Amplitude Load
Life Life
Nucleation Total Life
(N) (N) Life (Cycles) (Cycles) ( N nuc)exp ( N total ) exp
ε1, max ε1, min Rε ε1,max,0 (Cycles) (Cycles)
~0 1,025 -1,125 1.33 0.03 0.02 1.33 26,754 145,752 - >148,000 - 0.985
20,455 69,683 1.127 1.442
~0 1,100 -1,175 1.39 0.02 0.02 1.39 23,047 100,516
31,500 107,411 0.732 0.936
~0 1,300 -1,400 1.76 0.03 0.02 1.76 10,511 37,496 - 36,359 - 1.031
5,685 19,069 1.849 0.970
~0 1,500 -1,600 2.08 0.03 0.02 2.08 6,045 18,493 5,667 19,100 1.067 0.968
6,000 20,180 1.007 0.916
24,000 110,129 0.632 0.487
0.2 1,100 -1,650 1.81 0.21 0.11 1.58 15,178 53,687
35,189 161,447 0.431 0.329
1,500 20,414 2.428 0.450
0.2 1,500 -2,250 2.66 0.30 0.11 2.43 3,642 9,195
1,915 26,034 1.902 0.353

95
Table 5.3: Component fatigue test results and predictions under uniaxial variable
amplitude loading with R~0.

Predicted Predicted
Control Experimental Experimental ( Bnuc ) pred ( Btotal) pred
Nucleation Total
RP Range Nucleation1 Life Total Life
Life Life ( Bnuc )exp ( Btotal)exp
(N) (Blocks) (Blocks)
(Blocks) (Blocks)

3,653 2.48 1.17


0<B<1,500
~0 3,000 1,861 4,268
0<B<2,5982 6,327 1.43 0.67

2,454 5.58 0.66


0<B<300
~0 3,600 8,38 1,631
0<B<3202 2,615 5.27 0.62

1
First observed surface crack (with the length of 1 mm) by periodic test interruptions and visual
examination
2
Estimated initiation life ranges calculated based on the ratio of initiation life range to total life of the test
with the same loading which was stopped in regular intervals for crack growth monitoring

96
Figure 5.1: Maximum R = 0 equivalent engineering strain versus crack depth for load
amplitude of 1500 N and R = 0.2 from FEA simulation.

97
Figure 5.2: Component failure definition based on stiffness degradation in fatigue
tests.

98
Figure 5.3: Displacement amplitude versus cycles in fatigue tests.

99
Figure 5.4: Experimental failure locations and crack growth direction in component
tests.

100
(a)

(b)

Figure 5.5: Experimental versus predicted fatigue life based on crack initiation
approach (a) and based on crack growth approach (b).

101
Figure 5.6: Component displacement amplitude response as a function of applied blocks
of loading under variable amplitude fatigue loading.

Figure 5.7: Crack length versus applied load blocks for a test at R ~ 0 and with a load
range of 3,600 N.

102
Figure 5.8: Maximum principal strain contour results for R~0 and loading range of
3,000 N obtained from FE simulation.

103
Figure 5.9: Component fatigue testing and predictions for variable amplitude random
loading at R~0, both for crack initiation and total component lives.

104
Chapter 6

Component Multiaxial Fatigue Behavior

6.1 Introduction

At the critical location(s) of most elastomeric components and structures,

multi-axial states of stress and strain typically exist. Such components and structures are

also usually subjected to variable amplitude or random loading histories. The extreme

values for the strain and/or stress components are not usually coincide with each other

under complex loading conditions. This indicates the complex nature of strain history in

applications like vibration isolators. Therefore, the study of multiaxial fatigue and

deformation of elastomers under variable amplitude loading is an important issue for their

proper design and life prediction analysis.

In this chapter, first the life prediction methodology for general random and

multiaxial loading is discussed. Both crack initiation and crack growth approaches are

included and commonly used and more recently developed fatigue damage quantification

parameters are discussed. Then applications to constant amplitude axial-torsion in-phase

and out-of-phase loading as well as variable amplitude loading are demonstrated using

experimental results from a vehicle cradle mount made of natural rubber. Finally,

predictions for both approaches for all loadings including uniaxial CA and VA are

105
tabulated and the correlation results will be described. Material deformation and fatigue

behavior of the mount are characterized in Chapter 3. The component FEA and fatigue

life analysis and predictions for uniaxial constant and variable amplitude loadings are

presented in Chapters 4 and 5.

6.2 Life prediction methodology for general random and multiaxial

loading of components

Component fatigue damage is typically localized at points of high strains or

stresses. For complex component geometry, finite element analysis is often performed to

identify the critical location and obtain the stress and strain states for that location.

A cycle counting method and a cumulative damage rule are then used for damage

calculations. The cycle counting procedure relates the damage effect of variable

amplitude loading to constant amplitude material fatigue data and fits. The most popular

method is the Rainflow method proposed by Matsuishi and Endo [72], which is also used

here. Linear damage rule is the simplest form of cumulative damage rule, which is used

in this work. However, load sequence effect or interaction between cycles is not

accounted for in this rule.

A damage quantification parameter is also needed to relate the component

multiaxial stresses and strains to uniaxial specimen test and data. Harbour et al. [37]

evaluated maximum normal strain as a critical plane approach. Normal strain can be

defined from:

e( Nˆ )  Nˆ . C . Nˆ  1 (6.1)

106
where C = FT F is the Green’s deformation tensor, F is the deformation gradient tensor

and N̂ is the unit vector normal to the plane in space. By using a MATLAB script, the

maximum value of normal strain and its direction could be defined on all planes in space

in spherical coordinates. Then, the normal strain history on maximum normal strain

(MNS) plane is calculated.

The next step is using the strain-life equation obtained from uniaxial fatigue tests,

as shown in Figure 3.11 and represented by Equation (3.26). The strain R ratio effect

during each loading cycle at the critical location can be accounted for by using maximum

equivalent R = 0 engineering strain (  1,max, 0 ) based on Mars-Fatemi R ratio material

model [17] discussed in Chapter 3 and given by Equation (3.14). For variable amplitude

loading, cycle counting is performed on the critical plane (i.e. MNS) and damage is

calculated for each cycle using the linear damage rule (LDR).

Another successful damage quantification parameter mentioned previously is

cracking energy density (CED). The increment in cracking energy density dWc in the

spatial description is defined in terms of the traction vector T and the unit displacement

vector d on a given plane in the instantaneous deformed configuration and is given by:
 
dWc  T . d (6.2)

The normal to the plane is defined by the unit vector r . By converting from the spatial

description to the material description, the final expression for CED increment is given

by [71]:
 ~   ~ 
 R T C S dE R  R T (2 E  I ) S dE R
dWc       (6.3)
 0 RT C R  0 R T (2 E  I ) R

107
where /0 is the ratio of the deformed mass density (volume) to the undeformed mass
~
density (volume), F is the deformation gradient, S is the 2ndPiola-Kirchhoff stress tensor,

and E is the Green-Lagrange strain tensor. The relationship between unit vector in the
 
current configuration r and the unit vector in the undeformed configuration R is then

given by:

 FR
r  (6.4)
FR

Also C = 2E + I, where C is the Green deformation tensor. Equation (6.3) gives the

cracking energy density in terms of the stress and strain measures of the material

description and in terms of a unit vector in the undeformed configuration.

Two series of calculations were made, one for calculating cracking energy density

(CED) on maximum normal strain (MNS) plane, and another for calculating CED on

CED critical plane with the increment of 5° in spherical coordinates. Crack closure effect

is considered in each approach since the normal traction on each plane is used in the

calculation. Calculation of CED history on MNS plane is only on a single plane, while

calculation of CED history on the critical CED plane is performed on 1296 (= 36 × 36)

planes to find the critical plane with the highest damage value. Therefore, significant

difference in computation time exists between the two approaches.

SED (W) and CED values are the same for uniaxial loading. The SED or CED life

Equation is given by:

W  148 ( N f ) 0.47 (6.5)

This Equation is obtained by uniaxial crack initiation experiments.

108
Fatigue crack initiation approach is typically related to the fatigue life to grow a

crack from a natural flaw to a length on the order of about a millimeter. Crack growth

approach then considers fatigue life from this crack length to failure. Crack initiation is

often used for applications where micro-cracks dominate the total fatigue life. Crack

growth approach is usually considered when a large portion of the total life is involved

with macro-crack growth. In this study both approaches were considered. To calculate

total component fatigue life, the same methodology discussed in section 5.2 is used.

The cradle mount was initially made with a snubber part which stops application

of the high loads to the component. Moreover, the amount of load or displacement

applied to this component was much less than needed to produce fatigue failure.

Therefore, the stop or snubber parts of the mounts are removed prior to testing and in

simulations for analyzing this component in order to produce fatigue failure sooner.

This elastomeric cradle mount was not designed for applying torsion loads.

Therefore it is not appropriate to evaluate criteria like SED efficiently with this geometry.

The load levels chosen for testing were much higher than actual loading of the

component and experimental fatigue lives were much shorter than fatigue life of the

cradle mount in service.

The predictions are highly conservative, since they are based on initiation life of a

crack in the order of a millimeter. Almost all of the experiments showed that significant

life was left from this point to final failure of the component. Therefore, macro-crack

growth involves most of the cradle mount fatigue life.

109
6.3 Applications to multiaxial constant amplitude loading

6.3.1 Experimental program and results

Component level fatigue tests were conducted in order to validate the fatigue life

prediction methodology discussed above. Figure 4.1(a) shows the component used. The

finite element model of the component is shown in Figure 6.1. Axial loading with

amplitude of 1100 N and R ~ 0 and torsion loading with amplitude of 30 N.m. and R = -1

were used as the experimental loading conditions. Both in-phase and 90 out-of-phase

tests were conducted with two tests for each condition. One test of each condition was

stopped in regular intervals to monitor crack initiation and growth. Fatigue tests were

conducted with a sinusoidal waveform of 1 Hz using a servo-hydraulic axial-torsion load

frame and axial as well as rotational displacement responses of each test specimen were

periodically recorded.

The loading conditions chosen resulted in fatigue lives between about 21,500

cycles and 67,500 cycles. The maximum axial and rotational displacements increased

initially due to the Mullin’s effect and then stabilized after several hundred cycles. When

a fatigue crack initiated and then grew to a critical size, the displacement and rotation

amplitudes increased rapidly due to reduced stiffness. Failure was defined as the lower

number of cycles at which the axial or rotational displacement amplitudes increased

drastically from its gradual linear change, as illustrated for a typical test in Figure 6.2.

The experimental fatigue lives at different load levels are tabulated in Table 6.1.

Test results from uniaxial fatigue tests reported in Chapter 5 are also included in this

table. Experimental response for constant amplitude axial-torsion loading of components

is shown in Figure 6.3(a) for in-phase and 6.3(b) for 90° out-of-phase tests. The life

110
comparisons show shorter life for 90° out-of-phase loading, as compared to in-phase

loading. Experimental life scatter between duplicate tests was within a factor of three for

in-phase tests and within a factor of two for out-of-phase tests. This table also shows the

first crack observations for the tests which were stopped at regular intervals for

monitoring. The crack nucleation life for a duplicate test for which initiation was not

monitored was calculated based on the ratio of crack nucleation life to total life of the

monitored test of the same loading condition. Crack growth life was a significant portion

of the total fatigue life (about 75%) for in-phase test, while for out-of-phase test, this ratio

was about 30%.

In these load and torque controlled tests both displacement and rotation

amplitudes increased with increasing cycles. Displacement versus rotation at different

cycles during the test is shown for both In-phase and 90 out-of-phase loading in

Figure 6.4. By increasing the number of cycles more distinct change in maximum

displacement (in compression) and negative rotation value is observed. This is due to the

evolution of micro-cracks and subsequent coalescence of those cracks. This affects the

stiffness behavior of the component. Displacement versus rotation at midlife is shown in

Figure 6.5.

Crack length as well as its depth and their changes were measured at critical

(failure) location by periodic test interruptions and visual inspection. For in-phase

loading, crack length did not have a significant effect on fatigue life, as even at a length

of about 4 cm, the component was still at only about half of its total life. Crack depth had

a more dominant effect on life than crack length. When crack depth was on the order of

3 cm, the component was near failure. Stiffness drop also correlated better with crack

111
depth than with crack length. Crack length and depth versus cycles are shown in

Figure 6.6(a) for in-phase loading. For out-of-phase loading, crack length was a key

factor, rather than crack depth and this is shown in Figure 6.6(b). In out-of-phase loading,

the ratio of crack initiation life to total life increased about 3 times of the same loading

level for in-phase condition, therefore, less life is spent in crack growth for OP loading.

Since this ratio was much higher in OP tests compared to what was observed for uniaxial

load cases, phasing had much more effect on the ratio of initiation to total life. This is

shown in Figure 6.7.

To evaluate any difference between load control (LC) and displacement control

(DC) mode, IP and OP displacement-controlled tests were also performed based on

midlife displacements of the LC tests. In DC tests failure was defined based on 40% load

or torque drop, whichever occurred sooner. The midlife load-displacement and

torque-rotation was similar in both LC and DC tests. Cracks initiated more rapidly under

LC than under DC. This is the same as what was observed for CA uniaxial tests in

Chapter 5. The results are tabulated in Table 6.1. Crack initiation location was the same

for all cases of LC and DC mode tests, however.

6.3.2 Life predictions and validation of methodology

The nonlinear FE program ABAQUS was used to simulate the nonlinear and large

deformation hyper-elastic behavior of the mount. The Marlow strain energy density form

was used to model deformation of incompressible isotropic elastomers. The mount FE

mesh is shown in Figure 6.1. To adequately model the geometry, a relatively fine mesh

with about 30,000 elements was used. The element type used was the eight node brick

(hexahedral) element. For cyclic loading, the amplitude tool of the FE software was used,

112
so that a loading cycle could be exactly simulated. Stable deformation curve was applied

as an input to the model, therefore initial softening (Mullin’s effect) was inherently

considered. More details of FE model can be found in Chapter 4.

Maximum principal strain distributions of the mount under IP loading is shown in

Figure 6.8, which is similar to that for OP loading. The maximum principal strain occurs

at the interface between the inner bushing and the bulk rubber material at a location of

abrupt geometry change. Locations, at which the maximum principal strain occurred, as

indicated in Figure 6.8 for in-phase loading, were also observed to be the site(s) of

fatigue cracking during all fatigue tests of the component. The maximum principal strain

at the critical location determined from FEA was, therefore, one of the parameters used

for evaluating the fatigue damage.

By using the data from critical element (defined as the location with the highest

maximum principal strain) of the mount in each loading cycle, life predictions were

performed. Values of the maximum normal strain (MNS), critical cracking energy

density (CED), and CED on MNS plane are listed in Table 6.2. Through a tensor

transformation program developed in MATLAB, the maximum normal strain value and

direction was defined. Equations (3.14) and (3.26) based on uniaxial fatigue properties

were then used to perform life predictions. The correlation of the predicted and

experimental results is shown in Figure 6.9(a). The predictions are within a factor of two

for all experiments by using the maximum normal strain parameter. Since only strain is

involved in the calculations, this approach does not require the constitutive behavior of

the material under multiaxial loading. Therefore, numerically it is an efficient parameter

for determining the critical or failure plane(s) through a 3-D search process.

113
Maximum normal strain plane direction and Cauchy stress history were used in

another MATLAB program to obtain cracking energy density on the MNS plane.

Equation (6.5) for CED based on uniaxial fatigue properties was used to perform life

predictions. The correlation of predicted results with experimental results for CED is

shown in Figure 6.9(b) and the same plot for CED on MNS plane is shown in

Figure 6.9(c). The results show that using CED on MNS gives similar predictions to the

critical CED criterion. As discussed earlier, this is a computationally efficient approach

(by a factor of about 45), to predict crack initiation life of the component, since CED

history is calculated on a single plane.

Total life of the component was also predicted based on Equation (5.5).

The predicted and experimental results are tabulated in Table 6.2. The correlation of

predicted versus experimental total lives are shown in Figure 6.10. The predictions are all

within a factor of three of the experiments.

6.4 Applications to variable amplitude multiaxial loading

For evaluating more complex loading conditions, variable amplitude proportional

and non-proportional axial-torsion tests were also conducted on the elastomeric

component. Two in-phase and two out-of-phase tests were conducted. For in-phase

loading the peaks and valleys in both axial and torsion load signals are reached at the

same time, while for out-of-phase loading they are not (see Figure 6.11). Experiments

were conducted with the load range of 3,680 N in comparison with minimum load of near

zero and torque range of 50 N.m. with mean value of near zero. For one duplicate test of

each condition, the test was stopped at regular intervals (every 10% of the expected life)

to monitor crack initiation and growth.


114
Test results for variable amplitude tests are given in Table 6.1. Experimental life

scatter between duplicate tests was within a factor of two for both in-phase and

out-of-phase loading. Crack initiation range is also tabulated in Table 6.1. Crack growth

life constitutes a significant portion of the total life in all tests. In out-of-phase tests the

ratio of crack initiation life to total life increased about 1.5 times, as compared to

in-phase loading. The same observation was made in constant amplitude loading. This is

shown in Figure 6.7. Crack length and depth measurement results showed that crack

depth was key factor in component fatigue life for both in-phase and out-of-phase tests.

This was also observed in constant amplitude uniaxial tests.

The maximum principal strain occurs at the interface between the inner bushing

and the bulk rubber material at a location of abrupt geometry change, similar to all other

loading cases and corresponding to that observed during the fatigue tests. This crack

initiation site remained the same for all of the simulations, either IP or OP.

The same methodology used for constant amplitude axial-torsion loading

condition was used to define crack initiation life as well as total fatigue life of the

component. The only difference for variable amplitude loading is the use of Rainflow

cycle counting method for event identification and then using Miner linear rule to

accumulate the fatigue damage. Figures 6.12(a) and 6.12(b) show normal strain history

on maximum normal strain plane for in-phase and out-of-phase loadings, respectively. As

can be seen from this figure, the maximum normal strain history for both IP and OP

conditions at critical location is different from the nominal loading history applied to the

component (see Figure 6.11). Cycle counting was performed on the shown MNS history.

115
After event identification, by utilizing Miner rule, the fatigue life prediction was

performed.

Life prediction results, shown in Figure 6.9 for different damage parameters,

indicate that MNS criterion is not accurate for complex variable amplitude multiaxial

loading. Both critical CED and CED on MNS criteria give better predictions. Most of the

total fatigue life predictions are within a factor of three of experiments in this case

according to Figure 6.10.

6.5 Discussion of Results

Predicted crack nucleation locations based on the developed methodology

matched observed failure locations for all loading conditions. Crack initiation lives were

predicted well (mostly within a factor of two) based on the maximum normal strain and

the developed methodology for most of the loading conditions. Because only strain is

involved in this approach and no constitutive behavior is used, maximum normal strain

can be a very efficient parameter to determine critical or failure plane. Most of the

predictions were on non-conservative side (i.e. predicted life longer than experimental

life). This is, at least partly, because of the fact that in the FE model the strain and stress

at critical location were lower than in the component since a finer mesh could not be used

due to practical considerations.

Of the two CED criteria, the CED on MNS plane criterion gives quite acceptable

predictions with much less calculation time, compared to the critical CED plane criterion.

This is because CED history is calculated on just one MNS plane and not on many planes

in space. Calculation of CED requires knowledge of the constitutive behavior of the

116
material, however. Both of these approaches resulted in better life predictions than MNS

for multiaxial variable amplitude loading conditions.

Two criteria of Smith-Watson-Topper (SWT) commonly used for metallic

materials and strain energy density (SED) traditionally used for elastomers were also

evaluated. All of the predictions based on SWT criterion were overly conservative

predictions by up to a factor of 23. Predictions based on SED resulted mostly in life

within a factor of two for this particular component and the loading conditions

considered. SED criterion gives close predictions compared to CED criterion because of

the particular component geometry where based on the FE results the maximum principal

strain at the critical location is much higher than the other two principal strains. This

indicates essentially uniaxial strain state, although the loading of the component was

multiaxial. In addition, the principal strain at the critical location is tensile, while for

compression strain there would be a large deviation between SED and CED criteria

predictions. The difference between SED and CED criteria is also large when torsion

loading is dominant compared to axial loading, which was not the case for the loading

considered in this study. Differences between CED and SED criteria based on different

loading conditions are also discussed in [38]. Overall, of all fatigue crack initiation

criteria studied here, those which use critical plane approaches work better than the scalar

damage parameters or those damage parameters which do not take constitutive behavior

into account.

Crack growth constitutes a significant portion of the component total fatigue life.

The ratio of initiation to total fatigue life varied between 6 and 69 % for different loading

cases. For constant amplitude or variable amplitude uniaxial tests, by increasing the load

117
amplitude, the ratio of crack initiation life to total life decreased (i.e. more life spent in

crack growth). This is expected, since higher load level initiates a crack quicker and

fatigue life is dominated by crack growth.

Multiaxial load phasing had an important effect on the ratio of crack initiation to

total life. For constant amplitude out-of-phase loading, this ratio increased about 3 times

compared to in-phase loading at the same level. This means more life was involved in

crack initiation for out-of-phase loading, compared to in-phase loading. Total life

comparison showed shorter life for out-of-phase loading compared to

in-phase loading for both constant and variable amplitude loadings, indicating much

faster crack growth rate once a crack has initiated in out-of-phase loading.

The fracture mechanics approach was used for total fatigue life prediction for

each loading condition based on specimen crack growth data and FE simulation results.

The methodology used for component total fatigue life resulted in reasonable predictions

for nearly all the loading conditions. The prediction results are sensitive to initial crack

length used for analysis, but not very sensitive to the final crack length used. This was

also observed in uniaxial behavior of the component, as discussed in Chapter 5, where the

critical element of the component had the highest effect in the total life prediction.

6.6 Conclusions

1. Shorter total life was observed for out-of-phase loading compared to

in-phase loading at the same level for both constant and variable amplitude loadings.

In addition, for out-of-phase loading the ratio of crack initiation life to total life was

about 3 times higher than for in-phase loading. This indicates more life was involved

118
in crack initiation but much faster crack growth rate once a crack initiated in out-of-

phase loading, as compared to in-phase loading.

2. Crack initiation location was observed to remain the same for all of the loading

conditions used for fatigue tests of the component. This location corresponded to the

point at which the maximum principal strain occurred. Therefore, maximum normal

strain (MNS) is an effective parameter to determine the critical or failure plane.

3. Although the maximum normal strain (MNS) parameter correlated the constant

amplitude fatigue life data within mostly a factor of about two, it could not correlate

the data satisfactorily for the more complex case of variable amplitude loading.

4. Amongst the different approached evaluated, a computationally efficient and

relatively accurate fatigue crack initiation life prediction approach for complex

loading was found to be a critical plane approach based on MNS and quantifying

damage on this plane using the cracking energy density (CED) parameter.

5. For variable amplitude loading the rainflow cycle counting of the maximum normal

strain (MNS) history was found to be an efficient method. Miner linear damage rule

was then used to accumulate damage on the MNS plane based on constant amplitude

fatigue data, resulting in satisfactory life predictions for variable amplitude loading.

6. Crack growth constituted a significant portion of the component total fatigue life. The

fracture mechanics approach was used for total fatigue life prediction based on

specimen crack growth data and FE simulation results. The methodology used for

component total fatigue life resulted in reasonable predictions with nearly all the life

predictions being within a factor of three of the experimental lives.

119
Table 6.1: Component experimental conditions and test results.

Control RP /RM a / a Fatigue Life (Cycles or Blocks)


Test Load Load Control Frequency
Amplitude or or
Type Type Path Mode (Hz)
(Range) R/R Pa / Ma Initiation1 Total2
A3 CA N.A. LC 1025 N ~0 1 6.24 mm >148,000
A CA N.A. LC 1100 N ~0 1 7.16 mm 18,165<N<22,7064 69,683
A CA N.A. LC 1100 N ~0 1 7.52 mm 28,000<N<35,000 107,411
A CA N.A. LC 1300 N ~0 1 7.88 mm 36,359
A CA N.A. LC 1500 N ~0 1 10.1 mm 3,780<N<7,560 19,069
A CA N.A. LC 1500 N ~0 1 10.4 mm 4,000<N<8,000 20,180
A CA N.A. LC 1500 N ~0 3 8.85 mm 3,786<N<7,572 19,100
A CA N.A. LC 1100 N 0.2 1,0.5,0.75 8.04 mm 23,456<N<46,912 161,447
A CA N.A. LC 1100 N 0.2 1 7.39 mm 16,000<N<32,000 110,129
A CA N.A. LC 1500 N 0.2 0.75 9.44 mm N<3,826 26,034
A CA N.A. LC 1500 N 0.2 1 9.7 mm N<3,000 20,414
A VA N.A. LC (3000 N) ~0 16.5 S/B5 8.59 mm B<2,598 6,327
A VA N.A. LC (3000 N) ~0 16.5 S/B 8.77 mm B<1,500 3,653
A VA N.A. LC (3600 N) ~0 16.5 S/B 9.94 mm B<320 2,615
A VA N.A. LC (3600 N) ~0 16.5 S/B 9.87 mm B<300 2,454

120
Table 6.1 (Continued): Component experimental conditions and test results.

Control RP /RM a / a Fatigue Life (Cycles or Blocks)


Test Load Load Control Frequency
Amplitude or or
Type Type Path Mode (Hz)
(Range) R/R Pa / Ma Initiation1 Total2
A-T CA IP LC 1100 N/30 N.m ~0/-1 1 6.85 mm/19.26° 15,510 67,406
A-T CA IP LC 1100 N/30 N.m ~0/-1 1 7.04 mm/18.62° 7,200 31,290
A-T CA IP DC 6.95 mm/18.94° 0.1/-1 1 1044 N/28.2 N.m 0<N<3,600 74,230
A-T CA OP LC 1100 N/30 N.m ~0/-1 1 7.04 mm/19.68° 15,806<N<18,967 25,119
A-T CA OP LC 1100 N/30 N.m ~0/-1 1 8.14 mm/21.67° 13,500<N<16,200 21,454
A-T CA OP LC 1100 N/30 N.m ~0/-1 1 8.08 mm/21.97° >21,250
A-T CA OP DC 7.53 mm/20.68° ~0/-1 1 1085 N/30 N.m 43,750
A-T VA IP LC (3680 N/50 N.m) ~0/-1 43.5 S/B 11.32 mm/15.52° 122<B<244 1,393
A-T VA IP LC (3680 N/50 N.m) ~0/-1 43.5 S/B 9.81 mm/13.41° 150<B<300 1,710
A-T VA OP LC (3680 N/50 N.m) ~0/-1 47.6 S/B 11.22 mm/15.60° 319<B<638 1,501
A-T VA OP LC (3680 N/50 N.m) ~0/-1 47.6 S/B 11.99 mm/16.71° 170<B<340 800
A-T VA OP LC (3680 N/50 N.m) ~0/-1 47.6 S/B 7.63 mm/20.68° 340<B<510 2,000
1
1st observed crack (with the length of 1 mm) by periodic test interruptions and visual examination
2
In LC defined by sudden increase of displacement amplitude (log scale), in DC defined by 40% load drop
3
A- Axial, T- Torsion, CA- Constant Amplitude, VA- Variable Amplitude, LC- Load-Controlled,
DC- Displacement-Controlled, IP- In-Phase, OP- Out-of-Phase
4
Bold font indicates the estimated initiation life which is calculated based on the ratio of initiation to total life of the
test with the same loading which is stopped in regular intervals for crack growth monitoring
5
S/B is seconds per block

121
Table 6.2: Summary of component crack initiation and total life experiments and predictions.

Initiation Life (Cycles or Blocks) Total Life (Cycles or Blocks)


Control
Test Load Load
ε1,max,0 Amplitude RP /RM Experimental Prediction Prediction
Type Type Path Prediction
(Range) 1.Measured (CED on (CED Critical Experimental Prediction
(MNS)
2.Estimated MNS Plane) Plane)

A CA N.A. 1.33 1025 N ~0 1: >148,000 145,752

1: 28,000<N<35,000 1: 69,683
A CA N.A. 1.39 1100 N ~0 20,008 20,983 20,809 100,516
2: 18,165<N<22,706 2: 107,411

A CA N.A. 1.76 1300 N ~0 1: 36,359 37,496

1: 4,000<N<8,000 1: 19,069
A CA N.A. 2.08 1500 N ~0 2: 3,780<N<7,560 6,339 6,420 6,383 2: 20,180 18,493
2: 3,786<N<7,572 3: 19,100
1: 16,000<N<32,000 1: 110,129
A CA N.A. 1.58 1100 N 0.2 12,779 10,632 10,543 53,687
2: 23,456<N<46,912 2: 161,447
1: N<3,000 1: 20,414
A CA N.A. 2.43 1500 N 0.2 3,665 2,932 2,883 9,195
2: N<3,826 2: 26,034
1: B<1500 1: 6,327
A VA N.A. 2.01 (3000 N) ~0 1,951 1,338 979 4,268
2: B<2,598 2: 3,653
1: B<300 1: 2,615
A VA N.A. 2.47 (3600 N) ~0 932 662 865 1,631
2: B<320 2: 2,454

122
Table 6.2 (continued): Summary of component crack initiation and total life experiments and predictions.

Initiation Life (Cycles or Blocks) Total Life (Cycles or Blocks)


Control
Test Load Load
ε1,max,0 Amplitude RP /RM Experimental Prediction Prediction
Type Type Path Prediction
(Range) 1.Measured (CED on (CED Critical Experimental Prediction
(MNS)
2.Estimated MNS Plane) Plane)
1100 N, 1: 7,200 1: 67,406
A-T CA IP 1.45 ~0/-1 11,348 10,209 10,119 25,646
30 N.m 2: 15,510 2: 31,290
1100 N, 1: 13,500<N<16,200 1: 25,119
A-T CA OP 1.42 ~0/-1 21,585 25,389 18,677 55,765
30 N.m 2: 15,806<N<18,967 2: 21,454
(3680 N, 1: 150<B<300 1: 1,393
A-T VA IP 2.29 ~0/-1 910 513 563 3,680
50 N.m) 2: 122<B<244 2: 1,710
1: 170<B<340 1: 1,501
(3680 N,
A-T VA OP 2.30 ~0/-1 1: 340<B<510 1,606 746 770 2: 800 4,664
50 N.m)
2: 319<B<638 3: 2,000

123
Figure 6.1: Vehicle cradle mount FE model.

Figure 6.2 : Definition of failure illustrated by variation of axial and rotational


displacements in constant amplitude in-phase axial-torsion tests of the cradle
mount.

124
(a)

(b)

Figure 6.3: Experimental response for constant amplitude axial-torsion loading of


components for (a) IP and (b) OP tests.

125
(a)

(b)

Figure 6.4: Axial displacement versus rotation angle of the component for different
cycles throughout the CA and A-T tests of (a) IP and (b) 90 OP.

126
Figure 6.5: Axial displacement versus rotation angle of the component for mid-life
cycles for CA and A-T IP and 90° OP tests.

127
(a)

(b)

Figure 6.6: Evolution of crack length and depth for constant amplitude axial-torsion
(a) in-phase test with Nf = 31,290, and (b) out-of-phase test with
Nf = 21,454.

128
Figure 6.7: Crack initiation to total fatigue life ratio for all types of loadings.

129
Figure 6.8: Maximum principal strain location for constant amplitude in-phase loading
simulation.

130
(a)

(b)

131
(c)

Figure 6.9: Experimental versus predicted component initiation life for all loading
conditions based on (a) maximum normal strain, (b) critical CED, and
(c) CED on MNS plane.

132
Figure 6.10: Experimental versus predicted component total fatigue life for all loading
conditions.

133
Figure 6.11: Variable amplitude in-phase and out-of-phase axial-torsion loading history
used for component testing.

134
(a)

(b)

Figure 6.12: Normal strain on maximum normal strain (MNS) plane history of variable
amplitude and axial-torsion loading for (a) in-phase, and (b) out-of-phase
loading.

135
Chapter 7

Summary and Possible Future Research

7.1 Summary
The objective of this study was to evaluate a robust methodology for fatigue life

prediction of elastomeric components under complex loading conditions and validate it

by using an automobile cradle mount as an illustrative example. Material deformation

and fatigue behavior of the mount were characterized in Chapter 3. The material

properties needed for elastomeric component fatigue analysis and life prediction include

deformation behavior under different stress states, crack initiation life, and crack growth

rate properties. To obtain each of these material characterizations, specific experimental

procedures and data analysis techniques are used. It was then demonstrated how such

properties could be obtained from simple specimen geometries and experimental

techniques. For each type of characterization, first the experimental procedure was

described, followed by data analysis, and mathematical model representation of the

obtained results. A filled natural rubber material with 21% carbon black and 9.5%

plasticizer was used for both specimen and component experiments. Specimen test results

and life prediction under variable amplitude load history utilizing constant amplitude

material properties were presented.

136
The component FE analyses were presented in Chapter 4 including material

characterization, model definition, and obtained strain distributions. The component FEA

and fatigue life analysis and predictions for uniaxial constant amplitude and variable

amplitude loadings were presented in Chapter 5. The capabilities of Rainflow cycle

counting procedure and Miner’s linear damage rule were also evaluated with component

tests. Component experiments and life predictions including FE simulations were

presented where predicted fatigue lives were compared with experimental lives.

Finally, the life prediction methodology for general random and multiaxial

loading was discussed in Chapter 6. Both crack initiation and crack growth approaches

were included. Then applications to constant amplitude axial-torsion in-phase and

out-of-phase loading as well as variable amplitude loading were demonstrated using

experimental results from a vehicle cradle mount made of natural rubber. Validation of

life prediction methodology used was discussed with respect to both failure location and

fatigue life.

7.1.1 Constitutive behavior

1) Two simple specimen geometries were used for deformation characterizations of

elastomers: simple tension and planar tension specimens. In the simple tension specimen

a uniaxial state of stress with a multiaxial stretch state is present, while the planar tension

specimen is under plane stress condition with longitudinal and transverse stresses.

2) As monotonic and cyclic deformation behaviors can be vastly different, it is

important to obtain and use cyclic deformation properties in fatigue life analysis and

applications. Cyclic incremental tests on simple tension and/or planar tension specimen

were conducted to obtain the stabilized cyclic stress-strain curve. The choice of the curve

137
to use in FE simulations depends on the stress state at the critical location of the

component being analyzed.

3) Due to the load history dependence associated with the Mullin’s effect, peak

stress (or strain) is a key consideration in fatigue analysis of elastomers, in addition to the

stress (or strain) amplitude and the mean stress.

7.1.2 Fatigue material relations obtained from specimen experiments

4) Fatigue crack initiation tests were performed with the simple tension specimen

geometry in displacement control and with different strain ratios (Rε ratios) representing

tension-tension loading conditions. Best fit lines in log-log scale, where the fatigue life is

treated as the dependent variable, represent maximum strain versus crack nucleation life

for each strain ratio. The effect of Rε > 0 was found to increase the nucleation life and was

significant at low strains. This is in contrast to metals and this effect in natural rubber is

attributed to strain crystallization.

5) Energy release rate is often used as the crack driving parameter in characterizing

elastomers fatigue crack growth behavior. A pre-cracked planar tension specimen is

typically used for fatigue crack growth tests since energy release rate is independent of

crack length for this specimen geometry. Therefore, a single specimen can produce

results for multiple crack growth tests with different loading conditions and RT ratios. A

power-law relation was used to describe crack growth rates in terms of the maximum

energy release rate. The crack growth approach could be used as a total life approach,

based on growth of pre-existing flaws to failure.

6) The Mars-Fatemi model was used to correlate test results from different R ratio

conditions. This model was used to obtain an equivalent maximum strain in crack

138
initiation tests, or an equivalent maximum energy release rate in crack growth tests, to

account for the effect of R ratio.

7) Most of the life predictions for specimen random loading history tests based on

Rainflow cycle counting method in conjunction with the linear damage rule were within

the factor of two scatter bands. The R-ratio equation for fatigue life based on constant

amplitude material data was found to be applicable to random loading. The effect

associated with a rest period (dwell period) between blocks of cyclic loading on the

material fatigue behavior was found to be insignificant. Also, loading rate in the range

investigated did not have a significant effect on fatigue life in specimen fatigue crack

initiation tests.

7.1.3 Uniaxial constant and variable amplitude component fatigue behavior

and life predictions

8) A generalized fatigue analysis and life prediction approach for rubber components

was presented based on maximum principal strain as a fatigue damage quantification

parameter. Application of the approach was illustrated by analysis of a vehicle cradle

mount and verified by means of component tests.

9) Quasi-static simulations were performed for cyclic loading where by using the

stabilized cyclic stress-strain curve the initial softening associated with the Mullin’s

effect is inherently accounted for.

10) Due to large deformations, strain may be considered a natural choice to

characterize the fatigue behavior of rubber. Among the strain quantities, maximum

principal strain can be used for predictions since fatigue cracks in rubber typically initiate

and grow perpendicular to the maximum principal strain direction.

139
11) The R ratio can vary at different critical locations of the component even under

constant amplitude loading. As the R ratio can significantly influence fatigue life, its

effect can be taken into account by using a model such as the Mars-Fatemi R ratio model.

12) Experiments failure locations of the component agreed with the predictions based

on FEA for both CA and VA loadings. The predicted fatigue lives based on the crack

initiation approach used were within about a factor of two of the experimental lives for

CA loadings. Comparison of predicted versus experimental cradle mount fatigue lives

show satisfactory life predictions based on the linear cumulative damage rule and the

maximum principal strain as a damage criterion.

13) Crack growth constituted a significant portion of the component total fatigue life.

Life predictions for total component life based on a crack growth analysis approach were

within a factor of three of the experimental lives, with most of the predictions being on

the conservative side. The total life approach used resulted in more accurate fatigue life

predictions (within a factor of 2), as compared to the crack initiation approach.

7.1.4 Multiaxial constant and variable amplitude component fatigue

behavior and life predictions

14) Predicted crack nucleation locations based on the developed methodology

matched observed failure locations for all loading conditions. Crack initiation lives were

predicted well (mostly within a factor of two) based on the maximum normal strain and

the developed methodology for most of the loading conditions. Because only strain is

involved in this approach and no constitutive behavior is used, maximum normal strain

can be a very efficient parameter to determine the critical or failure plane. Most of the

predictions were on non-conservative side (i.e. predicted life longer than experimental

140
life). This is partly due to the fact that in the FE model the strain and stress at critical

location were lower than in the component since a finer mesh could not be used due to

practical considerations.

15) Of the two CED criteria of critical CED and CED on MNS plane, the CED on

MNS plane criterion gives quite acceptable predictions with much less calculation time,

compared to the critical CED plane criterion. This is because CED history is calculated

on just one MNS plane and not on many planes in space. Calculation of CED requires

knowledge of the constitutive behavior of the material, however. Both of these

approaches resulted in better life predictions than MNS for multiaxial variable amplitude

loading conditions.

16) The Smith-Watson-Topper (SWT) criterion which is commonly used for metallic

materials was also evaluated. All of the predictions based on SWT criterion were overly

conservative predictions by up to a factor of 23.

17) Predictions based on SED resulted mostly in life within a factor of two for this

particular component and the loading conditions considered. SED criterion gives close

predictions compared to CED criterion because of the particular component geometry

where based on the FE results the maximum principal strain at the critical location is

much higher than the other two principal strains. This indicates essentially uniaxial strain

state, although the loading of the component was multiaxial. In addition, the principal

strain at the critical location is tensile, while for compression strain there would be a large

deviation between SED and CED criteria predictions. The difference between SED and

CED criteria is also large when torsion loading is dominant compared to axial loading,

which was not the case for the loading considered in this study.

141
18) Crack growth constituted a significant portion of the component total fatigue life.

The ratio of initiation to total fatigue life varied between 6 and 69 % for different loading

cases. For constant amplitude or variable amplitude uniaxial tests, by increasing the load

amplitude, the ratio of crack initiation life to total life decreased (i.e. more life spent in

crack growth). This is expected, since higher load level initiates a crack quicker and

fatigue life is dominated by crack growth.

19) Multiaxial load phasing had an important effect on the ratio of crack initiation to

total life. For constant amplitude out-of-phase loading, this ratio increased about 3 times

compared to in-phase loading at the same level. This means more life was involved in

crack initiation for out-of-phase loading, compared to in-phase loading. Total life

comparison showed shorter life for out-of-phase loading compared to in-phase loading

for both constant and variable amplitude loadings, indicating much faster crack growth

rate once a crack has initiated in out-of-phase loading.

20) The fracture mechanics approach was used for total fatigue life prediction for

each loading condition based on specimen crack growth data and FE simulation results.

The methodology used for component total fatigue life resulted in reasonable predictions

for nearly all the loading conditions. The prediction results are sensitive to initial crack

length used for analysis, but not very sensitive to the final crack length used. This was

also observed in uniaxial behavior of the component where the critical element of the

component had the highest effect in the total life prediction.

21) Shorter total fatigue life was observed for out-of-phase loading compared to

in-phase loading at the same level for both constant and variable amplitude loadings.

142
22) Crack initiation location was observed to remain the same for all of the loading

conditions used for fatigue tests of the component. This location corresponded to the

point at which the maximum principal strain occurred. Therefore, maximum normal

strain (MNS) is an effective parameter to determine the critical or failure plane.

23) Although the maximum normal strain (MNS) parameter correlated the constant

amplitude fatigue life data within mostly a factor of about two, it could not correlate the

data satisfactorily for the more complex case of variable amplitude loading. Amongst the

different approaches evaluated, a computationally efficient and relatively accurate fatigue

crack initiation life prediction approach for complex loading was found to be a critical

plane approach based on MNS and quantifying damage on this plane using the cracking

energy density (CED) parameter.

24) For variable amplitude loading, the Rainflow cycle counting of the maximum

normal strain (MNS) history was found to be an efficient method. Miner linear damage

rule was then used to accumulate damage on the MNS plane based on constant amplitude

fatigue data, resulting in satisfactory life predictions for variable amplitude loading.

7.2 Possible future research

1) In order to predict fatigue crack growth life in the component more accurately, a

possible future study could be modeling a crack in the FE software and using Fracture

Mechanics (i.e. J-integral) for fatigue life predictions by using specimen fatigue crack

growth data.

2) For finding more accurate results for crack initiation life calculations, a refined

mesh at the critical location can be considered by using sub-modeling feature of the

143
software. This was not possible in current study, since the feature used was an orphan

mesh.

3) For acquiring more complete material properties to use in FE software as input,

equi-biaxial tension testing of the specimen could be performed, in addition the two

specimen tests in simple tension and planar tension. By utilizing three sets of test data,

more accurate hyper-elastic material characterization could be carried out by utilizing

Ogden strain energy density for input in FE software.

4) Choosing a component with appropriate geometry to apply higher torque is

another possible future study in order to be able to more efficiently evaluate the SED

criteria and its difference with the cracking energy density approach.

5) Most of the industrial structures experience long term rest period in service.

Therefore, testing with longer dwell period than what was used in this study could be the

issue of a further study.

6) Fatigue life in elastomers is highly affected by temperature. Therefore,

characterizing temperature effect on fatigue life for both material and component

behaviors is an important consideration in many applications.

7) In order to check the capability of the developed methodology for different types

of elastomers, studying different types of rubber (for instance non-strain crystallized

elastomers such as SBR) could more comprehensively demonstrate the versatility of

developed methodology.

144
References

[1] ASTM Standard E1823-05a, 2007, “Standard terminology relating to fatigue and
fracture testing,” ASTM, West Conshohocken, PA, Vol.03.01, pp.1147-1167.

[2] Woo, C.S., Kim, W.D., and Kwon, J.D., 2008, “A study on the material properties
and fatigue life prediction of natural rubber component,” Materials Science and
Engineering A 483-484, pp. 376-381.

[3] Laraba-Abbas, F., Ienny, P., and Piques, R., 2003, “A new ‘Tailor-made’
methodology for the mechanical behaviour analysis of rubber-like materials:
II. Application to the hyperelastic behaviour characterization of a carbon-black filled
natural rubber vulcanizates,” Polymer, Vol. 44, pp. 821–840.

[4] Choi, J.H., Kang, H.J., Jeong, H.Y., Lee, T.S., and Yoon, S.J., 2005, “Heat aging
effects on the material property and the fatigue life of vulcanized natural rubber, and
fatigue life prediction equations,” Journal of Mechanical Science and Technology (KSME
International Journal), Vol. 19, No. 6, pp. 1229-1242.

[5] Zarrin-Ghalami, T. and Fatemi, A., 2012, “Material deformation and fatigue behavior
characterization for elastomeric component life predictions,” Polymer Engineering and
Science, Vol. 52, pp. 1795-1805.

[6] Zarrin-Ghalami, T. and Fatemi, A., 2012, “Fatigue life predictions of rubber
components: applications to an automobile cradle mount,” Journal of Automobile
Engineering, in press, DOI: 10.1177/09544070112461863.

[7] Zarrin-Ghalami, T., and Fatemi, A., 2012, “Cumulative Fatigue Damage and Life
Prediction of Elastomeric Components,” Fatigue and Fracture of Engineering Materials
and Structures, in press, available online, DOI: 10.1111/j.1460-2695.2012.01720.x.

145
[8] Zarrin-Ghalami, T., and Fatemi, A., 2012, “Multiaxial Fatigue and Life Prediction of
Elastomeric Components,” Submitted to International Journal of Fatigue (under review).

[9] Zarrin-Ghalami, T., Fatemi, A., and Lee, Y-L, 2013, “Fatigue life predictions of an
automobile cradle mount,” Submitted to SAE 2013 Congress, April 16-18 2013, Cobo
Center, Detroit, Michigan, United States, accepted, Paper # 13M-0060.

[10] Saintier, N., Cailletaud, G., and Piques, R., 2006, “Multiaxial fatigue life prediction
for a natural rubber,” International Journal of Fatigue, Vol.28, pp. 530-539.

[11] Mars, W.V. and Fatemi, A., 2004, “Observations of the constitutive response and
characterization of filled natural rubber under monotonic and cyclic multiaxial stress
states,” Journal of Engineering Materials and Technology, Vol.126, pp. 19-28.

[12] Mars W.V., Fatemi A., 2006, “Multiaxial stress effects on fatigue behavior of filled
natural rubber,” International Journal of Fatigue, Vol.28, pp. 521-529.

[13] Mars, W. V. and Fatemi, A., 2002, “A Literature survey on fatigue analysis
approaches for rubber,” International Journal of Fatigue, Vol. 24, pp. 949-961.

[14] Stevenson, A., 1983, “A fracture mechanics study of the fatigue of rubber in
compression,” International Journal of Fracture, Vol. 23, pp. 47-59.

[15] Legorju-jago, K. and Bathias, C., 2002, “Fatigue initiation and propagation in
natural and synthetic rubbers,” International Journal of Fatigue, Vol. 24, pp. 85-92.

[16] Abraham, F., Alshuth, T., and Jerrams, S., 2005, “The effect of minimum stress and
stress amplitude on the fatigue life of non crystallizing elastomers”, Materials and
Design, Vol. 26, pp. 239-245.

[17] Mars, W. V. and Fatemi, A., 2003, “A phenomenological model for the effect of
R ratio on fatigue of strain crystallizing rubbers”, Rubber Chemistry and Technology,
Vol.76, pp. 1241-1258.

146
[18] Harbour, R. J., Fatemi, A., and Mars, W. V., 2007, “Fatigue crack growth of filled
rubber under constant and variable amplitude loading conditions,” Fatigue and Fracture
of Engineering Materials and Structures, Vol. 30, pp. 640-652.

[19] Wang, B., Lu, H., and Kim, G., 2002, “A damage model for the fatigue life of
elastomeric materials,” Mechanics of Materials, Vol. 34, pp. 475-483.

[20] Mars, W.V., 2010, “Identifying the damaging events in a multiaxial duty cycle,”
Constitutive Models for Rubber VI, Heinrich, G., Kaliske, M., Lion, A. and Reese, S.,
Eds., Proceedings of the 6th European Conference on the Constitutive Models for Rubber,
London, Taylor and Francis; 1st edition, pp. 261-267.

[21] Cadwell, S., Merrill, R.A., Sloman, C.M., and Yost, F.L., 1940, “Dynamic fatigue
life of rubber,” Rubber Chemistry and Technology, Vol.13, No.2, pp. 304-315.

[22] Saintier, N., Cailletaud, G., and Piques, R., 2006, “Crack initiation and propagation
under multiaxial fatigue in a natural rubber,” International Journal of Fatigue, Vol.28,
pp. 61-72.

[23] Mars, W.V. and Fatemi, A., 2005, “Multiaxial fatigue of rubber: Part II:
experimental observations and life predictions”, Fatigue and Fracture of Engineering
Materials and Structures, Vol. 28, pp. 523-538.

[24] Mars, W. V. and Fatemi, A., 2004, “Factors that affect the fatigue life of rubber: A
literature survey,” Rubber Chemistry and Technology, Vol. 77, No. 3, pp. 391-412.

[25] Kim, J. H. and Jeong, H. Y., 2005, “A study on the material properties and fatigue
life of natural rubber with different carbon blacks,” International Journal of Fatigue,
Vol. 27, pp. 263-272.

[26] Santangelo, P. G. and Roland, C. M., 1995, “Failure properties of natural rubber
double networks”, Rubber Chemistry and Technology, Vol. 68, pp. 124-131.

[27] Rivlin, R.S., and Thomas, A.G., 1953, “Rupture of Rubber. I. Characteristic energy
for tearing.”Journal of Polymer Science, Vol.10, pp. 291-318.

147
[28] Ayoub, G., Nait-Abdelaziz, M., Zairi, F., Gloaguen, J.M. and Charrier, P., 2011, “A
continuum damage model for the high-cycle fatigue life prediction of styrene-butadiene
rubber under multiaxial loading,” International Journal of Solids and Structures,
Vol.48, pp. 2458-2466.

[29] Wang, Y., Yu, W., Chen, X. and Yan, L., 2008, “Fatigue life prediction of
vulcanized natural rubber under proportional and non-proportional loading,” Fatigue and
Fracture of Engineering Materials and Structures, Vol.31, pp. 38-48.

[30] Mars, W.V. and Fatemi, A., 2004, “A novel specimen for investigating the
mechanical behavior of elastomers under multiaxial loading conditions”, Experimental
Mechanics, Vol. 44, No.2, pp. 136-146.

[31] Mars, W.V. and Fatemi, A., 2006, “Nucleation and growth of small fatigue cracks in
filled natural rubber under multiaxial loading,” Journal of Material Science, Vol.41,
pp. 7324-7332.

[32] Andre, N., Cailletaud, G. and Piques, R., 1999, “Haigh diagram for fatigue crack
initiation prediction of natural rubber components,” Kaut Gummi Kunstst, Vol. 52,
pp. 120-123.

[33] Brunac, J., Gerardin, O., and Leblond, J., 2009, “On the heuristic extension of
Haigh’s diagram for the fatigue of elastomers,” International Journal of Fatigue,
Vol. 31, pp. 859-867.

[34] Mars, W.V. and Fatemi, A., 2005, “Multiaxial fatigue of rubber: Part I: equivalence
criteria and theoretical aspects”, Fatigue and Fracture of Engineering Materials and
Structures, Vol. 28, pp. 515-522.

[35] Zine, A., Benseddiq, N., Nait-Abdelaziz, M., AitHocine, N. and Bouami, D., 2006,
“Prediction of rubber fatigue life under multiaxial loading,” Fatigue and Fracture of
Engineering Materials and Structures, Vol. 29, pp. 267-278.

[36] Harbour, R.J., Mars, W.V. and Fatemi, A., 2008, “Fatigue crack orientation in NR
and SBR under variable amplitude and multiaxial loading conditions,” Journal of
Material Science, Vol.43, pp. 1783-1794.

148
[37] Harbour, R.J., Mars, W.V. and Fatemi, A., 2008, “Fatigue life analysis and
predictions for NR and SBR under variable amplitude and multiaxial loading conditions,”
International Journal of Fatigue, Vol.30, pp. 1231-1247.

[38] Verron, E., Le Cam, J., and Gornet, L., 2006, “A multiaxial criterion for crack
nucleation in rubber,” Mechanics Research Communications, Vol.33, pp. 493-498.

[39] Eshelby, J.D., 1975, “The elastic energy-momentum tensor,” Journal of Elasticity,
Vol. 5, No. 3-4, pp. 321-335.

[40] Flamm, M., Steinweger, T., and Weltin, U., 2003, “Lifetime prediction of
multiaxially loaded rubber springs and bushings,” Constitutive Models for Rubber III,
Busfield, J.J.C. and Muhr, A.H., Eds., Proceedings of the 3rd European Conference on
the Constitutive Models for Rubber, London, Taylor and Francis; 1st edition, pp. 49-53.

[41] Wang, L.R., Lu, Z.H., and Hagiwara, I., 2002, “Finite element simulation of the
static characteristics of a vehicle rubber mount,” Journal of Automobile Engineering,
Vol. 216, Part D, pp. 965-973.

[42] Li, Q., Zhao, J., and Zhao, B., 2009, “Fatigue life prediction of a rubber mount based
on test of material properties and finite element analysis,” Engineering Failure Analysis,
Vol. 16, pp. 2304-2310.

[43] Kim, W.D., Lee, H.J., Kim, J.Y., and Koh, S.K., 2004, “Fatigue life estimation of an
engine rubber mount,” International Journal of Fatigue, Vol. 26, pp. 553-560.

[44] Luo, R.K. and Wu, W.X., 2006, “Fatigue failure analysis of anti-vibration rubber
spring,” Engineering Failure Analysis, No. 13, pp. 110-116.

[45] Luo, R.K., Cook, P.W., Wu, W.X., and Mortel, W.J., 2003, “Fatigue design of
rubber springs used in rail vehicle suspensions,” Journal of Rail and Rapid Transit,
Vol. 217, Part F, pp. 237–240.

[46] Luo, R.K., Mortel, W.J., and Wu, X.P., 2009, “Fatigue failure investigation on
anti-vibration springs,” Engineering Failure Analysis, Vol. 16, pp. 1366-1378.

149
[47] Luo, R.K., Wu, W.X., Cook, P.W., and Mortel, W.J., 2004, “An approach to
evaluate the service life of rubber springs used in rail vehicle suspensions,” Journal of
Rail and Rapid Transit, Vol. 218, Part F, pp. 173-177.

[48] Zhao, J., Li, Q., and Shen, X., 2008, “Finite element analysis and structure
optimization for improving the fatigue life of rubber mounts,” Journal of
Macromolecular Science, Vol. 45, pp. 542-547.

[49] Takeuchi, K., Nakagawa, M., Yamaguchi, H., and Okumoto, T., 1993, “Fatigue test
technique of rubber materials for vibration insulators and their evaluation,” International
Polymer Science and Technology, Vol. 20, No. 10, pp. T64-T69.

[50] Lee, J.S. and Kim, S.C., 2007, “Optimal design of engine mount rubber considering
stiffness and fatigue strength,” Journal of Automobile Engineering, Vol. 221 Part D,
pp. 823–835.

[51] Sundararaman, S., Liang, G., Chandrashekhara, K., Oliver, L.R., and Holmes, S.G.,
2007, “Mode-I fatigue crack growth analysis of V-ribbed belts,” Finite Elements in
Analysis and Design, Vol.43, pp. 870-878.

[52] Steinweger, T., Flamm, M., and Weltin, U., 2003, “A methodology for test time
reduction in rubber part testing,” Constitutive Models for Rubber III,
Busfield, J.J.C. and Muhr, A.H., Eds., Proceedings of the 3rd European Conference on
the Constitutive Models for Rubber, London, Taylor and Francis; 1st edition, pp. 27-32.

[53] Miner, M.A., 1945, “Cumulative damage in fatigue,” Journal of Applied Mechanics,
Vol. 6, pp. 159-164.

[54] Klenke, D. and Beste, A., 1987, “Ensurance of fatigue life of metal-rubber
components,” Kaut Gummi Kunstst, Vol. 40, pp. 1067-1071.

[55] Sun, C., Gent, A., and Marteny, P., 2000, “Effect of fatigue step loading sequence on
residual strength,” Tire Science and Technology, Vol. 28, pp. 196-208.

[56] Flamm, M., Spreckles, J., Steinweger, T., and Weltin, U., 2011, “Effects of very
high loads on fatigue life of NR elastomer materials,” International Journal of Fatigue,
Vol. 33, pp. 1189-1198.

150
[57] Flamm, M., Steinweger, T., and Weltin, U., 2002, “Schadensakkumulation bei
Elastomeren,” Kaustschuk Gummi Kunststoffe, Vol. 55 (12), pp. 665-668.

[58] Roland, C.M. and Sobieski, J.W., 1989, “Anomalous fatigue behaviour in
polyisoprene,” Rubber Chemistry and Technology, Vol. 62, pp. 683-697.

[59] Kim, W., Kim, W., and Hong, S., 2007, “Fatigue life prediction of automotive
rubber component subjected to a variable amplitude loading,” Elastomer, Vol. 42(4),
pp. 209-216.

[60] Wetzel, R. M., ed., 1977, Fatigue under Complex Loading: Analysis and
Experiments, AE-6, SAE, Warrendale, PA.

[61] ASTM Standard D4482-99, 2002, “Standard test method for rubber
property-extension cycling fatigue,” ASTM, West Conshohocken, PA, pp.661-667.

[62] Thomas, A. G., 1958, “Rupture of rubber. V. Cut Growth in Natural Rubber
Vulcanizates,” Journal of Polymer Science, Vol. 31, pp. 467-480.

[63] Colin, J. and Fatemi, A., 2010, “Variable amplitude cyclic deformation and fatigue
behaviour of stainless steel 304L including step, periodic, and random loadings,” Fatigue
and Fracture of Engineering Materials and Structures, Vol. 33, pp. 205-220.

[64] Bose, K., Hurtado, J. A., Snyman, M. F., Mars, W. V., and Chen, J. Q., 2003,
“Modeling of stress softening in filled elastomers,” Constitutive Models for Rubber III,
Busfield, J. J. C. and Muhr, A. H. Eds., Proceedings of the 3rd European Conference on
the Constitutive Models for Rubber, London, Taylor and Francis; 1st edition, pp. 223-230.

[65] Sharma, S., 2003, “Critical comparison of popular hyper-elastic material models in
design of anti-vibration mounts for automotive industry through FEA,” Constitutive
Models for Rubber III, Busfield, J.J.C. and Muhr, A.H., Eds., Proceedings of the 3rd
European Conference on the Constitutive Models for Rubber, London, Taylor and
Francis; 1st edition, pp. 161-167.

[66] Johnson, M. A. and Beatty, M. F., 1993, “The mullins effect in uniaxial extension
and its influence on the transverse vibration of a rubber string,” Continuum Mechanics
and Thermodynamics, Vol. 5, pp. 83-115.

151
[67] Marckmann, G., Verron, E., Gornet, L., Chagnon, G., Charrier, P., and Fort, P.,
2002, “A theory of network alteration for the Mullins effect,” Mechanics and Physics of
Solids, Vol. 50, pp. 2014-2018.

[68] Mars, W. V. and Fatemi, A., 2003, “Fatigue crack nucleation and growth in filled
natural rubber,” Fatigue and Fracture of Engineering Materials and Structures, Vol. 26,
pp. 779-789.

[69] Lindley, P. B., 1972, “Energy for crack growth in model rubber components,” Strain
Analysis for Engineering Design, Vol. 7, pp. 132-140.

[70] Lake, G. J., 1995, “Fatigue and Fracture of Elastomers,” Rubber Chemistry and
Technology, Vol. 68, pp. 435-460.

[71] Mars W. V., 2001, “Multiaxial Fatigue of Rubber,” Ph.D. Dissertation, The
University of Toledo, Toledo, Ohio.

[72] Matsuishi, M. and Endo, T., 1968, “Fatigue of metals subjected to varying stress,”
Japan Society of Mechanical Engineering, Fukuoka, Japan.

[73] Arruda, E.M. and Boyce, M.C., 1993, “A three-dimensional constitutive model for
the large stretch behavior of rubber elastic materials,” Journal of Mechanics Physics
Solids, Vol. 41, pp. 389-412.

[74] Rivlin, R.S., 1992, “The elasticity of rubber,” Rubber Chemistry and Technology,
Vol. 65-3, pp. G51–G66.

[75] ABAQUS 6.9.1 Analysis User’s Manual, 2009.

[76] Ogden, R.W., 1972, “Large Deformation Isotropic Elasticity - On the Correlation of
Theory and Experiment for Incompressible Rubberlike Solids,” Proceedings of the Royal
Society of London. Series A, Mathematical and Physical Sciences, Vol. 326, No. 1567,
pp. 565-584.

152
[77] Serban, D.A., Marsavina, L., and Silberschmidt, V., 2012, “Behaviour of
semi-crystalline thermoplastic polymers: Experimental studies and simulations,” Journal
of Computational Materials Science, Vol. 52, pp. 139-146.

[78] Kilian, H.G., 1981, “Equation of state of real networks,” Polymer, Vol. 22,
pp. 209-217.

[79] Romanov, K.I., 2001, “The Drucker stability of a material,” Journal of Applied
Mathematics and Mechanics, Vol. 65, No. 1, pp. 155-162.

[80] Mars, W.V. and Fatemi, A., 2007, “The correlation of fatigue crack growth rates in
rubber subjected to multiaxial loading using continuum mechanical parameters,” Rubber
Chemistry and Technology, Vol. 80, Issue 1, pp. 169-182.

[81] Ait-Bachir, M., Mars, W.V., and Verron, E., 2012, “Energy release rate of small
cracks in hyperelastic materials,” International Journal of Non-linear Mechanics,
Vol. 47, Issue 4, pp. 22-29.

[82] Lake, G.J. and Lindley, P.B., 1965 “The mechanical fatigue limit for rubber,”
Journal of Applied Polymer Science, Vol. 9, pp. 1233-1251 (reprinted in Rubber
Chemistry and Technology, Vol. 39, pp. 348-364, 1966).

[83] Ellul, M.D., 1992, “Mechanical Fatigue”, Chapter 6 in Engineering with Rubber,
How to Design Rubber Components, Gent, A., 1st Ed., published by Carl Hanser Verlag,
Munich.

[84] Young, D.G., 1986, “Fatigue crack propagation in elastomer compounds: effects of
strain rate, temperature, strain level, and oxidation,” Rubber Chemistry and Technology,
Vol. 59, pp. 809-825.

[85] Charrier, P., Ostoja-Kuczynski, E., Verron, E., Marckmann, G., Gornet, L. and
Chagnon, G., 2003, “Theoretical and numerical limitations for the simulation of crack
propagation in natural rubber components”, Constitutive Models for Rubber III, J.J.C.
Busfield and A.H. Muhr, Eds., Proceedings of the 3rd European Conference on the
Constitutive Models for Rubber, London, Taylor & Francis; 1st edition (2003), pp. 3-10.

153

Anda mungkin juga menyukai