You are on page 1of 264

ASPHALT MIXTURE FATIGUE TESTING

Influence of Test Type and Specimen Size

Ning LI
ASPHALT MIXTURE FATIGUE TESTING

Influence of Test Type and Specimen Size

Proefschrift

ter verkrijging van de graad van doctor


aan de Technische Universiteit Delft,
op gezag van de Rector Magnificus prof. ir. K.C.A.M. Luyben,
voorzitter van het College voor Promoties,
in het openbaar te verdedigen op maandag 18 november 2013 om 10:00 uur

door

Ning LI

Master of Science in Material Science and Engineering


Wuhan University of Technology, P.R. China
geboren te Hubei Province, P.R. China
Dit proefschrift is goedgekeurd door de promotoren:
Prof. dr. ir. A.A.A. Molenaar
Prof. S.P. Wu, BSc., MSc., PhD.

Copromotor
Ir. M.F.C. van de Ven

Samenstelling promotiecommissie:

Rector Magnificus, Technische Universiteit Delft, voorzitter


Prof. dr. ir. A.A.A. Molenaar Technische Universiteit Delft, promotor
Prof. S.P. Wu, BSc., MSc., PhD. Wuhan University of Technology, promotor
Ir. M.F.C. van de Ven Technische Universiteit Delft, copromotor
Prof. dr. ir. S.M.J.G. Erkens Technische Universiteit Delft
Prof. dr. A. Scarpas Technische Universiteit Delft
Prof. dr. ir. H.E.J.G. Schlangen Technische Universiteit Delft
Dr. A. Vanelstraete Belgian Road Research Centre

Published and distributed by:

Ning Li
Road and Railway Engineering Section
Faculty of Civil Engineering and Geosciences
Delft University of Technology
P.O. Box 5048, 2600 GA Delft, the Netherlands
E-mail: lining-008@hotmail.com, lining0618@gmail.com

ISBN: 978-94-6186-235-8

Key words: Fatigue Test; Endurance Limit; Test Type; Specimen Size; Yield Surface;
Safety Factor

Printing: Wohrman Print Service, Zutphen (the Netherlands)

©2013 by Ning Li

All rights reserved. No part of this publication may be reproduced, stored in a retrieval
system or transmitted in any form or by any means, electronic, mechanical, photocopying,
recording, or otherwise without the prior permission of the proprietor.
To my family
Acknowledgements

Looking back to the starting point is helpful for moving forward in a right way. Life is
not easy, especial for a PhD student living in a foreign country. During the five years’
study in Delft, many people gave me their guidance, encouragement and support when I
faced difficulties. Without their efforts, this research would never have been completed.
Therefore, this moment is a good opportunity for me to express my sincere gratitude to
all of them.

This PhD project was originated from the cooperation between the Delft University of
Technology (TUD) and Wuhan University of Technology (WHUT). The research
presented in this dissertation was carried out in the Road and Railway Engineering
Section of the Faculty of Civil Engineering and Geosciences at the TUD. Firstly, the
appreciation goes to those who built up the cooperative link between TUD and WHUT.
The author also would like to thank the financial supports from the China Scholarship
Council (CSC) and the TUD.

I would like to express my deepest appreciation to my promoter, Prof.dr.ir. A.A.A.


Molenaar. He always provided me with the valuable guidance and encouragement
throughout every stage of my PhD study. I enjoyed the great benefit from not only his
academic attainments but also his wisdom of life. His tireless efforts and constructive
comments on this dissertation are highly appreciated. At the same time, sincere gratitude
goes to my promoter Prof. Shaopeng Wu, who supervised my bachelor, master and PhD
study. He advised me to start my PhD study abroad and let me expand my field of vision.
His professional knowledge and experience in road industry led me into the road
engineering field.

I am grateful for my daily supervisors Associate Professor Martin van de Ven. He always
gave his patient guidance whenever I need. His careful review on my papers, reports and
dissertation are deeply appreciated. I would like to extend my sincere gratitude to Ir. A.C.
Pronk for his contribution on the calibration and the modeling work in this research.
Under his guidance, I adapted myself to the new study environment quickly and did not
feel fear of those complex equations. Even after his retirement, he still came over and
discussed with me when I need help. Thanks so much for all your efforts. My sincere
appreciation goes to Prof.dr. R.L. Lytton for his arrangement and guidance when I was
doing the project in Texas A&M University. Also many thanks to Dr. Rong Luo for her
kind help during my stay in Texas. I would like to appreciate Prof. Halil Ceylan from
Iowa State University. His valuable suggestions on the data analysis are highly
acknowledged. I also would like to thank Dr. Milliyon Woldekidan for the valuable
discussion and kind help with the finite element analysis.

I would like to thank all the colleges and former colleges of Road and Railway
Engineering Section. The extensive laboratory work could not have been successfully
finished without the arrangement and support provided by Associate professor Lambert
Houben, Abdol Miradi, Marco Poot, Jan-Willem Bientjes, Jan Moraal and Dirk Doedens.

i
Particularly, I am grateful to Marco Poot. His rich experience of the test equipments
makes the test program smooth and efficient. Many thanks to the secretariats, Jacqueline
Barnhoorn and Sonja van de Bos for their kind help with the daily administration affaires.
I would like to thank Prof. Tom Scarpas, Prof. Sandra Erkens, Prof. Rolf Dollevoet, Dr.
Rien Huurman, Accociate Professor Zili Li and Dr. Xueyan Liu. Their support and
guidance are deeply appreciated. Many thanks to all the PhD students, liantong Mo, Jian
Qiu, Gang Liu, Xin Zhao, Eyassu, Alemgena Araya, Yue Xiao, Jingang, Sadegh,
Mohamad, Oscar, Marija, Maider, Shaoguang, Nico, Dongya, Pengpeng, Chang, Haoyu,
Lizuo, Xiangyun and all the new PhD students. I really enjoy the time here with all of
them. Special thank goes to my officemates Diederik van Lent, Yuan Zhang and former
officemate Dongxing. I feel fortunate that I have sit in the Room 1.29 for 5 years. I will
never forget the moments when we were talking about our different cultures and
languages, when we were sharing our knowledge and experiences and when we were
playing table tennis.

I appreciate the wonderful time that I have spent together with all my friends: Associate
Prof. Ye, Xu Jiang, Xuhong Qiang, Quantao Liu, Hailing Zhang, Xuming Shan, Ying
Wang, Juan Tong, Lili Wu, Huanhuan Mao, Nannan Li, Lin Liu, Huisu Chen, Zhiwei
Qian, Yuguang Yang, Qi Zhang, Jinlong Li, Haoliang Huang, Bei Wu, Jiayi Chen, Yun
Zhang, Yong Zhang, Yuan Qiu, etc. Thanks all of them for giving me kind help and
enjoyable time.

Finally, my deepest gratitude goes to my family for their endless love and continues
support. My parents have never complained how less time I have spent with them since I
left home for my study. I appreciate their encouragement and understanding. My special
thanks to my wife Zhuqing Yu for her continuous support, patience and optimistic
attitude. No matter what happen, she always stand firm behind me and take good care of
my life even when she was also busy with her PhD study. This dissertation is dedicated to
my family.

Ning Li

李宁

October, 2013, Delft

ii
Summary

Fatigue characterization of an asphalt mixture is commonly estimated by laboratory


fatigue tests. Based on the classical fatigue analysis, fatigue lives obtained from different
test devices are not comparable even when they are performed at the same test conditions.
It is believed that there are two main reasons causing the difference in fatigue results
being the difference in stress-strain distribution of the different specimens and the fatigue
analysis approach. This research focuses on the harmonization of the fatigue results
obtained from the different methods, which are recommended by the European standard
EN 12697-24. The main goal is to find a correlation between the different fatigue test
methods and to improve the classical fatigue analysis approach to better represent the
actual fatigue characteristics of asphalt mixtures.

To realize the main objective of the study, an extensive fatigue testing program was
carried out on dense asphalt concrete 0/8 (DAC 0/8). In the program, uniaxial tension and
compression (UT/C) fatigue tests, four-point bending (4PB) fatigue tests and indirect
tensile (IT) fatigue tests were performed. For each fatigue test, specimens with different
sizes were tested to explore the size effect on the fatigue results. In order to limit the test
program, the tests were performed in two loading modes, at two temperatures and one
frequency.

For the 4PB fatigue test, the measured displacement highly depends on the properties of
the loading frame. Calibration tests on the 4PB test setup were conducted to obtain the
pure bending deflection of the beam.

Comparison of the fatigue results obtained with the different test methods at the same test
condition and loading mode shows that the fatigue life from the 4PB test is the longest
and from the IT test is the shortest. Because of the homogeneous tensile strain field, the
UT/C and IT fatigue results are not significantly influenced by the specimen size.
However, the 4PB test results depend on the dimension of the used specimen, because the
stress-strain field of the beam specimen varies along the length and cross section.

The partial healing (PH) model was used to determine the relationship between the UT/C
and 4PB fatigue results in strain-controlled mode. It is a material model that describes the
evolutions of the stiffness and phase angle for a unit volume during testing. This implies
that the model can be directly applied to the fatigue results obtained from the
“homogenous” tests, the UT/C test, because the strain is uniformly distributed throughout
the specimen. When analyzing the 4PB results, the backcalculated stiffness is not the
local stiffness of the material but the so-called weighted overall stiffness of the whole
specimen. Therefore a weighing procedure is required to calculate the weighted overall
stiffness modulus from the local stiffness by taking into account the dimensions and the
strain distribution of the specimen. By adjusting the model parameters, the PH model
provides a good simulation for the evolutions of the stiffness and phase angle. All the
model parameters can be expressed as functions of the applied strain level. The trends of
the parameter δγ1 and δγ2 indicate the existence of an endurance limit. The predicted

iii
endurance limit obtained by the UT/C test is around 68 µm/m for the tested DAC 0/8
mixture, which does not change with specimen size and temperature. For the 4PB fatigue
test, the local stiffness at different parts of the beam can be calculated by means of the PH
model. The evolution of the calculated stiffness at the surface in the midsection of the
beam is comparable with the UT/C fatigue results when the pure bending strain on the
beam surface is equal to the tensile strain in the cylinder. Therefore the PH model offers a
possibility to compare different fatigue results.

In the second part of this research, the yield surface concept was applied to develop a
new fatigue analysis approach. As a visco-elastic material, the yield surface of an asphalt
mixture highly depends on temperature and strain rate. Therefore, monotonic uniaxial
compression (MUC) and monotonic uniaxial tension (MUT) tests were performed at
different temperatures and strain rates to derive such yield surfaces. For the three fatigue
tests, the yield surface at the critical location of the specimen was determined in the I1-√J2
space. The fatigue results were then interpreted by comparing the actual stress condition
with the yield surface.

A new parameter R∆ was introduced as an indicator of the “safety against failure”. By


comparing the R∆ values at the different locations of the specimen for the IT test, the
weakest points are found at the locations with the maximum horizontal tensile strain,
which are close to the loading strips, instead of the center of the specimen. A straight line
was found by plotting R∆ at the critical location and the fatigue life on a log-log scale.
Compared to the traditional fatigue analysis, the size effect on the fatigue results was
excluded by using this new fatigue relation. For the stress-controlled mode, the fatigue
lines obtained from the UT/C test show a good agreement with the IT fatigue results. Of
course the influence of temperature and loading mode still exists in this new fatigue
method. In the normal coordinate, when the fatigue life tends to infinity, R∆ becomes a
constant value, denoted by Rlimit. The parameter Rlimit represents the endurance limit in a
three-dimensional state. This value does not change with specimen size and test type but
is influenced by the temperature and loading mode.

Based on all the test results and their analysis in this research, it was concluded that the
UT/C fatigue results are material properties and not influenced by specimen size. The PH
model provides a good simulation of fatigue behavior of the asphalt mixture and is able
to find the correlation between the UT/C and 4PB test results. The developed fatigue
analysis approach characterizes the fatigue performance of asphalt mixtures in three-
dimensional state, which is more close to the field situation. The endurance limit
predicted by the new fatigue approach is independent of the specimen size and fatigue
test type.

iv
Samenvatting

Het vermoeiingsgedrag van asfaltmengsels wordt in het algemeen bepaald door middel
van vermoeiingsproeven in het laboratorium. Wanneer de klassieke vermoeiingsanalyse
wordt gebruikt, zijn vermoeiingslevensduren verkregen met verschillende test types niet
vergelijkbaar, zelfs niet wanneer zij worden uitgevoerd bij dezelfde proefomstandigheden.
Aangenomen wordt dat er twee belangrijke oorzaken zijn voor het verschil in
vermoeiingsresultaten, namelijk het verschil in spanning-rek verdeling in de
verschillende proef types en de manier waaropn de analyse van de resultaten wordt
uitgevoerd. Dit onderzoek richt zich op de harmonisatie van de vermoeiings resultaten
van een aantal methoden die worden aanbevolen door de Europese norm EN 12697-24.
Het belangrijkste doel is om een correlatie tussen de verschillende vermoeiingsproeven te
vinden en de klassieke vermoeiingsanalyse te verbeteren teneinde de werkelijke
vermoeiingseigenschappen van asfaltmengsels zo goed mogelijk te kunnen bepalen.

Om het hoofddoel van de studie te realiseren is een uitgebreid vermoeiingsonderzoek


uitgevoerd op dicht asfaltbeton 0/8 (DAC 0/8). In het onderzoeksprogramma zijn
uniaxiale trek/druk vermoeiingsproeven (UT/C), vierpuntsbuigingsproeven (4PB) en
indirecte trekproeven (IT) uitgevoerd. Voor elk proeftype zijn proefstukken met
verschillende afmetingen getest om het effect van de afmetingen op de
vermoeiingsresultaten te onderzoeken. Om het testprogramma te beperken, zijn de
proeven uitgevoerd met twee verschillende belastingswijzen (constante spanning en
constante rek), bij twee temperaturen en met één frequentie.

Voor de 4PB vermoeiingsproef is de gemeten verplaatsing sterk afhankelijk van de


eigenschappen van het frame van de proeftopstelling. Om deze reden zijn
kalibratieproeven op de 4PB opstelling uitgevoerd om de zuivere doorbuiging van de
balk te kunnen bepalen.

Bij dezelfde proefomstandigheden en belastingwijze, is de vermoeiingslevensduur in de


4PB proef het langste en die in de IT-test de kortste. Vanwege het homogene
trekspanningsveld worden de UT/C en IT vermoeiingsresultaten niet significant
beïnvloed door de afmetingen van de proefstukken. Echter, de 4PB resultaten zijn wel
afhankelijk van de hoogte van de gebruikte balk, omdat het spanning-rekgebied in de
balk varieert over de lengte- en dwarsdoorsnede.

Het partial healing (PH) model is gebruikt om de relatie tussen de UT/C en 4PB
vermoeiingsresultaten met constante rek te bepalen. Het is een materiaalmodel dat de
ontwikkeling van de stijfheid en fasehoek van een volume-eenheid tijdens de
vermoeiingsproef beschrijft. Dit impliceert dat het model direct toepasbaar is op de
vermoeiingsresultaten van de "homogene" proeven, zoals de UT/C -test, omdat de rek
gelijkmatig over het gehele proefstuk verdeeld is. Bij het analyseren van de 4PB
resultaten, is de uit de maximale doorbuiging van de balk teruggerekende stijfheid niet de
werkelijke stijfheid van het materiaal, maar een gewogen stijfheid van het gehele
proefstuk. De werkelijke stijfheid van het materiaal varieert nl over de lengte en de

v
hoogte van de balk. Daarom is een wegingsmethode nodig om de gewogen
stijfheidsmodulus te berekenen uit de plaatselijke stijfheid door rekening te houden met
de afmetingen en de rekdistributie in het proefstuk. Door het aanpassen van de
modelparameters, kan met het PH-model de ontwikkeling van de stijfheid en fasehoek
worden gesimuleerd. Alle modelparameters kunnen worden uitgedrukt als functie van de
aangebrachte rekniveaus. De trends van de parameters δγ1 en δγ2 wijzen op het bestaan
van een vermoeiingsgrens. De voorspelde vermoeiingsgrens voor de UT/C proeven is
ongeveer 68 µm / m voor het geteste DAC 0/8 mengsel, en deze blijkt onafhankelijk te
zijn van de afmetingen van het proefstuk en de temperatuur. Voor de 4PB proeven kan de
ontwikkeling van de lokale stijfheid op iedere plaats in het proefstuk worden berekend
met behulp van het PH model. De ontwikkeling van de lokaal berekende stijfheid aan het
oppervlak in het middelste gedeelte van de balk is vergelijkbaar met de UT/C
vermoeiingsresultaten wanneer de zuivere buigingsrek aan het oppervlak van de balk
gelijk is aan de trekrek in de cilinder. Hiermeebiedt het PH-model de mogelijkheid om de
vermoeiingsresultaten van verschillende types proeven met elkaar te vergelijken.

In het tweede deel van dit onderzoek is het yield surface concept toegepast bij de analyse
van de vermoeiingsresultaten. Voor een visco-elastisch materiaal zoals een asfaltmengsel,
is de yield surface sterk afhankelijk van de temperatuur en de reksnelheid. Om deze reden
zijn monotone uniaxiale druk- (MUC) en monotone uniaxiale trek (MUT) proeven
uitgevoerd bij verschillende temperaturen en reksnelheden om de contouren van de yield
surface te bepalen. Voor de drie types vermoeiingsproeven zijn de kritieke locaties van
het proefstuk bepaald in de zogenaamde I1-√J2 ruimte. De vermoeiingsresultaten zijn
vervolgens geïnterpreteerd door de werkelijke spanningstoestand te vergelijken met de
driedimensionale yield surface voor de diverse proeftypes.

Een nieuwe parameter R∆ isgeïntroduceerd als indicator van de "veiligheid tegen


bezwijken". Door de R∆ waarden te vergelijken die voor verschillende locaties in het IT
proefstukr zijn berekend zijn de locaties gevonden waar falen het eerst op zal treden.
Deze bevinden zich dichtbij de belastingstrippen, in plaats van in het midden van het
proefstuk. Op log-log schaal is een rechte lijn gevonden bij het plotten van R∆ op de
kritieke locatie en de vermoeiingslevensduur. Vergeleken met de traditionele
vermoeiingsanalyse, ishet effect van de afmetingen op de vermoeiingsresultaten
geëlimineerd bij gebruik van deze nieuwe vermoeiingsrelatie. Voor de
spanningsgestuurde proef laten de vermoeiingslijnen verkregen uit de UT/C -proef een
goede overeenkomst zien met de IT vermoeiingsresultaten. Maar de invloed van de
temperatuur en belastingswijze blijft aanwezig, ook in deze nieuwe analysemethode.
Wanneer de vermoeiingslevensduur naar oneindig gaat blijkt R∆ een constante waarde
aan te nemen, die aangeduid is met Rlimit. De parameter Rlimit vertegenwoordigt de
endurance limit in de driedimensionale toestand. Deze waarde verandert niet met
proefstukrgrootte en type proef, maar wordt wel beïnvloed door de temperatuur en
belastingswijze.

Op basis van alle resultaten en hun analyse in dit onderzoek kan worden geconcludeerd
dat de eigenschappen zoals bepaald met de UT/C proef "echte" materiaaleigenschappen
zijn die niet beïnvloed worden door de proefstukafmetingen. Het PH model simuleert het

vi
vermoeiingsgedrag van het asfaltmengsel uitstekend en met dit model kan de correlatie
tussen de UT/C en 4PB proefresultaten worden bepaald. De yield surface benadering
maakt het mogelijk de vermoeiingsprestaties van asfaltmengsels in een driedimensionale
spanningstoestand te bepalen, vergelijkbaar met die welke in de praktijk optreden. De
endurance limit die wordt voorspeld met de nieuwe vermoeiingsaanpak is onafhankelijk
van de afmetingen van het proefstuk en het type vermoeiingsproef.

vii
viii
List of Abbreviations

HMA Hot Mix Asphalt


2PB Two-Point Bending
3PB Three-Point Bending
4PB Four-Point Bending
ITFT Indirect Tensile Fatigue Test
NAT Nottingham Asphalt Tester
UT/C Uniaxial Tension and Compression
IT Indirect Tension
S-N Stress amplitude-Life
AC Asphalt Concrete
TRRL Transport and Road Research Laboratory
CEN European Committee for Standardization
PI Penetration Index
CF Correction Factor
ER Energy Ratio
RDEC Ratio of Dissipated Energy Change
PV Plateau Value
GAC Gravel Asphalt Concrete
PH Partial Healing
2D Two-Dimensional
3D Three-Dimensional
ACRe Asphalt Concrete Response
DAC Dense Asphalt Concrete
SA Sand Asphalt
CDAS Control Data Acquisition System
UTM Universal Testing Machine
FFT Fast Fourier Transform
LVDT Linear Variable Differential Transformer
CDAS Control and Data Acquisition System
WLF Williams-Landel-Ferry
DER Dissipated Energy Ratio
MUC Monotonic Uniaxial Compression
MUT Monotonic Uniaxial Tension
EME Enrobé á Modele Elevé
SMA Stone Mastic Asphalt
GAC Gravel Asphalt Concrete
MPRE Mean Percent Relative Error
l.o.c. Level of Confidence
HISS Hierarchical Single Surface
FEM Finite Element Modelling
GUI Graphical User Interface
ABAQUS A Finite Element Package
RAW Dutch standard Specification for the Civil Engineering Sector

ix
PReSBOX A Laboratory Asphalt Compactor
COD Cracks Opening Displacement
ASTM American Society for Testing and Materials
AASHTO American Association of State Highway and Transportation Officials
RILEM International Union of Laboratories and Experts in Construction Materials

x
Table of Contents

Acknowledgements ............................................................................................................ i

Summary........................................................................................................................... iii

Samenvatting ..................................................................................................................... v

List of Abbreviations ....................................................................................................... ix

Chapter 1 Introduction.................................................................................................. 1
1.1 Fatigue Damage in Asphalt Pavements ................................................................ 1
1.2 Fatigue Failure Mechanism................................................................................... 2
1.3 Evaluation of Fatigue Properties........................................................................... 3
1.4 Problem Statement and Objectives ....................................................................... 4
1.5 Organization of the Dissertation ........................................................................... 6
References................................................................................................................... 8

Chapter 2 Literature Review ...................................................................................... 11


2.1 Background of Fatigue Research on Asphalt Mixtures ...................................... 11
2.2 Laboratory Fatigue Test Methods....................................................................... 12
2.2.1 Simple Flexure Test ................................................................................. 12
2.2.1.1 Two-Point Bending (2PB) Test .................................................... 12
2.2.1.2 Three-Point Bending Test ............................................................. 13
2.2.1.3 Four-Point Bending Test............................................................... 14
2.2.1.4 Rotating Bending Test .................................................................. 15
2.2.2 Direct Axial Loading Test ....................................................................... 16
2.2.3 Diametral Loading Test (Indirect Tensile Test)....................................... 17
2.3 Influence of Test Type, Specimen Size and Test Conditions on Fatigue Results
................................................................................................................................... 18
2.3.1 Influence of Test Type and Specimen Size.............................................. 19
2.3.1.1 Influence of Test Type .................................................................. 19
2.3.1.2 Influence of Specimen Size .......................................................... 21
2.3.2 Influence of Loading Mode ..................................................................... 22
2.4 Fatigue Analysis Approach................................................................................. 26
2.4.1 Results of Laboratory Fatigue Tests ........................................................ 26
2.4.2 Classical Fatigue Analysis ....................................................................... 29
2.4.3 Dissipated Energy Approach ................................................................... 34
2.4.3.1 Dissipated Energy Theory............................................................. 34
2.4.3.2 Cumulative Dissipated Energy ..................................................... 35
2.4.3.3 Dissipated Energy Ratio ............................................................... 37
2.4.4 Fracture Mechanics Approach ................................................................. 41
2.4.4.1 Theory ........................................................................................... 41
2.4.4.2 Determination of Fracture Mechanics Parameters........................ 43

xi
2.5 Fatigue Prediction Model.................................................................................... 45
2.5.1 Partial Healing (PH) Model ..................................................................... 45
2.5.2 Weibull’s Theory ..................................................................................... 47
2.5.2.1 Survival Probability ...................................................................... 47
2.5.2.2 Calculation of Survival Probability .............................................. 48
2.5.2.3 Application in Fatigue Test........................................................... 50
2.5.3 Mechanical Damage Model ..................................................................... 52
2.5.3.1 Theory ........................................................................................... 52
2.5.3.2 Size Effect of the Damage Model................................................. 53
2.5.3.3 Comparison between Model and Experiments ............................. 54
2.6 Summary ............................................................................................................. 55
References................................................................................................................. 56

Chapter 3 Research Methodology .............................................................................. 65


3.1 Introduction......................................................................................................... 65
3.2 Research Methodology ....................................................................................... 66
References................................................................................................................. 68

Chapter 4 Mixture Design and Specimen Preparation ............................................ 69


4.1 Selection of the Mixture...................................................................................... 69
4.2 Mixture Design ................................................................................................... 71
4.2.1 Materials .................................................................................................. 71
4.2.2 Mixture Design ........................................................................................ 72
4.3 Specimen Preparation ......................................................................................... 74
4.3.1 Mixture Compaction ................................................................................ 74
4.3.2 Selection of Specimen Size...................................................................... 76
4.3.3 Specimen Preparation .............................................................................. 77
References................................................................................................................. 82

Chapter 5 Different Laboratory Fatigue Experiments............................................. 83


5.1 Introduction......................................................................................................... 83
5.2 Test Equipment ................................................................................................... 83
5.2.1 Uniaxial Tension and Compression (UT/C) Test .................................... 83
5.2.2 Four-Point Bending (4PB) Test ............................................................... 84
5.2.3 Indirect Tensile (IT) Fatigue Test ............................................................ 90
5.3 Calibration of the 4PB Test Equipment .............................................................. 90
5.3.1 Theory ...................................................................................................... 91
5.3.2 Test Program............................................................................................ 93
5.4 Complex Modulus and Fatigue Tests ................................................................. 96
5.4.1 Complex Modulus Test............................................................................ 96
5.4.2 Fatigue Test.............................................................................................. 97
5.5 Data Processing................................................................................................... 98
5.5.1 UT/C Test................................................................................................. 98
5.5.2 4PB Test................................................................................................... 99
5.5.3 IT Test.................................................................................................... 101

xii
5.6 Test Results....................................................................................................... 102
5.6.1 Stiffness Results..................................................................................... 102
5.6.2 Fatigue Test Results............................................................................... 108
5.7 Summary ........................................................................................................... 130
5.7.1 Test Setup............................................................................................... 130
5.7.2 Test Results............................................................................................ 131
References............................................................................................................... 132

Chapter 6 Application of Partial Healing Model on Strain Controlled Fatigue


Tests .................................................................................................................... 133
6.1 Introduction....................................................................................................... 133
6.2 Application of PH Model on UT/C Test Results .............................................. 133
6.2.1 PH Model Theory .................................................................................. 133
6.2.2 Determination of PH Model Parameters................................................ 139
6.3 Application of PH Model on 4PB Test Results ................................................ 144
6.3.1 Weighing Procedure for 4PB Test ......................................................... 144
6.3.2 Determination of PH Model Parameters................................................ 148
6.4 Correlation between UT/C and 4PB Fatigue Test Results ............................... 153
6.5 Conclusions....................................................................................................... 155
References............................................................................................................... 157

Chapter 7 Monotonic Uniaxial Tension and Compression Tests .......................... 159


7.1 Introduction....................................................................................................... 159
7.2 Test Equipment ................................................................................................. 159
7.2.1 Monotonic Uniaxial Compression Test ................................................. 159
7.2.2 Monotonic Uniaxial Tension Test ......................................................... 161
7.3 Test Condition................................................................................................... 162
7.3.1 Central Composition Rotatable Design ................................................. 162
7.3.2 Test Conditions for MUC and MUT Tests ............................................ 163
7.4 Test Procedure .................................................................................................. 164
7.4.1 Test Procedure for MUC Tests .............................................................. 164
7.4.2 Test Procedure for MUT Tests .............................................................. 165
7.5 Data Processing................................................................................................. 166
7.5.1 Stress and Strain..................................................................................... 166
7.5.2 Strain Rate, Maximum Stress, Tangent Stiffness and Onset of Dilation168
7.6 Test Results....................................................................................................... 169
7.6.1 MUC Test Results.................................................................................. 169
7.6.2 MUT Test Results .................................................................................. 175
7.7 The Unified Model............................................................................................ 177
7.8 Prediction of the Unified Model Parameters .................................................... 180
7.9 Conclusions....................................................................................................... 195
References............................................................................................................... 196

xiii
Chapter 8 Yield Surface and Fatigue Tests ............................................................. 197
8.1 Introduction....................................................................................................... 197
8.2 Material Model Concept ................................................................................... 197
8.3 Determination of Model Parameters................................................................. 202
8.3.1 Model Parameters R and γ ..................................................................... 202
8.3.2 Model Parameter n ................................................................................. 205
8.3.3 Model Parameter α................................................................................. 205
8.3.4 Determination of Yield Surface ............................................................. 207
8.4 Critical Location ............................................................................................... 209
8.5 Determination of Yield Surface for Fatigue Test ............................................. 212
8.6 Relationship between R∆ and Fatigue Life ....................................................... 215
8.7 Conclusions....................................................................................................... 225
References............................................................................................................... 226

Chapter 9 Conclusions and Recommendations....................................................... 227


9.1 Conclusions....................................................................................................... 227
9.1.1 Conclusions Related to Literature Review ............................................ 227
9.1.2 Conclusions Related to Fatigue Test Equipment ................................... 227
9.1.3 Conclusions Related to Stiffness and Fatigue Results........................... 228
9.1.4 Conclusions Related to Partial Healing Model...................................... 228
9.1.5 Conclusions Related to Monotonic Test Results ................................... 229
9.1.6 Conclusions Related to Yield Surface Approach................................... 230
9.2 Recommendations............................................................................................. 230
9.2.1 Recommendations Related to Experimental Work................................ 230
9.2.2 Recommendations Related to Partial Healing Model ............................ 231
9.2.3 Recommendations Related to Yield Surface Approach......................... 231

Appendix A Calculations for 4-Point Bending Test................................................... 233

Appendix B Determination for Prony Series Model Parameters ............................. 239

xiv
Chapter 1 Introduction

Chapter 1 Introduction

1.1 Fatigue Damage in Asphalt Pavements


Fatigue is defined as the phenomenon of deterioration of a material (reduction in stiffness
and strength, ending in fracture) under repeated loading. Similar to other materials, the
stiffness and strength of asphalt concrete decrease when it is subjected to repetitive
loading [Pell, 1962].

Because of its cost efficiency, reduction in traffic noise, improved safety and comfort and
so on, asphalt concrete has been widely used in pavement structures since the beginning
of the last century [Hveem and Davis, 1950] [Hindley, 1971]. With the increase of traffic
volume and weight, fatigue cracking of the bituminous layer has become one of the major
distress modes in flexible road pavements associated with repeated traffic loads. Fatigue
cracks decrease the structural capacity of the pavement and increase the maintenance cost.
Furthermore, once fatigue cracks propagate through the entire asphalt thickness, water
and aggressive agents can infiltrate into the unbound layers, which greatly accelerates the
deterioration process. Therefore, understanding the fatigue cracking phenomenon and
measuring the fatigue properties of asphalt concrete is essential for the design of flexible
pavements. Figure 1-1 shows some examples of fatigue failure in pavements.

Figure 1-1 Fatigue cracks at the surface of asphalt pavement

Fatigue cracking is associated with repetitive traffic loading and pavement thickness
[Roberts et al., 1996] [McGennis et al., 1994]. Traditionally, it is believed that cracking
initiates at the bottom of the hot mix asphalt (HMA) layers where the tensile stress is the
highest, then migrates upward toward the surface where it shows up as one or more
longitudinal cracks. This is commonly referred to as "bottom-up" or "classical" fatigue
cracking. Researchers however found that in thick pavements (≥160 mm), the cracks
most likely initiate from the top in areas of high localized tensile stresses resulting from
tire-pavement interaction and asphalt binder aging. They then propagate down to a depth
of approximately 50 mm; this is called top-down cracking [Molenaar, 1983] [Gerritsen,
1987] [Molenaar, 2004] [Uhlmeyer, 2000]. Up till now, the mechanism of initiation and
progression of top-down cracks is not thoroughly understood. It is commonly assumed
that high concentration of stresses at the tire-pavement contact surface is the major cause
for these cracks.

1
Chapter 1 Introduction

1.2 Fatigue Failure Mechanism


In order to have a better understanding of the fatigue cracking mechanism in asphalt
pavements, it is necessary to look at how the wheel load is applied to a pavement
structure. Figure 1-2 shows the stress states developed in a particular element of the
pavement structure when subjected to a moving wheel load.

Moving Wheel Load


Stress
Vertical Stress
(compressive)
Shear
Stress
Horizontal Stress
Pavement Vertical Stress (compressive)
Structure

Typical
Element Time

Horizontal
Stress Horizontal Stress
(tensile at the bottom
Shear
of stiff layer)
Stress

(a) (b)

Figure 1-2 Stresses induced by a moving wheel load on a pavement element


[Brown, 1978]

The moving wheel causes vertical, horizontal and shear stresses on an element
underneath the wheel. As a result of the passing wheel load, the vertical compressive
stress changes following a half sine wave, but in horizontal direction the element is
subjected to a tensile or compressive stress alternatively. The bottom-up cracking is
mainly caused by the horizontal tensile stress at the bottom of the asphalt layer. Being a
visco-elastic material, the properties of asphalt mixtures are time dependent which will
have an effect on the magnitude of the tensile strains developed in the structure. The
loading time depends on vehicle speed and the depth below the pavement surface [Collop,
1995]. For example, at a velocity of 60 km/h, the loading time will be approximately
0.015 s at a depth of 150 mm. In addition to the loading frequency, environmental
conditions, engineering properties of the asphalt concrete, the condition of underlying
layers, and the pavement structure are all contributing factors to fatigue cracking.

Roads do not crack immediately after traffic starts to use the pavement. It usually takes
years or millions of load applications from vehicle tires. The sustained traffic loading
results in a decrease in the structural strength of the pavement. If the tensile stress
exceeds the local tensile capacity of the material, eventually cracking will occur. In the
beginning, cracking manifests itself as a series of parallel longitudinal cracks (cracks
along the direction of the flow of traffic) in the top layer of the asphalt pavement. These
cracks are initially thin and sparsely distributed. If further deterioration continues, these

2
Chapter 1 Introduction

longitudinal cracks are connected by transverse cracks forming many sided, sharp-angled
pieces. This interlaced cracking pattern resembles chicken wire or the skin of an alligator.
Figure 1-3 shows examples of the different levels of severity for fatigue cracking.

Figure 1-3 Alligator crack patterns of differing severity [Miller, 2003]

1.3 Evaluation of Fatigue Properties


The fatigue properties of asphalt mixtures are important parameters in pavement design.
In order to determine the fatigue resistance of asphalt mixtures, various fatigue tests are
carried out in the laboratory at the stress levels, loading times, rest periods and
temperature and moisture conditions as realistic as possible. Then the fatigue
characteristics of an asphalt mixture derived from the laboratory tests are used as input in
the design analysis to predict field performance.

Figure 1-4 Schematic demonstrating the main configurations of fatigue tests [Read, 1996]
(a) two point bending; (b) four point bending; (c) three point bending; (d) rotating
bending; (e) direct axial loading; (f) direct axial loading (necked specimen);
(g) diametral loading
3
Chapter 1 Introduction

Fatigue tests for asphalt mixtures were developed as early as the fifties of the last century,
and different test configurations have emerged since that time. According to the mode of
loading the most commonly used tests are classified as simple flexural tests (rotating
cantilever, 2-, 3- and 4-point bending), direct axial loading test and diametral loading test
[Read, 1996] [Tangella, 1990]. Figure 1-4 gives a schematic representation of the
different fatigue tests. The arrows in the figure indicate the direction of the applied
loading. A detailed description of each fatigue test will be given in the literature review.

During fatigue testing, the specimen is subjected to a repeated load at a certain


temperature and frequency. In the classical fatigue analysis, fatigue failure is determined
based on the stiffness reduction. For the strain-controlled mode, the failure point is
defined as the moment at which the stiffness of the specimen has reduced to 50% of its
initial value. For the stress-controlled mode, the failure point is defined as the moment
when the specimen has completely fractured. The results of fatigue tests can be
interpreted in terms of a relationship between the stress or strain and the number of cycles
to failure, which is represented by means of a straight line in a double logarithmic
coordinate system [Pell, 1962].

N f = k ⋅ δ −b (1-1)
where: Nf : number of cycles to failure;
δ : applied strain level [µm/m] or stress level [MPa];
k and b : coefficients related to the material properties.

1.4 Problem Statement and Objectives


As mentioned in Section 1.3, various fatigue test devices are currently used to evaluate
the fatigue performance of asphalt concrete. The two-point bending (2PB) test with
trapezoidal specimens was adopted by researchers from Shell [van Dijk, 1975] and LCPC
[Bonnot, 1986]. The Shell Laboratory at Amsterdam also has used the three-point
bending loading equipment to estimate the fatigue life [van Dijk, 1975]. In the USA
[SHRP, 1992] and the Netherlands [Pronk, 1996], the four-point bending test (4PBT) is
specified. In the UK and Sweden, the standard fatigue test is the indirect tensile fatigue
test (ITFT). The Nottingham Asphalt Tester (NAT) was specially designed for this test
[Brown, 1995]. In European standard [EN 12697-24], three bending tests and the indirect
tensile test are allowed as the standard fatigue test methods. Inevitably, they all give
different results for the same material. Di Benedetto et al. reported an inter-laboratory
investigation organized by the RILEM 182-PEB Technical Committee [Di Benedetto,
2004]. More than 150 fatigue tests were carried out using eleven different types of test
equipment, including uniaxial tension/compression (UT/C), bending and indirect-tension
tests. The classic fatigue relations obtained from the different fatigue tests are shown in
Figure 1-5.

4
Chapter 1 Introduction

Nf50 = α*ε^(-β)

1,E+09

1,E+08
3PB

1,E+07
Nf50
1,E+06
T/C 4PB
1,E+05
2PB

1,E+04
ITT
1,E+03
80 100 110 ε0 (µm/m)

Figure 1-5 Classic fatigue relations obtained from the different fatigue tests
[Di Benedetto, 2004]

Fatigue results are very sensitive to load conditions and the used test type. The Indirect
Tension Test (ITT) shows the shortest fatigue life due to accumulation of permanent
deformation in addition to fatigue damage. For a given strain amplitude, the beam tests
generally result in longer fatigue life compared to T/C (Tension/Compression) tests. The
conclusion was that fatigue test results obtained from different test equipment are not
comparable to each other. It is therefore desirable to find a way to harmonize the fatigue
results obtained from the different test methods and improve the classic fatigue analysis
approach to better represent the actual fatigue characteristics of an asphalt mixture. In
order to achieve these goals, the following objectives were defined in this research:

(1) Compare the fatigue results obtained from representative fatigue test methods and
analyze their failure mechanisms.

(2) Explore the influence of the specimen size on fatigue results. Three different sizes
(size 0.5, 1.0 and 1.5) are selected in this research, in which size 1.0 corresponds to
the standardized size. Size 0.5 is a smaller one, a half of the size 1.0 and size 2 is
twice larger than the standardized size.

(3) Develop a material model which is able to describe the fatigue behavior obtained
from the different fatigue tests.

(4) Develop a new fatigue analysis approach to decrease if possible exclude the
influence of the test type and specimen size on the fatigue result.

5
Chapter 1 Introduction

1.5 Organization of the Dissertation


This thesis consists of nine chapters. Chapter 1 gives a general introduction of the fatigue
failure of asphalt pavements and the assessment of the fatigue characteristics in the
laboratory. Problem statements and objectives are described.

Chapter 2 provides a literature review on laboratory fatigue test methods and fatigue
analysis models.

In Chapter 3, the research methodology is presented.

Chapter 4 gives a detailed description of the materials used in this study. Ample attention
is given to the mixture design and specimen preparation.

Chapter 5 describes the experimental work. Three different fatigue configurations,


uniaxial tension and compression test, four-point bending test and indirect tension test,
were conducted with different specimen sizes. The results of the fatigue tests and the
interpretation techniques are also discussed.

In Chapter 6, the Partial Healing model is applied to simulate the evolution of the
complex modulus and the phase angle for the uniaxial tension and compression and the
four-point bending fatigue tests in strain controlled mode.

In Chapter 7, the results of the monotonic uniaxial tension and compression test are given
at different strain rates and temperatures. In order to determine the yield surface, the
unified model is used to calculate the strength and failure strain in both compression and
tension.

Chapter 8 describes the yield surface concept and its application on the fatigue results. A
new parameter R∆ is introduced as safety factor and a new fatigue analysis approach is
developed to reduce or exclude the influence of the test type and specimen size on the
fatigue life.

Chapter 9 presents the conclusions and recommendations.

6
Chapter 1 Introduction

Chapter 1 Introduction

Chapter 2 Literature Review

Chapter 3 Research Methodology

Chapter 4 Mixture Design and


Specimen Preparation

Chapter 5 Different Laboratory Chapter 7 Monotonic Uniaxial


Fatigue Experiments Tension and Compression Test

Chapter 6 Application of Partial Healing Model Chapter 8 Yield Surface and


on Strain Controlled Fatigue Tests Fatigue Tests

Chapter 9 Conclusions and


Recommendations

Figure 1-6 Structure of the dissertation

7
Chapter 1 Introduction

References

Bonnot, J., Asphalt aggregate mixtures, Transportation Research Record 1096,


Transportation Research Board, 1986, pp.42-50.

Brown, S.F., Practical test procedures for mechanical properties of bituminous materials,
Proceeding of ICE Transport 111, 1995, pp. 298-297.

Collop, A.C, Cebon, D., A theoretical analysis of fatigue cracking in flexible pavements,
J. Mech. Eng. Sci., IMech E, 1995, Vol 209, No C5, pp. 345-361.

Di Benedetto H., de la Roche C., Baaj H., Pronk A., Fatigue of bituminous mixtures.
Materials and Structures, 2004, vol. 37, n. 3, pp. 202-216.

European committee for standardization, Bituminous Mixtures-Test Methods for Hot Mix
Asphalt, BS EN 12697: Part 24: Resistance to Fatigue. CEN, Brussels, 2004.

Gerritsen, A.H.; van Gurp, C.A.P.M.; van der Heide, J.P.J.; Molenaar, A.A.A. and Pronk,
A.C., Prediction and Prevention of Surface Cracking in Asphaltic
Pavements. Proceedings, 6th International Conference Structural Design of Asphalt
Pavements, The University of Michigan. Ann Arbor, Michigan, July 1987, pp. 378-391.

Hindley, G., A History of Roads, the Chausser Press Ltd., Bungay, Suffolk, ISBN 432-
06-7361, 1971.

Hveem, F. and Davis, H., Some Concepts Concerning Triaxial Compression Testing of
Asphalt Paving Mixtures and Subgrade Materials, ASTM Special Technical Publication
No. 106: Triaxial Testing of Soils and Bituminous Mixtures, American society for
Testing Materials (ASTM), 1916 Race Street, Philadelphia, 1950.

McGennis R.B., Anderson R.M., Kennedy T.W., Solaimanian M., Background of


Superpave Asphalt Mixture Design and Analysis.Publication, No.FHWA-SA-95-003,
1994.

Miller, J.S., Bellinger, W.Y., Distress Identification Manual for the Long-Term Pavement
Performance Program, Publication No. FHWA-RD 03-031, June 2003.

Molenaar, A.A.A., Bottom-Up Fatigue Cracking: Myth or Reality? in Proc of 5th


International RILEM Conference, Limoges, France, 2004. pp. 275-282.

Molenaar, A.A.A., Structural Performance and Design of Flexible Road Construction and
asphalt Concrete Overlays, Ph.D Thesis, Delft University of Technology, the Netherlands,
1983.

Pell, P. S., Fatigue Characteristics of Bitumen and Bituminous Mixes, Proceedings,


International Conference on the Structural Design of Asphalt Pavements, 1962.

8
Chapter 1 Introduction

Pronk, A.C., The theory of the four point dynamic bending test-Part 1. Report P-DWW-
96-008, Rijkswaterstaat, the Netherlands, 1996.

Roberts FL, Kandhal PS, Brown ER, Lee DY, Kennedy TW, Hot Mix Asphalt Materials,
Mixture Design, and Construction. NAPA Education Foundation, Second Edition. 1996,
pp. 603.

Strategic Highway Research Program (SHRP), Fatigue response of asphalt-aggregate


mixes, Executive summary, National Research Council, 1992.

Tangella, S. R., Craus, J., Deacon, J. A. and Monismith, C. L., Summary report on
fatigue response of asphalt mixtures, SHRP Report No. TM-UCB-A-003-A, 1990.

Uhlmeyer, Jeff S., Willoughby, Kim. Top-down Cracking in Washington State Asphalt
Concrete Wearing Courses.". Journal of the Transportation Research Board, 2000, Vol.
1730, pp.110-116.

Uhlmeyer, J.S.; Willoughby, K.; Pierce, L.M. and Mahoney, J.P., Top-Down Cracking in
Washington State Asphalt Concrete Wearing Courses. Transportation Research Record
1730. Transportation Research Board, National Research Council, Washington, D.C.
2000. pp. 110-116.

van Dijk, W., Practical fatigue characterization of bituminous mixes, Proceedings of the
Association of Asphalt Paving Technologists, 1975, pp. 38.

9
Chapter 1 Introduction

10
Chapter 2 Literature Review

Chapter 2 Literature Review


The literature survey on fatigue testing of asphalt mixtures in the laboratory is presented
in this chapter. Since in this project, the test type and specimen size effect on the
laboratory fatigue results are involved, this Chapter is mainly focused on the following
issues.

Section 2.1 provides the background regarding the fatigue research on asphalt mixture. In
Section 2.2, the widely used laboratory fatigue tests for measuring fatigue behavior of
asphalt mixtures are discussed. Some factors affecting fatigue results are summarized in
Section 2.3. In Section 2.4, the different analysis approaches to characterize fatigue are
reviewed, including the classical analysis, the dissipated energy approach and the fracture
mechanics approach. Section 5 discusses some fatigue models which are used to describe
the influence of test type and specimen size on the fatigue behavior of asphalt mixtures.

2.1 Background of Fatigue Research on Asphalt Mixtures


The first fatigue study was initially done on metal. Richard [1988] stated that Wöhler
[1860] conducted systematic investigations of fatigue failure in railroad axles for the
German railway industry. His work also led to the characterization of fatigue behavior in
terms of stress amplitude-life (S-N) curves.

Fatigue considerations in the design of bituminous pavements date back to the early
1940s due to a significant increase in the traffic volume and the magnitude of wheel loads.
Many flexible pavement researchers and designers expressed their concern over the
fatigue failure of pavements under various loading conditions. Porter noted that flexible
pavements failed under deflections as small as 0.02 to 0.03 inch (0.5-0.8 mm) [Porter,
1942]. In 1953, Nijboer and van der Poel showed that cracks which often appeared in the
later stages of the life of an asphaltic concrete could be related to the bending stress
induced by the moving traffic, exceeding the flexural strength of the material [Nijboer,
1953]. Examinations of the results of the AASHTO Road Test also revealed that cracking
and initial failure of the pavement were primarily caused by repeated bending of the
bituminous layers. Hveem F.N. [1955] reported fatigue failure caused by repeated
loading on asphalt pavements built on highly resilient soils. Hveem concluded that there
was a correlation between observations of cracking, fatigue type failures in bituminous
pavements, and the measured repeated deflections that the pavement undergoes with each
passing wheel.

However, it is not fully clear which chemical, physical, and mechanical processes exactly
occur in fatigue. It is generally known that, during the fatigue process, successive stages
of deterioration occur. These are often named: initiation of microcracks, propagation of
microcracks, initiation/formation of macrocracks or coalescence of microcracks into
macrocracks, propagation of macrocracks, disintegration or rupture/failure. The use of
fundamental, non-empirical mechanical tests to characterize the fatigue behavior of these

11
Chapter 2 Literature Review

materials is particularly important when designing and predicting the in-service


performance of asphalt mixtures.

2.2 Laboratory Fatigue Test Methods


Over the past 60 years, a number of test methods have been developed to simulate the
fatigue behavior of bituminous road construction materials. According to the mode of
loading the most commonly used tests are classified in three groups, simple flexure,
direct uniaxial loading and diametral loading tests, as mentioned in Section 1.3 [Tangella,
1990]. According to the stress-strain distribution in the specimen, the fatigue tests are
divided into two types, the homogenous type and inhomogeneous type. If the stress-strain
distribution is uniform throughout the specimen, the test is called homogeneous test. An
example is the uniaxial loading test. For non-homogeneous tests, such as bending tests
and the indirect tensile test, the stress-strain field is not uniform along the specimen and
cross section.

2.2.1 Simple Flexure Test


2.2.1.1 Two-Point Bending (2PB) Test
Two-point bending tests on trapezoidal specimens were conducted by researchers of
Shell [van Dijk, 1975], the Center of Road Research in Belgium [Verstraeten, 1972], and
LCPC researchers [Bonnot, 1986]. Figure 2-1 illustrates the LCPC equipment. The
smaller end is subjected to either a sinusoidal displacement [Bonnot, 1986] [van Dijk,
1975] [Verstraeten, 1972] or load [Kunst, 1989]. By properly selecting the dimensions of
the trapezoid, the specimen will fail at about mid height where the bending stress is
largest and not at the base where the bending moment is largest and the boundary
conditions might adversely affect interpretation of test the results. Specimens tested by
van Dijk had a base cross section of 55 mm by 20 mm, a top cross section of 20 mm by
20 mm, and a height of 250 mm.

Displacement
sensor
Specimen

Figure 2-1 2PB fatigue test machine with a trapezoidal specimen [Chkir, 2009]

According to EN 12697-24 [EN, 2003], the specimens shall be of an isosceles trapezoidal


shape as shown in Figure 2-2, for which the dimensions are given in Table 2-1. The large

12
Chapter 2 Literature Review

base of each specimen should be glued in a groove (about 2mm deep) of a metal base
having a minimum thickness of 20mm, as shown in Figure 2-1. The specimen shall be
moved sinusoidally at its head at the imposed displacement amplitude until the failure
criterion has been reached. The deformation shall be such that at least one third of the
element tests provide results with N ≤ 106 and at least one third of the element tests
provide results with N ≥ 106.
b
e Table 2-1 Dimensions of the 2PB specimen
Type of mixture
D ≤ 14mm 14 < D ≤ 20mm 20 < D ≤ 40mm
h B 56 70 70
b 25 25 25
e 25 25 50
h 250 250 250
e
D: maximum aggregate size of asphalt mixture
B

Figure 2-2 Geometry of the specimens

2.2.1.2 Three-Point Bending Test


At the Shell Laboratory in Amsterdam, van Dijk [1972] used the center-point loading
equipment. The specimen dimensions are 30 mm (1.2 in.) × 40 mm (1.6 in.)× 230 mm
(9.2 in.), tests were done in the controlled-deflection (strain) mode. Figure 2-3 shows the
test apparatus and the scheme of three point bending test.

(a)
F0

F0/2 L/2 L/2 F0/2

(b)
Figure 2-3 Three point bending test apparatus scheme (a) and load characteristics (b)

13
Chapter 2 Literature Review

According to EN 12697-24 [EN, 2003], the dimensions of the test beams shall be (300 ±
10) mm × (50 ± 3) mm × (50 ± 3) mm. The specimen shall be clamped to the support
mechanism through the two metallic tubes glued to one of its faces and to the piston rod
through the tube glued to the opposite face. The support mechanism shall be capable of
moving and tilting its axes. Specimens and extensometer are assembled and brought to
the test temperature, 20 ºC. A cyclic sinusoidal displacement of the piston rod shall be
applied. The wave frequency shall be 10Hz, and the values of the total amplitude usually
range from 80 µm to 350 µm depending on the mixture.

2.2.1.3 Four-Point Bending Test


According to Molenaar [Molenaar, 1987], the specimens for the four-point bending test,
can be obtained in the following way. Slabs of about 0.6 × 0.6 m2 are sawn from the AC
pavement. From the bottom layer of these slabs, beams are sawn with dimensions of 450
mm × 50 mm × 50 mm, with the length perpendicular to the direction of traffic. For the
four-point bending tests, a servo-hydraulic testing rig is used. The setup is schematically
shown in Figure 2-4. The distance between the outer supports is 400 mm, between the
inner supports 130 mm. The tests are executed with a displacement-controlled fully
sinusoidal load signal at 30 Hz and the range of the test temperature is 0~20 ºC.
Generally, different preset displacements are chosen, which will result in an expected
fatigue life of 105 to 106 load repetitions.

Figure 2-4 Schematic diagram of the four-point bending test

According to the ASTM standard D 7460 [ASTM, 1996], the dimensions of the test beam
are 380 (length) × 50 (height) × 63 (width) mm. The horizontal spacing of the clamps is
119mm. Before testing, the specimen is placed in an environment which is at 20 ± 0.5 ºC
for two hours. The desired initial strain (250 to 750 microstrain) and loading frequency (5
to 10Hz) are selected. The initial stiffness of the specimen is determined at the 50th load
cycle, which is used as a reference for determining specimen failure. A deflection level
(strain level) is selected such that the specimen will undergo a minimum of 10000 load
cycles before its stiffness is reduced to 50 percent or less of the initial stiffness.

14
Chapter 2 Literature Review

Figure 2-5 Scheme of the four-point bending test and load characteristics [ASTM, 1996]

2.2.1.4 Rotating Bending Test

pulleys

bearing
Loading
head

specimen
wire

chuck
seat
housing

shaft tank
bearings
& electric
motor

weights
rev. counter stop switch

(a) Controlled-Stress Rotating Flexure (b) Failure specimen


Figure 2-6 Flexure apparatus used by Pell [Pell, 1965] and specimen before and after
testing [Saal, 1960]

At the University of Nottingham, U. K. [Pell et al., 1975 and 1973a] a rotating cantilever
machine (Figure 2-6a) was used in which the specimen is mounted vertically on a

15
Chapter 2 Literature Review

rotating cantilever shaft and a single point load is applied through a bearing at the top.
This loading system results in a sinusoidally varying bending stress of constant amplitude
at any particular cross section of the specimen. Under this system of loading there is no
shear stress on the specimen. The specimens were tested in a controlled temperature bath.
The majority of the tests were carried out at a temperature of 10°C and a speed of 1,000
rpm. Specimens of the shape have a minimum diameter at the neck of 2.5 in. Figure 2-6b
shows a demoulded specimen ready for testing and also a broken specimen which has
failed in fatigue after testing.

2.2.2 Direct Axial Loading Test


Raithby [Raithby, 1972a] developed this form of fatigue testing at the Transport and
Road Research Laboratory (TRRL) in the United Kingdom. Axial tensile and
compressive loading was applied using a servo-controlled electro-hydraulic machine.
Specimens were prismoidal, with a 75 mm2 cross section and a length of 225 mm (Figure
2-7). Aluminum caps are glued upon the ends of the specimens, in such a way that these
can be mounted in a servo-controlled hydraulic MTS testing machine. Loading
frequencies were 16.7 and 25 Hz, and the effects of rest periods, shape of wave form, and
the sequence of load application (compression/tension, tension/compression, compression
only, and tension only) were evaluated.

Temperature chamber

C.R.O.

Summing LVDT
junction

Command
signal Hydraulic Load cell
actuator
Error signal
amplifier

Servo-valve

Hydraulic
Power
supply

Figure 2-7 Schematic representation of a direct axial fatigue test [Raithby, 1972a]

Molenaar [1983] selected the direct tensile test for crack growth experiments. The sketch
and picture of the test set-up as used are given in Figure 2-8.

16
Chapter 2 Literature Review

Pressure cell

steel plate
epoxy resin
LVDT
Range ± 1 mm specimen

holder

epoxy resin
steel plate

Figure 2-8 Set-up of direct tensile test [Molenaar, 1983]

The specimens used in the test were about 0.15 m long and the cross section was
approximately 0.05 × 0.05 m. At mid length, an artificial crack was sawn at two opposite
sides which had a depth of about 0.005 m. The beam specimens were sawn from slabs.
An epoxy resin was used to glue the specimens to the top and bottom loading plate. To
ensure a proper alignment of the specimens, hardening of the glue took place while the
beam was positioned under the ram of the dynamic loading system. The shape of the load
pulse was a haversine and a load to rest period ratio of 1 to 7 was used. During the tests
the elastic vertical displacement was continuously recorded.

2.2.3 Diametral Loading Test (Indirect Tensile Test)


Indirect tension testing is done by applying a compressive force to a cylindrical specimen
along its vertical diameter to produce tensile stresses perpendicular to the loading axis.
Figure 2-9 shows a schematic of the indirect tensile test.

Compressive force

Specimen Loading strip

Figure 2-9 Schematic of indirect tension test

17
Chapter 2 Literature Review

The test was developed and has been used since the early 1970s to characterize
bituminous materials in terms of strength and elastic stiffness [Kennedy, 1968] [Schmidt,
1971]. Since that time the repeated load indirect tensile fatigue test is widely used to
evaluate the fatigue properties of asphalt materials by a number of investigators
[Kennedy, 1983] [Khosla and Omer, 1985] [Tangella, 1990]. In the UK this test has also
been developed as a simple fatigue tool conducted as one of the test modules in the
Nottingham Asphalt Tester (NAT) [Brown, 1995].

Normally a haversine load pulse is employed in the test and the diametral specimen is
assumed to fail near the load line. Three types of failure patterns were usually observed
[Sousa, 1991]: (1) crack initiation at or near the center of the specimen, resulting in
complete splitting of the specimen; (2) crack initiation at the top of the specimen,
progressively spreading downward in a V-shape, the arms of which originate from the
outside edges of the loading platen; and (3) no real cracking occurs, with the specimen
being plastically deformed beyond the limiting vertical deformation.

Due to the absence of stress reversal, the accumulation of permanent deformation


increases. The possibility that under high loads and/or high temperatures either
compressive or shear failure occurs in the specimen. Fatigue fracture under indirect
tensile loading ideally should occur by splitting of the specimen in two halves with
minimum permanent deformation. Read and Collop recommended that the loading time
should be 120 ms and test temperatures should be less than 30°C [Read, 1997].

2.3 Influence of Test Type, Specimen Size and Test Conditions


on Fatigue Results
The European Committee for Standardization (CEN) specifies several tests for
characterizing the fatigue of bituminous mixtures, as follows:

1) Two-point bending test on trapezoidal shaped specimens

2) Two-point bending test on prismatic shaped specimens

3) Three-point bending test on prismatic shaped specimens

4) Four-point bending test on prismatic shaped specimens

5) Indirect tensile test on cylindrical shaped specimens

Figure 2-10 shows a schematic illustrating all of the allowed fatigue test methods. Most
of these test methods are already discussed in Section 2.2. Because the European
Standard does not impose a particular type of testing device, the choice of the test
methods and the test conditions depends on the possibilities and the working range of the
used device. This might lead to incomparable results obtained from the different test
methods. Therefore it is necessary to investigate the influence of the test type, specimen
size and some important test conditions on the fatigue results.

18
Chapter 2 Literature Review

(1) (2) (3)

(4) (5)

Figure 2-10 Schematics of the laboratory fatigue tests recommended by the European
Committee for Standardization [EN 12697-24, 2003]

2.3.1 Influence of Test Type and Specimen Size


2.3.1.1 Influence of Test Type
A certain bituminous material may show different fatigue results dependent on the
various geometries (beam, trapezoidal, cantilever beam, cylinder and other special shapes)
and the different types of loading (bending, tension-compression and shear). Many
researchers compared the results from different types of fatigue tests. Aguirre et al, found
that fatigue lives are shorter in tension-compression than in bending at the same nominal
strain, as shown in Figure 2-11 [Aguirre, 1981].

0.001
Flexion (labo1)

Flexion (labo2)
Initial strain

Traction Compression
0.0001 (labo3)
Traction Compression
(labo4)
Traction Compression
(labo5)

0.00001
1000 10000 100000 1000000 10000000
N cycles

Figure 2-11 Fatigue behavior at constant stress amplitude (10°C, 10Hz) [Aguirre, 1981]

19
Chapter 2 Literature Review

Di Benedetto H. et al, reported results from an inter-laboratory investigation organized by


the RILEM 182-PEB Technical Committee [Di Benedetto, 2004]. Eleven (11) different
test types, including uniaxial tension/compression, 2-, 3- and 4-point bending and
indirect-tension tests, were used to evaluate fatigue characteristics of a dense grade
asphalt concrete mixture. The loading signal was specified as sinusoidal at a frequency of
10 Hz and a test temperature of 10 °C was chosen.

Figure 2-12 shows fatigue results of the used 11 test types. The obtained fatigue lives are
significantly influenced by the test method used. The Indirect Tension Test (ITT) shows
the shortest fatigue life due to accumulation of permanent deformation in addition to the
fatigue damage. For a given strain amplitude, the beam tests generally result in longer life
durations compared to T/C (Tension/Compression) tests. The strain value, ε6, which
represents the failure at one million cycles, is plotted in Figure 2-13. The ITT has the
smallest strain value, ε6, but this value can not be compared with the other strain-
controlled tests.

Nf50 = α*ε^(-β)

1,E+09

1,E+08
3PB

1,E+07
Nf50
1,E+06
T/C 4PB
1,E+05
2PB
1,E+04
ITT
1,E+03
80 100 110 ε0 (µm/m)

Figure 2-12 Fatigue lives for strain controlled fatigue tests (except ITT)

190
170
150
130
Average
110 all
90 2PB 4PB
70 T/C
50 ITT
30

Figure 2-13 Strain amplitude “ε6” giving failure at 106 cycles from different fatigue tests

20
Chapter 2 Literature Review

According to de la Roche et al., the fatigue life is longer in three-point bending than in
two-point bending [de la Roche, 1993]. The relationship of ε6 between these two tests is:

ε 6(3 PB )
= 1.4 ∼ 2 (2-1)
ε 6(2 PB )

2.3.1.2 Influence of Specimen Size


Doan carried out two-point bending tests to study the influence of the size effect on
fatigue evaluation of trapezoidal specimens with different sizes [Doan, 1973]. He found
that the specimen size effect was negligible leading to the same average fatigue life for
similar specimen series.

Didier Bodin from LCPC, however, showed a significant effect of the trapezoidal beam
size in bending fatigue tests [Bodin, 2006]. The fatigue line plots and ε6 of different sizes
are presented in Figure 2-14 (a) and (b). Size 1.0 corresponds to the standardized size.
Size 0.5, the smaller beam, has a longer fatigue life compared to size 1.0 and size 2.0, the
bigger beam. The value of ε6 is also a function of specimen size and decreases with
increasing size.

10000000 200

180
Fatigue life NF (50%)

1000000

160
ε6 [10-6]

100000
140
size0.5
10000
size1.0
120
size2.0
1000
100
100 140 180 220 260 0 0.5 1 1.5 2 2.5
Strain anplitude (10 -6) Size D/Do

(a) (b)
Figure 2-14 Fatigue line fitted for each sample size (a) and values of ε6 versus the size of
the specimen sets in two-point bending test [Bodin, 2006]

Figure 2-15 shows the influence of the width b of a specimen on the fatigue life, Nf
[Jacobs, 1995]. At a chosen strain amplitude of 400 µm/m the ratio in fatigue life between
a 50 and a 20 mm wide specimen is around 3.3; for a 250 mm wide specimen the ratio
Nf(b=250)/Nf(b=20) is about 19.

21
Chapter 2 Literature Review

b=250mm
b=150mm
b= 50 mm
b= 20 mm

Figure 2-15 Influence of the specimen width b on the fatigue life Nf [Jacobs, 1995]

For the four-point bending tests on beams with a notch subjected to half sine load,
Groenendijk developed the function of the number of load repetitions to failure based on
the Paris’ law [Groenendijk, 1998]:

n
1−
d ( c / h)
2 cf / h
h
⋅ ε −n
) ∫
N=
A ( Smix
n 4.5 n
c0 / h  0.5
c
1.5
c c
2.5 3.5
c c  (2-2)
1.99   − 2.47   + 12.97   − 23.17   + 24.8   
 h h h h h 

Where: c0 : initial crack length, [mm];


cf : maximum crack length for which stable crack occurs,
[mm];
A,n : parameters depending on the material and on the
experimental conditions;
c : crack length, [mm];
Smix : stiffness of asphalt mixture, [MPa];
ε : applied strain, [m/m];
h : height of the specimen, [mm].

2.3.2 Influence of Loading Mode

Normally one of two main modes of loading are applied in a fatigue test:

 The load (stress) controlled mode – the amplitude of the applied load is held constant
during the test.

22
Chapter 2 Literature Review

 The displacement (strain) controlled mode – the amplitude of the applied


deformation is held constant.

Strictly speaking it is the load or displacement which is controlled and the stress and
strain are calculated from them. However, controlled stress and strain are commonly used
terms.

Brown indicated that in the controlled stress mode, a repeated stress or load of constant
amplitude is applied to a specimen which causes a gradual increase in strain at midspan
and a gradual decrease of the stiffness [Brown, 1978]. The increase of strain is rather
rapid at the end of the test, till complete fracture occurs. In controlled strain tests, the
loading is applied such that the repeated deflection or strain stays constant during the test
(see Figure 2-16). During the test the load will gradually decrease because of damage
development. Most of the time no real failure will occur and it is therefore assumed that
the specimen has failed if the load level has decreased to 50% of the original load level.

(a) (b)

Figure 2-16 Graphical representation of controlled stress (a) and controlled strain (b)
modes of loading [Epps and Monismith, 1971]

Tayebali, et al [Tayebali, 1994] showed a typical plot (Figure 2-17) of the stiffness ratio
(defined as quotient of stiffness at the ith load repetition to the initial stiffness, Si/S0)
versus the number of load repetitions for flexural beam fatigue tests in both controlled-
stress and controlled-strain modes of loading. Obviously the stiffness ratio in the
controlled-stress mode decreases more rapidly compared to the stiffness ratio results in
controlled-strain mode.

23
Chapter 2 Literature Review

1,2
1,0

Stiffness Ratio
0,8
0,6
0,4
Controlled-Stress Test
0,2 Controlled-Strain Test

0,0
1,E+02 1,E+03 1,E+04 1,E+05
Number of cycles

Figure 2-17 Stiffness ratio versus number of cycles, flexural beam fatigue controlled-
stress and controlled-strain tests [Tayebali, 1994]

Figure 2-18 shows the variation of dissipated energy per cycle with the number of load
repetitions. The dissipated energy per cycle decreases with an increasing number of load
repetitions in the controlled-strain fatigue tests; whereas, for the controlled-stress tests,
the dissipated energy per cycle increases as the number of load repetitions increases.

0,04
Dissipated Energy (psi)

0,02

Controlled-Stress Test
Controlled-Strain Test
0
1,E+02 1,E+03 1,E+04 1,E+05
Num be r of cycle s

Figure 2-18 Dissipated energy per cycle versus number of cycles, flexural beam fatigue
controlled-stress and controlled-strain tests [Tayebali, 1994]

As shown in Figure 2-19, the results from stress and strain controlled tests are different
for the same asphalt mixture. Brown [1978] explained this phenomenon as follows:
“cracks are initiated at the points of high stress concentration and propagate through the
specimen until fracture occurs. The propagation of cracks depends on the stress intensity
at the crack tip. In a strain controlled test, the stress gradually decreases with the decrease
of the stiffness. Therefore strain controlled tests have a long period of crack propagation.
In a stress controlled test, however, the crack propagation is very rapid because the stress
is constant in the test”. Each of these modes has been linked to a particular pavement

24
Chapter 2 Literature Review

construction and represents realistic service condition [Pell, P. S., 1973]. Monismith and
Salam analyzed pavement structures by using multilayer elastic theory and found that:
(1) The strain controlled mode of loading is characteristic for thin asphalt pavements (≤
50mm),
(2) The stress controlled mode of loading is more realistic for thick layers (≥ 150mm),
(3) For a pavement of average thickness, an intermediate mode is appropriate [Monismith,
1979].

Figure 2-19 Graphical representation of the difference in fatigue response to various


modes of loading for one mixture [Read, J. M. 1996]

Many researchers have evaluated the characteristics of these two modes of loading. A
brief summary is given in Table 2-2. The results summarized in this table have been
obtained from research reported in the following references [Pell, 1972] [Pell, 1973],
[Tangella, 1990] [Kim, 1992], [Brennan, 1992], [Tayebali, 1994], [Monismith, 1971],
[Francken, 1998].

Table 2-2 Comparison of results obtained with controlled stress and strain loading
Variables Controlled-stress test Controlled-strain test
Usual failure criterion Failure of specimen Loss of half of initial force
Magnitude of fatigue life shorter Longer
Scatter in fatigue test data Less scatter More scatter
Thinner asphalt-bound
Thickness of asphalt Thicker asphalt-bound
layers,
concrete layer layers
< 3 inches
Evolution during test Increase of displacement Reduction of force
Rate of crack propagation Faster Slower

25
Chapter 2 Literature Review

Attempts have been made to determine the mode of loading that best simulates actual
pavement conditions [Monismith and Deacon, 1969] [Monismith et al., 1977]. The type
of loading is expressed by means of a mode factor (MF) defined in Equation 2-3:

A− B
MF = (2-3)
A+ B

where MF is the mode factor; A is the percentage change in stress; B is the percentage
change in strain due to a fixed reduction stiffness. The MF assumes a value of -1 for
controlled-stress conditions and + 1 for controlled-strain conditions.

2.4 Fatigue Analysis Approach


2.4.1 Results of Laboratory Fatigue Tests
The measurements obtained during the test are the signals of the applied force and the
displacement. The places where they are measured depend on the test device (see Section
2.2). Figure 2-20 shows the force and displacement signals for one loading cycle. The
stiffness is calculated based on the amplitudes of the force and defection and the phase
angle is obtained from the time lag between the load and displacement.

Amplitude Time lag

Time

Displacement
Force

Figure 2-20 Relationship between force and displacement in one cycle

Generally, asphalt materials are considered to be linear visco-elastic materials under


relatively low strain applications. For the homogenous fatigue tests, the value of the
stiffness is directly obtained from the stress, σ, and strain, ε, and for non-homogenous
fatigue tests, the stiffness is calculated from the load and displacement using the specific
factors. Table 2-3 shows the calculations of stiffness in different fatigue tests.

26
Chapter 2 Literature Review

Table 2-3 Calculation of the stiffness in different fatigue tests


Fatigue test Calculation of stiffness (kPa) Remarks
ω: horizontal load (kN);
l: length of specimen (m);
b: the base width (m);
ω ⋅l  l2  δ0: deflection at top of beam
Two-point bending test E= ⋅  K1 ⋅ + K 2 (1 + ν )  (mm); d: top width (m);
[EN 12697-24, 2003] δ0 ⋅ d ⋅ b  d 2
 b: thickness of the specimen (m);
ν: assumed Poisson’s ratio;
K1, K2: coefficients depending
on the geometry of the specimen;
ω: horizontal load (kN);
L: length of specimen (m);
Three- point bending test σt ωL δ: maximum deflection at center
E= =
[Read, J. M. et al, 2003] ε t 4bh3δ of beam (m);
b: thickness of the specimen (m);
h: average specimen height (m);
σt: maximum tensile stress (Pa);
εt: maximum tensile strain
(m/m);
P: load applied by actuator (N);
Four- point bending test σ t Pa (3L2 − 4a 2 ) b: average specimen width (m);
[ASTM standard D E= = h: average specimen height (m);
7460,1996] εt 4bh3δ
δ: maximum deflection at center
of beam (m);
a: space between inside clamps;
L: length of beam between
outside clamps;
σt σt: maximum tensile stress (Pa);
Uniaxial test E= εt: maximum tensile strain
εt (m/m);
L: peak value of applied vertical
load (N);
D: peak horizontal diametral
resulting from the applied load
Indirect tensile test L
E= × (υ + 0.27) (mm);
[EN 12697-24, 2003] D ⋅t t: mean thickness of the test
specimen (mm);
ν: assumed Poisson’s ratio.

The stiffness decreases under repeated loading during the fatigue tests. An example of
stiffness evolution with the number of cycles is given in Figure 2-21. A typical fatigue
process for asphalt mixtures is generally consisting of three distinctive response phases as
reported in previous fatigue studies [Di Benedetto H. et al, 1996] [Di Benedetto, 2004].

27
Chapter 2 Literature Review

Adaptation Quasi-stationary Failure

Number of cycles
Figure 2-21 Stiffness modulus versus load cycles evolution curve

Phase I -- adaptation phase: During this phase, the stiffness (or stiffness ratio) of the test
specimen decreases rapidly due to repetitive loading. The decrease is possibly caused by
the combined effects of fatigue damage, heating of the specimen and thixotropy.
Microcracks are created in the binder films between the aggregates.

Phase II -- quasi-stationary phase: The stiffness is decreasing slowly in this period where
the role of fatigue on the stiffness decrease is dominant. Any artifact effects such as
thermal heating and thixotropy can be considered to be small compared to the dominant
effect of fatigue damage. Microcracks are in a stable growth phase.

Phase III -- failure phase: Local crack propagation occurs in this phase. Macrocracks
start to develop and failure is obtained at the end of this phase [Jacobs 1995].

The phase angle, φ, is calculated from the time lag between the load and displacement:

ϕ = 360 ⋅ f ⋅ ∆t (2-4)

where
φ = phase angle, in degrees;
f = load frequency, Hz, and
∆t = time [s] lag between the maximum applied load, Pmax, and maximum
deflection, δmax, (see Figure 2-20).

For a purely elastic material the phase angle is 0°, and 90° for a purely viscous material.
Figure 2-22 illustrates the fatigue test results plotted in a phase angle vs. stiffness
diagram [Di Benedetto H. et al, 2004]. The gradual increase in phase angle is due to an
increased viscous response of the material as damage accumulates in the form of micro-
cracks and the test specimen experiences a reduction of its elasticity. However, once the
specimen experiences significant structural changes due to macrocracking, it can no
longer accumulate damage and consequently the phase angle decreases rapidly [Kim Y.
et al. 2006].

28
Chapter 2 Literature Review

Figure 2-22 stiffness vs. phase angle during fatigue test [Di Benedetto H. et al, 2004]

2.4.2 Classical Fatigue Analysis


In order to evaluate and compare the fatigue properties of a bituminous material, it is
necessary to define the failure of the specimen in a consistent way. The stiffness at the
beginning of the test (N=50 to 100 cycles) is defined as the initial stiffness. For the
controlled stress mode the specimen has a relatively short crack propagation period, so
the failure point is close to the moment when the specimen has completely fractured, as
shown in Figure 2-23. However, for the controlled strain mode the failure point is not
easy to define, due to the long time of crack propagation during the test. Hence an
arbitrary specimen fatigue life criterion has been defined. The point of failure is normally
defined as the moment at which the stiffness has reduced to 50% of its initial value or
when the applied stress is half of the initial stress [van Dijk, W., 1975, Moutier, F. et al,
1988, Bonnaure, F. et al, 1980].

Controlled strain mode

Controlled stress mode

Failure point Crack propagation time associated with


controlled strain mode

Stiffness

Failure point in controlled


strain mode
Half initial stiffness

Figure 2-23 Idealized representation of the failure points for the two modes of loading
[Read, 1996]

29
Chapter 2 Literature Review

Rowe, et al, proposed another fatigue failure criterion [Rowe, 2000]. They stated that
fatigue failure occurs at the maximum value of the nE/Einitial against n plot, illustrated in
Figure 2-24. This definition of fatigue life is considered to be more accurate and
reasonable than simply determining failure as an arbitrary condition such a 50% or 90%
reduction of initial stiffness.

The phase angle increases until a certain maximum value followed by a drop as loading
continues (see Figure 2-24). This peak in phase angle has been successfully used as an
indication of the fatigue failure point by some researchers [Reese, R. 1997] [Lee, H. J. et
al.].

The failure point can also be identified based on the slope of stiffness change with load
cycles. There are two rates of change in stiffness modulus against load cycles
representing micro-cracking and macro-cracking, respectively, and a transition point
between these two rates. This transition point represents the shift from micro- to macro-
cracking [Rowe, G.M. et al. 2000] [Kim Y. R. et al. 2002]. This is also shown in Figure
2-24.

Figure 2-24 Comparison of different fatigue life criterions [Cocurullo, A. et al. 2008]

Early fatigue research found that the fatigue life could be described using a Wöhler curve
described by:

log N = n log λ + k1 (2-5)

Where λ is the loading amplitude, either force (F) (or stress) amplitude or displacement
(D) (or strain) amplitude.

The most important variables from the fatigue test are the intercept and the slope of the
fatigue curve, k1 and n respectively. The fatigue coefficients k1 and n are experimentally
determined and are used in fatigue based mechanistic design procedures. A large range of
values for n is reported. Typically n values are between 3 and 14 based on the database of

30
Chapter 2 Literature Review

fatigue testing. Using the Wöhler relationship, as a basis for the description of fatigue
results, some researchers found material depending characteristics for the parameters k1
and n.

Finn et al. and Monismith et al. proposed a more general formula for the fatigue life as a
function of maximum tensile strain and initial mix stiffness [Finn, 1977] [Monismith,
1985]:

a b
1  1 
N f = A    (2-6)
 ε t   S mix 

Where: ε t = tensile strain, S mix = initial mix stiffness, N f = fatigue life, and A, a and b are
experimentally determined coefficients.

This equation was used in the Shell [Shell, 1978] [Shell, 1985] and the Asphalt Institute
[AI, 1981] design procedures; the coefficients were determined by considering the
amount of cracking, mix type and the thickness of the asphalt-bound layer.

The effects of the volumetric bitumen content (Vb) and the air void (Va) content on the
fatigue performance of hot mix asphalt (HMA) were introduced by Pell and Cooper [Pell,
1975] as follows:
K2 K3 K4
 1   1   Vb 
N f = K1       (2-7)
 ε t   S mix   Va + Vb 
Where:
S mix = stiffness modulus of the HMA mixtures,
Vb = volume percentage bitumen content,
Va = air void content,
K1 , K 2 , K3 and K 4 = experimentally determined coefficients.

Different models have been proposed by the Nottingham researchers [Brown et al., 1982]
(Equation 2-8), Shell [Shell, 1978] (Equation 2-9), and the Asphalt Institute [AI, 1981]
(Equation 2-10) to account for the effects of other factors on fatigue life:

( 5.13 logV + 8.63log T −15.8)


 1 
b R&B
−40.7
Nf =V 14.39
T 10 24.2
 6 
(2-8)
 ε t ⋅10 
b R&B

ε t = ( 0.856 × Vb + 1.08 ) S mix


−0.036
× N −f 0.2 (2-9)
N f = 18.4C ( 4.325 × 10 ε −3
t
−3.291 −0.854
mix S ) (2-10)
Where:
TR & B = ring and ball softening point temperature (°C),

31
Chapter 2 Literature Review

 Vb 
C = 10M, M = 4.84 ⋅  − 0.69  ,
 Vb + Vv 
Vv = volume of air voids.

Bonnaure et al. employed a statistical approach using 146 fatigue curves from various
fatigue tests and found the following equations in the form of initial strain and fatigue life
(Equation 2-11 and 2-12) to determine the fatigue resistance of a mix. [Bonnaure, 1980]

For displacement controlled tests:


5
1
N f = (4.102 PI − 0.205 PI ⋅ Vb + 1.094Vb − 2.707) ( S mix ⋅10
5
)
6 −1.8
  (2-11)
 εt 
For force controlled tests:
5
1
N f = (0.300 PI − 0.015 PI ⋅ Vb + 0.080Vb − 0.198) ( S mix ⋅10
5
)
6 −1.4
  (2-12)
 εt 
Where:
PI = penetration index of the bitumen.

Based on the results of 38 fatigue tests including tests on polymer modified mixtures,
Medani developed the equation to predict the slope of the fatigue line, n [Medani, 2000].

nmas
n= (2-13)
CF
2 2
nmas = = (2-14)
m d ( log S m )
d ( log t )
CF = 0.541 + 0.173nmas − 0.03524Va (2-15)

Where:
nmas = n-value determined from the master curve,
m = slope of the master curve of stiffness;
Sm = mix stiffness, [MPa];
t = loading time, [s];
CF = correction factor.

The equation to predict the intercept of the fatigue line k1 (Equation 2-16) was also
developed using the results of 108 fatigue tests. The equation has successfully been used
to predict the k1 values of 10 asphalt mixtures that were tested in the Strategic Highway
Research Program [Medani, 2000].

3209 Vb
log k1 = 6.589 − 3.762n + + 2.332 log Vb + 0.149
+ 0.928 PI − 0.0721TR & B (2-16)
Sm Va
These fatigue equations, mentioned above, are applicable for a specific value of the

32
Chapter 2 Literature Review

applied tensile strain level, εt. For pavements, strains induced in the structure vary widely
as a result of variations in the type of axles, their loaded weights, tire pressures, lateral
wandering and so on. Accordingly, following the development of these laboratory fatigue
relationships, Deacon [1965] demonstrated the applicability of a cumulative fatigue
damage hypothesis by using it to consider the development of damage resulting from a
range of strains. This relationship is termed the linear summation of cycle ratios, or
Miner’s Law [Miner, M.A., 1945]. Miner’s hypothesis for cumulative fatigue damage has
been widely used to predict fatigue cracking. The hypothesis is stated as follows,

n1 n2 ni nm
+ + ... + + ... + (2-17)
N1 f N2 f N if N mf
Where
i = the i th level of applied strain at a critical point within the pavement
structure,
ni = number of actual traffic load applications at strain level i ,
N if = number of allowable traffic load applications at strain level i .

Fatigue failure is defined when the linear summation of cycle ratios reaches one. In spite
of the fact that Miner’s law is difficult to verify for bituminous materials, it is widely
used for pavement design purposes because of its simplicity [Peyronne et al, 1984],
[Francken, 1979 and 1987].

Loading
amplitude Curve sometimes found

Endurance limit

Life duration N
Figure 2-25 Wöhler curve: Loading amplitude vs. number of loading cycles
[Di Benedetto H. et al, 1997]

As shown in Figure 2-25, the material will have infinite fatigue life if the loading
amplitude is maintained at levels below the endurance limit. It is also believed by some
researchers that an endurance limit exists for an asphalt mixture. Monismith et al.
suggested that an endurance limit exists for HMAC and proposed 70 µ-strain as a likely
value [Monismith, 1971]. Carpenter et al. concluded that the endurance limit is in the
range of 70 to 90 µ-strain at 20°C for a loading frequency of 10 Hz [Carpenter, 2003].
Based on the test results obtained from uniaxial tension-compression testing, Soltani
postulated the presence of an endurance limit for two mixtures at 10°C and 10 Hz. The
estimated endurance limits are 30 and 80 µ-strain for unmodified binder and modified
binder, respectively.

33
Chapter 2 Literature Review

2.4.3 Dissipated Energy Approach


2.4.3.1 Dissipated Energy Theory
Energy is dissipated in asphalt mixtures during loading and relaxation because the
material behaves substantially visco-elastic at ambient temperatures. The dissipation of
energy is demonstrated in Figure 2-26, which shows a comparison between a linear
elastic material and a visco-elastic material.

Figure 2-26 Linear elastic versus visco-elastic behavior

For an elastic material, the energy stored in the system is equal to the area under the load-
deflection curve and during unloading, all the energy is recovered. By contrast, a visco-
elastic material, traces a different path when unloaded to that when loaded. This
phenomenon is commonly known as “Hysteresis” and the energy dissipated is equivalent
to the area within the loop. When a visco-elastic material, is sinusiodally loaded around a
zero position, as in a bending beam fatigue test, a phase lag is observe between the load
and measured deflection. If the load is plotted against the deflection, a hysteresis loop is
obtained, as shown in Figure 2-27 [Francken L. and Clauwaert C. 1987].

Figure 2-27 Hysteresis loop obtained from plotting load versus deflection
[Francken, 1987]

34
Chapter 2 Literature Review

The area of the loop can be calculated and if the load-deflection relationship is expressed
as stress-strain, the dissipated energy per loading cycle is obtained as follows:

wi = π σ i ε i sin φi (2-18)
Where: wi = dissipated energy in cycle i , σ i = stress amplitude in cycle i , ε i = strain
amplitude in cycle i , φi = phase lag in cycle i .

The dissipated energy is largely associated with viscous flow of the binder which
dissipates the energy as heat. The dissipation of energy, more importantly, also relates to
the formation of micro-cracks/crack surfaces [Little, 1995].

The amount of energy dissipated per loading cycle changes throughout a fatigue test. In a
controlled stress test, the dissipated energy per loading cycle increases whereas in a
controlled strain test it decreases, as illustrated in Figure 2-28. This is due to the fact that
the dependent variable (strain or stress) and the phase angle φi in Equation 2-18 do not
remain constant.

Fatigue under controlled stress Fatigue under controlled strain

Figure 2-28 Variation of dissipated energy per load cycle during controlled stress and
strain fatigue tests [Francken, 1987]

2.4.3.2 Cumulative Dissipated Energy


The earliest work using dissipated energy with asphalt materials was reported by
Chomton [Chomton, 1972] and van Dijk et al. [van Dijk, 1972]. Chomton and Valayer
[1972] presented a relationship in terms of “cumulative dissipated energy” versus number
of loading cycles, as follows:
i=N
W = π ∑ σ iε i sin φi = AN z (2-19)
i=0
Where: W = Cumulative dissipated energy to failure, A and z: experimentally determined
coefficients.

35
Chapter 2 Literature Review

Chomton and Valayer presented the results for three materials and suggested that the
parameters A and z could be independent of the mixture formulation. With this approach,
the fatigue life could be predicted if the dissipated energy was determined for a given mix
formulation. Dissipated energy captures both elastic and viscous effects and thus it is
possible to predict the relative fatigue behavior of mixes in the laboratory from the results
of fatigue tests when strain is the only test variable [Tayebali et al., 1994].

Van Dijk reported further work which demonstrated several important aspects [van Dijk,
1975] [van Dijk 1977]. It was clear that a single relationship could not be used for
different materials, as shown in Figure 2-29. For a given mixture the relationship between
the dissipated energy and the number of load repetitions to fatigue is valid, independent
of the testing method and temperature. In addition, a ratio was developed which was
related to the mixture stiffness.

Winitial
Ψ= (2-20)
W
Winitial = π ⋅ N ⋅ σ 0 ⋅ ε 0 ⋅ sin φ0 (2-21)

WFAT, J/m3 1: Asphaltic concrete


2: Lean sand asphalt
3: Dense bitumen macdam
4: Gravel sand asphalt
5: Lean bitumen macdam
6: Dense asphaltic concrete
7: Dense bitumen macdam
8: Rolled asphalt base course
9: Grave bitumen
10: asphalt base course
11: Rich sand asphalt
12: Gravel sand asphalt
13: Bitumen sand base course

NFAT

Figure 2-29 Accumulative dissipated energy versus fatigue life for a series of mixture
[van Dijk 1977]

This ratio was found to be larger than 1 for controlled strain tests and lower than 1 for
controlled stress tests (see Figure 2-30). This is a result of decreasing dissipated energy
per cycle in a controlled strain test compared to increasing dissipated energy per cycle in
a controlled stress test. It can be seen from Figure 2-30 that as the stiffness of the
materials increases, the ratio tends to approach unity. This is probably related to the
increasing rate of crack growth that occurs with stiffer materials. Clearly, Ψ is a function
of the mode of testing and the mixture stiffness.

36
Chapter 2 Literature Review

Figure 2-30 Stiffness versus Ratio Ψ

This work has been extended by Himeno et al. [1987] who developed the dissipated
energy concept for three-dimensional stress conditions and applied it to the failure of an
asphaltic layer in a pavement and by Rowe who has shown that dissipated energy can be
used to accurately predict the life to crack initiation [Rowe, 1993].

2.4.3.3 Dissipated Energy Ratio

Figure 2-31 Illustration of energy ratio versus number of cycles in a controlled strain
fatigue test

As discussed earlier, the number of cycles to failure is defined differently depending on


the mode of loading. Hopman et al. proposed the use of an “Energy Ratio” to define the
number of cycles (N1) in a controlled strain test, to a point where cracks are considered to
initiate (defined as the merging of micro-cracks to form a sharp crack, which then
propagates), as shown in Figure 2-31 [Hopman, 1989]. The “Energy Ratio” is defined as

37
Chapter 2 Literature Review

follows:

n w0
Energy Ratio = (2-22)
wi
Where: n = number of cycles, w0 = dissipated energy at the start of test, wi = dissipated
energy in cycle i .

The “Energy Ratio” plotted against number of cycles reveals a change in behavior at a
number of cycles, N1. This point is considered to be the formation of a sharp crack.
However, the “Energy Ratio” can be written as Rowe [Rowe, 1993]:

n (π σ 0 ε 0 sin φ0 )
ER= (2-23)
(π σ i ε i sin φi )
Where:
σ 0 , ε 0 and φ0 = stress, strain and phase angle at the start of test,
σ i , ε i and φi = stress strain and phase angle in cycle i .

It is considered that σ can be replaced by E * ⋅ ε , in which E * is the complex stiffness


modulus and the strain level is a constant for a controlled strain test. Then Equation 2-23
can be written as follows:

n (π E0* ε 02 sin φ0 ) n E0* sin φ0


ER= = (2-24)
(π E ε* 2
i i sin φi ) Ei* sin φi

The above equation has a constant term, E0* sin φ0 , which can be removed without
changing the shape of the curve in Figure 2-31. Thus the equivalent ratio, Rε , is reduced
to the following form:

n n
Rε = or (2-25)
Ei* sin φi Ei"
Where: Ei" is the loss modulus at cycle i .

The change in sin φ is very small compared to the change in E * and does not significantly
alter the shape of the curve in Figure 2-31. Hence, the equation of “Energy Ratio” can be
further simplified [Rowe, 1993]:

n
Rε ≃ (2-26)
Ei*

For a controlled stress test, the same method is used to simplify the equation, as follows:

38
Chapter 2 Literature Review

Rσ ≃ n Ei* (2-27)

For controlled strain test data N1 is defined as the point at which the slope of the energy
ratio versus number of load cycles deviates from a straight line, as shown in Figure 2-32
(a), (b) and (c). The load amplitude remains constant in a controlled stress test and after
crack initiation the stress at the crack tip increases rapidly. Fatigue life, N1 , can be
determined from the peak of Rσ , as shown in Figure 2-32 (d), (e) and (f). It can be seen
that the N1 condition is very hard to define for controlled strain tests compared to
controlled stress tests. N f is the fatigue life defined by classical analysis. These two
parameters, N1 and N f , correspond to “crack initiation” and “failure”.

* *
E (MPa) CONTROLLED STRAIN E (MPa) CONTROLLED STRESS
5000 7000

6000
4000
5000
3000
4000

2000 3000

2000
1000
1000

0 0
0 1,E+05 2,E+05 3,E+05 4,E+05 5,E+05 0 1,E+04 2,E+04 3,E+04 4,E+04 5,E+04
Load Cycles Load Cycles

(a) (d)
Rε CONTROLLED STRAIN Rσ CONTROLLED STRESS
1,E+03 1.4E+08

1.2E+08
8,E+02
1.0E+08

6,E+02 8.0E+07

6.0E+07
4,E+02
4.0E+07

2,E+02 2.0E+07

0.0E+00
0 0 1,E+04 2,E+04 3,E+04 4,E+04 5,E+04
0 1,E+05 2,E+05 3,E+05 4,E+05
Load Cycles Load Cycles

(b) (e)

39
Chapter 2 Literature Review

3
wi (J/m ) CONTROLLED STRAIN wi (J/m )
3
CONTROLLED STRESS
250 1000

200 800

150 600

100 400

50 200

0 0
1,E+05 0 1,E+04 2,E+04 3,E+04 4,E+04 5,E+04
0 2,E+05 3,E+05 4,E+05 5,E+05
Load Cycles Load Cycles

(c) (f)
Figure 2-32 E*, Rσ and wi versus load cycles; trapezoidal controlled strain (left) and
controlled stress (right) fatigue tests [Rowe, 1993]

The N1 point occurs generally in a range between 40% and 50% reduction of the initial
extensional complex modulus. In a controlled strain test, the dissipated energy, which is
associated with crack propagation, is usually small. In a controlled stress test, the crack
growth accelerates rapidly after initiation so life associated with crack propagation is also
relatively small. This is illustrated in Figure 2-32 (c) and (f). The shaded area in the
figures is the amount of dissipated energy associated with crack propagation. Thus, both
fatigue tests give similar relationships for N1 and N f . In addition the dissipated energy
after the N1 stage should only be considered as an apparent energy. Consequently, the N1
point is considered to be the preferred failure point, since the energy computed to
the N f criterion will contain a certain amount of error.

Carpenter and Jansen [1997] showed in their study that the change in dissipated energy is
responsible for fatigue damage. They found that the rate of change in dissipated energy
versus load cycles to failure produces a unique curve regardless of loading mode and test
condition. Therefore, Ghuzlan and Carpenter [Ghuzlan, 2000] defined a ratio of
dissipated energy change (RDEC) based on the change in dissipated energy between load
cycle i and load cycle i +1 divided by the dissipated energy in load cycle i :

( DEn+1 − DEn )
RDEC = (2-28)
DEn

Where:
RDEC = the ratio of dissipated energy change;
DEn = the dissipated energy produced in load cycle n;
DEn+1 = the dissipated energy produced in load cycle n+1.

40
Chapter 2 Literature Review

The RDEC represents the percentage of dissipated energy causing damage to the material.
Using this approach, the percent of input dissipated energy that goes into damage for a
cycle can be directly determined during the fatigue test. This ratio is calculated
approximately every 100 cycles.

As introduced by Ghuzlan and Carpenter [Ghuzlan, 2000], the damage curve represented
by RDEC vs. loading cycles can be distinctively divided into three stages. In stage II
(plateau stage) the RDEC is almost constant until the dramatic increase in stage III which
is the onset of true failure. The schematic is shown in Figure 2-33.

Figure 2-33 Typical RDEC plot with three behavior zones [Carpenter 2003]

The Plateau Value (PV), the nearly constant value of RDEC, characterizes a period where
there is a constant percent of input energy being turned into damage. The failure is
defined as the number of load cycles at which this ratio begins to increase rapidly.
Carpenter and Shen [Carpenter, 2006] developed the relation between the PV value and
the number of load cycles to failure Nf:

PV = cN df (2-29)

Where: c and d are regression constants.

2.4.4 Fracture Mechanics Approach


2.4.4.1 Theory
The basic concepts of fracture mechanics as introduced by Griffith [Griffith, 1921]
provide an alternative approach to define the fatigue properties of asphalt mixtures. In
this method, fatigue is considered to develop in three phases [Ewalds and Wanhill, 1986]:
(1) crack initiation; (2) stable crack growth; and (3) unstable crack propagation, as shown
in Figure 2-34. It is assumed that the second phase consumes most of the fatigue life and,
consequently, it is for this phase that quantitative models based on fracture mechanics
have been proposed. The crack propagation phase was first described by Paris and

41
Chapter 2 Literature Review

Erdogan [Paris, 1963]. From experimental data they found by means of regression
analysis that the crack propagation rate dc/dN can be described following:

dc
= AK n (2-30)
dN

Where: c = crack length (in mm); N = number of load repetitions; K = stress intensity
factor describing the stress conditions near the crack tip. A, n = parameters, depending on
the material and on the experimental conditions (waveform, temperature, frequency etc.)

Figure 2-34 Fatigue crack growth

In phase 1, Kth is a threshold value for crack initiation. Above this value the crack growth
rate increases rapidly to a constant value, as shown in Figure 2-34. Equation 2-30 is used
to describe the stable crack growth process. In phase 3 the crack growth rate becomes
infinite. The critical value Kc, called the fracture toughness, is a material property which
is not dependent on the type of loading or on the dimensions of the specimen. The stress
intensity factor K is dependent on the overall stress conditions and the geometry of the
crack [Jacobs 1995].

Based on the experimental data of various Dutch mixtures, Molenaar [Molenaar, 1983]
found that the A-value varies between 5×10-3 and 5×10-9 N/mm1.5 and the n-value
between 2 and 5 for uniaxial tensile tests, frequencies between 1 and 10 Hz and
temperatures between 5 °C and 25 °C.

Generally three crack modes are considered, as shown in Figure 2-35 [Broek 1991].

(a) Mode I, the opening mode (tensile or pure bending stress);


(b) Mode II, the sliding or shearing mode (shear stresses);
(c) Mode III, the tearing mode (torsional stresses).

42
Chapter 2 Literature Review

Mode I Mode II Mode III


Figure 2-35 Three modes of loading to describe the crack growth phenomenon
[Broek 1991]

In most cases combinations of crack modes occur in practice, especially for mode I and
mode II. For the simple pure tension or bending case the stress intensity factor can be
written as [Tseng and Lytton, 1990]:

q
c
K =σ c ⋅r  (2-31)
b

Where: σ = nominal tensile or bending stress = E ⋅ ε ; E = the elastic stiffness of the beam;
ε = maximum strain in the specimen; c = crack length; b = width or height of the
specimen; r, q = regression constants.

At the moment of failure, K has reached a critical value, Kc. The equation for K will
change if the displacement is kept constant during the test. Furthermore K might be
influenced if a crack is initiated both at the bottom and the top of the specimen
(sinusoidal loading).

2.4.4.2 Determination of Fracture Mechanics Parameters


Asphaltic materials show a significant viscous contribution to their behavior. To use the
fracture mechanics approach in practice, it is necessary to determine A and n of Paris’
law by means of simple tests or nomographs. Schapery presented a crack growth theory
for visco-elastic media for mode I cracking using the so-called generalized J-integral
theory [Schapery, 1973] [Schapery, 1975] [Schapery, 1978]. Schapery developed a
relationship between A and n and the measurable properties of the material:

π  (1 − µ 2 ) D2  m ∆t  1
( )
2 1+ 

 ∫0
A= ⋅   ⋅ w t  m  dt (2-32)
6σ m2 I12  2Γ
 
π K1
I1 = (2-33)
2ασ m

43
Chapter 2 Literature Review

Where:
σm = maximum tensile stress the material can withstand before failure, MPa;
I1 = a factor dependent on the stress conditions at the crack tip, the failure
stress and length of failure zone;
α = size of the failure zone in front of the crack tip, mm;
K1 = stress intensity factor for mode I loading, N/mm1.5;
D2 = intercept of a line drawn tangent to the double-log creep compliance
(D(t)-D0) vs. time plot at t = 1 sec;
D(t) = creep compliance, D(t)=D0+D2tm , MPa-1;
D0 = initial creep compliance at t=0 s, MPa-1;
µ = Poisson’s ratio;
Γ = fracture energy defined as the work done on a material to produce a unit
area of crack surface, N/mm;
w (t ) = pulse shape of the stress intensity factor;
m = slope of the compliance curve;
t = time, s;
∆t = period of one loading cycle, s.

According to Schapery’s theory, the constants A and n from Paris’ law can be determined
from simple tests or from nomographs [Heukelom and Klomp, 1964] [Bonnaure et.al
1977]. Also fatigue data at different temperatures and for different geometries can be
predicted directly. Furthermore the influence of various material properties can be
evaluated explicitly without using extensive series of fatigue tests.

Several researchers have verified Schapery’s theory. Molenaar considered that


differences exist between the experimentally determined n-values and the theoretical ones
[Molenaar, 1983]. These differences are mainly caused by the limitations of the model.
Schapery derived his theory for a homogenous material which is not consistent with
asphalt mixtures because of the presence of air voids and aggregates. Especially air voids
strongly influence the crack growth rate. For mixtures with a void content more than 3%
n-values need to be corrected by regression analysis [Molenaar, 1983]. Based on the
evaluation of crack growth tests on COD-measurements, similar findings were reported
by Jacobs [Jacobs, 1995] who found that correction factors were needed in Schapery
relationships to predict values for the parameters A and n adequately.

2
n= (2-34)
m ⋅ CF
n n
log A = d − 2a log σ m − b log 2Γ m − c log S mix (2-35)
2 2
d (log S mix )
m=− (2-36)
d (log t )
CF = exp ( b0 + b1 ⋅ Smix + b2 ⋅ Sbit + b3 ⋅ Sbit ln ( Sbit ) ) (2-37)

44
Chapter 2 Literature Review

where:
a, b, c, d = regression coefficients depending on mix type (a=0.39, b=-0.04,
c=0.62, d=-1.29 for dense asphalt concrete and stone matrix
asphalt);
b0, b1, b2, b3 = regression coefficients depending on mix type (b0=0.34,
b1=-3.58×10-4, b2=-6.67×10-3, b3=1.01×10-4 for dense asphalt
concrete and stone matrix asphalt);
Smix = stiffness modulus of the mix determined from master curves;
Sbit = stiffness modulus of bitumen backcalculated from Smix.

Erkens et al. and Sabha et al. did measurements on gravel asphalt concrete (GAC) and
renewed the regression analysis. A new correction factor for GAC was expressed as
[Erkens, 1995] [Sabha, 1995]:

CF = 0.89981 + 7.2814 × 10 −5 ⋅ S mix + 7.1487 ×10 −9 ⋅ S mix


2
(2-38)

2.5 Fatigue Prediction Model


2.5.1 Partial Healing (PH) Model
The PH model is originally a material model and is based on the viscous elastic
dissipated energy per cycle, ∆W fat , which is used to create micro defects and cracks
[Pronk, 2001]. ∆W fat can be regarded as a small portion of the dissipated energy ∆Wdis .
Pronk stated that the stiffness damage term Q is related to the fatigue consumption
∆W fat [Pronk, 2005].

d d ∆Wdis
Q ∴ ∆W fat ∴∆Wdis ⇒ Q= δ ⋅ Wdis ≈ δ (2-39)
dt dt T
With T= time of one cycle.

It is assumed that the stiffness damage Q affects both the loss modulus S mix ⋅ sin ϕ and the
storage modulus S mix ⋅ cos ϕ in which S mix is the complex stiffness modulus and ϕ is the
phase angle. The two expressions of the PH model are given in Equation 2-40 and 2-41:

dQ {t}
(α e )
t
S mix {t} sin (ϕ {t}) = S mix {0} sin (ϕ {0} ) − ∫
− β ( t −τ )
+ γ 1 dτ (2-40)

1
0

dQ {τ }
(α e )
t
S mix {t} cos (ϕ {t} ) = S mix {0} cos (ϕ {0} ) − ∫
− β ( t −τ )
+ γ 2 dτ (2-41)

2
0

The terms with the parameters α1,2 in the above equations will be ‘healed’ in time with
time decay β . The terms with the parameter γ 1,2 will not heal. Pronk gave the complete
solutions of Equation 2-40 and 2-41 [Pronk, 2000]. From earlier tests [Pronk 1997], it

45
Chapter 2 Literature Review

was found that in continuously loaded 4 point bending tests no or negligible healing
occurs for the loss modulus. Therefore, the parameter α1 can be taken as equal to zero. An
example of a simulation of a 4PB controlled strain test is presented in Figure 2-36.

Figure 2-36 Data and model fitted for a 4PB controlled strain test at 172 µm/m
[Pronk, 2005]

Table 2-4 shows parameter values, obtained from regression analysis. It can be seen
that β is a function of the applied strain amplitude level and its value increases with
increasing strain amplitude, as shown in Figure 2-37.

Table 2-4 Results of the regression from 4PB [Pronk, 2005]

Strain (µm/m) α 2 ⋅ δ (-) γ 2 ⋅ δ (-) β (10-5 s-1) γ 1 ⋅ δ (-)


137 618 36.42 111 6.6
177 618 36.42 146 6.6
217 618 36.42 248 6.6

Figure 2-37 Parameter β as a function of the strain amplitude (4PB) [Pronk, 2005]

46
Chapter 2 Literature Review

The PH model is also applied in strain controlled Tension/Compression (T/C) test on


cylindrical specimens. According to Pronk [2005], the model parameters of T/C tests
could be deduced from those obtained using 4PB tests. Therefore, the PH model can
predict stiffness and phase angle evolution during fatigue tests for both 4PB and T/C tests.

2.5.2 Weibull’s Theory


In probability theory and statistics, the Weibull distribution is a continuous probability
distribution. It is named after Waloddi Weibull who described it in detail in 1951. The
Weibull distribution is one of the most widely used lifetime distributions in reliability
engineering. Pronk introduced Weibull’s theory into the research on fatigue tests of
asphalt concrete [Pronk, 1998].

2.5.2.1 Survival Probability

Pronk [1999] defined the survival probability S1 ( xi ) for a unit volume of material in an
uniaxial fatigue test. The parameter xi stands for the loading conditions at a certain
moment during the fatigue test. Survival probability means that the change from the
fatigue crack initiation phase into the crack propagation phase has not yet occurred at that
moment. This parameter will be between 0 and 1. In case of the Weibull distribution, the
survival probability of a unit volume can be expressed as:

− Cunit ,volume ( xi )
Sunit ,vol . ( xi ) = e (2-42)

Where: Cunit ,volume ( xi ) is defined as the "risk of change".

In ceramics rupture tests (increasing the stress to a level at which the specimen breaks)
the function C, called the "risk of rupture", is defined as [Weibull, 1951]:

σ − σ1 
m

C ( xi ) = R (σ i ) =  i  (2-43)
 σ0 

Where: σ is the applied stress, MPa; σ1 is the stress which there is a zero probability of
failure, MPa; σ0 is the mean strength of the material, MPa; m is the Weibull modulus.

In case of uniaxial fatigue tests in controlled strain mode, the following expression is
supposed to be valid based on Equation 2-43:

m
 N i ⋅ ε ib 
Cunit ,volume ( xi ) =   (2-44)
 Qv 

47
Chapter 2 Literature Review

Where: ε i is the applied strain level; N i is the fatigue life under strain level ε i ; Qv is a
reference constant.

The ‘risk of change’ for a volume of material is given by:

dCvolume ( xi ) = Cunit ,volume ( xi ) ⋅ dVolume (2-45)


m
 N ⋅ε b 
Cvolume ( xi ) = ∫ Cunit ,volume ( xi ) ⋅ dVolume = ∫  i i  ⋅ dVolume (2-46)
volume volume 
QV 

The survival probability of a certain material volume is given by:


m
 Ni ⋅ε ib 
− ∫   ⋅dVolume
Svolume ( xi ) = e volume  QV 
(2-47)

2.5.2.2 Calculation of Survival Probability


In Equation 2-47, the parameter m is called the Weibull modulus. If in the volume which
is subjected to a certain stress, the loading conditions stay constant, such as in a uniaxial
push-pull test, the survival probability can be integrated leading to:
m
 N ⋅ε b 
−  i i  ⋅Volume
Svolume ( xi ) = e
 QV 
(2-48)

The value of the parameter m can be determined by taking twice the natural logarithm of
the reverse value of the survival probability:

  1   N i ⋅ ε ib 
log log   = log [ volume ] + m ⋅ log   (2-49)
  S volume ( xi )    QV 

The survival probability itself can be calculated from the ranking sequence of the
measurements. Two equations are possible:

j j − 0.3
S ( j) = 1 − ; S ( j) = 1 − (2-50)
T −1 T + 0.4

Where: T is the total number of measurements; j is the ranking according to the loading
N i ⋅ ε ib
condition xi = .
Qv

48
Chapter 2 Literature Review

  1 
  against log  N i ⋅ ε i  , the slope of the linear regression is the
b
By plotting log log 
  S ( j )  
Weibull modulus m.

If the survival probabilities of two different uniaxial tests are the same, the following
requirement is fulfilled:

Volume1 ⋅ x1m = Volume2 ⋅ x2m (2-51)

In dynamic bending tests, such as a four point bending test, the maximum strain occurs at
the surface of the beam and decreases linear to zero at the ‘neutral’ zone in the middle of
the beam for tests in constant strain mode. Taking into account that the surface cracks
have a significant effect on the initiating of microcracks, the "risk of change" can be
calculated by:

C = Carea + Cvolume = ∫C
area
unit , area dArea + ∫
Volume
Cunit ,volume dVolume (2-52)
m m
 N i ⋅ ε ib   N i ⋅ ε ib 
Cunit ,volume ( xi ) =   and Cunit ,area ( xi ) =  
 QV   QA 

In order to calculate the "risk of change" of beams in a four point bending test, two
assumptions are made:

(1) For a beam with a total length 2 L (x axis), a height 2 H (z axis), a width 2 H (y axis)
and the first inner clamp at the distance x = A = α L , the strain distribution is shown in
Figure 2-38.

y
z
compressive
strain
H
0
x
-H tensile
A strain

L L
x z
ε = ε0 ⋅
⋅ for 0 ≤ x ≤ A and 0 ≤ z ≤ H
A H
z
ε = ε0 ⋅ for A ≤ x ≤ L and 0 ≤ z ≤ H
H
Figure 2-38 The strain distribution in the beam in four point bending test

49
Chapter 2 Literature Review

(2) The tensile strains and compression strains equally in creating of fatigue damage.
Integration over the total volume of the beam will give the following relationship:

 y =2 B z=H x= A y=2B z=H x=L 


4 ∫ ∫ ∫ Cunit ,volume ( xi ) dxdzdy + ∫ ∫ (
∫ unit ,volume i
C x ) dxdzdy 
 y = 0 z =0 x =0 y =0 z =0 x= A 
(2-53)
 mb (1 − α ) + 1   N i ⋅ ε 0b 
m

= 8 BHL ⋅  ⋅ 
 ( mb + 1)   Qv 
2

If taking into account the stressed surface area, the following equation is obtained:
m
8L  N i ⋅ ε 0b 
⋅  ( mb ⋅ (1 − α ) + 1)(
⋅ H + ( mb + 1) ⋅ B )  Q 
 (2-54)
( mb + 1) 
2
 A 

Therefore the "risk of change" of beams in a four point bending test can be calculated by
combining the two equations above.

 mb (1 − α ) + 1   N i ⋅ ε 0b 
C = 8 BHL ⋅  ⋅ +
 ( mb + 1)   Qv 
2

(2-55)
m
8L  N ⋅ε  b

⋅ ( mb ⋅ (1 − α ) + 1) ⋅ ( H + ( mb + 1) ⋅ B )   i 
0

( mb + 1)
2
 QA 

2.5.2.3 Application in Fatigue Test


Pronk carried out the four-point bending fatigue tests in controlled strain mode and
compared the test results of beams with different mid span lengths of 90 mm, 130 mm
and 190 mm [Pronk, 1999]. The beam dimensions were: a total length of 450 mm (with
an effective length between two outer clamps of 400 mm), height of 50 mm and also a
width of 50 mm.

Figure 2-39 represents the relationships between fatigue life N1 and the applied constant
strain amplitudes. Here the fatigue life N1 was determined by the dissipated energy ratio,
as discussed in Section 2.4.3. Pronk found that there is little difference between the two
Wohler curves for a mid span of 90 and 190 mm and the slope b of the Wohler curve for
the standard mid span length of 130 mm is a little bit lower. The slopes of Wohler curves
for the mid span of 90 and 190 mm are b1 = 4.471 and b2 = 4.455 respectively.

The survival probabilities (calculated by Equation 2-48) are plotted for the measurements
with a mid span length of 90 mm and 190 mm, as shown in Figure 2-40. The slope of the
curves corresponds to the value for the Weibull modulus m.

50
Chapter 2 Literature Review

1000

100

10
100 200

Figure 2-39 Fatigue lines for different mid span lengths of beams [Pronk, 1999]

T = 15

j=15

….
j=3
j=2

j=1

(a)

(b)
Figure 2-40 Survival probability as a function of the parameter N ⋅ ε b Fatigue
measurements with a mid span length of 90 mm (a) and 190 mm (b) [Pronk, 1999]

51
Chapter 2 Literature Review

In order to be able to compare the fatigue lives for the two different mid span lengths, the
survival probabilities should be the same. For the Weibull modulus m a mean value of
4.1 is adopted and for the coefficient b of the Wohler curves a mean value of 4.46 is
taken.

A. For the volume damage the following equation is obtained (Equation 2-53):

 155   105 
18.3 ⋅ 1 −  +1 18.3 ⋅  1 −  +1
 200  ⋅  N ⋅ ε 4.46  4.1 =  200  ⋅  N ⋅ ε 4.46  4.1 (2-56)
 i 0   i 0 
(18.3 + 1) (18.3 + 1)
2 2

(mid span length of 90 mm) (mid span length of 190 mm)

According to this equation, at the same strain level the ratio in fatigue lives will be 1.15
for the same survival probability.

B. For the surface damage the ratio in fatigue lives will be nearly the same: 1.17
(calculated by Equation 2-54).

2.5.3 Mechanical Damage Model


In order to describe the whole damage process during a fatigue test, LCPC has developed
an approach based on mechanical damage modeling. The presented modeling and
simulation have been performed by Bodin [Bodin, 2002]. This damage approach has been
used to describe the local complex modulus decrease induced by microcrack
development.

2.5.3.1 Theory
The mathematical model used to describe mechanical damage is an elasticity based
damage model for fatigue, which is inspired by the non-local damage model proposed by
Mazars and Pijaudier-Cabot [Mazars, 1984] [Pijaudier-Cabot, 1987]. Compared to
existing proposals, e.g., by Lee et al. [Lee, 2000], the constitutive relations are simpler
since they rely on a scalar, isotropic, damage model which is widely used for concrete
and other quasi brittle materials. The influence of microcracking is introduced via a scalar
damage parameter, d, that ranges from 0 to 1, which affects the Young’s modulus.

The evolution of damage, d, depends on the amount of tensile stress that the material
experiences during mechanical loading. The equivalent strain εɶ induced by a principal
stress σ i is assumed to lead to damage growth as follows:

3  σi 
εɶ = ∑  E (1 − d ) 
+
(2-57)
i =1  

where +
= Macauley bracket; σ i = principal stress; E = Young’s modulus.

52
Chapter 2 Literature Review

For structural computation, the weighted average value of equivalent strain, ε , is used to
replace the equivalent strain εɶ .

1
ε ( x) = ψ ( x − s ) εɶ ( s ) ds , with Vr ( x ) = ∫ ψ ( x − s ) ds
Vr ( x ) ∫Ω
(2-58)

where Ω = volume of the structure; Vr(x) = representative volume at point x; and Ψ(x-s)
= weight function:

 4 x−s 2 
ψ ( x − s ) = exp  −  (2-59)
 l 2

 c 
and lc = internal length of the nonlocal continuum. For concrete, lc is often specified as
equal to three times the size of the largest aggregate fraction.

The damage growth rate is defined as a function of the equivalent strain rate [Mazars,
1989]:

1−α 3 α3
α2  d   d 
dɺ = f ( d ) ε β
εɺ with f ( d ) =   exp   (2-60)
+ α1 ⋅ α 3  α 2   α2 

where α1, α2, α3 and β are material parameters, and it can be demonstrated that –(β+1)
corresponds to the slope of the Wohler fatigue curve. f(d) is an original law modeling the
three phases of the loss of modulus curve during fatigue testing.

The non-local damage model has been implemented into a finite element code "Castem
2000", and a specific so-called "jump in cycle" procedure is computed to allow
calculations involving large number of cycles. More details on the construction of the
model can be found in [Di Benedetto, 1999] and [Bodin, 2002 and 2003].

2.5.3.2 Size Effect of the Damage Model


For the size effect study, Bodin conducted two point bending tests with different
specimen sizes [Bodin, 2002]. The specimens are 2D geometrically similar and their
thickness is the same as shown in the Figure 2-2 and Table 2-1. Size 1 corresponds to the
standardized size.

The size effect is illustrated with the stiffness decrease curves. Figure 2-41 shows the
results simulated at a strain level of 140×10-6 m/m. Model parameters α1, α2, α3 and β are
fixed to take into account the experimental scatter of experimental data. The mean
parameter set had been fitted for the average experimental data. The maximum and
minimum parameters set had been fitted on experimental data. Compared to the standard
size (size = 1), the stiffness decrease is faster for large samples (size = 2) and much
slower for small samples (size = 0.5).

53
Chapter 2 Literature Review

Size = 1
Size = 2 Size = 0.5

Figure 2-41 Simulated stiffness decrease calculated with the damage model fitted for the
strain level 140 µ ε [Bodin, 2002]

2.5.3.3 Comparison between Model and Experiments

Model

Model

Model

Figure 2-42 Comparison of the model prediction with experimental data for the three
specimen series for the strain level 140 µ ε [Bodin, 2006]

54
Chapter 2 Literature Review

To validate the nonlocal damage model, Bodin repeated the two-point bending tests using
similar granular materials but manufactured with a different binder [Bodin, 2006]. Non
local fatigue damage simulations are plotted in a gray zone superimposed to experimental
data for the three tested sizes, as shown in Figure 2-42. It is observed that the fatigue life
is well predicted for the size 2.0 and the difference between model and experiment grows
with the decrease of the size. For the smallest size the model predicts failure at a larger
number of cycles than it is observed in the lab.

Figure 2-43 illustrates the model fatigue lines with a visualization of the prediction zone
given by the error bars for the three loading levels. Model agreement is better for large
samples with a fatigue line very close to experimental data. For the other sizes the
differences between model and experimental data grows with the decrease of the size.

Figure 2-43 Comparison of model predictions and experiment fatigue line for each size

2.6 Summary
Based on this literature review on fatigue of asphalt mixtures, the following findings can
be reported:

1. Fatigue damage, a form of cracking resulting from repeated traffic loading, is


recognized as one of the main failure modes of pavement structures. The fatigue
characteristics of asphalt concrete mixtures should be known for the thickness design of
asphalt pavements.
2. The fatigue response of a asphalt mixture is a complicated phenomenon and has not
been understood completely. When asphalt mixtures are subjected to repeated loading,
the fatigue cracks propagate gradually, leading to failure at last. However other
phenomena are also involved in this process, like self-healing, thermal effects, permanent
deformation and so on. It is important to describe these phenomena in a correct way.

55
Chapter 2 Literature Review

3. Laboratory fatigue tests are widely used to simulate the situation in the field. However,
accurate prediction and evaluation of fatigue is a difficult task not only because of the
complex nature of the fatigue phenomena but also because of the fatigue testing itself,
which is expensive and time-consuming. In principle, fatigue resistance should be a
material property, but fatigue behavior tends to be very sensitive to material type, test
characteristics and environmental conditions. Based on the fact that type of test, specimen
size and mode of loading all affect the fatigue testing result, laboratory fatigue tests only
provide the properties of specimens rather than of materials. Therefore large shift factors
have to be applied to ‘adjust’ the laboratory data for practical pavement design purposes.
4. Different countries or institutions now apply different fatigue tests as the standard
method for design. The experimental fatigue results obtained for the same asphalt
mixture are different and difficult to compare and exchange with each other. For this
reason, it is necessary to harmonize the existing test methods so that the evaluation of
fatigue performance is more accurate and the test method can be reproduced.
5. The phenomenological approach, widely used to evaluate fatigue performance in the
lab, has its limits, because the definition of fatigue life is based on an arbitrary choice.
Failure in controlled strain has been defined as a 50% reduction in the initial stiffness
modulus. The fatigue life defined by this traditional method is material and test
conditions dependent and is not an intrinsic property of the materials. Therefore a new
fundamental definition is needed for the fatigue life in order to harmonize the different
fatigue tests.

References

Aguirre, Morot, De La Taille, Doan Tu Ho, Bargiacchi, Smadja, Udron, Guay et Roncin,
Etude comparée des essais de module complexe et de résistance à la fatigue des enrobés
bitumineux - Bulletin de Liaison des Laboratoires des Ponts et Chaussées, N° 116, pp.33-
43, 1981.

ASTM stander D 7460, Standard test method for Determining the fatigue life compacted
Hot-Mix Asphalt (HMA) subjected to repeated flexural bending, 1996.

Bonnot, J., Asphalt Aggregate Mixtures. Transportation Research Record 1096,


Transportation Research Board, Washington, D. C., 1986, pp: 42-50.

Bonnaure, F.P., Gest G., Gravoir A. and Ugé P., A new Method of Predicting the
stiffness of Asphalt Paving Mixtures, Proceedings AAPT, 1977, Vol.46, pp: 64-105.

Bonnaure, F.P., Gravois, A. and Udron, J., A New Method for Predicting the Fatigue Life
of Bituminous Mixes, the Association of Asphalt Paving Technologists (AAPT), 1980,
Volume 49, pp 499-528.

Brennan M.J., Lohan G. and Golden J.M., A laboratory study of the effect of bitumen
content, bitumen grade, nominal aggregate grading and temperature on the fatigue

56
Chapter 2 Literature Review

performance of dense bitumen macadam, Proceeding of the IVth International Rilem


Symposium Budapest, Ed. Chapman & Hall. 1990.

Brown, S. F., Pell P.S. and Stock A.F., The Application of Simplified, Fundamental
Design Procedures for Flexible Pavements, Proceedings fourth ICSDAP, Ann Arbor,
Michigan, 1977, Vol.1, pp: 327-341.

Brown, S. F., Material characteristics for analytical pavement design, Developments in


Highway Pavement Engineering-1, ed. P.S. Pell, London, 1978 .

Brown, S.F., Gibb, J.M., Read, J.M., Scholz, T.V., Cooper, K.E., Design and Testing of
Bituminous Material. Volume 2: Research Report, Submitted to DOT/EPSRC LINK
Programme on Transport Infrastructure and Operations, 1995.

Carpenter, S. H. and Jansen, M., Fatigue behavior under new aircraft loading conditions,
In Aircraft/Pavement Technology: In the Midst of Change, Seattle, Washington. Edited
by F.V. Hermann, American Society of Civil Engineers, New York, 1997, pp:259-271.

Carpenter, S.H. and S. Shen., A Dissipated Energy Approach to Study HMA Healing in
Fatigue. In Transportation Research Record: Journal of the Transportation Research
Board, No.1970, TRB, National Research Council, Washington D. C., 2006, pp. 178-185.

Chkir, R., Bodin, D., Piau, J.M., Pijaudier-Cabot, G., Gauthier, G. and Gallet, T., An
inverse analysis approach to determine fatigue performance of bituminous mixes, Mech
Time-Depend Mater, 2009, 13: 357–373.

Chomton, G. and Valayer, P.J., Applied Rheology of Asphalt Mixes-Practical


Applications, Proceedings, Third International Conference on the Structural Design of
Asphalt Pavements, London, England, 1972, pp: 214-225.

Di Benedetto H., Ashayer Soltani M.A. and Chaverot P., Fatigue damage for bituminous
mixtures: a pertinent approach. Journal of the Association of Asphalt Paving
Technologists, 1996, 65, 141–176.

Di Benedetto H., de la Roche C., Baaj H., Pronk A. and Lundstrom R., Fatigue of
bituminous mixtures. Materials and Structures, 2004, 37, No. 3, 202–216.

Di Benedetto H., de la Roche C., Francken L., fatigue of bituminous mixtures: Different
approaches and RILEM interlaboratory tests, Proceedings of the 5th International RILEM
Symposium, Mechanical Tests for Bituminous Materials—Recent Improvements and
Future Prospects, Lyon, 1997, 15–18.

Di Benedetto H., Ashayer Soltani M.A. and Chaverot P., Étude rationnelle de la fatigue
des enrobés: annulation des effets parasites de premier et second ordre, Int. Eurobitume
Workshop, Performance related properties for bituminous binders 4p, 1999.

57
Chapter 2 Literature Review

Dider Bodin, Size effect regarding fatigue evaluation of asphalt mixtures, 2nd European
Asphalt Technology Association Meeting, 2006.

Dider Bodin, Modèle d’endommagement cyclique: Application à la fatigue des enrobés


bitumineux, PhD thesis, EC Nantes, 2002.

Dider Bodin, de La Roche, C. and Piau, J.M., A damage approach for asphalt mixture
fatigue tests, Paper proposed to the 9th International Conference on Asphalt Pavements-
Copenhagen, 2002.

Dider Bodin, Chabot, A., de La Roche, C. and Pijaudier-Cabot, G., Endommagement par
fatigue des matériaux bitumineux-Recherche de lois d’évolution de l’endommagement,
Matériaux, Tours, France, 2002, pp:21-25.

Dider Bodin, de La Roche, C., Piau, J.M. and Pijaudier-Cabot, G., Prediction of the
intrinsic damage during bituminous mixes fatigue tests, 6th Internal RILEM Symposium:
Performance Testing and Evaluation of Bituminous Materials, 2003.

Erkens, S.M.J.G., and Moraal, J., A practical method to determine the fatigue
characteristics of asphalt concrete, Report 7-95-117-1, DUT, Faculty of Civil engineering,
Delft, the Netherlands, 1995.

EN 12697-24: 2003 Bituminous mixture – Test methods for hot mix asphalt - Part 24:
Resistance to fatigue. European committee for standardization, Brussels, 2003.

Ewalds H.L. and Wanhill R.J.H., Fracture Mechanics, Co-publication of the Delftse
Uitgevers Maatschappij and Edward Arnold Publishers, London, 1986.

Francken L., Fatigue performance of a bituminous road mix under realistic test conditions.
Transportation Research Record, 712, Washington, 1979, pp: 30-37.

Francken L. and Clauwaert C. Characterization and structural Assessment of Bound


Materials for Flexible Road Structures, Proceedings of the Sixth ICSDAP, Ann Arbor,
Michigan, 1987 pp: 130-144.

Francken L., Bituminous Binders and Mixes, Rilem Report 17, E&FN Spon. 1998.

Finn, F.N., Saraf, C., Kulkarni, R., Nair, K., Smith, W. and Abdulah, A., The use of
Distress Prediction Subsystems for the Design of Pavement Structures, Proceedings,
Fourth International Conference on the structural Design of Asphalt Pavements,
University of Michigan, Ann Arbor, 1977, Vol.1, pp: 3-38.

Germann F.P. and Lytton R.L., Methodology for Predicting the Reflection Cracking Life
of Asphalt Concrete Overlays, Report No. TTI-2-8-75-207-5, Texas Transportation
Institue of the Texas A&M University, College Station, 1979.

58
Chapter 2 Literature Review

Ghuzlan, K, and S. H. Carpenter, Energy-Derived/Damage-Based Failure Criteria for


Fatigue Testing, In Transportation Research Record: Journal of the Transportation
Research Board, No. 1723, TRB, National Research Council, Washington D.C., 2000, pp.
141-149.

Griffith, A.A., The Phenomena of Rupture and Flow in Solids, Philosophical


Transactions of the Royal Society, London, Series A, 1921, Vol. 221.

Hertzberg, Richard W., Deformation and fracture mechanics of engineering materials, 3rd
edition, New York, publisher John Wiley and Son, 1989.

Heukelom W. and Klomp A.J.G., Road Design and Dynamic Loading, Proceedings
AAPT, 1964, Vol.33, pp: 92-125.

Himeno, K., Watanabe, T., and Maruyama, T., Estimation of Fatigue Life of Asphalt
Pavements, 6th International Conference on Structural Design of Asphalt Pavements
(ISAP), Ann Arbor, 1987, pp 272-288.

Hveem, F.N., Pavement Deflections and Fatigue Failures, Highway Research Board,
Bulletin 114, Washington D.C., 1955.

Hopman P.C., Kunst P.A. and Pronk A.C., A renewed interpretation method for fatigue
measurements-Verification of Miner’s rule, Proc. of the 4th Eurobitume Symposium,
Madrid, Spain, 1989.

Jacobs, M. M. J., Crack Growth in Asphaltic Mixes, Ph.D. dissertation, Delft University
of Technology, Delft, the Netherlands, 1995.

Kennedy, T. W. and Hudson, W. R., Application of the Indirect Tensile Test to Stabilized
Materials, Highway Research Record 235, Highway Research Board, Washington, D.C.,
1968.

Kennedy, T. W., Characterization of Asphalt Pavement Materials Using the Indirect


Tensile Test, Proceedings of the Association of Asphalt Paving Technologists, 1977, Vol.
56.

Kennedy, T. W. and Anagnos, J. N., Procedures for the Static and Repeated-Load
Indirect Tensile Tests. Research Record 183-14, Center for Transportation Research,
University of Texas at Austin, 1983.

Khosla, N.P., Omer, M.S., Characterization of Asphaltic Mixtures for Prediction of


Pavement Performance, TTR No. 1034, Transportation Research Board.
Kim Y., Lee H. J., Little D. N. and Kim Y. R., A simple testing method to evaluate
fatigue fracture and damage performance of asphalt mixtures, Journal of the Association
of Asphalt Paving Technologists, 2006, Volume 75, 755-788.

59
Chapter 2 Literature Review

Kim Y. R., Little D. N. and Lytton R. L., Use of dynamic mechanical analysis (DMA) to
evaluate the fatigue and healing potential of asphalt binders in sand asphalt mixtures,
Journal of the Association of Asphalt Paving Technologists, 2002, Volume 71, 176-206.

Kunst, P.A.J.C, Surface Cracking on Asphalt Layers, Working Committee B12,


Hoevelaken, Holland, 1989.

Lee, H. J., Kim, Y. R. and Lee, S. W., Prediction of Asphalt Mix Fatigue Life with
Viscoelastic Material Properties. Transportation Research Record, No. 1832, 139-147.

Lee, H. J., Daniel, J.S. and Kim, Y. R., Continuum damage mechanics-based fatigue
model of asphalt concrete, Journal of Material in Civil Eng. 105, 2000, pp: 1005-1012.

Little, D.N., Investigation of Microdamage Healing in Asphalt and Asphalt Concrete,


Task K, Semi-Annual Technical Report Western Research Institute, FHW A Project
DTFH61-92-C-00170-Fundamental Properties of Asphalt and Modified Asphalt, 1995.

Mazars, J., Application de la mécanique de l’endommagement au comportement non


linéaire et à la rupture des bétons de structure, Thèse de Doctorat d’État, Université Pierre
et Marie Curie, France, 1984.

Medani, T.O., Molenaar, A.A.A., Estimation of fatigue characteristics of asphaltic mixes


using simple tests. Heron, 2000, Vol. 45(No. 3): pp. 155-166.

Mazars, J. and Pijaudier-Cabot, G., Continuum damage theory – Application to concrete,


International Journal of Engineering mechanics, 1989, vol. 115, num.2, pp: 345-365.

Molenaar, A.A.A., Fatigue and Reflection Cracking due to Traffic Loads, Proceedings
AAPT, 1984, Vol. 53, pp: 440-474.

Molenaar, A.A.A., Structural Performance and Design of Flexible Road Construction and
asphalt Concrete Overlays, Ph.D Thesis, Delft University of Technology, the Netherlands,
1983.

Molenaar, J.M.M., Standard execution of the dynamic four-point bending test; Report
MAO-R-87060; RHED; Delft, 1987.

Monismith, C. L. and Deacon, J. A., Fatigue of Asphalt Paving Mixtures, ASCE


Transportation Engineering Journal, 1969, Vol. 95:2, pp: 317-346.

Monismith, C.L., Epps, J.A., Kasianchuk, and McLean, D.B., Asphalt Mixture Behavior
on Repeated Flexure. Report No. TE 70-5, University of California, Berkeley, 1971.
Monismith, C. L., Inkabi, K., McLean, D. B., and Freeme, C. R., Design Considerations
for Asphalt Pavements, Report No. TE 77-1, University of California, Berkeley, March,
1977.

60
Chapter 2 Literature Review

Monismith, C. L., Fatigue Characteristics of Asphalt Paving Mixtures and Their Use in
Pavement Design, Proceedings, 18th Paving Conference, University of New Mexico,
Albuquerque, 1981.

Monismith, C. L. and Salam, Y.M., Distress Characteristics of Asphalt Concrete mixes,


Proceedings of the Association of Asphalt Paving Technologists (AAPT), 1979, Volume
42, pp: 320-349.

Monismith, C.L., J.A. Epps and F.N. Finn, Improved Asphalt Mix Design, Proceedings,
Association of Asphalt Paving Technologists Technical Sessions, San Antonio, Texas,
1985, pp: 347-406.

Moutier, F., Duan, T.H. and Chauvin, J.J., The Effects of Formulation Parameters on the
Mechanical Behaviour of Mixes, Proceeding of the Association of Asphalt Paving
Technologists (AAPT), 1988, Volume 57, pp 213-242.

Nijboer and van der Poel, A study of vibration phenomena in asphaltic road constructions,
Proceeding of the Association of Asphalt Paving Technologists, 1953, No. 22, pp:197-
231.

Paris P.C., and Erdogan K. A Critical Analysis of Crack Propagation Laws, from:
Transactions of the ASME, Journal of basic Engineering, Series D, 85, No. 3, 1963.

Pijaudier-Cabot, G. and Bazant, Z.P. Non local damage theory, Journal of Engineering
Mechanics 113 (10), 1987, pp: 1512-1533.

Priest, A., Timm D., Methodology and Calibration of Fatigue Transfer Functions For
Mechanistic-Empirical Flexible pavement design, NCAT Report 06-03, Auburn
University, Alabama, 2006.

Pell, P. S., McCarthy , P.F. and Gardner, R.R., Fatigue of Bitumen and Bituminous
Mixes, International Journal of Mechanical Science, Pergamon Press Ltd, 1961, pp: 247-
267.

Pell, P. S., Fatigue Characteristics of Bitumen and Bituminous Mixes, Proceedings,


International Conference on the Structural Design of Asphalt Pavements, 1962.

Pell, P.S. Fatigue of Bituminous Materials in Flexible Pavements, Proc. Institution of


Civil Engineers, 1965, Vol. 31.

Pell, P. S. and Brown, S. F., The Characteristics of Materials for the Design of Flexible
Pavement Structures, Proceedings, Third International Conference on the Structural
Design of Asphalt Pavements, London, 1972.
Pell, P. S. and Hanson, J. M., Behavior of Bituminous Road Base Materials under
Repeated Loading, Proceedings Association of Asphalt Paving Technologists, 1973a.

61
Chapter 2 Literature Review

Pell, P. S., Characterization of Fatigue Behavior, in Structural Design of Asphalt


Concrete Pavements to Prevent Fatigue Cracking. Special Report 140, Highway Research
Board, 1973b, 49-64.

Pell P. S. and Cooper, K. E., The Effect of Testing and Mix Variables on the Fatigue
Performance of Bituminous Materials," Proc. The Association of Asphalt Paving
Technologists, 1975, Vol. 44.

Porter, O.J., Foundations for Flexible Pavements, Proceedings, Highway Research Board,
Washington, D.C., 1942.

Pronk, A. C., Evaluation of the dissipated energy concept for the interpretations of
fatigue measurements in the crack initiation phase, The Road and Hydraulic Engineering
Division, Ministry of Transport, Public Work and Water Management, The Netherlands,
1995

Pronk, A. C., Theory of the Four Point Dynamic Bending Test, Research Report,
Ministerie van Verkeer en Waterstaat Dienst Weg en Waterbouwkunde, The Netherlands,
1996.

Pronk, A. C., Healing during fatigue in 4 point dynamic bending tests, Report: W-DWW-
97-095, DWW, Delft, The Netherlands, 1997.

Pronk, A. C., Harmonisation of Bending Fatigue Tests: A(n) (Im)possibility? (In Dutch)
Proc., Wegbouwkundige Werkdagen 1998. CROW, Ede, Netherlands.

Pronk, A. C., Fatigue lives of asphalt beams in 2 and 4 point dynamic bending tests based
on a ‘new’ fatigue life definition using the ‘Dissipated Energy Concept’ Phase IV: DAB
0/8 Weibull stressed volume effect, Report: W-DWW-99-070, DWW, Delft, The
Netherlands, 1999.

Pronk, A. C., Partial healing model-Curve fitting, Report: W-DWW-2000-047, DWW,


Delft, The Netherlands, 2000.

Pronk, A. C., Partial Healing in Fatigue Tests on Asphalt Specimen, Road Materials and
Pavement Design, Vol. 4, Issue 4/2001.

Pronk, A. C., Partial Healing, A new approach for the damage process during fatigue
testing of asphalt specimen, Symp. Amer. Soc. Civ. Eng., Baton Rouge, USA, 2005.

Raithby, K. D. and Sterling, A. B. Some Effects of Loading History on the Performance


of Rolled Asphalt, TRRL-LR 496, Crowthorne, England, 1972.

Read, J. M. and Collop, A. C., Practical fatigue characterisation of bituminous paving


mixtures. Journal of the Association of Asphalt Paving Technologists, 1997, 66, pp: 74-
108.

62
Chapter 2 Literature Review

Read, J. M., Whiteoak D. The Shell Bitumen Handbook, Fifth Edition, Shell Bitumen,
U.K., 2003.

Reese, R. Properties of aged asphalt binder related to asphalt concrete life, Journal of the
Association of Asphalt Paving Technologists, 1997, Volume 66, pp: 604-632.

Richard C.Rise, Brian N.Leis, Drew V.Nelson, Henry D.Berns, Dan Lingenfelser, M.R.
Mitchell, Fatigue Design Handbook, Society of Automotive Engineers, Inc. 1988.

Rowe, G.M., Performance of Asphalt Mixtures in the Trapezoidal Fatigue Test,


Proceedings of the Association of Asphalt Paving Technologists (AAPT), 1993, Volume
62, pp: 344-384.

Rowe, G.M. and Bouldin, M.G. Improved techniques to evaluate fatigue resistance of
asphaltic mixes, Proceedings of the Second Euraphalt and Eurobitume Congress,
Barcelona. Foundation Eurasphalt, Breukelen, The Netherlands, 2000, vol. 1, pp. 754–
763.

Sabha, H., Groenendijk, J. and Molenaar, A.A.A., Estimation of Crack Growth


Parameters and Fatigue Characteristics of Asphalt Mixes Using Simple Tests, Delft
University of Technology, Road and Railroad Research Laboratory, the Netherlands,
1995.

Schapery R.A., A Theory of Crack Growth in Visco-Elastic Media, Report MM 2764-73-


1, Mechanics and Materials Research Centre, Texas A&M University, 1973.

Schapery R.A., A Theory of Crack Initiation and Growth in Visco-Elastic Media; I:


Theoretical Development, II: Approximate methods of analysis, III: Analysis of
Continuous Growth, from: International Journal of Fracture, Sijthoff and Noordhoff
International Publishers, 1975, Vol.11, No.1, pp: 141-159, Vol.11, No.3, pp: 369-388,
and Vol.11, No.4, pp: 549-562.

Schapery R.A., A Method for Predicting Crack Growth in Non-homogeneous Visco-


Elastic Media, from: International Journal of Fracture, Sijthoff and Noordhoff
International Publishers, 1978, Vol.14, No.3, pp: 293-309.

Schmidt, R. J., "A Practical Method for Measuring the Resilient Modulus of Asphalt -
Treated Mixes," Highway Research Record No. 404, Highway Research Board, pp 22-32.

Shell pavement design manual – asphalt pavements and overlays for road traffic, Shell
International Petroleum Company Limited, London, 1978.

Soltani A. Comportement en Fatigue des Enrobes Bitumineux, Doctoral Dissertation,


Ecole Nationale des Travaux Publics de l’Etat-INSA, Lyon, France, 1998.

63
Chapter 2 Literature Review

Symposium on flexible pavement behavior as related to deflection, Proceedings of the


Association of Asphalt Paving Technologists, 1962, Vol. 31, pp: 208-399.

Tangella S. R., Craus J., Deacon J. A. and Monismith C. L., Summary report on fatigue
response of asphalt mixtures. SHRP Report TM-UCB-A-003A-89-3 for project A-003-A,
University of California, Berkeley, 1990.

Tayebali A. A., J. A. Deacon, J.S. Coplantz, F.N. Finn, and C.L. Monismith, Fatigue
Response of Asphalt-Aggregate Mixes, SHRP-A-404, Strategic Highway Research
Program, National Research Council, Washington, D. C., 1994.

Tseng K.H. and Lytton R.L., Fatigue Damage Properties of Asphaltic Concrete
Pavements, Texas Transportation Institute of the Texas A&M University, submitted to
the 69th Annual Meeting of the Transportation Research Board, Washington D.C., 1990.

van Dijk, W. Moreaud, H., et al., The Fatigue of Bitumen and Bituminous Mixes, 3th
International Conference on Structural Design of Asphalt Pavements (ISAP), London,
1972, Volume 1.

van Dijk, W., Practical Fatigue Characterization of Bituminous Mixes, Proceedings The
Association of Asphalt Paving Technologists, 1975.

van Dijk, W. and Visser, W. The Energy Approach to Fatigue for Pavement Design,
Association of Asphalt Paving Technologists (AAPT), 1977, Volume 46, pp: 1-41.

Verstraeten, J., Moduli and Critical Strains in Repeated Bending of Bituminous Mixes,
Application to Pavement Design, Proceedings 3rd International Conference on the
Structural Design of Asphalt Pavements, London, 1972.

Weibull, W., A statistical Distribution of Wide Applicability. J. Appl. Mech., 18, 1951,
pp. 293-296.

Wohler, A., Versuche uber die Festigkeit der Eisenbahnwagenachsen, Zeitschrift fur
Bauwesen,10; English summary (1867), Engineering, 1860, Vol.4, pp.160-161.

64
Chapter 3 Research Methodology

Chapter 3 Research Methodology

3.1 Introduction
The literature review on the fatigue of asphalt concrete presented Chapter 2 reveals that
the assessment of the fatigue characteristics of asphalt mixtures is a very complicated
process and affected by many influence factors. The fatigue characteristics of an asphalt
mixture essentially relate to the material properties and volumetric composition of the
mixture. In mechanistic pavement design procedures, fatigue tests are normally used to
simulate realistic traffic loading and environmental conditions, hence the fatigue
characteristics are estimated in the laboratory in a simplified way. To evaluate the
fatigue life in the field, more influence factors should be taken into account, such as
pavement structure, rest period, lateral wandering, environmental factors, etc [Shell, 1978]
[Groenendijk, 1998] [Al-Qadi, 2003]. Shift factors are therefore applied to laboratory
results to be able to predict the fatigue life in the field [Lytton, 1993] [Molenaar, 1994].
The process of the fatigue estimation in pavement design is schematically described in
Figure 3-1.

Asphalt mixture Laboratory fatigue test Field fatigue damage

Shift factor
Lab Nf Field Nf
Influence factors:

• aggregate properties • test type • pavement thickness


• bitumen properties • specimen size • rest period
• gradation • test conditions • environment
• volumetric properties: • loading mode • lateral wander
density, void content

Figure 3-1 Assessment of fatigue properties of asphalt mixture and the influence factors

From earlier statements (see Chapter 2), it is clear that laboratory fatigue results depend
on the type of fatigue test, specimen size and mode of loading. Until now there is no
unified fatigue test method according to testing specifications implemented by many
countries or organizations [EN 12697-24:2004] [ASTM D7460-10]. Various types of
fatigue tests with their advantages and disadvantages have been used to obtain a

65
Chapter 3 Research Methodology

laboratory fatigue life. With these different fatigue tests, similar conclusions might be
found in qualitative analysis for a series of asphalt materials, but there is a large
difference in quantitative results for the same mixture and this difference is difficult to be
compared with each other. The goal of this research is therefore to determine the
influence of the test type and specimen size on laboratory fatigue test results,
understanding why these differences occur and developing a procedure to make the
results, obtained from different tests, comparable. A major achievement of this study
would be if the results help in harmonizing the current European standard on fatigue
testing.

3.2 Research Methodology


In order to achieve this goal, it was decided to perform uniaxial tension and compression
(UT/C) fatigue tests, four-point bending (4PB) fatigue tests and indirect tensile (IT)
fatigue tests. Furthermore it was decided that for each fatigue test specimen with different
sizes had to be tested to explore the size effect on the fatigue results. It was also decided
to conduct the tests in two different loading modes, namely constant load and constant
displacement. In order to limit the test program, the tests were performed at only two
temperatures and one frequency.

By means of the testing program, the influence of test type, testing mode and specimen
size can be determined but in itself the test results do not give an explanation why
different tests give different test results. Therefore a specific model called the “Partial
Healing model” developed by Pronk will be used in an attempt to explain differences
between the displacement controlled UT/C and 4PB fatigue test [Pronk, 2001] [Pronk,
2012]. As the name of the model implies, it claims to qualify and quantity healing of a
mixture and therefore the PH model will also be used to determine at which stress/strain
level damage development is balanced with loading. At such a strain level no damage
would occur as a result of repeated loading. This strain level, called “endurance limit”, is
a very important parameter in the design of heavy loaded, long life pavements, for which
no structural maintenance is needed.

Fatigue results are normally obtained from tests in which the specimens are subjected to a
uniaxial or at best bi-axial stress condition. Such uni- and bi-axial stress conditions
however never occur in pavement where the stress conditions are always tri-axial. It was
therefore decided to devote part of this research to apply the yield surface concept to the
fatigue results in a similar way as it is done for stone and unbound granular materials.
Extensive research at the Road and Railway Engineering Section of the Delft University
of Technology on unbound materials has shown that permanent deformation will be
limited if the occurring stress condition represented by Mohr’s circles are limited to
around 40% of the stress at failure which can be derived from the yield surface. Since it is
believed that a similar approach is also applicable to characterize the fatigue behavior of
asphalt mixtures, it was decided to perform monotonic compression and tension tests to
derive such yield surface as a function of temperature and strain rate. The failure results
will then be interpreted by comparing the actual stress condition with the yield surface

66
Chapter 3 Research Methodology

and an endurance limit will be developed. This endurance limit is then also valid for the
3D state of stress at which no fatigue will occur.

Based on the selected research approach, the general research layout is shown in Figure
3-2.

Mixture design and


specimen preparation

Monotonic testing

Dynamic testing

• Uniaxial tension/compression test Monotonic uniaxial Monotonic uniaxial


• Four-point bending test tension test compression test
• Indirect tensile test

Unified model

Stiffness test Fatigue test


Yield surface

The PH model: A new fatigue relationship R∆ and fatigue life


• Influence of specimen size • Influence of specimen size
• Influence of temperature • Influence of temperature
• Influence of loading modes

Figure 3-2 Schematic of the research program framework

67
Chapter 3 Research Methodology

References

ASTM D7460-10 Standard Test Method for Determining Fatigue Failure of Compacted
Asphalt Concrete Subjected to Repeated Flexural Bending, 2010.

Al-Qadi, I.L., Nassar, W.N., Fatigue Shift Factors to Predict HMA Performance,
International Journal of Pavement Engineering, Volume 4, Number 2, June 2003 , pp. 69-
76.

EN 12697-24: 2003 Bituminous mixture – Test methods for hot mix asphalt - Part 24:
Resistance to fatigue. European committee for standardization, Brussels, 2003.

Groenendijk, J., Crack Growth in Asphaltic Mixes, Ph.D. dissertation, Delft University of
Technology, 1995.
Lytton, R.L., Uzan, J., Fernando, E.M., Roque, R., Hiltunen, D. and Stoffels, S.M.,
Development and validation of performance prediction models and specifications for
asphalt binders and paving mixes; SHRP Report A-357; SHRP/NRC, Washington DC,
USA, 1993.

Molenaar, A.A.A. Road Material, Part III, Asphalt Materials, lecture notes e52 (CT4850),
Delft University of Technology, the Netherlands, 1994.

Pronk, A.C., Partial healing in fatigue tests on asphalt specimen, Road Materials and
Pavement Design, vol. 4, n. 4, pp 433-445, 2001.

Pronk, A.C., Description of a procedure for using the Modified Partial Healing model
(MPH) in 4PB test in order to determine material parameters. To be published at the 3rd
4PB conference, Davis, USA, 2012.

Shell, Shell pavement design manual: asphalt pavements and overlays for road traffic,
London: Shell International Petroleum, 1978.

68
Chapter 4 Mixture Design and Specimen Preparation

Chapter 4 Mixture Design and Specimen


Preparation

4.1 Selection of the Mixture


The aim of this research was to investigate the effect of the specimen size and test type
on the laboratory fatigue test results of asphalt mixtures. In the study, the test results for
the different specimen sizes and fatigue tests were compared to each other. It was
considered important to minimize the variability of the specimen as much as possible.
This implied that air void content, distribution of particles, etc. had to be kept under strict
control. Therefore it was decided to use a relatively homogeneous and isotropic mixture
with a small maximum aggregate size. From a previous PhD project it was known that
the ACRe mixture, which is a kind of sand asphalt mixture with a maximum grain size of
4 mm, fulfilled these requirements and was specially designed as such [Erkens 2002].

On the other hand, the mixture to be used should be a realistic one implying that it is used
in practice. The ACRe mixture is not used in practice. Dense graded asphalt concrete 0/8
and 0/11 (DAC 0/8 and 0/11) mixtures are commonly used as wearing course and this
mixture type is durable because of its low voids content and relatively high bitumen
content. Furthermore they have a good resistance to deformation, water damage and
ageing. Figure 4-1 represents the comparison of the aggregate gradations of the ACRe
mixture, DAC 0/8 mixture and DAC 0/11 mixture. More details about the composition of
these mixtures are shown in Table 4-1.

100
Percentage passing [%]

80

60

40
ACRe
20 DAC 0/8
DAC 0/11
0
0.01 0.1 1 10 100
sieve size [mm]

Figure 4-1 Grading curves of dense asphalt concrete, DAC [Data from: Erkens, 2002 and
CROW 2005]

69
Chapter 4 Mixture Design and Specimen Preparation

Table 4-1 Compositions of ACRe, DAC 0/8 and DAC 0/11 mixture [Erkens, 2002 and
CROW 2005]
Percentage passing by mass [%]
Sieve size [mm]
ACRe DAC 0/8 DAC 0/11
11 100 100 97
8 100 97 86
5.6 100 80 68.5
4 96.8 - -
2.8 94.4 - -
2 90.8 45 45
1 65.3 - -
0.5 43.4 - -
0.18 25.2 - -
0.063 14.6 8.23 6.5
Ratio of bitumen and aggregate 9.4 6.6-7 -

Furthermore numerous fatigue tests are time consuming and expensive. Due to the high
bitumen content and low air void content, fatigue tests on the ACRe mixture would take
much more time compared to the other two mixtures. After considering the homogeneity,
time needed for testing, the aim to use realistic mixtures and the influence of the
maximum grain size on the specimen size (Figure 4-2), it was decided that the tests
would be performed on a single mixture, being a DAC 0/8 mixture.

Suitable for results comparison

Sand asphalt Dense asphalt concrete 0/8 Dense asphalt concrete 0/11
(SA) (DAC 0/8) (DAC 0/11)

Suitable for application in practice

Figure 4-2 Scheme of the selection of mixture

70
Chapter 4 Mixture Design and Specimen Preparation

4.2 Mixture Design


4.2.1 Materials
The basic components used for the asphalt mixture DAC 0/8 include aggregates, filler
and bitumen. The aggregates consisted of crushed Scottish granite (2~8 mm), Norwegian
bestone (2~6 mm), and crushed sand (≤ 2 mm). The Scottish granite and Norwegian
bestone were obtained from BAM contracting company. To achieve the gradation
according to the Dutch RAW specifications [CROW 2005], the Scottish granite and
crushed sand were separated into two fractions respectively by using a laboratory sieving
machine. For each fraction, the apparent particle density was measured by the saturated
surface dry technique [EN 12697-05, 2003].

m2 − m1
ρa = (4-1)
1000 × V p − ( m3 − m2 ) / ρ w
where: ρa : density of aggregate, [kg/m3];
m1 : mass of pyknometer plus head piece, [g];
m2 : mass of pyknometer plus head piece and aggregate, [g];
m3 : mass of pyknometer plus head piece, aggregate and water,
[g];
Vp : volume of pyknometer, when filled up to the reference mark,
[m3];
ρw : density of water at test temperature, [kg/m3].

The percentage passing a certain sieve size and the apparent density of each fraction are
presented in Table 4-2.

Table 4-2 Percentage passing by mass of each fraction of the aggregates


Norwegian
Scottish crushed granite Crushed sand
Sieve size [mm] Bestone
8-5.6 mm 5.6-4 mm 6-2 mm 2-1 mm 1-0 mm
8 87.6 100.0 100.0 100.0 100.0
5.6 2.0 100.0 96.7 100.0 100.0
4 0.0 51.2 71.9 99.9 100.0
2.8 0.0 30.0 30.0 98.0 100.0
2 0.0 13.9 13.8 74.3 100.0
1 0.0 6.9 2.7 2.0 100.0
0.5 0.0 4.6 1.7 0.0 55.5
0.18 0.0 2.6 1.3 0.0 16.1
0.063 0.0 0.3 0.2 0.0 5.8
Apparent density [kg/m3] 2629.9 2700.7 2702.6 2682.7 2647.8

The used filler was Wigras 40K filler, produced by Ankerpoort NV. Table 4-3 shows the
properties of this filler.

71
Chapter 4 Mixture Design and Specimen Preparation

Table 4-3 Properties of Wigras 40K filler


Bitumen Mass loss Solvability in
Ca(OH)2 [%] Density Void [%]
number at 110 ºC H2O [%] by
by mass [kg/cm3] by volume
ml/100g [%] mass
46 8.9 0.5 9.8 2620 43
Cumulative retained [%] by mass
0.063 mm 0.125 mm 2 mm
18 5 0

Bitumen with a 40/60 pen paving grade was provided by Q8 / Kuwait Petroleum B.V.
The properties of the virgin 40/60 pen bitumen are given below:

Table 4-4 Properties of bitumen 40/60


Penetration @ 25 ºC Softening point Density
Penetration index
[0.1 mm] [ºC] [kg/cm3]
45 53.6 -0.59 1035

4.2.2 Mixture Design

The aim of the DAC mixture design was to obtain a DAC 0/8 that satisfied the mixture
specifications for Vehicle Class 4. In order to optimize the binder content and the
gradation of the DAC 0/8 mixture, a Marshall mixture design was carried out in
accordance with the Dutch RAW specifications [CROW 2005]. Based on the tolerance
limits for the DAC 0/8 mixture, several bitumen contents and aggregate gradations were
selected. The specimens were compacted with the Marshall hammer with 50 blows on
each side. The following properties of the mixture specimen were measured for the
optimization:

- Marshall stability (Pm), Marshall flow (Fm), Marshall quotient (Qm);


- Apparent density and air void content of specimen.

Marshall stability and Marshall flow of the mixture were obtained by conducting the
Marshall test [J.M.M. Molenaar 2003]. The apparent density of the mixture specimen was
measured by the saturated surface dry technique [EN 12697-06, 2003].

m1
ρa = × ρw (4-2)
m3 − m2
where: ρa : apparent density of specimen, [kg/m3];
m1 : mass of dry specimen, [g];
m2 : mass of specimen in water, [g];
m3 : mass of saturated surface-dried speicmen, [g];
ρw : density of water at test temperature, [kg/m3].

The air voids content was calculated from Equation 4-3.


72
Chapter 4 Mixture Design and Specimen Preparation

 ρ 
Va = 1 − a  × 100% (4-3)
 ρ max 

where: Va : air void content of specimen, [%];


ρa : apparent density of specimen, [kg/m3];
ρmax : maximum density of asphalt mixture, [kg/m3];

With the densities of all the fractions in the mixture, the theoretical maximum density
was calculated by Equation 4-4 [EN 12697-05, 2003].

100
ρ max =
 pa1   pa 2   pb  (4-4)
 +  + ... +  
 ρa1   ρ a 2   ρb 
pa1 + pa 2 + ... + pb = 100% (4-5)
ρmax
: maximum density of mixture, [kg/m3];
where:
pa1 : the percentage of aggregate 1 in mixture, [%];
ρa1 : the apparent density of aggregate 1, [kg/m3];
pa2 : the percentage of aggregate 2 in mixture, [%];
ρa2 : the apparent density of aggregate 2, [kg/m3];
pb : the percentage of bitumen in mixture, [%];
ρb : the density of bitumen, [kg/m3];

According to the specification [CROW 2005], the optimal material composition for the
mixture DAC 0/8 was determined, as shown in Table 4-5. Figure 4-3 shows the aggregate
gradation used in the mixture.

100
DAC 0/8 Desired
Min Max
Percentage passing [%]

80

60

40

20

0
0.063 0.18 0.5 1 2 2.8 4 5.6 8
sieve size [mm]

Figure 4-3 Aggregate gradation of the mixture DAC 0/8

73
Chapter 4 Mixture Design and Specimen Preparation

Table 4-5 Composition of the DAC 0/8 mixture


Scottish crushed Norwegian Wigras
Sieve Crushed sand
granite Bestone 40K Binder Total
(mm)
8-5.6 5.6-4 6-2 2-1 1-0 0.063-0
Wt. % 11.2 19.6 21.5 16.8 16.4 7.9 6.5 100

Table 4-6 Properties of the Marshall specimen


Specification
Measured value
Min Max
Marshall stability [N] 9200 7000 -
Marshall flow [mm] 3.2 2 4
Marshall quotient 2875 2500 -
maximum density [kg/m3] 2428 - -
Air void content [%] 3.5 4

4.3 Specimen Preparation


4.3.1 Mixture Compaction
Cylindrical specimens and rectangular beam specimens were fabricated for the uniaxial
tension and compression test, the indirect tensile test and the four-point bending test. All
specimens were cored or sawn from blocks prepared with the PReSBOX compactor. The
rectangular block prepared by the PReSBOX compactor has a fixed length and width of
450 mm × 150 mm. The height can be varied from 145 mm to 185 mm. As one of the
input parameters for the block preparation, the height of the block is determined by the
maximum density of the mixture and the target air void content. During compaction,
asphalt mixtures are compacted with a constant vertical force and a cyclic shear force
with a constant maximum shear angle. It is believed that this way of compaction
simulates the real compaction process in the field [Qiu, Xuan et al. 2009].

Figure 4-4 illustrates the compaction procedure of the hot asphalt mixture; it consists of
the following steps.

(1) Heating: The aggregates and filler were placed in an oven at a temperature of 175ºC
for at least 4 hours. In addition to the materials, the needed accessories for mixing and
compaction (i.e. the mixing bowl, dough hook, trays, steel plates and chute box) were
also kept in the oven at the same temperature and for the same duration. Bitumen was
preheated for 2 hours in an oven at a temperature of 150ºC.

(2) Mixing: Before starting the mixing, the bitumen was poured into the bowl and then
the coarse aggregates were poured in the bowl followed by the crushed sand and after
that the filler. The total weight of the mixture was around 25 kg. All the materials were
mixed using a Hobert H600 mixer for a duration of 3 minutes to make sure that a well
mixed material was obtained.

74
Chapter 4 Mixture Design and Specimen Preparation

Figure 4-4 Illustration of the block production process in the compactor: mixing (top left),
laying (top right), compaction (bottom left) and demoulding (bottom right)

(3) Laying: The hot asphalt mixture was transferred from the bowl into two trays. A steel
plate was placed at the bottom of the compaction mold, and the chute box was placed
afterwards with the bottom gate closed (Figure 4-4 (top right)). When the temperature
of the mixture was around 155ºC, the mixtures were fed carefully into the chute box
(Figure 4-4 (bottom left)). Then the bottom gate of the chute box was opened to allow
the asphalt mixture to fall into the shear box mold. After that the chute box was
removed (Figure 4-4 (bottom right)). The mixture was poked vertically at the four
corners of the mold. Then the other steel plate was placed on the surface. Finally, the
mold was moved into the frame and ready to be compacted.

(4) Compaction: All the required input data, such as the maximum density, mixture
weight, target height etc. were added to the computer, after that the compaction was
started. The compaction process is shown in Figure 4-5 [Qiu, 2012].
.

Start
Vertical stress Vertical + Shear stress Cycle Centre Eject

Figure 4-5 Illustration of the compaction process [Qiu, 2012]

75
Chapter 4 Mixture Design and Specimen Preparation

Firstly, a vertical force was applied to provide a constant vertical stress of 0.75 MPa.
Then cyclic shearing of the framework was applied with a shear angle of 4º. Figure 4-6
shows the development of the vertical and shear stress during the compaction process.
Once the target height was reached (in this study the target height was 155 mm), the
compaction program was terminated automatically. The framework centered itself and
the loads were released. Finally, the asphalt block was ejected by an air pump and
removed after it was cooled down.

2000

1000
Stress [kPa]

-1000

Vertical stress
Shear stress
-2000
0 50 100 150 200
Loading time [s]

Figure 4-6 Loading curves during the compaction process [Qiu, 2012]

4.3.2 Selection of Specimen Size


With regard to the determination of the specimen dimensions for each fatigue test, the
following principles were taken into account.

• From small to large, three specimen sizes were selected, denoted as size 0.5, size 1.0
and size 1.5. The difference between the three specimen sizes had to be large enough
to distinguish the size effect on the fatigue results.
• The test equipment to be used could handle all the specimen sizes.
• The minimum length of the specimen should be at least three times larger than the
maximum aggregate size of the mixture [EN 12697-24].
• To obtain comparable results from the different fatigue tests, the cross-sectional area
in the critical location of specimen size 1.0 should be close to that of the other two
specimen types as explained in Figure 4-7.

Figure 4-7 shows the cross-sectional area in the critical location for the three different
specimen types. The cross-sectional area of the cylindrical specimen for the UT/C test
and the beam specimen for the 4PB can be simply determined. For the IT test, the tensile
stress is induced by the compressive stress along the vertical diametrical plane. Fatigue

76
Chapter 4 Mixture Design and Specimen Preparation

failure is initiated in the region of a relatively uniform tensile stress. According to the
stress distribution of the IT specimen [Hondros, 1959] [Kennedy, 1977], this critical area
is around 70% of the entire cross section of the cylindrical specimen (see Figure 4-7).

UT/C specimen 4PB specimen IT specimen

: Cross-sectional area

Figure 4-7 Cross-sectional area in the critical location for different specimen types

Based on the above considerations, the dimensions for the specimen types are presented
in Table 4-7.

Table 4-7 Dimensions of the different specimen types


Cross area in critical
Test type Specimen size Dimension [mm]
location [mm2]
Size 0.5 ø × h = 25×62.5 491
UT/C Size 1.0 ø × h = 50×125 1964
Size 1.5 ø × h = 75×187.5 4418
Size 0.5 l × w × h = 400×50×25 1250
4PB Size 1.0 l × w × h = 400×50×50 2500
Size 1.5 l × w × h = 400×50×75 3750
Size 1.0 ø × h = 100×30 2100
IT
Size 1.5 ø × h = 150×45 4725
Note: size 0.5 = small size; size 1.0 = medium size; size 1.5 = large size
ø × h = diameter × height, l × w × h = length × width × height

The UT/C specimens with size 1.0 are also used for monotonic tension and compression
tests.

4.3.3 Specimen Preparation


One day after the asphalt mixture block was made, it was cut into cylindrical and beam
specimens with different sizes. Table 4-8 gives the schematic plan for cutting the
different specimens. About 20 mm from the edges of the slab was trimmed-off to avoid
inadequately compacted parts of the block; the thickness of the sawing blade was taken
into account to get the wanted dimensions.

77
Chapter 4 Mixture Design and Specimen Preparation

The rectangular beam specimens were cut by using a water-cooled masonry saw and the
cylindrical specimens were cored by a water-cooled core drill, as shown in Figure 4-8.
After coring, the cylindrical specimens for the UT/C test were polished with a polishing
machine to make the upper and lower surfaces smooth and parallel to each other.

Table 4-8 Sawing plan for the different specimens


Sawing plan Specimen number per
Specimen name
(left: front view, right: top view) slab
Φ=25

450

160

UT/C-size0.5 h=62.5 450 30

150

Φ=50
450

160

UT/C-size1.0 450 12
h=125
150

Φ=74.5
450

160

UT/C-size1.5 2
450
h=186

150

h=25
450

160

4PB-size0.5 400 450


6

150

h=50
450

160

4PB-size1.0 400 450 4

150

78
Chapter 4 Mixture Design and Specimen Preparation

h=75
450

160

4PB-size1.5 400 450


2

150

30

150

R=100
IT-size1.0 170
R=100 450
9

450

45

150

R=150 R=150

IT-size1.5 170 450 4

450

(a) Masonry saw; (b) Core drill


Figure 4-8 Masonry saw and core drill for preparing the different specimens

A numbering system was used to trace from which part of the compacted block the
specimens came. The first letter corresponds to the type of sample (C: cylindrical sample
for the UT/C test, B: beam, I: cylindrical sample for the IT test), the second number is the
number of the slab, and the last number corresponds with the position in the block, which
starts from top to bottom, from left to right and from front to back. Figure 4-9 gives an
example of the specimen labeling for the cylinders with size 1.0. If the block shown in
Figure 4-9 was the first one prepared to obtain cylindrical specimens, then the specimens
were named as C-1-1, C-1-2, ······, C-1-12. Specimen C-1-1 means a cylindrical specimen
for the UT/C test which is cut from the first block prepared for the UT/C tests and taken
from position number 1.

79
Chapter 4 Mixture Design and Specimen Preparation

9 10

Directions in 5 6
1 2 compactor
3 4 1 2

(a) Front view (b) Top view


Figure 4-9 Positions of the labeled cylindrical specimens

Bulk density and voids content of all the specimens were measured by the saturated
surface dry technique and calculated by Equations 4-2 and 4-3. Figure 4-10 compares the
air voids of the specimens from the different positions of a block. It shows that the upper
part of the block has a high density than the lower part. The same results were also found
by Qiu [Qiu, 2012]. Figure 4-11 gives the simulation results of the shear and the vertical
stress distribution of the cross-sectional area of the block during the compaction. It can be
seen that the material is compacted by both shear and compressive force. The upper part
experienced more shear stress than the lower part, which will have consequences on the
distribution of the air voids of the specimens taken from different positions. But the
difference of the air voids content at the different positions was lower than 1 % in the
same slab.

lower part
4
upper part
Air voids content [%]

0
C-8-1 C-8-2 C-8-3 C-8-4 C-8-5 C-8-6 C-8-7 C-8-8 C-8-9 C-8-10 C-8-11 C-8-12

Figure 4-10 Air voids content of the cylinders with size 1.0 in one slab

Table 4-9 gives an overview of the measured void contents of totally 190 specimens. The
average value of the voids content is very close to the target voids content of the block,
3.5%. The standard deviations show that the variability of the voids content of the
specimens is low. Some irregular specimens with extreme air voids of 2.2 % or 4.88 %
can also be observed. This might be related to abnormal compacting conditions such as
too low or too high compaction temperature, material loss during the laying process and
so on.

80
Chapter 4 Mixture Design and Specimen Preparation

(a) (b)
Figure 4-11 Shear stress (a) and vertical stress (b) distribution in asphalt mixtures at the
maximum shear angle during compaction [Qiu, 2012]

Table 4-9 Statistical analysis of volumetric properties of specimens


Max. air Min. air Standard
Numbers Target air Average air
voids voids deviation COV [%]
of beams voids [%] voids [%]
[%] [%] [%]
190 3.5 3.45 4.88 2.20 0.538 15.6

Figure 4-12 gives the distribution of the air voids content of all 190 specimens. The voids
content of the specimens should be between 2.75 % and 4.25 % in order to minimize the
influence of the volumetric properties on the fatigue test results.

5
Air voids content [%]

4.25%
4

3
2.75%

1
0 50 100 150 200
Specimen number
Figure 4-12 Selection of the specimens based on the air voids contents

Before testing, all the selected samples were stored in a climate room at a temperature of
15 ºC. Figure 4-13 shows the beam and cylindrical specimens with different sizes.

81
Chapter 4 Mixture Design and Specimen Preparation

(a) Cylinders for UT/C test; (b) Beams for 4PB test; (c) Cylinders for IT test
Figure 4-13 Beam and cylindrical specimens with different sizes

References

CROW. In: Standaard RAW Bepalingen, C.R.O.W., Ede (In Dutch). 2005.

EN 12697-05: 2003 Bituminous mixture – Test methods for hot mix asphalt - Part 5:
Determination of the maximum density. European committee for standardization,
Brussels, 2003.

EN 12697-06: 2003 Bituminous mixture – Test methods for hot mix asphalt - Part 6:
Determination of bulk density of bituminous specimens. European committee for
standardization, Brussels, 2003.

Erkens, S.M.J.G., Asphalt Concrete Response (ACRe)-Determination, Modelling and


Prediction, PhD Thesis, Delft University of Technology, the Netherlands, 2002.

Jacobs, M. M. J., Crack Growth in Asphaltic Mixes, Ph.D. dissertation, Delft University
of Technology, Delft, the Netherlands, 1995.

Molenaar, J.M.M., Performance related characterisation of the mechanical behaviour of


asphalt mixtures, PhD Thesis, Delft University of Technology, the Netherlands. 2003.

Qiu, J., Xuan D., van de Ven, M.F.C. and Molenaar, A.A.A.,, Evaluation of the shear box
compactor as an alternative compactor for asphalt mixture beam specimens. AES -
ATEMA'2009, 3rd Int. Conf. on Advances and Trends in Engineering Materials and their
Applications. Montreal, Canada, 2009.

Qiu, J., Li, N., Pramesti, F.P., van de Ven, M.F.C. and Molenaar, A.A.A.et al., Evaluating
Laboratory Compaction of Asphalt Mixtures Using the Shear Box Compactor, Journal of
Testing and Evaluation, Vol. 40, No. 5, 2012.

82
Chapter 5 Different Laboratory Fatigue Experiments

Chapter 5 Different Laboratory Fatigue


Experiments

5.1 Introduction
In this chapter, the uniaxial tension and compression (UT/C), four-point bending (4PB)
and indirect tensile (IT) fatigue tests, which were performed at different temperatures and
loading modes, will be described. Details of the experimental work are presented in this
chapter. The Chapter begins with a description of the used test equipments and the
calibration work for the 4PB test setup. Then the stiffness and fatigue test programs are
given for the different specimen sizes. Finally the test results from all three test methods
are compared and discussed.

5.2 Test Equipment


5.2.1 Uniaxial Tension and Compression (UT/C) Test
The uniaxial tension and compression test was performed on a IPC Universal Testing
Machine (UTM 25 kN) electro-hydraulic servo system. This system consists of three
main components: the testing frame, the environmental chamber, and the control data
acquisition system (CDAS); all are shown in Figure 5-1. The temperature range of the
environmental chamber is between -15°C and 60°C. All specimens were glued to two
steel plates at both ends using a two-component fast curing adhesive glue (X60). Firstly,
the specimen was glued to the top platen that is fastened to the actuator. After sufficient
glue was placed on the bottom platen, the top platen with the specimen was slowly
lowered till the bottom end of the specimen contact with the bottom platen. Due to the
stress concentration near the ends of the specimen, enough glue needs to be used to
ensure sufficient bond and prevent cracking near the ends. The thickness of the glue was
around 3 mm.

During the test, the load is applied to the top platen through the actuator. The force is
measured with a load cell and the axial deformation is registered by means of three
spring-loaded LVDT’s with a range of ±2 mm The LVDT’s are mounted vertically
around the side of the specimen at an interval of 120° between two aluminum rings
fastened on the top and bottom platens. In terms of data acquisition, the CDAS
automatically controls the valve to apply the requested loading during the test. The
CDAS with the personal computer collects the data at the same time.

83
Chapter 5 Different Laboratory Fatigue Experiments

Testing frame

Aluminum ring

Environmental
chamber
Adhesive LVDT’s

Figure 5-1 Uniaxial tension and compression test setup with the temperature chamber
(left) and a close-up of a cylindrical specimen for size 0.5 (right)

5.2.2 Four-Point Bending (4PB) Test


To allow the use of beam specimens with different dimensions, a four-point bending
(4PB) test setup was developed by the Laboratory of the Road and the Railway
Engineering Section; it is shown in Figure 5-2a. Like the UT/C test, the test setup was
placed in a temperature chamber capable of maintaining a temperature between -10°C to
60°C. Four clamps fix the beam specimen in the bending bed. The distance between the
two outer is 400 mm and the distance between the inner clamps is 133.3 mm. The load
was applied to the beam specimen through the inner clamps by means of two loading
jacks. The two outer clamps prevent vertical movement of the specimen. They consist of
two thin steel sheets around the beam with a groove. A small cylindrical spindle lies in
the two grooves between clamp and the frame of which the radius is larger than that of
the spindle. The clamp allows free rotation and horizontal translation of the specimen at
the four supports, as shown in Figure 5-2b.

LVDT

Bolts

Spring
Spindle

(a) (b)
Figure 5-2 Set-up of the 4PB fatigue test (a) and local view of inner clamps (b)

84
Chapter 5 Different Laboratory Fatigue Experiments

The clamping force is applied by tightening the bolts with the spring on top of the clamps.
A torque wrench was used to apply a specific clamping force, see Figure 5-3a. Figure 5-
3b shows the measurement of the clamping force. The force measured by the load cell
placed between the clamps was recorded at the applied torque level.

Load cell

(a) Torque wrench (b) Measurement of the clamping force


Figure 5-3 Measurement of the clamping force

Figure 5-4 gives the relations between the clamping force and the applied torque at the
inner and outer clamps. Two different linear relations were found at low torque 0≤ τ ≤2.
lb-in and at high torque 2.5 lb-in ≤ τ ≤ 7.5 lb-in (The torque unit “lb-in” is the product of
pound and inch, 1 lb-in=0.1129 Nm). The functions representing the relation between
torque and clamping force at the inner and outer clamps are listed in Table 5-1.

800
Inner clamping force
Outer clamping force_right
600 Outer clamping force_left
Force [N]

400 2

1
200

0
0 2 4 6 8 10
Torque [LB-IN]

Figure 5-4 Relationship between clamping force and torque

85
Chapter 5 Different Laboratory Fatigue Experiments

Table 5-1 Linear equations of the clamping force


Linear equation 1 Linear equation 2
(0≤ τ ≤2.5LB-IN) (2.5≤ τ ≤7.5LB-IN)
Inner clamping force Fin (N) Fin = 48.5 ⋅ τ + 90.2 Fin = 82.4 ⋅ τ + 24.3
Outer clamping force_right For (N) For = 28 ⋅ τ + 49.2 For = 62.8 ⋅ τ − 34.5
Outer clamping force_left Fol (N) Fol = 21 ⋅ τ + 40.4 Fol = 58 ⋅ τ − 48.5

The clamping force has a significant influence on the stresses and strains in the beam
specimen. In theory the clamping force should be as low as possible to prevent damage
near the supports. However, for the test setup shown in Figure 5-2, the sum of the two
inner clamping forces should be a little larger than the applied force. If the force supplied
by the springs is smaller than the applied force from the actuator, the springs will be
deformed and the specimen will be separated from the clamps during the test. The
disadvantage of this setup is that the amplitude of the clamping force varies with the
applied strain level.

At high strain level, implying high load levels and high clamping forces restrain the
horizontal movement and rotation of the spindles between grooves and cause high shear
stresses on the beam near the supports. In order to determine the stress and strain
distributions in the beam specimen, finite element simulations were performed by means
of the ABAQUS software. The model geometry of the 4PB specimen is presented in
Figure 5-5.

Loading Loading
Clamp

Beam specimen

20 mm
133.3 mm

400 mm

450 mm

Figure 5-5 Model geometry used in the finite element simulations of the 4PB test

Figure 5-6 shows the horizontal strain and shear stress distribution in the beam during the
test. The clamping force on the model is 300N. At the same time, a force of 250 N is
applied on top of each inner clamp. The viscoelastic properties of the mixture are
described by Prony series. The value of the Prony elements were calculated from the
results of the 4PB stiffness tests. Details of this analysis are presented in Appendix B. For
the Poisson’s ratio of the mixture, a value of 0.3 was used. The steel clamps on the
supports are considered to be linear elastic material with a modulus of 200 GPa and a
Poisson’s ratio of 0.27.

86
Chapter 5 Different Laboratory Fatigue Experiments

As shown in Figure 5-6, the computational results show that the highest horizontal tensile
strain and shear stress occur around the inner clamps. After observing the location of
failure in numerous specimens, it was found that most of the specimens failed at the
location where the highest stress was computed (see Figure 5-6c). In this case, the cracks
on the beam are mainly caused by the combination of horizontal tensile strain and shear
stress and not because of pure bending.

(a) Horizontal strain ε11

(b) Shear stress s12 (c) Cracking near the inner clamps

Figure 5-6 Horizontal strain and shear stress distributions in the beam

To avoid the unwanted horizontal strain and shear stress concentrations near the inner
clamps, the springs on top of the inner clamps were removed from the test setup, as
shown in Figure 5-7. After the modification, a certain small clamping force is allowed to
be applied for different strain levels. Under a small clamping force, the cylindrical
spindle can move and rotate freely between the two grooves.

(a) (b)
Figure 5-7 Modified setup of the 4PB fatigue test (a) and local view of inner clamps (b)

87
Chapter 5 Different Laboratory Fatigue Experiments

Also for this setup, finite element simulations were performed with the ABAQUS
software. The same model geometry and loading mode were used. The difference is that
the inner clamping forces reduced to 120 N and the clamps rotated with the bended
specimen surface. Figure 5-8 shows the horizontal strain and shear stress distribution of
the beam at the lower clamping force of 120 N. It can be seen that the maximum
horizontal compressive and tensile strains distribute uniformly at the specimen surface
between two inner clamps. This implies that in most of the cases, the fatigue cracks will
initiate at the surface in the middle section of the beam specimen.

(a) Horizontal strain ε11

(b) Shear stress s12


Figure 5-8 Horizontal strain and shear stress distribution of the beam

clamping force=120N clamping force=300N


600
Horizontal strain [µm/m]

500
Inner clamps
400

300

200

100

0
0 100 200 300 400 500
Length of beam [mm]

Figure 5-9 Computed horizontal strain at beam surface along the length

88
Chapter 5 Different Laboratory Fatigue Experiments

clamping force=120N clamping force=300N


1
0.8
0.6 Inner clamps

Shear stress [MPa]


0.4
0.2
0
-0.2 0 100 200 300 400 500
-0.4
-0.6
-0.8
-1
Length of beam [mm]

Figure 5-10 Computed shear stress at beam surface along the length

clamping force=120N clamping force=300N


0.2

0.15 Inner clamps


0.1
Shear stress [MPa]

0.05

-0.05 0 100 200 300 400 500

-0.1

-0.15

-0.2
Length of beam [mm]

Figure 5-11 Computed shear stress in the middle of the beam along the length

Figure 5-9, 5-10 and 5-11 compare the distributions of the horizontal strain at the surface,
shear stress at the surface and the shear stress in the middle of the beam. When the
applied force is 250 N and the clamping force is 120 N, the horizontal strain increases
gradually from the outer supports to the inner supports, in the middle section the
horizontal strain takes a constant value (≈160µm/m) and the maximum shear stress at
two inner supports is around 0.29 MPa. At the high clamping force of 300 N, the
maximum horizontal strain at the inner supports is 500µm/m, while in the middle section
of the specimen the strain is almost zero. The maximum shear stress at the inner supports
is 0.6 MPa, about 2 times higher than that at the low clamping force. Therefore based on
the model analysis, it can be concluded that the fatigue failure of the beam tested in the
modified 4PB setup is mainly caused by pure bending when the two clamping forces are
as small as possible.

89
Chapter 5 Different Laboratory Fatigue Experiments

5.2.3 Indirect Tensile (IT) Fatigue Test


Like the UT/C test, the indirect tensile (IT) test is also performed in the IPC Universal
Testing Machine (UTM 25 kN). Figure 5-12 shows the loading configuration of the IT
test.
F
Loading strip

Loading strip
Vertical
LVDT

Horizontal
LVDT

Figure 5-12 The set-up of the IT test and loading configuration

The cylindrical sample is placed between two arc-shaped steel loading strips. During the
test, a repeated compressive load is applied by means of the load actuator along the
vertical diametrical plane. This load causes both vertical compressive and horizontal
tensile stresses in the specimens. Two LVDT's (range ±2 mm) at either side of the
specimen are mounted on the frame to measure the horizontal displacement along the
horizontal diameter. The vertical deformations are measured by two LVDT's with a range
of ±10 mm glued on the upper loading strip. For specimens with different sizes, different
loading strips have been used. The dimensions of specimens and the loading strips are
shown in table 5-2.

Table 5-2 Dimensions of the loading strips for IT [EN 12697-26]


Nominal specimen Width of
Nominal depth Radius of the load side
diameter loading strip
concave segment [mm] of the strip [mm]
[mm] [mm]
100 12 ± 1 0.4 ± 0.05 50.5
150 19 ± 1 0.6 ± 0.05 76

5.3 Calibration of the 4PB Test Equipment

In the UT/C and IT test, the loading is applied directly to the specimen. So the measured
force and displacement are not influenced by the test equipment. However, the 4PB test
setup used in this study was specially designed by the Road and Railway Engineering
Laboratory and differed from existing commercial test equipment. Furthermore the
loading is not directly applied on the beam specimen but through a loading frame. The

90
Chapter 5 Different Laboratory Fatigue Experiments

measured displacement strongly depends on the properties of the loading frame. Three
aspects should be taken into account: pure bending deflection of the beam, shear
deflection of the beam at the supports and deflection of the frame. However, only the
deflection due to pure bending is taken into account in the stiffness calculations. To
obtain reliable measurement results, it is necessary to perform calibration tests on the
4PB test setup.

5.3.1 Theory
In the whole test setup, the beam and frame from the test setup can be represented by two
serial springs, as shown in Figure 5-13.

vm

vbeam vframe
F F
Sbeam Sframe

Figure 5-13 Theoretical model for the deflection of 4PB setup consisting of beam and
loading frame

Based on this two serial spring’s model, the following relations can be obtained:

F F F
vm = , v frame = , vbeam = (5-1)
Sm S frame Sbeam
1 1 1
vm = v frame + vbeam ⇒ = + (5-2)
Sm S frame Sbeam

Where: vm : measured deflection, [m];


vframe : frame deflection, [m];
vbeam : beam deflection between inner clamps, [m];
F : applied force, [N];
Sm : stiffness of whole system, [N/m];
Sframe : frame stiffness, [N/m];
Sbeam : beam stiffness, [N/m].

(1) Correction for shear deflection


The beam deflection includes two parts, bending deflection and shear deflection
[Huurman, 2009] [Pronk, 2009]. In the midsection between the two inner clamps, the
solution for the bending deflection is given by Equation 5-4 [Pronk, 2009].

vbeam = vs + vb ,vs = C s ⋅ vb (5-3)

91
Chapter 5 Different Laboratory Fatigue Experiments

2
h
4 ⋅ (1 + µ ) ⋅  
v L
⇒ vb = beam ,Cs = (5-4)
1 + Cs   A 
2

3− 4  ⋅α
  L  

Where: vs : shear deflection, [m];


vb : bending deflection between inner clamps, [m];
Cs : correction coefficient;
Μ : Poisson’s ratio;
H : beam height, [m];
L : distance between two outer clamps, [m];
A : distance between inner and outer clamps, [m];
Α : shear factor, 0.859.

(2) Correction for frame deflection


The beam stiffness in [MPa] is related to the applied force and the bending deflection, as
given in Equation 5-5.

F (5-5)
E=Z⋅
vb
6 AL2 − 8 A3
Z= (5-6)
8bh3

Where: E : theoretical value of the beam stiffness, [MPa];


Z : form factor, [1/m].

The beam stiffness in [N/m] is transformed from the beam stiffness in [MPa] by
substituting Equation 5-3 and 5-5 into Equation 5-1.

F F E
Sbeam = = = (5-7)
vbeam (1 + Cs ) ⋅ vb Z ⋅ (1 + Cs )

Therefore the frame stiffness is calculated by the following equation:

1 1
S frame = =
1 1 F Z ⋅ (1 + Cs ) (5-8)
− −
S m Sbeam vm E

By substituting Equation 5-8 into Equation 5-2, the beam deflection can be calculated
from the measured force and deflection if the frame stiffness is known.

92
Chapter 5 Different Laboratory Fatigue Experiments

F
vbeam = vm − v frame = vm − (5-9)
S frame

Finally, the pure bending deflection can be computed by Equation 5-4.

5.3.2 Test Program


In order to calculate the frame stiffness of the used 4PB test equipment, a frequency
sweep was conducted with an aluminum beam (theoretical stiffness E = 71 GPa). The
dimension of the aluminum beam is 450 mm long, 35 mm wide and 34 mm high (see
Figure 5-14). During the test, a sinusoidal loading was applied on the beam at different
force levels and frequencies (1, 2, 4, 6, 8 and 10 Hz).The inner clamping force was 120 N
and kept constant for all force levels. The outer clamping force depended on the applied
force level. Table 5-3 presents the applied force level for each outer clamping force.

Figure 5-14 Calibration test for the 4PB setup with aluminum beam

Table 5-3 Applied force levels and used clamping forces


Inner clamping force [N] Outer clamping force [N] Applied force level [N]
100
120
200
100
180 200
300
120 200
230 300
400
300
270 400
500

93
Chapter 5 Different Laboratory Fatigue Experiments

Figure 5-15 shows the measured stiffness of the aluminum beam at various applied force
and clamping force levels. It can be seen that the measured stiffness of the aluminum
beam is independent of the frequency but is strongly influenced by the applied force and
clamping force. At a same outer clamping force, a higher applied force leads to a smaller
measured stiffness, while at a certain applied force the measured stiffness increases with
an increasing clamping force. For each clamping force, the maximum applied force gives
in all cases a stiffness of approximately, about 56.4 GPa. This implies that the outer
clamping forces should be as low as possible relative to the applied force.

1Hz-120N 2Hz-120N 4Hz-120N 6Hz-120N 8Hz-120N 10Hz-120N


1Hz-180N 2Hz-180N 4Hz-180N 6Hz-180N 8Hz-180N 10Hz-180N
1Hz-230N 2Hz-230N 4Hz-230N 6Hz-230N 8Hz-230N 10Hz-230N
1Hz-270N 2Hz-270N 4Hz-270N 6Hz-270N 8Hz-270N 10Hz-270N
6.2E+04
Measured stiffness [MPa]

6.0E+04

5.8E+04

5.6E+04
Outer clamping force: 120N 180N 230N 270N
5.4E+04

5.2E+04

5.0E+04
0 100 200 300 400 500 600
Applied force [N]

Figure 5-15 Measured stiffness of aluminum beam at various applied force and outer
clamping force levels

1Hz-120N 2Hz-120N 4Hz-120N 6Hz-120N 8Hz-120N 10Hz-120N


1Hz-180N 2Hz-180N 4Hz-180N 6Hz-180N 8Hz-180N 10Hz-180N
5.5E+07 1Hz-230N 2Hz-230N 4Hz-230N 6Hz-230N 8Hz-230N 10Hz-230N
Calculated frame stiffness [N/m]

1Hz-270N 2Hz-270N 4Hz-270N 6Hz-270N 8Hz-270N 10Hz-270N


5.0E+07

4.5E+07

4.0E+07

3.5E+07

3.0E+07

2.5E+07

2.0E+07
0 100 200 300 400 500 600
Applied force [N]

Figure 5-16 Calculated frame stiffness values at various applied force and outer clamping
force levels

94
Chapter 5 Different Laboratory Fatigue Experiments

As mentioned in the last section, the frame stiffness can be calculated using Equation 5-8.
The calculated Sframe values are presented in Figure 5-16; they show a similar trend with
respect to the applied force as the measured stiffness. The figure shows that the frame
stiffness at the maximum applied force for each clamping force is more or less same.

Figure 5-17 gives the measured phase angle at the different applied force levels and
clamping forces. Like the stiffness, the measured phase angle does not change with
frequency. The phase angle however does increase with increasing applied force and
decreases with increasing outer clamping force. For each clamping force, the phase angle
measured at the maximum applied force has a linear relationship with the applied force
level. Table 5-4 and Figure 5-18 give the average stiffness and phase angle values at the
maximum applied force for each outer clamping force. Compared to the stiffness of the
aluminum beam 7.47×106, the average stiffness of the 4PB set up is about 4 times higher.

1Hz-120N 2Hz-120N 4Hz-120N 6Hz-120N 8Hz-120N 10Hz-120N


3 1Hz-180N 2Hz-180N 4Hz-180N 6Hz-180N 8Hz-180N 10Hz-180N
1Hz-230N 2Hz-230N 4Hz-230N 6Hz-230N 8Hz-230N 10Hz-230N
1Hz-270N 2Hz-270N 4Hz-270N 6Hz-270N 8Hz-270N 10Hz-270N
2.5
Phase angle [deg]

1.5

0.5

0
0 100 200 300 400 500 600
Applied force [N]

Figure 5-17 Calculated phase angle at various applied force and outer clamping force

Table 5-4 Average value of frame stiffness and phase angle


Inner clamping Outer clamping Applied force Calculated Sframe Phase angle
force [N] force [N] level [N] [N/m] [deg]
120 200 3.14E+07 2.0
180 300 3.18E+07 1.5
150
230 400 3.08E+07 1.6
270 500 3.05E+07 1.3

95
Chapter 5 Different Laboratory Fatigue Experiments

4.5E+07 3.0
y = -0.0021x + 2.3374 Sframe

Frame stiffness [N/m]


R2 = 0.8073 2.5
Phase angle

Phase angle [deg]


4.0E+07
2.0

3.5E+07 1.5

1.0
3.0E+07 y = -3733.6x + 3E+07
0.5
R2 = 0.697
2.5E+07 0.0
0 100 200 300 400 500 600
Applied force [N]

Figure 5-18 Average stiffness and phase angle at the maximum applied force for each
outer clamping force

In the 4PB tests the following regression equations were used to correct the measured
stiffness and phase angle.

Frame stiffness: S frame = −3733.6 ⋅ F0 + ( 3.24 E + 07 ) (5-10)


System phase angle: θ frame = −0.0021 ⋅ F0 + 2.34 (5-11)

where F0 = applied force, N.

5.4 Complex Modulus and Fatigue Tests


In this study complex modulus tests and fatigue tests are performed. These experiments
are conducted according to the test procedures described below.

5.4.1 Complex Modulus Test


The complex modulus tests are performed at different temperatures and frequencies. The
test programs for the different test equipments are listed in Table 5-5. Due to the capacity
of the different test equipments, the used test conditions for UT/C, 4PB and IT tests are
slightly different. The testing order always was from low to high temperatures and from
high to low frequencies; this is done in order to minimize the damage to the specimens. A
five-minute rest period is allowed between each two adjacent different frequencies and at
least two and a half hours acclimatization was allowed after changing the testing
temperature to achieve thermal equilibrium. All tests are performed according to
EN12697-26. The loading levels were carefully adjusted until the sample deformations
reach the target value which was considered low enough not to cause damage in the
specimen. This implies that the resulting complex modulus represents the modulus of the
intact, not damaged specimen. The complex modulus was determined from the average

96
Chapter 5 Different Laboratory Fatigue Experiments

values of load and displacement obtained during the final five cycles of each loading
series.

Table 5-5 Complex modulus test programs on different test equipments


Test type UT/C 4PB IT
Control mode Strain controlled Strain controlled Displacement control
Horizontal displacement
Strain level [µm/m] 40-70 40-70
=3-5µm
Waveform Sine wave Sine wave Haversine wave
Frequency [Hz] 0.5, 1, 2, 4, 6, 8, 10 1, 2, 4, 6, 8, 10 0.5, 1, 2, 4, 6, 8,
0, 5, 10, 15, 20, 0, 5, 10, 15, 20, 25, 30,
Temperature [°C] 0, 5, 10, 20, 30, 40
25, 30, 35, 40 35, 40

5.4.2 Fatigue Test


The test conditions for the fatigue tests are presented in Table 5-6. Eight strain or stress
levels are chosen in each test series. The specimens are conditioned at the target testing
temperature for a minimum of 3 hours before conducting the test. The load, vertical and
horizontal deformations are monitored and recorded continuously during the test.

Table 5-6 Test conditions for different fatigue test methods


Test type Loading mode Specimen size Test condition
size 0.5 Temp.: 20 ºC
strain controlled size 1.0 Freq.: 10 Hz
size 1.5 Waveform: sinusoidal
UT/C
Temp.: 5 ºC
stress and strain
size 1.0 Freq.: 10 Hz
controlled
Waveform: sinusoidal
size 0.5 Temp.: 20 ºC
strain controlled size 1.0 Freq.: 10 Hz
size 1.5 Waveform: sinusoidal
4PB
Temp.: 5 ºC
stress and strain
size 1.0 Freq.: 10 Hz
controlled
Waveform: sinusoidal
size 1.0 Temp.: 5 ºC
IT stress-controlled Freq.: 10 Hz
size 1.5
Waveform: haversine

For each set of experiments, 8 strain or stress levels are selected to obtain the relationship
between loading level and fatigue life. The loading levels were selected in such a way
that the fatigue life would cover a wide range from 5×103 to 2×107 cycles.

97
Chapter 5 Different Laboratory Fatigue Experiments

5.5 Data Processing


5.5.1 UT/C Test
For both the stiffness and fatigue tests, the complex stiffness and phase angle are the
important test parameters to characterize the material properties. However, there was no
test program for the UT/C test data in the software package from the UTM-25 testing
machine. In this study, all the UT/C test results were computed from the raw data (time,
force and displacement) recorded during the test. The force and displacement in each
cycle were described as a sinusoidal function of time t and loading frequency, f, as
expressed by Equation 5-12 and 5-13.

F = F0 sin(2π f ⋅ t + ϕ1 ) + a (5-12)
L = L0 sin(2π f ⋅ t + ϕ2 ) + b (5-13)

Where:
F : applied force, [N];
F0 : amplitude of the applied force in a cycle, [N];
f : loading frequency, [Hz];
t : testing time[s];
L0 : amplitude of the displacement in a cycle, [mm];
a, b, φ1, φ2 : regression coefficients.

Based on the measured results, the regression coefficients in the equations can be simply
obtained by the solver function in Excel. Figure 5-19 shows the measured force and the
average displacement of the three LVDT’s in a loading cycle. After regression, the stress
σ and strain ε were calculated according to Equation 5-14 and 5-15. The phase angle was
converted into degrees by using Equation 5-16.

4F0
σ0 = (5-14)
π D2
L0
ε0 = (5-15)
H

θ=
(ϕ2 − ϕ1 ) ⋅180 (5-16)
π

Where: σ0 : stress amplitude, [MPa];


D : diameter of specimen, [mm];
ε0 : strain amplitude, [mm/mm];
H : height of specimen, [mm];
θ : phase angle in degree, [º].

98
Chapter 5 Different Laboratory Fatigue Experiments

Prediction
1600 Measured force 0.015
θ Prediction
1200 Measured displacem ent
0.01
800 F0

Displacment [mm]
0.005
400 L0
Force [N]

0 0
0 0.02 0.04 0.06 0.08 0.1 0.12
-400
-0.005
-800
-0.01
-1200

-1600 -0.015
Time [s]

Figure 5-19 Force and displacement signal in a cycle

The absolute value of the complex modulus was then calculated as the ratio of the stress
to strain amplitudes.

σ0
E = (5-17)
ε0
σ
E ' = 0 cos (θ ) (5-18)
ε0
σ
E '' = 0 sin (θ ) (5-19)
ε0

Where: E : absolute value of complex stiffness, [MPa];


E' : storage modulus, [MPa];
E '' : loss modulus, [MPa].

5.5.2 4PB Test


The 4PB software was designed and developed in the National Instruments LabView
environment to collect and calculate the 4PB test results. The sample frequency is 1000
Hz. Therefore the interval between two data points is 1 ms. In this study the highest test
frequency is 10 Hz which means that at least 100 data points are captured per cycle
during the test. To minimize noise in the signal, Fast Fourier Transform (FFT) is applied
to determine the amplitudes and phase difference between force and deflection. A typical
frequency spectrum is displayed in Figure 5-20. Taking into account the specimen shape
and the moving mass, the complex stiffness and phase angle are expressed by Equation 5-
20 and 5-21. The detailed description of the derivation of these formulas is described in
Appendix A.

99
Chapter 5 Different Laboratory Fatigue Experiments

Figure 5-20 Fast Fourier Transform of the measured force and deflection at 10 Hz.

2
ZF vb v 
E = 1 + 2 cos (ϕ s ) ⋅ ⋅ meqω +  b ⋅ meqω 2 
2
(5-20)
vb F F 
 
 sin (ϕs ) 
ϕmix = arctan   (5-21)
 cos (ϕs ) + vb ⋅ meqω 2 
 F 
6 AL2 − 8 A 3
Z= (5-22)
8bh3

Where: E : absolute value of complex stiffness, [MPa];


Z : shape factor, [m-1];
vb : bending deformation, [m];
F : amplitude of force [N];
φs : measured phase lag [º];
meq : equivalent mass [kg] ;
ω : circular frequency [rad/s]
φmix : phase angle of specimen [º];
A : distance between clamps, next to each other [m];
L : effective length=distance between outer clamps [m];
b : Width of specimen[m]
h : height of specimen [m]

100
Chapter 5 Different Laboratory Fatigue Experiments

5.5.3 IT Test
During the test, the control and data acquisition system (CDAS) automatically controls
the applied loading and records all the signals, such as vertical load, horizontal and
vertical deformation. With the specimen dimension and an estimated value of the
Poisson’s ratio for the material, the stiffness modulus can be calculated by the following
equations [Hondros, 1959].

2F0
σ0 = (5-23)
π ×h× D
 2 ⋅ ∆H r   1+ 3⋅ v 
ε0 =  ×  (5-24)
 D   4 +π ⋅v −π 
σ0
E = (5-25)
ε0

Where: σ0 : tensile stress at specimen center, [MPa];


F0 : peak value of applied vertical force [N];
h : mean thickness of specimen [mm];
D : specimen diameter [mm];
ε0 : tensile strain at the specimen center [µm/m];
v : Poisson’s ratio;
∆ Hr : recovered horizontal deformation, [mm];
E : absolute value of complex stiffness, [MPa];

Figure 5-21 shows the total horizontal deformation determined from the horizontal
displacement transducers. Table 5-7 gives the assumed values for Poisson’s ratio of
asphalt mixtures at several temperatures.

12 0.04
Force
Horizontal LVDT
10
Deformation [mm]

8 0.035
Force [kN]

4 ∆ Hr 0.03
F0
2

0 0.025
100.2 100.25 100.3 100.35 100.4 100.45
time [s]

Figure 5-21 Determination of the horizontal deformation

101
Chapter 5 Different Laboratory Fatigue Experiments

Table 5-7 Estimated Poisson’s ratio at different temperatures

Temperature [ºC] 5 10 15 23 35

Poisson’s ratio 0.22 0.24 0.27 0.32 0.4

5.6 Test Results


5.6.1 Stiffness Results
Figure 5-22 and 5-23 give typical results from the frequency sweep tests performed at
various temperatures for the UT/C test using specimen size 1.5.

UT/C_size1.5
30000
0ºC 10ºC 20ºC
30ºC 40ºC
Complex modulus [MPa]

25000

20000

15000

10000

5000

0
0 2 4 6 8 10 12
Frequency [Hz]

Figure 5-22 Complex modulus at various temperatures and frequencies

UT/C_size1.5
100
0ºC 10ºC 20ºC
80 30ºC 40ºC
Phase angle [deg]

60

40

20

0
0 2 4 6 8 10 12
Frequency [Hz]

Figure 5-23 Phase angle at various temperatures and frequencies

102
Chapter 5 Different Laboratory Fatigue Experiments

@10ºC
20000

Complex modulus [MPa]


16000

12000

8000

UT/C_Size 0.5 UT/C_Size 1.0 UT/C_Size 1.5


4000 4PB_Size 1.0 4PB_Size 1.5 IT_Size 1.0
IT_Size 1.5
0
0 2 4 6 8 10 12
Frequency [Hz]

(a)
@20ºC
12000

10000
Complex modulus [MPa]

8000

6000

4000
UT/C_Size 0.5 UT/C_Size 1.0 UT/C_Size 1.5
2000 4PB_Size 1.0 4PB_Size 1.5 IT_Size 1.0
IT_Size 1.5
0
0 2 4 6 8 10 12
Frequency [Hz]

(b)
@30ºC
6000
UT/C_Size 0.5 UT/C_Size 1.0 UT/C_Size 1.5
5000 4PB_Size 1.0 4PB_Size 1.5 IT_Size 1.0
Complex modulus [MPa]

IT_Size 1.5
4000

3000

2000

1000

0
0 2 4 6 8 10 12
Frequency [Hz]

(c)
Figure 5-24 Comparison of complex modulus at 10, 20 and 30 ºC

103
Chapter 5 Different Laboratory Fatigue Experiments

From these figures it can be seen that, like all viscoelastic materials, the complex
modulus and phase angle of the asphalt mixture are significantly influenced by
temperature and frequency. An increase in modulus is observed with increasing load
frequency and decreasing temperature while the phase angle data show the opposite trend.
To make a comparison of the stiffness values obtained with the different test methods, the
complex stiffness obtained at temperature of 10, 20 and 30 ºC were plotted vs. the
loading frequency; the results are shown in Figure 5-24. In general, the values of complex
modulus measured by UT/C and 4PB test are very close at the same test condition and
these values are not significantly influenced by specimen size.

However the complex stiffness measured by the IT test is always smaller than those
measured by the other two tests. The difference might be caused by the loading mode.
The UT/C and 4PB test were conducted under a sinusoidal waveform loading (including
both tension and compression), while in the IT test the haversine waveform was used.
Furthermore, the IT specimen is subjected to a bi-axial stress condition.

The rheological properties of the asphalt mixture can be described by a master curve
using the time-temperature superposition principle. This principle allows shifting of the
data obtained at various temperatures with respect to time or frequency to a selected
reference temperature. The curve obtained in this way is plotted as a function of reduced
time or frequency. The amount of shifting required at each temperature to the reference
temperature is determined using the Williams-Landel-Ferry (WLF) equation [Ferry,
1980].

- C1 (T − T0 )
LogaT = (5-26)
C2 + (T − T0 )

Where: C1 and C2 : model parameters;


T : test temperature, ºC;
T0 : reference temperature, ºC;
aT : shift factor.

The resulting master curves for the complex modulus and phase angle are described using
a sigmoidal model similar to the one described by Pellinen, [Pellinen and Witczak 2002].

 
 1 
log ( Emix ) = log ( Emin ) + log ( Emax ) − log ( Emin )  ⋅ 1 −  (5-27)
+
b
  10 log fr  
   
 e a  
 
 1 
log ( θmix ) = log ( θmin ) + log ( θmax ) − log ( θmin )  ⋅ 1 −  (5-28)
 10+ log f r 
b
  

 
 e a  

104
Chapter 5 Different Laboratory Fatigue Experiments

Where: Emix : mixture complex modulus, MPa;


Emin : minimum modulus, MPa;
Emax : maximum, MPa;
θmix : mixture phase angle, degree;
θmin : minimum phase angle, degree;
θmax : maximum, degree;
fr : reduced frequency, Hz;
a, b : shape parameters.

To construct the master curves of the complex modulus and phase angle, a reference
temperature of 20ºC was chosen. The experimental data were then fitted to the sigmoidal-
shape function given by Equation 5-27 and 5-28 in combination with a shift factor as
determined by means of Equation5-26. All the model parameters or constants can
automatically be obtained by minimizing the mean relative error using the Solver
function in the Excel spreadsheet. All the model parameters for complex modulus and
phase angle are presented in Table 5-8 and 5-9. The master curves shown in Figure 5-25,
5-26 and 5-27 give a perfect fit for all test data over a wide range of test conditions.

Table 5-8 Master curve parameters for complex modulus


Sigmoidal model WLF factors
Model parameters R2
Emin [MPa] Emax [MPa] a b C1 C2
UT/C-size0.5 0 23058 6.42 2.08 16.30 115.8 0.998
UT/C-size1.0 0 23058 6.39 2.08 15.10 111.2 0.998
UT/C-size1.5 0 18687 7.56 2.79 10.33 76.4 0.991
4PB-size0.5 0 23062 5.87 1.80 15.20 89.7 0.996
4PB-size1.0 0 23058 7.58 2.44 18.77 132.4 0.999
4PB-size1.5 4 23060 8.48 3.11 15.47 111.9 0.998
IT-size1.0 100 16061 9.97 4.86 10.37 72.6 0.998
IT-size1.5 72 22887 9.81 3.93 21.78 150.3 0.998

Table 5-9 Master curve parameters for phase angle


Sigmoidal model WLF factors
Model parameters R2
θmin [º] θmax [º] a b C1 C2
UT/C-size0.5 3.6 76.3 10.45 -4.37 16.30 115.8 0.997
UT/C-size1.0 2.4 77.7 10.97 -3.87 15.10 111.2 0.997
UT/C-size1.5 1.8 73.0 11.57 -3.68 10.33 76.4 0.997
4PB-size0.5 1.9 52.6 11.13 -3.71 11.94 75.8 0.996
4PB-size1.0 2.4 58.1 10.91 -5.10 18.77 132.4 0.999
4PB-size1.5 3.6 51.5 10.63 -5.85 15.47 111.8 0.996

105
Chapter 5 Different Laboratory Fatigue Experiments

1.E+05 100

1.E+04 80
Complex modulus [MPa]

Phase angle [deg]


UTC-size0.5-E
1.E+03 UTC-size1.0-E 60
UTC-size1.5-E
UTC-size0.5-θ
1.E+02 UTC-size1.0-θ 40
UTC-size1.5-θ

1.E+01 20

1.E+00 0
1.E-04 1.E-02 1.E+00 1.E+02 1.E+04 1.E+06
Reduced Frequency [Hz]

Figure 5-25 Master curves obtained from UT/C tests at a reference temperature of 20ºC

1.E+05 60
Complex modulus [MPa]

1.E+04

Phase angle [deg]


40
4PB-size0.5-E
4PB-size1.0-E
1.E+03 4PB-size1.5-E
4PB-size0.5-θ
4PB-size1.0-θ
4PB-size1.5-θ 20
1.E+02

1.E+01 0
1.E-04 1.E-02 1.E+00 1.E+02 1.E+04 1.E+06
Reduced Frequency [Hz]

Figure 5-26 Master curves obtained from 4PB tests at a reference temperature of 20ºC

Figure 5-25 shows that the master curves of the different UT/C specimens overlap each
other. It means that the complex stiffness and phase angle measured from the UT/C test
are not dramatically influenced by the specimen size. The complex modulus measured
with the 4PB test is also identical for different specimen sizes (see Figure 5-26). From the
comparison of the master curves obtained by the IT test which are shown in Figure 5-27,
one can conclude that the complex modulus of the size 1.5specimen is slightly higher
than that of the size 1.0 specimen.

106
Chapter 5 Different Laboratory Fatigue Experiments

1.E+05

Complex modulus [MPa]


1.E+04

1.E+03

IT-size1.0 IT-size1.5

1.E+02
1.E-04 1.E-02 1.E+00 1.E+02 1.E+04 1.E+06
Reduced Frequency [Hz]

Figure 5-27 Master curves obtained from IT test at a reference temperature of 20ºC

1.E+05
Complex modulus [MPa]

1.E+04

1.E+03

UTC-size0.5 UTC-size1.0
1.E+02 UTC-size1.5 4PB-size0.5
4PB-size1.0 4PB-size1.5
IT-size1.0 IT-size1.5
1.E+01
1.E-04 1.E-02 1.E+00 1.E+02 1.E+04 1.E+06
Reduced Frequency [Hz]

Figure 5-28 Comparison of complex modulus master curves for different test methods at
a reference temperature of 20ºC

The comparison of the complex modulus master curves for the different test methods and
specimen sizes is shown in Figure 5-28. In most cases, the complex stiffness measured
with the IT test is lower than the values measured with the UT/C and 4PB test, but at the
lower reduced frequency the IT stiffness becomes larger than the UT/C and 4PB stiffness.
To make the difference clear, Table 5-10 presents the predicted modulus from the master
curve at several reduced frequencies.

As mentioned before, the ABAQUS program was used to simulate the 4PBT. In order to
be able to do so prony series parameters were required in the ABAQUS software. The

107
Chapter 5 Different Laboratory Fatigue Experiments

determination of the prony parameters based on the complex modulus master curves is
demonstrated in Appendix B.

Table 5-10 Comparison of the Predicted modulus values at the reference temperature of
20 ºC
Reduced frequency 0.1 Hz 1 Hz 10 Hz 100 Hz
UT/C-size0.5 2054 5270 9840 14494
UT/C-size1.0 1909 5101 9734 14480
UT/C-size1.5 1740 4668 9257 13856
4PB-size0.5 2365 5170 9724 13813
4PB-size1.0 1960 4772 9018 13708
4PB-size1.5 1872 4615 9340 14512
IT-size1.0 1019 2562 5854 10430
IT-size1.5 1365 3240 6916 12181

5.6.2 Fatigue Test Results


5.6.2.1 Fatigue Life Definition
In this study, two different fatigue criteria are used to determine the fatigue life.

(1) Classical fatigue definition. For the strain-controlled mode, the point of failure is
defined as the moment at which the back calculated stiffness of the specimen has reduced
to 50% of its initial value (stiffness at the 100th cycle), denoted as Nf,50. For the stress-
controlled mode, the failure point is when the specimen has completely fractured,
denoted as Nf.

(2) Dissipated energy ratio. In Section 2.4.3, the dissipated energy theory was introduced
in detail. The fatigue life is defined as the point at which the slope of the dissipated
energy ratio (DER) versus number of load cycles deviates from a straight line [Pronk,
1991]. DER is the ratio of the accumulated dissipated energy up to cycle N and the
dissipated energy in cycle N, given as Equation 5-29.

n =N

∑w i
(5-29)
DER = n =1

wN
wi = π σ i ε i sin θi (5-30)

Where: DER : dissipated energy ratio;


wi : dissipated energy in cycle i , J/m3;
σi : stress amplitude in cycle i ;
εi : strain amplitude in cycle i ;
θi : phase lag in cycle i ;
wN : dissipated energy in cycle N, J/m3.

108
Chapter 5 Different Laboratory Fatigue Experiments

Figure 5-29, Figure 5-30 and Figure 5-31 show examples of the determination of the
classical fatigue life and the fatigue life based on DER for the different fatigue tests.

UT/C_ε=110µε, 20ºC
10000
1400000
Stiffness DER
8000 1200000
Stiffness [MPa]

1000000
6000

DER
800000

4000 50% of initial stiffness 600000

400000
2000
NR Nf,50 200000

0 0
0 100000 200000 300000 400000 500000 600000 700000 800000
Number of cycle

Figure 5-29 Fatigue life determination for the UT/C fatigue test in strain-controlled mode

4PB_σ=2.3 MPa, 5ºC


30000 50000
Stiffness
25000 DER
40000
Stiffness [MPa]

20000
30000
15000 DER
20000
10000

10000
5000 NR Nf

0 0
0 100000 200000 300000 400000 500000 600000 700000
Number of cycle

Figure 5-30 Fatigue life determination for 4PB fatigue test in stress-controlled mode

109
Chapter 5 Different Laboratory Fatigue Experiments

IT-size1.0_σ=1.06 MPa
12000 60000
Stiffness DER
10000 50000

Stiffness [MPa] 8000 40000

DER
6000 30000

4000 20000

2000 10000
NR Nf
0 0
0 20000 40000 60000 80000 100000
Number of cycles

Figure 5-31 Fatigue life determination for IT fatigue test in stress-controlled mode

5.6.2.2 UT/C Fatigue Test Results

Figure 5-32 shows the stiffness evolutions at different test conditions during the UT/C
fatigue test. At 20 ºC, the stiffness curves are rather close for the different specimen sizes
at similar strain level. It indicates that the effect of specimen size has no influence on the
stiffness reduction for the UT/C fatigue test. The blue dots in the figure represent the
stiffness evolution in controlled stress and strain mode at 5 ºC. It can be seen that, under
similar initial strain, the initial stiffness in both loading modes is almost same, but the
fatigue failure in the controlled stress mode occurred considerably earlier than in the
controlled strain mode.

Similar results were also found based on the dissipated energy ratio. The dissipated
energy ratio (DER) was plotted against number of cycles, as shown in Figure 5-33. For
all the cases, the DER of the specimen increases linearly with the number of cycles
following the same slope in the beginning. At 20 ºC, the deviation points of the three
specimen sizes are rather close to each other. At the test temperature of 5 ºC, the
deviation of the DER curve occurs much earlier at the same initial strain level and the
specimen tested in stress controlled mode has the lowest fatigue life determined by DER.

110
Chapter 5 Different Laboratory Fatigue Experiments

size0.5_ε control_116µm /m size1.0_ε control_118µm /m


size1.5_ε control_116µm /m size1.0_ε control_118µm /m
size1.0_σ control_117µm /m
20000
18000
16000
T=5 ºC
14000
Stiffness [MPa]

12000
10000
8000 T=20 ºC
6000
4000
2000
0
10 100 1000 10000 100000 1000000
Number of cycle

Figure 5-32 Stiffness evolutions from the UT/C fatigue tests

size0.5_ε control_116µm /m size1.0_ε control_118µm /m


size1.5_ε control_116µm /m size1.0_ε control_118µm /m
size1.0_σ control_117µm /m
1000000

T=20 ºC
800000
DER

600000

400000
T=5 ºC

200000

0
0 100000 200000 300000 400000 500000 600000
Number of cycle

Figure 5-33 DER evolutions from the UT/C fatigue tests

The fatigue lives based on different definitions are presented in Table 5-11. In nearly all
cases the fatigue life NR is smaller than the classical fatigue life. At 20 ºC the differences
between these two numbers are smaller than at 5 ºC.

As discussed in Chapter 2, fatigue relationships for an asphalt mixture are usually


represented by means of a straight line in a double logarithmic coordinate system.

111
Chapter 5 Different Laboratory Fatigue Experiments

N f = k ⋅ ε 0b (5-31)

Where: Nf : fatigue life;


ε0 : strain amplitude, [µm/m];
k, b : material coefficients

Table 5-11 Fatigue life of the different specimen sizes in the UT/C fatigue test

Specimen Loading Sample Strain level N f ,50 − N R


Nf,50 or Nf NR
size mode code [µm/m] N f ,50
C-3-18 88 1.6×106 1.5×106 6.3%
C-3-25 117 5.4×105 5.2×105 3.7%
C-3-26 130 3.0×105 2.7×105 10.0%
Size 0.5 C-3-9 136 2.8×105 2.6×105 7.1%
at 20 ºC C-3-24 138 2.3×105 2.2×105 6.9%
C-3-4 146 1.3×105 1.0×105 24.2%
C-3-11 168 9.0×104 7.5×104 16.7%
C-3-2 186 5.6×104 5.4×104 3.6%
C-10-12 68 6.0×106 5.2×105 13.3%
C-10-10 84 1.9×106 1.8×105 2.7%
Strain-
C-10-9 110 5.5×105 4.8×105 11.9%
Size 1 at controlled
C-10-8 131 2.8×105 2.3×105 17.9%
20 ºC
C-10-5 155 1.1×105 1.1×105 1.9%
C-10-11 172 7.2×104 7.0×104 2.1%
C-10-7 199 3.2×104 3.4×104 -6.3%
C-14-1 68 4.3×106 3.8×106 12.2%
C-13-1 89 1.3×106 1.2×106 5.5%
Size 1.5
C-11-2 115 4.1×105 3.4×105 16.0%
at 20 ºC
C-9-1 133 1.9×105 1.5×105 23.7%
C-12-2 161 7.0×104 8.0×104 -14.3%
C-11-1 182 4.0×104 4.6×104 -15.0%
C-15-4 69 2.7×106 2.3×106 15.1%
C-15-11 87 1.5×106 1.2×106 20.7%
C-16-5 94 6.6×105 5.6×105 15.2%
Strain-
C-16-10 105 2.8×105 2.3×105 18.1%
controlled
C-15-6 118 1.6×105 1.4×105 31.7%
C-15-9 124 1.6×105 1.2×105 28.6%
Size 1 at C-16-10 136 8.3×104 8.2×104 0.6%
5 ºC C-15-5 68 1.5×106 1.2×106 23.6%
C-16-9 73 1.2×106 7.0×105 40.0%
C-16-2 75 8.4×105 5.2×105 38.3%
Stress-
C-15-7 79 4.7×105 3.6×105 24.1%
controlled
C-15-8 86 2.9×105 1.9×105 34.5%
C-15-3 94 1.6×105 1.1×105 35.4%
C-15-10 118 6.3×104 4.4×104 29.7%

112
Chapter 5 Different Laboratory Fatigue Experiments

Figure 5-34 shows the plot of the classical fatigue life Nf,50 against the initial strain for all
the UT/C fatigue tests. The regression coefficients in Equation 5-31 are presented in
Table 5-12. It is clear that the fatigue lines of the different specimen sizes are almost
identical at the same temperature. With decreasing temperature, the fatigue life becomes
shorter at the same initial strain level and the slope of the fatigue lines, b, becomes
somewhat larger.

1.E+07
Size0.5_ε control_20C
Size1.0_ε control_20C
Size1.5_ε control_20C
Size1.0_ε control_5C
1.E+06 Size1.0_σ control_5C
Nf,50

1.E+05

1.E+04
10 100 1000
Strain level [µm/m]

Figure 5-34 Fatigue life vs. initial strain level for the UT/C fatigue tests

Table 5-12 Regression coefficients of the fatigue lines


Material coefficients
Specimen size Loading mode R2
k b
15
Size 0.5@20ºC 1.5×10 -4.60 0.98
Size 1.0@20ºC Strain-controlled 5.7×1014 -4.43 0.99
15
Size 1.5@20ºC 1.4×10 -4.64 1.00
16
Size 1.0@5ºC Strain-controlled 2.8×10 -5.40 0.97
Size 1.0@5ºC Stress-controlled 1.6×1017 -6.03 0.94

Compared to the strain controlled mode, the specimens have a shorter fatigue life in the
stress controlled mode. This phenomenon can be explained by the dissipated energy
theory. In controlled strain tests, a strain with constant amplitude is applied to the
specimen and the stress decreases gradually. The dissipated energy per cycle decreases
during the testing, as shown in Figure 5-35. In the stress controlled mode, the amplitude
of the applied stress stays constant, which causes an increase of the strain amplitude and
an increase in the dissipated energy. The strain increases rapidly at the end of the test, till
complete fracture occurs.

113
Chapter 5 Different Laboratory Fatigue Experiments

Figure 5-35 Dissipated energy per cycle versus number of cycles in the stress and strain
controlled UT/C test

5.6.2.3 4PB Fatigue Test Results

As mentioned in Section 5.3.2, the measured displacement was corrected to calculate the
complex stiffness. The measured stiffness calculated from the measured deflection vm
and the stiffness corrected from the bending deflection, vb, are presented in Table 5-13.
Also Figure 5-36 gives information about this.

24000
Size0.5_m easured_20°C
Size0.5_corrected_20°C
20000 Size1.0_m easured_20°C
Initial stiffness [MPa]

Size1.0_corrected_20°C
16000 Size1.5_m easured_20°C
Size1.5_corrected_20°C
Size1.0_m easured_5°C
12000 Size1.0_corrected_5°C

8000

4000

0
50 100 150 200 250
Bending strain [µm/m]

Figure 5-36 Comparison of measured and corrected initial stiffness in the 4PB test

114
Chapter 5 Different Laboratory Fatigue Experiments

Figure 5-36 shows that the initial stiffness is independent of the strain level. The
measured stiffness is always lower than the corrected value. For size 0.5, the bending
deflections are close to the measured values, which indicates that during the test the
frame and shear deflection are very small. This is because the specimen size 0.5 has a low
h/L ratio. With the increase of specimen size, the shear deflection becomes larger and
also the deformation of the test frame is playing a more important role because of the
higher loads that need to be applied to obtain a certain strain level. So for size 1.5, the
differences between corrected and measured stiffness become significant, more than 30
%. At lower temperatures, also a larger force is needed to obtain a certain strain level, so
the frame deformation has a significant influence on the measurement. That is why the
corrected stiffness is 25 % higher than the measured value. After the deflection correction,
the difference of the calculated stiffness between the three specimen sizes is much
smaller compared to the measured values.

Table 5-13 Comparison of measured and corrected results in strain controlled mod
Specimen Sample Stress vm vb Measured Corrected Ecorrected − Emeasured
size code [MPa] [mm] [mm] E [MPa] E [MPa] Ecorrected
B-19-5 0.97 0.135 0.131 9410 9989 5.8%
B-14-5 1.17 0.160 0.156 9456 9987 5.3%
Size 0.5 B-19-2 1.36 0.191 0.187 9514 9908 4.0%
@20ºC B-14-6 1.50 0.218 0.213 9062 9586 5.5%
B-19-6 1.69 0.238 0.233 9210 9601 4.1%
B-19-4 1.81 0.272 0.266 8954 9246 3.2%
B-15-2 0.72 0.067 0.058 7282 8857 17.8%
B-16-1 0.91 0.082 0.071 7590 8908 14.8%
Size 1.0 B-16-2 1.06 0.097 0.085 7555 8879 14.9%
@20ºC B-17-4 1.22 0.110 0.096 7600 8976 15.3%
B-16-4 1.40 0.124 0.109 7803 9103 14.3%
B-17-3 1.57 0.140 0.123 7787 9173 15.1%
B-20-2 0.61 0.050 0.033 6105 8545 28.5%
B-23-1 0.70 0.058 0.041 5658 8430 32.9%
Size 1.5
B-22-2 0.80 0.063 0.042 5833 9077 35.7%
@20ºC
B-22-1 0.80 0.063 0.041 5955 9099 34.6%
B-20-1 0.77 0.066 0.046 5134 8225 37.6%
B-21-2 0.89 0.077 0.053 5365 8146 34.1%
B-7-2 1.75 0.079 0.060 14527 19622 26.0%
B-18-3 1.80 0.086 0.065 14183 18855 24.8%
B-15-3 1.84 0.092 0.070 13987 18517 24.5%
Size 1.0
B-18-1 2.12 0.099 0.074 14491 19496 25.7%
@5ºC
B-15-4 2.25 0.105 0.078 14069 19601 28.2%
B-18-4 2.48 0.112 0.085 15025 20467 26.6%
B-15-1 2.46 0.117 0.090 14298 19639 27.2%

115
Chapter 5 Different Laboratory Fatigue Experiments

The stiffness evolutions of specimens with different sizes at a similar strain are shown in
Figure 5-37. It can be seen that the fatigue behavior in the 4PB fatigue tests also depends
on temperature and loading mode. In contrast to the UT/C test results, the effect of
specimen size becomes distinct. At the same loading level, the stiffness of a larger
specimen drops more quickly. These results also can be found on the plot of the
dissipated energy ratio, as shown in Figure 5-38.

size0.5_ε control_100µm/m size1.0_ε control_105µm/m


size1.5_ε control_116µm/m size1.0_ε control_109µm/m
size1.0_σ control_108µm/m
25000

T=5 ºC
20000
Stiffness [MPa]

15000

10000
T=20 ºC

5000

0
10 100 1000 10000 100000 1000000 10000000
Number of cycle

Figure 5-37 Stiffness evolutions from the 4PB fatigue test in strain and stress controlled
mode

size0.5_ε control_100µm/m size1.0_ε control_105µm/m


size1.5_ε control_116µm/m size1.0_ε control_109µm/m
size1.0_σ control_108µm/m
800000

700000

600000
T=20 ºC
DER

500000

400000

300000

200000

100000 T=5 ºC
0
0 1000000 2000000 3000000 4000000 5000000 6000000
Number of cycle

Figure 5-38 DER evolutions from the 4PB fatigue tests in strain and stress controlled
mode

116
Chapter 5 Different Laboratory Fatigue Experiments

The classical fatigue life and the fatigue life based on DER, NR, are presented in Table 5-
14. As was the case for the UT/C results, the value of NR obtained with the 4PB test is
always smaller than the classical fatigue life, but the differences between them are much
larger in most of the cases compared to the differences obtained in the UT/C test.

Table 5-14 Fatigue life of the different specimen sizes in the 4PB fatigue test

Specimen Loading Sample Strain level N f ,50 − N R


Nf,50 or Nf NR
size mode code [µm/m] N f ,50
6 6
B-19-5 100 5.7×10 3.1×10 44.8%
B-14-5 119 1.3×106 6.8×105 45.9%
Size 0.5 B-19-2 140 8.5×105 4.3×105 49.5%
at 20 ºC B-14-6 160 3.8×105 1.6×105 57.5%
B-19-6 179 2.2×105 9.5×104 56.4%
B-19-4 201 2.1×105 7.3×104 64.8%
B-15-2 87 4.0×106 3.2×106 20.6%
B-16-1 105 1.9×106 1.0×106 47.8%
Strain-
Size 1 at B-16-2 124 6.6×105 4.2×105 36.5%
controlled
20 ºC B-17-4 142 4.2×105 2.7×105 36.8%
B-16-4 161 3.1×105 1.8×105 43.0%
B-17-3 180 1.9×105 1.1×105 45.6%
B-20-2 74 7.9×106 6.2×106 22.1%
Size 1.5 B-23-1 88 1.7×106 1.5×106 11.7%
at 20 ºC B-22-2 93 1.6×106 1.0×106 35.9%
B-20-1 103 2.5×106 1.7×106 30.2%
B-21-2 119 5.8×105 5.0×105 14.0%
B-7-2 89 2.0×106 1.7×106 14.5%
B-15-3 105 1.6×106 1.0×106 37.8%
Strain- B-18-1 109 1.3×106 9.1×105 27.2%
controlled B-15-4 116 3.4×105 2.8×105 18.0%
B-18-4 124 4.7×105 4.2×105 10.7%
B-15-1 135 4.1×105 3.4×105 16.4%
Size 1 at
B-11-4 87 2.4×106 1.6×106 34.2%
5 ºC
B-11-1 95 1.7×106 8.9×105 48.8%
B-11-2 93 1.9×106 8.3×105 55.5%
Stress-
B-7-3 98 9.1×105 3.2×105 65.1%
controlled
B-7-4 97 7.3×105 2.5×105 65.4%
B-7-1 99 6.0×105 2.1×105 64.6%
B-16-3 108 5.8×105 1.9×105 67.3%

The fatigue lines obtained for size 0.5, 1.0 and 1.5 specimens are plotted in Figure 5-39
and all the regression coefficients are gathered in Table 5-15. The effect of specimen size
can be observed from the comparison of the fatigue lines. A specimen with a larger
height has a shorter fatigue life at the same loading level. One of the reasons is that the
shear deflection is related to the height/length (h/l) ratio of the beam specimen, as shown
in Equation 5-4. This implies that at the same bending strain level, the shear strain

117
Chapter 5 Different Laboratory Fatigue Experiments

between the inner supports is larger for the specimen with a high h/l ratio, resulting in
more damage during the test. The slope and the intercept of the fatigue line increase with
a decrease of temperature. In terms of the correlation coefficients, the variation of fatigue
results for specimen size 1.5 and low temperature is relatively high. The reason might be
that the range of applied loading levels in these cases is not wide enough.

1.E+07
Size0.5_ε control_20C
Size1.0_ε control_20C
Size1.5_ε control_20C
Size1.0_ε control_5C
Size1.0_σ control_5C
Nf,50

1.E+06

1.E+05
10 100 1000
Bending strain [µm/m]

Figure 5-39 Fatigue life vs. initial strain level for the 4PB fatigue tests

Table 5-15 Regression coefficients of the fatigue lines for the 4PB test
Material coefficients
Specimen size Loading mode R2
k b
16
Size 0.5@20ºC 1.1×10 -4.72 0.95
Size 1.0@20ºC Strain-controlled 5.3×1014 -4.20 0.98
15
Size 1.5@20ºC 3.6×10 -4.69 0.78
Size 1.0@5ºC Strain-controlled 2.0×1015 -4.58 0.74
22
Size 1.0@5ºC Stress-controlled 1.2×10 -8.08 0.76

As mentioned in section 5.2.2, the 4PB test setup was modified to avoid stress
concentration near the inner clamps. Figure 5-40 gives the comparison of the fatigue lines
from the modified and unmodified test setup for each specimen size.

It can be seen from the figures, that for size 0.5 the fatigue lines from the two setups are
very close. However, with the increase of the specimen size, the fatigue life from the
unmodified setup decreases dramatically at lower strain levels compared to that from the
modified setup at the low loading range so that the slope of the fatigue line decreases
with the increase of the specimen size.

118
Chapter 5 Different Laboratory Fatigue Experiments

1.E+07
y = 1E+16x -4.723
R2 = 0.9532
y = 5E+15x -4.6029
R2 = 0.9601
Nf,50 1.E+06

Size0.5-unmodified
Size0.5-modified
1.E+05
10 100 1000
Strain level [µm/m]

(a) Specimen size 0.5


1.E+08
Size1.0-unmodified
Size1.0-modified

1.E+07
y = 5E+14x -4.2037
Nf,50

y = 3E+11x -2.7383 R2 = 0.9838

1.E+06 R2 = 0.9454

1.E+05
10 100 1000
Strain level [µm/m]

(b) Specimen size 1.0


1.E+07

y = 4E+15x -4.6946
R2 = 0.7783
Nf,50

1.E+06
y = 2E+10x -2.1199
R2 = 0.919

Size1.5-unmodified
Size1.5-modified
1.E+05
10 100 1000
Strain level [µm/m]

(c) Specimen size 1.5


Figure 5-40 Comparison of fatigue results between modified and unmodified 4PB setup

119
Chapter 5 Different Laboratory Fatigue Experiments

5.6.2.4 IT Fatigue Test Results


During the IT fatigue test, a compressive load is applied on the specimen along its
vertical diameter. Therefore permanent deformation will develop even at a low
temperature of 5ºC. Figure 5-41 shows the load and deformation responses in the 50th
and the 60th cycle. In the figure, Dr is defined as the recoverable deformation in the 50th
cycle and Dp represents the permanent deformation that developed between the 50th and
60th loading cycle.
10 0.01
Force
Radial deformation

Radial deformation [mm]


8 0.008
Force [kN]

6 0.006

4 Dr 0.004

2 Dp 0.002

0 0
4.8 5 5.2 5.4 5.6 5.8 6 6.2 6.4
time [s]

Figure 5-41 Force and deformation response in the 50th and the 60th cycle

Figure 5-42 shows the horizontal recoverable deformation and vertical permanent
deformation for size 1.5. The horizontal recoverable deformation Dr at the various stress
levels stays stable in the beginning of a test, because in this period, the material behavior
can be considered linear viscoelastic. After a certain number of cycles, Dr increases
rapidly and the specimen is damaged in a short time. The vertical permanent deformation,
Dp always increases during the test.

0.018
I-3-4_0.56MPa I-2-6_0.61MPa
0.016
Horizontal recoverable deform. [mm]

I-4-4_0.64MPa I-3-6_0.69MPa
0.014 I-2-7_0.83MPa I-4-7_0.88MPa
I-4-6_0.93MPa
0.012

0.01

0.008

0.006

0.004

0.002

0
10 100 1000 10000 100000 1000000
Number of cycles

(a) Horizontal recoverable deformation

120
Chapter 5 Different Laboratory Fatigue Experiments

7
I-3-4_0.56MPa I-2-6_0.61MPa
6 I-4-4_0.64MPa I-3-6_0.69MPa

Vertical permanent deform. [mm]


I-2-7_0.83MPa I-4-7_0.88MPa
5 I-4-6_0.93MPa

0
10 100 1000 10000 100000 1000000
Number of cycles

(b) Vertical permanent deformation


Figure 5-42 Recoverable and permanent deformation for size 1.5 at 20 ºC

Figure 5-43 gives the evolution of the resilient modulus for the specimen with size 1 and
size 1.5 at the various stress levels. Similar to the recoverable deformation results, the
material behaves linear elastic in the beginning, as the stiffness stays constant with
increasing number of load repetitions. When cracks start to initiate and propagate inside
the specimen, the stiffness decreases quickly. Table 5-16 presents the initial stiffness and
fatigue life at all the stress levels for both specimen sizes. In general, the initial stiffness
of the specimen size 1.5 is a little higher than that of specimen size 1.0, which was also
found in the stiffness tests.

10000

8000
Resilient Stiffness [MPa]

6000
I-4-3_0.46MPa
I-3-1_0.49MPa
4000 I-1-4_0.58MPa
I-1-9_0.71MPa
I-1-5_0.89MPa
2000 I-1-7_1.06MPa
I-1-6_1.28MPa
0
10 100 1000 10000 100000 1000000 10000000
Num ber of cycles

(a) Size 1.0

121
Chapter 5 Different Laboratory Fatigue Experiments

12000

10000

Resilient stiffness [MPa]


8000
I-3-4_0.56MPa
6000 I-2-6_0.61MPa
I-4-4_0.64MPa
I-3-6_0.69MPa
4000
I-2-7_0.83MPa
I-4-7_0.88MPa
2000 I-4-6_0.93MPa

0
1 10 100 1000 10000 100000 1000000
Number of cycles

(b) Size 1.5


Figure 5-43 Stiffness evolutions for the specimen size 1 (a) and size 1.5 (b) at 5 ºC

The fatigue results of the two specimen sizes, based on the initial tensile strain plotted
against the number of cycles to failure are shown in Figure 5-44. The regression
coefficients of the fatigue functions are presented in Table 5-17. The fatigue lines of size
1.0 and size 1.5 appear to have a similar slope. The fatigue life of the size 1.5 specimen is
slightly lower than that of the size 1.0 specimens, but the effect of specimen size is not as
significant as for the 4PB fatigue results.

Table 5-16 Initial stiffness and fatigue life for two specimen sizes at 5 ºC

Sample Initial εt Initial E N f − NR


σt [MPa] Nf NR
code [µm/m] [MPa] Nf
6 5
I-4-3 0.46 51 9361 1.8×10 9.0×10 50.4%
I-3-1 0.49 57 9010 7.4×105 4.0×105 45.8%
I-1-4 0.58 61 9551 1.0×106 4.5×105 56.8%
Size 1 I-1-9 0.71 78 9381 4.4×105 2.0×105 55.0%
I-1-5 0.90 97 9282 1.5×105 1.1×105 28.9%
I-1-7 1.06 116 9170 8.4×104 5.5×104 34.3%
I-1-6 1.28 143 9005 3.0×104 2.0×104 32.5%
I-3-4 0.55 60 9536 7.7×105 4.0×105 48.1%
I-2-6 0.61 59 10213 6.9×105 4.3×105 37.7%
I-4-4 0.64 70 8474 4.5×105 2.2×105 51.6%
Size
I-3-6 0.69 72 9620 3.4×105 1.8×105 47.2%
1.5
I-2-7 0.83 87 9493 1.6×105 8.0×104 50.2%
I-4-7 0.88 89 9820 1.3×105 6.0×104 53.7%
I-4-6 0.93 95 9893 9.4×104 6.0×104 36.1%

122
Chapter 5 Different Laboratory Fatigue Experiments

1.E+07
Size1
Size1.5

Number of cycles
1.E+06

1.E+05

1.E+04
10 100 1000
Strain level [µm/m]

Figure 5-44 Fatigue lines for specimen size 1.0 and size 1.5 in IT fatigue tests

Table 5-17 Regression coefficients of the fatigue lines for the IT test
Material coefficients
Specimen size Loading mode R2
k b
Size 1.0@5ºC Stress-controlled 6.4×1012 -3.84 0.99
Size 1.5@5ºC Stress-controlled 4.0×1013 -4.33 0.98

5.6.2.5 Comparison of Different Fatigue Results


In the above discussion, the UT/C, 4PB and IT fatigue tests results were presented
obtained on specimens with different specimen sizes. To explore the effect of the test
type on the fatigue tests, the fatigue lines at the same test conditions obtained with the
different fatigue tests are compared. Figure 5-45, Figure 5-46 and Figure 5-47 summarize
the fatigue results obtained from the UT/C, 4PB and IT fatigue tests. To make a clear
comparison, Table 5-18 presents the fatigue life at an initial strain of 100 µm/m.

1.E+07

1.E+06
Nf,50

1.E+05
UT/C_Size0.5 UT/C_Size1.0
UT/C_Size1.5 4PB_Size0.5
4PB_Size1.0 4PB_Size1.5
1.E+04
10 100 1000
Strain level [µm/m]

Figure 5-45 Comparison of fatigue lines in strain-controlled mode at 20ºC and 10 Hz

123
Chapter 5 Different Laboratory Fatigue Experiments

1.E+07

1.E+06
Nf

1.E+05 UT/C_Size1.0@5C
4PB_Size1.0@5C
ITT_Size1.0@5C
ITT_Size1.5@5C
1.E+04
0.1 1 10
Stress level [MPa]

Figure 5-46 Comparison of fatigue lines in stress-controlled mode at 5ºC from UT/C and
4PB and IT fatigue tests

1.E+07
UTC Size0.5_ε_20°C
UTC Size1.0_ε_20°C
UTC Size1.5_ε_20°C
UTC Size1.0_ε_5°C
1.E+06
UTC Size1.0_σ_5°C
4PB Size0.5_ε_20°C
Nf

4PB Size1.0_ε_20°C
4PB Size1.5_ε_20°C
1.E+05 4PB Size1.0_ε_5°C
4PB Size1.0_σ_5°C
IT Size1.0_σ_5°C
IT Size1.5_σ_5°C
1.E+04
10 100 1000
Initial strain [µm/m]

Figure 5-47 Comparison of fatigue lines obtained from all the three fatigue tests

Table 5-18 Comparison of the fatigue life at the initial strain of 100 µm/m
Test type Fatigue life Test type Fatigue life
UT/C-size0.5_ε_20 ºC 8.7×105 4PB-size0.5_ε_20 ºC 4.1×106
UT/C-size1.0_ε_20 ºC 7.8×105 4PB-size1.0_ε_20 ºC 2.1×106
UT/C-size1.5_ε_20 ºC 7.3×105 4PB-size1.5_ε_20 ºC 1.5×106
UT/C-size1.0_ε_5 ºC 4.5×105 4PB-size1.0_ε_5 ºC 1.4×106
UT/C-size1.0_σ_5 ºC 1.4×105 4PB-size1.0_σ_5 ºC 8.3×105
IT-size1.0_ε_5 ºC 1.3×105
IT-size1.5_ε_5 ºC 8.6×104

124
Chapter 5 Different Laboratory Fatigue Experiments

For both strain-controlled and stress-controlled mode, the 4PB test gives the longest
fatigue life for all the test types and the IT test shows the shortest life duration. Similar
results were also found in previous research [Porter, 1975] [Di Benedetto, 2004]
[Molenaar, 1983]. The difference in fatigue life obtained from different test types can be
explained by the stress-strain distribution in the specimen during the test. In the 4PB test,
the stress and strain values vary along the length and height. The maximum strain only
occurs at the specimen surface between the two inner supports, while in the UT/C test,
the stress-strain field is relatively uniform. The target stress or strain level is distributed
throughout most of the cylindrical specimen. In the IT test, a haversine signal was used,
while the stress signal in both the 4PB and UT/C test was a full sine, implying alternating
tension/compression stresses. Furthermore, the specimen in the IT test is subjected to a
biaxial stress state (vertical and horizontal stresses). Therefore, more damage was caused
by this complicated loading configuration. In addition to the fatigue damage
accumulation of permanent deformation occurs in the IT fatigue test.

5.6.2.6 Dissipated Energy Theory


As discussed in the literature review, the area inside the stress–strain hysteresis loop for a
loading–unloading process indicates the amount of dissipated energy for one loading
cycle. Figure 4-48 shows the development of the stress–strain hysteresis loop for the
UT/C, 4PB and IT fatigue tests in the stress-controlled mode.

For the UT/C and 4PB fatigue test, the stress-strain loop rotates around the origin and its
area becomes larger with the number of loading cycles. In contrast, during the IT fatigue
test, the center of the specimen is only subjected to a tensile stress in the horizontal
direction, the loop moves along the strain axis due to the irreversible deformation.

3
100 cycle
2
80000 cycle
160000 cycle
1
Stress [MPa]

0
-0.0002 -0.0002 -0.0001 -5E-05 0 5E-05 0.0001 0.0002 0.0002
-1

-2

-3
Strain [µm/m]

(a) UT/C fatigue test: C-15-3

125
Chapter 5 Different Laboratory Fatigue Experiments

3
100 cycle
2
800000 cycle
1700000 cycle
1
Stress [MPa]
0
-0.0002 -0.0002 -0.0001 -5E-05 0 5E-05 0.0001 0.0002 0.0002
-1

-2

-3
Strain [µm/m]

(b) 4PB fatigue test: B-11-1

1.5
50 cycle 100 cycle 300 cycle
Stress [MPa]

0.5

0
0 0.0001 0.0002 0.0003 0.0004
Strain
Horizontal [µm/m]
[µm/m]
strain

(c) IT fatigue test: I-1-5


Figure 5-48 Development of strain-stress hysteresis loops for the different fatigue tests

It is generally believed that the fatigue life is related to the accumulated dissipated energy.
In this section the relationship between fatigue life NR and the total dissipated energy till
NR is investigated. NR is determined by the evolution of the dissipated energy and
represents the number of cycles to crack initiation. As shown in Figure 5-32, the
evolution of the dissipated energy ratio DER in the stress controlled and strain controlled
mode follows the same slope till crack initiation. Therefore it is believed that the fatigue
life NR for the different loading modes should be comparable. The total dissipated energy
is given as follows:

NR NR
W fat = ∑ Wi = π ∑ σ iε i sin φi (5-32)
i =1 i =1

126
Chapter 5 Different Laboratory Fatigue Experiments

Where: Wfat : total dissipated energy till NR, J/m3;


Wi : dissipated energy at cycle i, J/m3;
NR : fatigue life based on DER;
σi : stress amplitude at cycle i, MPa;
εi : strain amplitude at cycle i, m/m
Φi : phase angle at cycle i.

Based on previous research [van Dijk, 1972], the following relation was obtained:

W fat = A ⋅ ( N R )
z
(5-32)

Where: A and z : experimentally determined coefficients

Figure 5-49 shows the relationships between total dissipated energy and fatigue life NR
obtained from the different fatigue tests. Table 5-19 presents all the regression
coefficients.
In Figure 5-49a, the NR-Wfat lines from the UT/C test are not influenced by the specimen
size like the traditional fatigue analysis, but the specimen size still has influence on the
NR-Wfat lines obtained from the 4PB test. From Figure 5-49b, it can be seen that the
fatigue relations based on the total dissipated energy obtained from the UT/C and 4PB
test are close and both of them are influenced by the loading mode. Compared to the
strain controlled mode, the slope of the curves in the stress-controlled mode is a little
larger, but the effect of temperature on the fatigue line is negligible. While the fatigue life
NR from the IT fatigue test is much shorter at consuming the same amount of total
dissipated energy. One of the reasons is that the damage of the IT specimen is not only
caused by the dissipated energy, but also by the irreversible deformation (see Figure 5-
48c).

(a) Fatigue lines from the UT/C and 4PB test with different specimen sizes

127
Chapter 5 Different Laboratory Fatigue Experiments

(b) Fatigue lines from the different tests at different temperatures


Figure 5-49 Total dissipated energy versus fatigue life NR

Table 5-19 Coefficients of the NR-Wfat lines from the different fatigue tests
Coefficients
Specimen size Loading mode R2
A Z
UTC-Size 0.5@20ºC 0.042 0.561 0.99
UTC-Size 1.0@20ºC 0.062 0.523 1.00
Strain-controlled
UTC-Size 1.5@20ºC 0.030 0.578 1.00
UTC-Size 1.0@5ºC 0.015 0.541 0.99
UTC-Size 1.0@5ºC Stress-controlled 0.008 0.658 1.00
4PB-Size 0.5@20ºC 0.025 0.624 0.99
4PB-Size 1.0@20ºC 0.053 0.547 0.98
Strain-controlled
4PB-Size 1.5@20ºC 0.005 0.691 0.96
4PB-Size 1.0@5ºC 0.039 0.558 0.96
4PB-Size 1.0@5ºC Stress-controlled 0.001 0.796 0.99
IT-Size 1.0@5ºC Stress-controlled 0.108 0.304 0.96
IT-Size 1.5@5ºC Stress-controlled 0.030 0.406 0.96

The question now is why the NR-Wfat relation of the IT test is so different from the ones
of the other tests. In the UT/C and 4PB test, the outer layer of the specimen is alternately
subjected to tension and compression. The dissipated energy calculated for these test
consists of a "tensile" and "compressive" part. The dissipated energy calculated from the
IT test results is only due to tension. The effect of the dissipated energy due to the high
compressive strains in the vertical direction is not taken into account. The influence of
this component was investigated and is reported hereafter.

Figure 5-50 gives a comparison of the products of the stress and strain in horizontal and
vertical direction for the IT test. The values of the horizontal stress σx, horizontal stress εx,

128
Chapter 5 Different Laboratory Fatigue Experiments

vertical stress σy and vertical strain εy are calculated with the ABAQUS software. The
details are discussed in Chapter 8. From the figure, it can be seen that the product of the
vertical stress and strain is much larger than that of the tension part. Assuming the phase
angle in vertical direction is the same as that in horizontal direction, the total dissipated
energy in vertical and horizontal direction together is around 7 times larger than that only
in horizontal direction. A correction of NR vs. Wfat relationship of the IT test with this
factor moves it very close to the other relationships, but the slopes of the curves are still
smaller than the other two tests (see Figure 5-51).

Furthermore, the low dissipated energy in the IT tests is caused by the measured phase
angle. Figure 5-52 gives a comparison of the initial phase angle measured for the
different test. At various stress levels, the phase angles measured by the IT test are
always smaller than those measured by the UT/C and 4PB test. The lower phase angle
causes a lower total dissipated energy as calculated by means of Equation 5-32.
0.00025

0.0002

0.00015
σ*ε

0.0001

0.00005

0
σx*εx σy*εy σx*εx+σy*εy

Figure 5-50 Products of the stress and strain at the center of the specimen I-3-1

Figure 5-51 Total dissipated energy versus fatigue life NR after correcting the total
dissipated energy for the IT test

129
Chapter 5 Different Laboratory Fatigue Experiments

20
UT/C-size1.0@5°C
4PB-size1.0@5°C

Phase angle [deg]


15
IT-size1.0@5°C

10

0
0 0.5 1 1.5 2 2.5
Stress [MPa]

Figure 5-52 Comparison of the phase angle from the different test in stress-controlled
mode

5.7 Summary
5.7.1 Test Setup

In this Chapter, asphalt specimens with different specimen sizes were tested using the
UT/C, modified 4PB and IT fatigue test. With regard to the test setups, the following
conclusions were reached:

1. The test configuration of the UT/C test is relatively simple, but the requirements for
specimen preparation and test setup are very strict. To prevent bending failure, the
upper and lower steel platens should be parallel to each other and perpendicular to the
loading direction. Likewise the polished top and bottom ends of the cylindrical
specimen should be exactly parallel to each other and perpendicular to the axis of the
cylinder.
2. Compared to the UT/C and IT test, the measurements obtained from the 4PB test are
highly influenced by the test equipment. To avoid high shear and tensile stress
concentrations near the supports, the clamps should be allowed to rotate freely as well
as translate freely at the four supports. Therefore the clamping force should be as low
as possible. Another important issue is the calibration of the 4PB test setup. To
calculate the stiffness of the specimen, shear deflection of beam and deflection of the
frame should be taken into account to get the pure bending deflection of the beam.

130
Chapter 5 Different Laboratory Fatigue Experiments

5.7.2 Test Results

Based on the results discussed in this chapter, the following conclusions can be drawn:

1. The complex stiffness measured by means of the UT/C and 4PB test is not
dramatically influenced by the specimen size and the stiffness mastercurves from
these two tests are very close to each other. For the IT test, the larger specimens
showed have a higher complex stiffness. However the complex stiffness measured by
the IT test is always smaller than the complex stiffness measured by the other two
tests. The difference is most probably caused by the difference in the mode of loading
and the state of stress in the specimen
2. In the UT/C test, the fatigue life NR based on the dissipated energy ratio is smaller
than the classical fatigue life Nf,50. Due to the homogeneous stress-strain field, the
effect of specimen size on the fatigue life of the cylindrical specimen at the same
temperature is not significant. At a lower temperature, the fatigue life becomes
shorter and the slope of the fatigue lines becomes steeper.
3. The difference between the NR and Nf,50 fatigue values obtained from the 4PB test is
larger compared to the UT/C test. The 4PB fatigue results are significantly influenced
by the specimen size. The beam with a higher height/length ratio has a shorter fatigue
life at the same loading level. The slope and the intercept of the fatigue line increase
at a lower temperature.
4. For the IT test, the effect of specimen size on the fatigue results is not significant.
5. Based on the comparison of the fatigue results between the different test methods, at
the same test condition and loading mode, the fatigue life from the 4PB test is the
longest and from the IT test is the shortest. The differences can be explained by the
difference in stress-strain distributions of the specimens in the different fatigue tests.
6. With regard to the NR-Wfat relation, the effect of specimen size on the 4PB test is
obvious, but not for the UT/C and IT test. The NR-Wfat line is independent of
temperature but influenced by the loading mode. Compared to the strain controlled
mode, the slope of the curves in the stress-controlled mode is a little steeper. The total
dissipated energy measured by the IT test is lower compared to the UT/C and 4PB
tests at the same fatigue life NR. This difference is caused by the irreversible
deformation, biaxial stress state at the center of the specimen and the smaller phase
angle measured by the IT test.

131
Chapter 5 Different Laboratory Fatigue Experiments

References

Di Benedetto H., de la Roche C., Baaj H., Pronk A. and Lundstrom R., Fatigue of
bituminous mixtures. Materials and Structures, 2004, 37, No. 3, 202–216.

EN 12697-24: 2003 Bituminous mixture – Test methods for hot mix asphalt - Part 24:
Resistance to fatigue. European committee for standardization, Brussels.

Ferry, J.D., Viscoelastic properties of Polymers. 3rd edition. New York, 1980.

Hondros, G.. Evaluation of Poisson’s Ratio and the Modulus of Materials of a Low
Tensile Resistance by the Brazilian (Indirect Tensile) Test with Particular Reference to
Concrete. Austr. J. Appl. Sci. 1959, 10(3): 243-268.

Huurman, M., Pronk, A.C., Theoretical analysis of the 4 point bending test Proceedings
of the 7th Int. RILEM Symposium Advanced Testing and Characterization of Bituminous
Materials, May 2009, Rhodes, Greece.

Li, N., Molenaar, A.A.A., Pronk A.C., van de Ven, M.F.C. and Wu, S., Effect of
Specimen Size on Fatigue Behavior of Asphalt Mixture in Laboratory Fatigue Tests,
Proceedings of the 7th RILEM International Conference on Cracking in Pavements, Delft,
the Netherlands, pp 827-836, 2012.

Li, Ning, Molenaar, A.A.A., Pronk A.C., van de Ven, M.F.C.. Investigation into the Size
effect on Four Point Bending Fatigue Tests. Proceedings of the 3rd Workshop on 4PB.
Davis, California. 2012.

Pronk A.C. and Huurman M., Shear deflection in 4PB tests. 2nd Workshop on Four Point
Bending, University of Minho, Guimaraes, Portugal, 2009.

Pronk, A. C. and Hopman P. C., Energy dissipation: the leading factor of fatigue,
Proceedings of the Conference on the United States Strategic Highway Research Program,
1991, pp. 255–267.

Porter, B. W., Kennedy, T. W., Comparison of the fatigue test methods for asphalt
materials. Research report 183-4, Center for Highway Research, University of Texas at
Austin. 1975.

Pellinen, T.K., and Witczak, M.W.. “Stress Dependent Master Curve Construction for
Dynamic (Complex) Modulus.” Journal of the Association of Asphalt Paving
Technologists, Volume 71. Colorado Spring, CO, 2002.

van Dijk, W. Moreaud, H., et al. The Fatigue of Bitumen and Bituminous Mixes, 3th
International Conference on Structural Design of Asphalt Pavements (ISAP), London,
Volume 1. 1972.

132
Chapter 6 Application of Partial Healing Model on Strain Controlled Fatigue Tests

Chapter 6 Application of Partial Healing


Model on Strain Controlled Fatigue Tests

6.1 Introduction
In Chapter 5 it was shown that the different fatigue tests give different fatigue results.
The main reason for this is the difference in the stress-strain field in the specimen during
the test. In this chapter, the partial healing (PH) model is used to simulate the fatigue
behavior. With the PH model the differences between the UT/C and 4PB test are
explained and the so-called endurance limit is determined. This model was developed and
modified by A.C. Pronk and is based on the dissipated energy theory [Pronk, 2001]. It is
a material model that describes the evolution of the complex modulus and phase angle for
a unit volume during a fatigue test. This implies that the model can be applied directly to
the fatigue results obtained from “homogenous” tests [Pronk, 2006]. By taking into
account the dimensions and the stress-strain distribution of the specimen, a weighted
complex modulus can be obtained for so-called “inhomogeneous” tests, such as bending
tests [Pronk, 2010]. Therefore the PH model offers a possibility to compare different
fatigue results.

6.2 Application of PH Model on UT/C Test Results


6.2.1 PH Model Theory
As discussed in Chapter 2, energy is dissipated in asphalt mixtures during repeated
loading, because the material behaves substantially visco-elastic. An example of a stress–
strain hysteresis loop during one loading cycle is shown in Figure 6-1. The area inside of
the hysteresis loop indicates the amount of dissipated energy in one cycle, which can be
calculated with Equation 6-1 [van Dijk, 1972] [Pronk, 1991] [Rowe, 1996].

2
σi
1.5

1
εi
Stress [MPa]

0.5

0
-200 -150 -100 -50 0 50 100 150 200
-0.5

-1

-1.5

-2
Strain [µm/m]

Figure 6-1 Stress-strain hysteresis loop in one cycle

133
Chapter 6 Application of Partial Healing Model on Strain Controlled Fatigue Tests

wi = π σ i ε i sin ϕi (6-1)

Where: wi : dissipated energy in cycle i , [MPa];


σi : stress amplitude in cycle i , [MPa];
εi : strain amplitude in cycle i , [m/m];
φi : phase lag in cycle i , [°]

Normally it is believed that the total energy dissipated into the test device (system losses
∆Wsys) and the specimen (visco-elastic losses ∆Wdis and fatigue damage ∆Wfat),
consisting of three components [Pronk, 1996] [Pronk, 2001]:

(1) System losses ∆Wsys:


The energy is dissipated into the test device. By calibrating the test set-up, the system
loss can be diminished.

(2) Visco-elastic losses ∆Wdis:


Most of the dissipated energy is associated with viscoelastic loss, which is completely
transformed into heat and results in an increase of specimen temperature. This amount of
energy is approximately equal to the area of the stress-strain loop. So ∆Wdis for a unit
volume dV can be calculated by Equation 6-2 for a sinusoidal loading.

∆Wdis = π ⋅ σ i ⋅ ε i ⋅ sin (ϕi ) (6-2)

Due to the increase of the temperature in the specimen, the visco-elastic losses are
responsible for a small part of the decrease in complex stiffness in the beginning of the
fatigue test. If a good forced convection is used in the climate chamber, equilibrium in
temperature distribution is normally reached within 10000 loading cycles. Therefore the
decrease in stiffness modulus due to the increase in temperature is negligible after 10000
loading cycles [Pronk, 1996].

(3) Fatigue consumption ∆Wfat:


Only a small part of the total energy relates to the formation of microcracks, which is
denoted as fatigue consumption ∆Wfat. This part of the energy is the main reason for the
decrease of stiffness and the increase of the phase angle during the fatigue test. In the
proposed model, it is assumed that the mathematical formulation of this part is modeled
as a very small portion of the visco-elastic losses ∆Wdis. In a strain controlled mode test,
∆Wfat is expressed by Equation 6-3:

∆W fat = δ ⋅ ∆Wdis = δ ⋅ π ⋅ σ i ⋅ ε i ⋅ sin (ϕi ) = δ ⋅ π ⋅ Si ⋅ ε 02 ⋅ sin (ϕi ) (6-3)

Where: ∆Wfat : dissipated energy related to fatigue damage at cycle i , [MPa];


δ : very small parameter (<<1);
Si : stiffness modulus at cycle i, [MPa];
ε0 : strain amplitude, a constant in the strain controlled mode, [m/m].

134
Chapter 6 Application of Partial Healing Model on Strain Controlled Fatigue Tests

A new parameter, the stiffness damage Q, is introduced, which relates to the fatigue
consumption ∆Wfat. The damage factor Q reduces the stiffness modulus, including the
loss modulus F and the storage modulus G, expressed by Equation 6-4 and 6-5,
respectively.

t
dQ {τ }
F {t} = S {t} ⋅ sin {ϕ ( t )} = F0 − ∫ α1 e − β (t −τ ) + γ 1  ⋅ dτ
  (6-4)
0

t
dQ {τ }
G {t} = S {t} ⋅ cos {ϕ ( t )} = G0 − ∫ α 2 e − β (t −τ ) + γ 2  ⋅ dτ
  (6-5)
0

Where: t : test duration, [s];


F{t} : loss modulus at testing time t, [MPa];
G{t} : storage modulus at testing time t, [MPa];
S{t} : stiffness modulus at testing time t, [MPa];
φ{t} : phase angle at testing time t, [°];
F0 : initial loss modulus, [MPa];
G0 : initial storage modulus, [MPa];
Q{t} : stiffness damage at testing time t, [MPa];
α1, α2, γ1, γ2, β : model parameters.

In the damage part (the integral terms in Equation 6-4 and 6-5), the term with α1 and α2
will vanish with time according to the time decay parameter β. So this term represents the
reversible damage which will heal in time. It can also be seen as a mathematical
formulation for the thixotropic behavior of asphalt mixtures. The terms describe partial
healing of the stiffness modulus which is not always the same as partial healing of the
strength. The terms with γ1 and γ2 correspond to the irreversible damage which
accumulates in time. Because the stiffness damage part Q directly relates to the fatigue
damage ∆Wfat. The rate of change of stiffness damage Q is equal to that of ∆Wfat.

d d ∆W fat
Q (t ) = W fat ( t ) ≈ (6-6)
dt dt ∆t

The change rate of Q in a strain controlled test can be expressed by substituting Equation
6-3 in Equation 6-6.

d π ⋅ S ( t ) ⋅ ε 02 ⋅ sin {ϕ ( t )}
Q (t ) = δ ⋅ (6-7)
dt ∆t

1
Since the loss modulus F ( t ) = S ( t ) ⋅ sin {φ ( t )} and the frequency f = , Equation 6-7
∆t
can be rewritten as follows:

135
Chapter 6 Application of Partial Healing Model on Strain Controlled Fatigue Tests

d π ⋅ ε 02 ⋅ S ( t ) ⋅ sin {φ ( t )}
Q =δ ⋅ = δ ⋅ π ⋅ f ⋅ ε 02 ⋅ F ( t ) (6-8)
dt ∆t

Where: ∆t : test duration in one cycle, [s];


f : frequency [Hz].

To simplify the equation, the following abbreviations are used:

α1* = δ ⋅ π ⋅ f ⋅ ε 02 ⋅ α1 ; α 2* = δ ⋅ π ⋅ f ⋅ ε 02 ⋅ α 2 (6-9)
γ 1* = δ ⋅ π ⋅ f ⋅ ε 02 ⋅ γ 1 ; γ 2* = δ ⋅ π ⋅ f ⋅ ε 02 ⋅ γ 2 (6-10)

Then Equation 6-4 and 6-5 can be rewritten as follows:


t
F ( t ) = F0 − ∫ F (τ ) α1*e− β (t −τ ) + γ 1*  ⋅ dτ (6-11)
0
t
G ( t ) = G0 − ∫ F (τ ) α 2*e − β ( t −τ ) + γ 2*  ⋅ dτ (6-12)
0

For a unit volume, the solutions of the local loss and storage stiffness modulus are given
by Equation 6-13 and 6-14, respectively.

F ( t ) = F0 e− Bt Cosh ( Ct ) + DSinh ( Ct )  (6-13)


 α 2* − Bt γ 2* 
(
G ( t ) = G0 − F0  e ⋅ Sinh ( Ct ) + * 1 − e − Bt ⋅ Cosh ( Ct ) + ESinh ( Ct )  
γ1
) (6-14)
C 
α + β + γ1
* *
β −B B − γ1
*

where B = 1 , C = B 2 − βγ 1* , D = and E =
2 C C

In Equation 6-9 and 6-10, π is a mathematical constant and the values of f and ε0 are from
the test conditions, but the parameter δ is an unknown constant value. It is not possible to
calculate the parameters α1, α2, γ1 and γ2 from the parameters α1*, α2*, γ1* and γ2*.
Therefore, in the following analysis, the item of δα1 is considered as one parameter. This
is also applied for the items of δα2, δγ1 and δγ2. From previous research, it is noticed that
the value of the parameter δα1 is always zero or small enough to be ignored for the UTC
and 4PB fatigue test.

Figure 6-2, Figure 6-3 and Figure 6-4 show the influence of the model parameters on the
evolution curves of the stiffness and phase angle. In each figure, only one parameter is
varied and all the other parameters are kept constant. The starting point of the curve is
determined by the parameters F0 and G0. The major influence of the parameter δα2 is on
the shape and steepness of the evolution curves. With a decrease of the parameter δα2, the
curve is more close to a straight line. Figure 6-3 shows that the stiffness is not influenced
by the parameter δγ1 and the phase angle grows slower with a larger δγ1. The parameter
δγ2 controls both the stiffness and phase angle. A larger δγ2 results in a steeper slope of
the evolutions curve.

136
Chapter 6 Application of Partial Healing Model on Strain Controlled Fatigue Tests

S-δα2=100 S-δα2=1000 S-δα2=2000 S-δα2=3000


φ-δα2=100 φ-δα2=1000 φ-δα2=2000 φ-δα2=3000
10000 100
δα2 δα2
8000 80

Stiffness [MPa]

Phase angle [°]


6000 60

4000 40

2000 20

0 0
0 10000 20000 30000 40000 50000 60000 70000
Number of cycles

Figure 6-2 Influence of δα2 on the evolution of the stiffness and phase angle

S-δγ1=40 S-δγ1=60 S-δγ1=80 S-δγ1=100


φ-δγ1=40 φ-δγ1=60 φ-δγ1=80 φ-δγ1=100
10000 100

δγ1
8000 80

Phase angle [°]


Stiffness [MPa]

6000 60

4000 40

2000 20

0 0
0 10000 20000 30000 40000 50000 60000 70000

Number of cycles

Figure 6-3 Influence of δγ1 on the evolution of the stiffness and phase angle

S-δγ2=50 S-δγ2=100 S-δγ2=150 S-δγ2=200


φ-δγ2=50 φ-δγ2=100 φ-δγ2=150 φ-δγ2=200
10000 80
δγ2
60
8000
Stiffness [MPa]

Phase angle [°]

40
6000
20

4000
0

δγ2
2000 -20
0 10000 20000 30000 40000 50000 60000 70000
Number of cycles

Figure 6-4 Influence of δγ2 on the evolution of the stiffness and phase angle

137
Chapter 6 Application of Partial Healing Model on Strain Controlled Fatigue Tests

Figure 6-5 gives the side view and the top view of the tensile strain distribution of a
cylindrical specimen (size 1.0) as obtained from a finite element simulation. In the model,
it is assumed that the cylindrical specimen behaves as a homogeneous material. The
viscoelastic properties of the mixture are described with Prony series representation and
are determined with the UT/C stiffness test. The boundary conditions are such that the
movement of the bottom surface of the cylindrical specimen is fully restricted; the
specimen is subjected to a tensile strain of 165 µm/m.

x
x=0 x=25

(a) (b)
Figure 6-5 Tensile strain distribution of a specimen in the UT/C test:
(a) side view; (b) top view

125
Glue area
Height of cylinder [mm]

100
x=0 x=12.5mm
75
x=20mm x=21mm Applied strain,
165 µm/m
x=22mm x=23mm
50
x=24mm x=25mm

25
Glue area
0
0 50 100 150 200 250
Tensile strain [µm/m]

Figure 6-6 Tensile strain along the height at different places from the center of the
cylinder (x=0) to the side (x=25)

Figure 6-6 gives the tensile strain distributions along the height at the different positions
from the centre to the side of the specimen (from x=0 to x=25). As shown, due to the side
effect, the tensile strain differs slightly from the applied strain near the upper and lower

138
Chapter 6 Application of Partial Healing Model on Strain Controlled Fatigue Tests

ends. To avoid the unwanted failure near the ends, extra glue was used between the
specimen and the steel caps to ensure sufficient bond and confinement. The thickness of
the glue is around 5 mm. Except for the ends of the specimen, the tensile strain is
uniformly distributed in the vertical and radial directions and the value is very close to
the target value.

In theory, the stress-strain distribution for the UT/C test is uniform throughout the
specimen. It means that the evolution of the stiffness modulus for a small unit volume is
equal to that of the whole specimen as measured during the test. So Equation 6-11 and 6-
12 can be directly used to fit the measured values for the homogenous fatigue test.

6.2.2 Determination of PH Model Parameters


To calculate the parameters F0 and G0, the initial stiffness and phase angle were measured
at the 50th loading cycle, as shown in Figure 6-7.

30000
UT/C_Size0.5@20°C UT/C_Size1.0@20°C
25000
Initial stiffness [MPa]

UT/C_Size1.5@20°C UT/C_Size1.0@5°C

20000

15000

10000

5000

0
0 50 100 150 200 250
Strain [µm/m]

(a) Initial stiffness


40
UT/C_Size0.5@20°C UT/C_Size1.0@20°C
35
UT/C_Size1.5@20°C UT/C_Size1.0@5°C
30
Phase angle [°]

25
20
15
10
5
0
0 50 100 150 200 250
Strain [µm/m]

(b) Initial phase angle


Figure 6-7 Initial stiffness and phase angle of the different UT/C specimen sizes

139
Chapter 6 Application of Partial Healing Model on Strain Controlled Fatigue Tests

At 20 ºC, the initial stiffness does not depend on the strain amplitude and the specimen
size. When the temperature reduces to 5 ºC, the initial stiffness of specimen size 1.0
becomes two times higher and the phase angle shows a reverse trend. The parameters F0
and G0 can be calculated from the initial stiffness and phase angle.

The model parameters δα1, δα2, δγ1, δγ2 and β are determined by minimizing the
differences between the measured and fitted values for S, φ, F (S×sinφ) and G (S×cosφ)
in the interval from N=10000 to N=NR (fatigue life determined by DER). This procedure
can be simply completed by the Solver function in Excel. In Figure 6-8, the fitted lines of
stiffness and phase angle correspond well with the measured values in the first and
second phase, because the PH model only describes the crack initiation phase and does
not take into account any plastic damage. It is believed that plastic effects are more
relevant in the crack propagation phase. In the crack initiation phase the fatigue damage
consists of the creation of micro-cracks, dislocations etc, which can be captured by the
proposed dWfat term.

Stiffness F(S*sinφ) G (S*cosφ)


Phase angle Model fit
12000 36
Phase I Phase II Phase III
10000 34
Stiffness [MPa]

Phase angle [°]


8000 32

6000 30

4000 28

2000 26

0 24
0 100000 200000 300000 400000 500000 600000 700000 800000
Number of cycles

Figure 6-8 Measured and fitted stiffness and phase angle of cylinder C-10-9 in the UT/C
fatigue test while using a strain level of 110 µm/m

The values of the model parameters are given in Table 6-1. It appears that in most cases
the values of the parameter δα1 are nil for the three specimen sizes. The other parameters
can be expressed as function of the applied strain. The relationships between the model
parameters and the strain levels are illustrated in Figure 6-9. The parameters δα2, δγ1 and
δγ2 can be expressed as a linear relationship with the strain level. The parameter β, the
time decay, is dependent on the squared value of the strain. The parameters δγ1 and δγ2
represent irreversible damage of the specimen. Its intersection point with the x-axis
means that irreversible damage would not occur when the strain level is lower than a
certain value. This indicates the existence of an endurance limit. Table 6-2 presents the
regression equations of the model parameters and the predicted endurance limit.

140
Chapter 6 Application of Partial Healing Model on Strain Controlled Fatigue Tests

Table 6-1 PH model parameters from regressions for the UT/C test
Specimen Sample Strain level β
δα1 δα2 δγ1 δγ2
size code [µm/m] [10-5 s-1]
C-3-18 88 0 964 11.4 25.5 63
C-3-25 117 0 1109 14.9 49.9 138
C-3-26 130 0 1333 19.5 56.8 171
Size 0.5 at C-3-9 136 0 1263 28.6 64.2 187
20 ºC C-3-24 138 0 1339 23.4 74.6 191
C-3-4 146 0 1704 32.5 98.1 217
C-3-11 168 0 1773 37.5 114.8 289
C-3-2 186 0 2042 43.3 145.4 397
C-10-12 68 0 755 2.1 8.9 24
C-10-10 84 0 892 9.5 24.1 45
C-10-9 110 0 1134 21.2 48.1 98
Size 1 at
C-10-8 131 0 1367 24.6 60.5 175
20 ºC
C-10-5 155 0 1655 39.0 99.9 244
C-10-11 172 0 1980 46.0 127.8 302
C-10-7 199 0 2142 55.9 171.1 402
C-14-1 68 0 627 3.8 12.1 37
C-13-1 89 0 1060 11.0 23.5 63
Size 1.5 at
C-11-2 115 0 1427 24.4 47.4 106
20 ºC
C-9-1 133 0 1383 36.5 91.7 138
C-12-2 161 0 1719 48.8 110.7 259
C-11-1 182 0 2009 63.2 149.2 318
C-15-4 69 0 606 2.1 18.9 49
C-15-11 87 0 578 7.4 40.8 77
C-16-5 94 0 685 0.9 55.6 61
Size 1 at
C-16-1 105 0 617 9.1 90.0 112
5 ºC
C-15-6 118 0 866 8.9 138.9 141
C-15-9 124 0 968 12.2 135.0 156
C-16-10 136 0 1089 21.9 198.4 226

From Figure 6-9 and Table 6-2, it can be seen that the parameter β does not change with
the specimen size and temperature and it only relates to the used strain amplitude. The
parameter δα2 is also independent of the specimen size, but the value of δα2 becomes
smaller with decreasing temperature at the same strain level. For the parameters δγ1 and
δγ2, the slopes of the curves are slightly influenced by the specimen size. The smaller
specimen has a steeper slope, but the difference between these three specimen sizes is not
significant at the same temperature. When the temperature reduces to 5 ºC, the value of
δγ1 becomes smaller and the value of δγ2 becomes larger at the same strain level. It
indicates that the more irreversible damage happens at a higher strain level or at a lower
testing temperature.

141
Chapter 6 Application of Partial Healing Model on Strain Controlled Fatigue Tests

4000
Size 0.5@20ºC Size 1.0@20ºC
Size 1.5@20ºC Size1.0@5ºC
3000

δα2
2000

1000

0
0 50 100 150 200 250
Strain level [µm/m]

(a) Parameter δα2 vs. strain amplitude


800
Size 0.5@20ºC Size 1.0@20ºC
Size 1.5@20ºC Size1.0@5ºC
600
β [10-5/s]

400

200

0
0 50 100 150 200 250
Strain level [µm/m]

(b) Parameter β vs. strain amplitude

250 δγ1_Size0.5@20ºC δγ2_Size0.5@20ºC δγ1_Size1.0@20ºC


δγ2_Size1.0@20ºC δγ1_Size1.5@20ºC δγ2_Size1.5@20ºC
δγ1_Size1.0@5ºC δγ2_Size1.0@5ºC
200
δγ 1 or δγ 2

150

100

50

0
0 50 100 150 200 250
Strain level [µm/m]

(c) Parameter δγ1 and δγ2 vs. strain amplitude


Figure 6-9 PH model parameters as functions of strain amplitude for the UT/C test

142
Chapter 6 Application of Partial Healing Model on Strain Controlled Fatigue Tests

The range of the endurance limit as presented in Table 6-2 is predicted based on the
trends of the parameters of δγ1 and δγ2. It can be seen that the predicted endurance limit
does not vary with specimen size and temperature for the UT/C fatigue test. The range of
the endurance limit is from 66 to 72 µm/m for the different specimen sizes and
temperatures.

Table 6-2 Functions of the PH model parameters for the UT/C test

Predicted
Specimen size Functions of the PH model parameters εlimit
[µm/m]

δγ1=0.36×(ε-66)
Size0.5@20 ºC δα2=11.0·ε-69.6 66~72
δγ2=1.17×(ε-72)

δγ1=0.41×(ε-63)
Size1@20 ºC δα2=11.2·ε-64.0 63~69
δα1=0, δγ2=1.32×(ε-69)
β=(1.01×10-7)·ε2 δγ1=0.51×(ε-66)
Size1.5@20 ºC δα2=10.8·ε+10.8 66~67
δγ2=1.36×(ε-67)

δγ1=0.24×(ε-68)
Size1@5 ºC δα2=7.81·ε-45 66~68
δγ2=2.39×(ε-66)

Table 6-2 shows that all the PH model parameters are related to the applied strain level.
The constants in the functions of the parameter δα2 are much lower than the values of the
parameter δα2 presented in Table 6-1. For simplification, it is assumed that the fitted line
of the parameter δα2 goes through the origin (0, 0). The endurance limits predicted by the
parameter δγ1 and δγ2 are much closer to each other and should be a constant value for
one mixture type at a certain test condition. Therefore the PH model parameters can be
expressed by the following equations.

β = β * ⋅ε 2 (6-15)
δα 2 = α 2** ⋅ ε (6-16)
δγ 1 = γ 1** ⋅ ( ε − ε lim it ) (6-17)
δγ 2 = γ 2** ⋅ ( ε − ε lim it ) (6-18)
Where: β*, α 2** , γ 1** , γ 2** : regression constants;
εlimit : predicted endurance limit, µm/m;
ε : applied strain level, µm/m.

143
Chapter 6 Application of Partial Healing Model on Strain Controlled Fatigue Tests

6.3 Application of PH Model on 4PB Test Results


6.3.1 Weighing Procedure for 4PB Test

In an inhomogeneous fatigue test, for example in the 4PB test, the stresses and strains
vary along the length and height of the beam specimen.

Figure 6-10a gives the horizontal strain distribution of a beam specimen in the 4PB test.
This finite element simulation is the same as the one mentioned in Chapter 5 (see Figure
5-8). Figure 6-10b shows that the horizontal strain at the bottom surface increases linearly
from the sides to the middle section. In the middle section the horizontal strain is constant.
As can be seen from Figure 6-10c, the horizontal strain varies linearly with the height.
The maximum strain that occurs at the surface in the middle section is the same as the
target value. It indicates that the stiffness varies at the different positions during the 4PB
fatigue test. The measured stiffness is not the local stiffness but a weighted overall
stiffness for the whole specimen. A weighing procedure is required to calculate the
weighted stiffness modulus per cross sectional area (along the height) and the weighted
overall stiffness modulus (along the length) from the local stiffness.

Height

Length

(a)
200 60
Horizontal strain [µm/m]

Height of beam [mm]

160 50

40
120
30
80
20
40
10
0 0
0 100 200 300 400 500 -200 -100 0 100 200
Length of beam [mm] Horizontal strain [µm/m]

(b) (c)
Figure 6-10 Schematic illustration of the 4PB test modeling (a), variation of the
horizontal strain with the length (b) and the linear variation of the horizontal strain with
the height at the centre point (c)

In the numerical analysis demonstrated by Pronk [Pronk, 2012], it is assumed that the
strain and material properties do not depend on the coordinate y (width of the beam).

144
Chapter 6 Application of Partial Healing Model on Strain Controlled Fatigue Tests

Because of symmetry, only a quarter part of the beam was considered (0<x<L/2;
0<z<H/2), as shown in Figure 6-11.

Neutral line

Height
x=L/2

z10
…….

j=1 j=2 j=3 ………. j=11 z=H/2

z2
z1
0 x1 x2 ……. x11 Length
x10
x=A

Figure 6-11 Schematic diagram of the numerical analysis

It is assumed that the origin of the coordinate system is at the location of the outer
support and the inner clamp is at the location x=A. The interval 0<x<A is divided into 10
sections and the interval x=A to x=L/2 is divided into 5 sections, because the weighted
stiffness is constant in the middle section (A<x<L/2), which is considered as being only
one section. The beam height from z=0 to z=H/2 is also divided into 10 sections with a
unit length of H/20. For a unit volume at the point (xj, zi), the local loss and storage
stiffness modulus at loading cycle n are given by Equation 6-19 and 6-20.

F ( x j , zi , n ) = F ( 0 ) e − Bt Cosh ( Ct ) + DSinh ( Ct )  (6-19)


α * γ* 
( )
G ( x j , zi , n ) = G ( 0 ) − F ( 0 )  2 e − Bt ⋅ Sinh ( Ct ) + 2* 1 − e − Bt ⋅ Cosh ( Ct ) + ESinh ( Ct )  
γ1
(6-20)
C 

Based on the strain distribution shown in Figure 6-10, the local strain distributions in a
beam are given by Equation 6-21 and 6-22.

 x j H − 2 zi 
ε { x j , zi } = ε 0  1 − ⋅  when 0<x<A and 0<z<H/2 (6-21)
 A H 
 H − 2 zi 
ε { x j , zi } = ε 0  1 −  when A<x<L/2 and 0<z<H/2 (6-22)
 H 

Where: ε0 : maximum strain in the midsection of the beam

Because of the variation of the bending strain level in the beam specimen, the evolutions
of the stiffness and the phase angle are different for each unit volume. This implies that
145
Chapter 6 Application of Partial Healing Model on Strain Controlled Fatigue Tests

the PH model parameters in Equation 6-19 and 6-20 are not constant but vary with the
local strain level. As discussed in section 6.2.2, all the parameters can be expressed as
function of the strain level (Equation 6-15 to 6-18). With the strain distribution and the
regression constants β*, α 2** , γ 1** and γ 2** , the evolution of the local stiffness with the
loading cycles was simulated.

Then the weighted loss and storage stiffness modulus per cross sectional area for a given
value of x are calculated by Equation 6-23 and 6-24. This function is based on a weighing
function with respect to the distance to the neutral axis. In the case of bending, the
w ⋅ h3
product of stiffness modulus (Smix) times the moment of inertia ( ) is relevant.
12

  H
10
   i   i − 1  
3 3

F { x, n} = ∑  F  x, i, n     −    (6-23)
i =1   20    10   10   
 
10 
   i   i − 1  
3 3
 H
G { x, n} = ∑ G  x, i, n     −    (6-24)
i =1   20    10   10   
 

Where: F { x, n} G { x, n} : weighted loss and storage stiffness modulus per cross


area at location x at the cycle n, [MPa];

The weighted stiffness modulus of each section j on the interval 0<x<A is the mean value
of the weighted stiffness at the borders of the section j. The ratios of these weighted
stiffness and the initial weighted stiffness are denoted as α j ,loss and α j , storage (Equation 6-
25 and 6-26).

Fj ( n ) =  F ( x j , n ) + F ( x j −1 , n )  / 2 = α j ,loss ⋅ F ( 0 ) (6-25)
G j ( n ) = G ( x j , n ) + G ( x j −1 , n )  / 2 = α j ,storage ⋅ G ( 0 ) (6-26)

Where: Fj ( n ) G j ( n ) : weighted loss and storage stiffness modulus of each


section j, [MPa];
α j ,loss , α j , storage : ratio of weighted loss and storage stiffness and the
initial weighted stiffness [Hz].

In the middle section (A<x<H/2), the weighted loss and storage stiffness per section are
constant and denoted as:

F11 ( n ) = α11,loss ⋅ F ( 0 ) = α11,loss ⋅ S0 sin (ϕ0 ) (6-27)


G11 ( n ) = α11, storage ⋅ G ( 0 ) = α11, storage ⋅ S0 cos (ϕ0 ) (6-28)

146
Chapter 6 Application of Partial Healing Model on Strain Controlled Fatigue Tests

Vb,loss and Vb,storage are defined as the deflections corresponding to the loss and storage
stiffness, respectively. By ignoring the influence of any mass on the phase lag, the
bending deflections as a result of loss stiffness along the length are determined from the
differential equations (Equation 6-29 and 6-30) for a slender beam.

d2 P
Fj ( n ) ⋅ I ⋅ 2
Vb ,loss ( x, n ) = − ⋅x when 0<x<A (6-29)
dx 2
d2 P
Fj ( n ) ⋅ I ⋅ Vb ,loss ( x, n ) = − ⋅A when A<x<H/2 (6-30)
dx 2 2

Where: Vb,loss ( x, n ) : bending deflection for loss stiffness at location of x at the


number of cycle n, [mm];
P : amplitude of applied force [N].

Because similar equations are also valid for the storage stiffness, in the following analysis
the part of the loss stiffness is only mentioned. The general solution for the sections from
j=1 to 10 is given by Equation 6-31.

 1 x
3
x 
Vb,loss ( x, n ) = β loss  −   + C j ,loss   + D j ,loss  (6-31)
 6α j ,loss L L 
 

Where: βloss, Cj,loss, Dj,loss : constants, [mm];

The boundary conditions are that both the deflection and its first derivative are
j
continuous at the cross sections of x j = A . The recursion formulas for the constant
10
Dj,loss are given by the following equations:

 D1,loss = 0

 1 
3
 1 j  1 (6-32)
 D j +1,loss = D j ,loss + 3  10 L A  − 
    α j ,loss α j +1,loss 

Over the interval of A<x<H/2, the bending deflection has a parabolic shape as shown in
equation 6-33.

 A  
2
1 x 1 x
Vb ,loss ( x, n ) = β loss    −   +   + D  (6-33)
 L   2α11,loss  L  2α11,loss 
11, loss
L 

At the cross section x=A both the deflection Vb,loss and its first derivative have to be
continuous leading to a relation between the constants D10,loss and D11,loss.

147
Chapter 6 Application of Partial Healing Model on Strain Controlled Fatigue Tests

1  j
9
  1
3
1 
D10 ,loss = ∑  A  −  (6-34)
3 j =1  10 L   α j ,loss α j +1,loss  

 A  
3
 A  1 1
D11,loss =    D10,loss −    −   (6-35)
 L   
 L   2α11,loss 3α10,loss  

The amplitude of the maximum deflection at x=L/2 is given by Equation 6-36.

 A  1 
Vb ,loss ( n ) = βloss    + D11,loss  (6-36)
 L   8α11,loss 

The expressions of the weighted loss, storage and overall modulus for the whole beam
are given by Equation 6-37, 6-38 and 6-39.

 1 1  A 2 
 −   
 8 6  L  
F ( n) =  ⋅ S0 sin (ϕ0 ) (6-37)
 1 
 
 8α11,loss + D11,loss 
 1 1  A 2 
 −   
 8 6  L  
G ( n) = ⋅ S0 cos (ϕ0 ) (6-38)
 1 
 
 8α11,storage + D11,storage 

( ) ( )
2 2
S {n} = F ( n) + G ( n) (6-39)

The model parameters are determined in such a way that the calculated weighted overall
modulus stays consistent with the measured stiffness of the beam.

6.3.2 Determination of PH Model Parameters


Similar to the procedure for analyzing the UT/C test results, the initial stiffness and phase
angle were used to calculate the initial values for the parameter F0 and G0. By means of
the weighing procedure, the weighted overall storage modulus, loss modulus, complex
modulus and phase angle were fitted with the measured results by varying the regression
constants β*, α2**, γ1**, γ2** and εlimit. An example of the regression fits is shown in Figure
6-12.

148
Chapter 6 Application of Partial Healing Model on Strain Controlled Fatigue Tests

12000 35
Stiffness F(S*sinφ )
G(S*cosφ ) Phase angle
10000 Model fit 30

Stiffness [MPa]

Phase abgle [º]


8000 25

6000 20

4000 15

2000 10

0 5
0 50000 100000 150000 200000 250000
Number of cycles

Figure 6-12 Measured and calculated 4PB results for a strain level of 179 µm/m

Table 6-3 PH model parameters for the specimen size 0.5 at 20 ºC


β*
Sample Strain level εlimit
α [10 ] [10-8 s-1] γ 1** [106]
** 6
γ 2** [106]
code [µm/m] 2
[µm/m]
B-19-5 100 11.89 7.55 0.38 1.36 73
B-14-5 119 11.89 8.20 0.32 1.38 66
B-19-2 140 11.89 8.58 0.41 1.35 80
B-14-6 160 11.89 8.96 0.39 1.36 72
B-19-6 179 11.89 9.91 0.41 1.35 69
B-19-4 201 11.89 10.17 0.44 1.34 88
Mean 11.89 8.89 0.39 1.36 73.5
St.dev. 0.0003 1.01 0.04 0.01 5.8

Table 6-4 PH model parameters for the specimen size 1.0 at 20 ºC


Strain β*
Sample εlimit
level α [10 ]
**
2
6
[10-8 s-1] γ 1** [106] γ 2** [106]
code
[µm/m] [µm/m]
B-15-2 87 11.89 8.87 0.38 1.40 61
B-16-1 105 11.89 10.05 0.34 1.41 61
B-16-2 124 11.89 10.70 0.37 1.40 61
B-17-4 142 11.89 10.12 0.40 1.39 60
B-16-4 161 11.89 9.51 0.45 1.37 66
B-17-3 180 11.89 12.04 0.40 1.39 69
Mean 11.89 10.22 0.39 1.39 63.1
St.dev. 0.001 1.09 0.04 0.01 3.7

149
Chapter 6 Application of Partial Healing Model on Strain Controlled Fatigue Tests

Table 6-5 PH model parameters for the specimen size 1.5 at 20 ºC


Strain β*
Sample εlimit
level α 2** [10 ]
6
[10-8 s-1] γ 1** [106] γ 2** [106]
code
[µm/m] [µm/m]
B-20-2 74 11.86 11.91 0.58 1.30 57
B-23-1 88 10.89 8.85 0.51 1.41 60
B-22-2 93 11.89 10.11 0.31 2.81 61
B-22-1 90 11.89 11.23 0.30 2.81 61
B-20-1 103 11.89 10.21 0.53 1.21 61
B-21-2 119 11.87 12.93 0.47 2.54 63
Mean 11.71 10.87 0.45 2.02 60.3
St.dev. 0.40 1.45 0.12 0.79 1.8

Table 6-6 PH model parameters for the specimen size 1.0 at 5 ºC


Strain β*
Sample εlimit
level α [10 ]
**
2
6
[10-8 s-1] γ 1** [106] γ 2** [106]
code
[µm/m] [µm/m]
B-7-2 89 5.29 5.02 0.27 2.88 53
B-15-3 105 5.27 13.29 0.06 2.91 69
B-18-1 109 6.29 7.56 0.46 3.56 84
B-15-4 116 5.58 8.50 0.22 3.19 49
B-18-4 124 5.29 12.38 0.0011 2.92 60
B-15-1 135 5.29 11.71 0.08 3.08 65
Mean 5.50 9.74 0.18 3.09 63.2
St.dev. 0.40 3.23 0.17 0.26 12.5

Table 6-3, Table 6-4, Table 6-5 and Table 6-6 present the regression constants β*, α2**,
γ1**, γ2** and εlimit at each strain level for all the specimen sizes and their average values
and standard deviations.

Similar to the UT/C test, for the 4PB test, the parameter δα2 does not change with the
specimen size and becomes lower with a decrease in temperature. The average values of
the constant β* for the different specimen sizes and temperatures are very close to each
other. It means that the parameter β is independent of the specimen size and temperature
and is only related to the used strain amplitude. For the parameters δγ1 and δγ2, the slopes
of the curves slightly increase with the specimen size. When the temperature reduces to 5
ºC, the value of γ1** becomes smaller and the value of γ2** becomes larger. The standard
deviations of these regression constants are relatively larger for the size 1.5 specimens
and those tested at the low temperature of 5 ºC.

The average values of the regression constants and the predicted endurance limits
obtained by means of the UT/C and 4PB test are compared in Figure 6-13, Figure 6-14,
Figure 6-15, Figure 6-16 and Figure 6-17, respectively. It can be seen that the values of
the constants α2**, β*, γ1** obtained from the 4PB test are not very much different from
those obtained from the UT/C test. However, the constant γ2** and the predicted

150
Chapter 6 Application of Partial Healing Model on Strain Controlled Fatigue Tests

endurance limit for a unit volume vary with the 4PB specimen size. The larger specimen
has a lower endurance limit. All the data are listed in Table 6-7.
18
4PB UT/C
15

12
α2** [106]
9

0
Size 0.5@20ºC Size 1.0@20ºC Size 1.5@20ºC Size 1.0@5ºC

Figure 6-13 Comparison of the regression constant α2** for 4PB and UT/C test

18
4PB UT/C
15

12
β* [10-8]

0
Size 0.5@20ºC Size 1.0@20ºC Size 1.5@20ºC Size 1.0@5ºC

Figure 6-14 Comparison of the regression constant β* for the 4PB and UT/C test

4PB UT/C
0.8
γ 1** [106]

0.6

0.4

0.2

0
Size 0.5@20ºC Size 1.0@20ºC Size 1.5@20ºC Size 1.0@5ºC

Figure 6-15 Comparison of the regression constant γ1** for the 4PB and UT/C test

151
Chapter 6 Application of Partial Healing Model on Strain Controlled Fatigue Tests

3.5

3 4PB UT/C

2.5

γ 2** [106]
2

1.5

0.5

0
Size 0.5@20ºC Size 1.0@20ºC Size 1.5@20ºC Size 1.0@5ºC

Figure 6-16 Comparison of the regression constant γ2** for the 4PB and UT/C test

100
4PB UT/C
80
εlimit [µm/m]

60

40

20

0
Size 0.5@20ºC Size 1.0@20ºC Size 1.5@20ºC Size 1.0@5ºC

Figure 6-17 Comparison of the predicted endurance limit for the 4PB and UT/C test

Table 6-7 Comparison of the predicted endurance limit for 4PB and UT/C test
β*
Test εlimit
Specimen size α [10 ]
**
2
6
[10-8 s-1] γ 1** [106] γ 2** [106]
type
[µm/m]
Size 0.5@20 ºC 11.0 10.1 0.36 1.17 69
Size 1.0@20 ºC 11.2 10.1 0.41 1.32 66
UT/C
Size 1.5@20 ºC 10.8 10.1 0.51 1.36 67
Size 1.0@5 ºC 7.81 10.1 0.24 2.39 67
Size 0.5@20 ºC 11.89 8.89 0.39 1.36 74
Size 1.0@20 ºC 11.89 10.22 0.49 1.39 63
4PB
Size 1.5@20 ºC 11.71 12.21 0.45 2.02 60
Size 1.0@5 ºC 5.50 9.74 0.18 3.09 63

152
Chapter 6 Application of Partial Healing Model on Strain Controlled Fatigue Tests

6.4 Correlation between UT/C and 4PB Fatigue Test Results

In this section the correlation between the UT/C and 4PB test results for the size 1.0
specimens at 20 ºC is investigated. Due to the inhomogeneous stress-strain distribution in
the 4PB test, the stiffness of the beam specimen decreases at different rates along the
length and height. With the PH model parameters, the stiffness evolutions from different
locations can be calculated. Figure 6-18 shows the comparison of the weighted stiffness
evolution in different cross sectional areas.

B-16-1@105µm/m
9000
x=0.1A
x=0.3A
W eig h ted stiffn ess [M P a]

x=0.5A
8000
x=0.7A

7000
x=0 x=A x=L-A x=L

6000 x=0.9A

overall
5000
A<x<L-A
4000
0 500000 1000000 1500000 2000000 2500000
Number of cylces

Figure 6-18 Weighted stiffness per cross sectional area and overall weighted stiffness.

It is clear that the stiffness in the midsection (A<x<L-A) decreases dramatically with the
number of loading cycles. The fatigue damage becomes less and less from the inner
clamps (x=A) to the outer clamps (x=0). Almost no damage develops near the outer
clamps. Taking into account the stress-strain distribution, the weighted overall stiffness is
calculated, which is higher than the weighted stiffness in the midsection. Along the
height direction, the strain reaches the maximum value at the surface and decreases
linearly from the surface area to the neutral line at which the strain level becomes zero.
During testing, only the surface area in the midsection is always subjected to the target
strain. It means that the stiffness evolution at the surface in the midsection of a beam
should be comparable to the results from the UT/C test.

Three pairs of the cylinders and beams are selected and for each pair the applied strain on
the cylinder is similar to the pure bending strain at the surface in the midsection of the
beam specimen. Figure 6-19 gives the comparison between the measured stiffness of the
cylinder, the weighted overall stiffness of the beam and the calculated stiffness at the
surface in the midsection of the beam.

153
Chapter 6 Application of Partial Healing Model on Strain Controlled Fatigue Tests

12000
C-10-10@84µε
10000 B-15-2@87µε_Surface
B-15-2@87µε_overall

Stiffness [MPa]
8000

6000

4000

2000

0
0 1000000 2000000 3000000 4000000 5000000
Number of cylces

12000
C-10-9@110µε
10000 B-16-1@105µε_Surface
B-16-1@105µε_overall
8000
Stiffness [MPa]

6000

4000

2000

0
0 500000 1000000 1500000 2000000 2500000
Number of cycles

12000
C-10-5@155µε
10000 B-16-4@161µε_Surface
B-16-4@161µε_overall
Stiffness [MPa]

8000

6000

4000

2000

0
0 100000 200000 300000 400000 500000
Number of cylces

Figure 6-19 Comparison of stiffness evolutions obtained from UT/C and 4PB test.

From these figures, it can be seen that the local stiffness at the surface of the beam drops
more rapidly compared to the weighted overall stiffness, which means that the most
serious fatigue damage occurs at the beam surface. It is noted that the stiffness evolution
obtained in the UT/C test is consistent with the one valid for the beam surface. For the

154
Chapter 6 Application of Partial Healing Model on Strain Controlled Fatigue Tests

UT/C fatigue test, the classical fatigue life can be easily determined by the measured
stiffness, but for the 4PB test the problem is that the changes in the stiffness are not
uniform inside the specimen. The fatigue lives determined at the different locations are
also different.

Figure 6-20 gives a comparison of the fatigue lines determined from the overall stiffness
and the local stiffness on the surface of the beam. As expected, the fatigue line based on
the local surface stiffness of the beam is lower than that based on the overall weighted
stiffness and it shows a good agreement with the fatigue line obtained from the UT/C
fatigue test.

1.E+07

1.E+06
Nf,50

1.E+05 UT/C_Nf,50
4PB_overall
4PB_surface
1.E+04
10 100 1000
Strain [µm/m]

Figure 6-20 Comparison of classical fatigue life determined by different stiffness curves.

Table 6-8 Regression coefficients of the fatigue lines


Material coefficients
R2
k b
14
UT/C_size1.0 5.7×10 -4.43 0.995
4PB_ size1.0_overall 5.3×1014 -4.20 0.984
14
4PB_ size1.0_surface 2.9×10 -4.25 0.989

6.5 Conclusions
In this chapter, the PH model was applied on the strain-controlled UT/C and 4PB fatigue
test results. The main findings, remarks and conclusions are summarized below:

Application of the PH model on the UTC fatigue test:

1. The PH model describes the evolutions of the complex modulus and phase angle for
a unit volume in a strain-controlled fatigue test. The solution can be directly applied
on the UT/C test results. By adjusting the model parameters, the PH model provides

155
Chapter 6 Application of Partial Healing Model on Strain Controlled Fatigue Tests

good simulations for the evolutions of the complex stiffness and phase angle in phase
I and phase II (see figure 6-8).
2. All the model parameters can be expressed as function of the applied strain level.
The parameter β (time decay) is independent of the specimen size and temperature
and is only related to the square of the used strain amplitude. The parameter δα1
(reversible damage in loss modulus) is always nil or close to nil in any case. The
parameter δα2 (reversible damage in storage modulus) does not change with the
specimen size, but becomes smaller with a decrease in temperature. For the
parameters δγ1 and δγ2 (irreversible damage in loss and storage modulus), the
difference between the different specimen sizes is not significant. When the
temperature reduces to 5 ºC, the value of δγ1 becomes smaller and the value of δγ2
becomes larger at the same strain level. This indicates that in the UT/C test, the
healing ability of the asphalt mixture decreases at a lower temperature and more
irreversible damage happens at a higher strain level or at a lower testing temperature.
The relationships between the PH model parameters and the applied strain level can
be expressed by Equations 6-15 to 6-18.
3. The trends of the values of the parameter δγ1 and δγ2 indicate the existence of an
endurance limit. The predicted endurance limit obtained by the UT/C test is not
influenced by the specimen size and temperature and can therefore be considered as
a material property. The range of the endurance limit for the used mixture is from 66
to 69 µm/m at 20 ºC.

Application of the PH model on the 4PB fatigue test:

1. In the 4PB test, the strain is not uniformly distributed throughout the length and
height of the beam specimen. The functions of the PH model parameters are used to
calculate the local stiffness for each unit volume. By a weighing procedure, the
weighted overall stiffness was calculated and by adjusting the regression constants
the weighted overall stiffness was fitted to the measured stiffness.
2. The average values of the regression constants α2**, β*, and γ1** obtained from the
4PB test are close to those obtained from the UT/C test. However, the constant γ2**
and the predicted endurance limit for a unit volume are influenced by the specimen
size. The larger specimen has a larger γ2** value and a lower endurance limit.
3. For the specimen size 1.0, the functions of the PH model parameters obtained from
the UT/C and 4PB test are similar to each other. It indicates that the 4PB test results
can be predicted by means of the PH model parameters determined from the UT/C
test results.
4. By means of the PH model, the local stiffness at different parts of the beam can be
calculated. The most serious fatigue damage occurs at the surface in the midsection of
the beam. The stiffness evolution of this area is comparable with the UT/C fatigue
test results when the pure bending strain on the beam surface is equal to the strain
applied to the cylinder. The fatigue line based on the local surface stiffness of the
beam is in good agreement with the fatigue line obtained from the UT/C fatigue test.

156
Chapter 6 Application of Partial Healing Model on Strain Controlled Fatigue Tests

References

Li, N., Molenaar, A.A.A., Pronk A.C., van de Ven, M.F.C. and Wu, S., Effect of
Specimen Size on Fatigue Behavior of Asphalt Mixture in Laboratory Fatigue Tests,
Proceedings of the 7th RILEM International Conference on Cracking in Pavements, Delft,
the Netherlands, pp 827-836, 2012a.

Li, Ning, Molenaar, A.A.A., Pronk A.C., van de Ven, M.F.C., Investigation into the Size
effect on Four Point Bending Fatigue Tests, Proceedings of the 3rd Workshop on 4PB,
Davis, California, 2012b.

Li, Ning, Molenaar, A.A.A., Pronk A.C., van de Ven, M.F.C. and Wu, S., Comparison of
Uniaxial and Four Point Bending Fatigue Tests for asphalt mixtures. The Transportation
Research Board (TRB) 92nd Annual Meeting. Washington, D.C. January 13-17, 2013.

Pronk, A. C. and Hopman P. C., Energy dissipation: the leading factor of fatigue,
Proceedings of the Conference on the United States Strategic Highway Research
Program, pp. 255–267, 1991.

Pronk, A.C., Analytical description of the heat transfer in an asphalt beam, tested in the 4
point dynamic bending apparatus, DWW report W-DWW-96-006, 1996.

Pronk, A. C., Comparison of 2 and 4 point fatigue tests and healing in 4 point dynamic
test based on the dissipated energy concept. In: Proceedings of the 8th International
Conference on Asphalt Pavements, Seattle, pp. 987–994, 1997.

Pronk, A.C., Partial healing in fatigue tests on asphalt specimen, Road Materials and
Pavement Design, vol. 4, n. 4, pp 433-445, 2001.

Pronk, A.C., Partial Healing, A new approach for the damage process during fatigue
testing of asphalt specimen, Proceedings of the Symposium on Mechanics of Flexible
Pavements, ASCE, Baton Rouge, 2006.

Pronk, A.C. and Molenaar, A.A.A., The Modified Partial Healing Model used as a
Prediction Tool for the Complex Stiffness Modulus Evolutions in Four Point Bending
Fatigue Tests based on the Evolutions in Uni-Axial Push-Pull Tests, Proceedings of the
11th Int. Conf. on Asphalt Pavements, Nagoya, Japan, 2010.

Pronk, A.C., Description of a procedure for using the Modified Partial Healing model
(MPH) in 4PB test in order to determine material parameters. Proceedings of the 3rd
Workshop on 4PB, Davis, USA, 2012.

Rowe, G.M., Application of the dissipated energy concept to fatigue cracking in asphalt
pavements, PhD thesis, University of Nottingham, Nottingham, England. 1996.

157
Chapter 6 Application of Partial Healing Model on Strain Controlled Fatigue Tests

van Dijk, W. Moreaud, H., et al. 1972 The Fatigue of Bitumen and Bituminous Mixes,
3th International Conference on Structural Design of Asphalt Pavements (ISAP), London,
Volume 1.

158
Chapter 7 Monotonic Uniaxial Tension and Compression Tests

Chapter 7 Monotonic Uniaxial Tension and


Compression Tests

7.1 Introduction
As mentioned in Chapter 3, representing fatigue relations in a 3-Dimensional state of
stress rather than a 1-dimensional state of stress was defined as one of the objectives of
this research. This was considered necessary since in reality a unit volume of pavement
material is always subjected to a 3D state of stress. Therefore it was decided to determine
the yield surface for the asphalt mixture, similar to what is done for granular materials.

The laboratory tests including triaxial tests, monotonic uniaxial compression (MUC) and
monotonic uniaxial tension (MUT) tests are required to determine the yield surface of the
asphalt mixture. In this research only the MUC and MUT test are used to develop yield
surfaces. The response to the applied load in terms of stress and strain (axial and radial
strain) is measured during the tests, when the cylindrical specimens were subjected to
uniaxial loading. Because the response of an asphalt mixture is strongly influenced by
loading time and temperature, both the MUC and MUT tests are conducted at different
temperatures and strain rates. The set-ups of the MUT and MUC test were designed and
built at the Delft University of Technology, Road and Railway Engineering Section. A
detailed description of these two tests is given in Section 7.2 and 7.4. To reduce the
number of tests, the central composite rotatable design method was used for optimization
of the experimental program. Section 7.3 briefly describes this experimental design
technique and gives the test conditions. The experimental results are shown in Section 7.6

7.2 Test Equipment


7.2.1 Monotonic Uniaxial Compression Test
The uniaxial compression test set-up (see Figure 7-1) was designed and built by
considering several influence factors on the state of stress and deformation, such as
specimen alignment, frame stability, temperature effect, boundary friction, etc [Erkens
and Poot, 2000].

The compression test set-up consists of a 3D-space frame in which an MTS 150 kN
hydraulic actuator is mounted. The actuator is rigidly connected to the upper loading
plate. The frame itself is placed on an elastically supported concrete block. The load is
transmitted from the actuator to the specimen through two steel plates placed at the top
and bottom of the specimen. The bottom and top plate are kept parallel by using three
guidance bars (Φ16 mm) made of Fortal (a strong aluminium alloy), which are connected
to the bottom plate and pass through linear bearings in the top plate [Erkens, 2002].
Without any precautions at the contact surface between specimen and the loading plates,
the radial deformation would be restrained due to the fact that the plates and the specimen
have a different ν/E ratio. The resulting friction would act as a confinement for the top
159
Chapter 7 Monotonic Uniaxial Tension and Compression Tests

and bottom of the specimen, causing the well-known barrel-shape of specimens in


compression. To avoid these stress concentrations, a friction reduction system is applied
to the top and bottom ends of the specimen. The friction reduction system consists of two
thin steel plates and two pieces of rubber, which are greased with a soft soap at both sides.
The specimen is placed between two of these metal-soap-rubber-soap sandwiches.

Frame Hydraulic LVDT’s


Actuator

String

Temperature
cabinet Load cell

Extensometer

Platen Extension

(A) (B)
Figure 7-1 Complete compression test setup (A) and a close-up of a specimen inside of
the temperature cabinet (B)

An insulated cabinet with dimensions 0.6×0.5×0.6 m is placed within the frame, which
allows the tests to be performed at temperatures in the range of 0 to 45 °C with an
accuracy of ± 0.5 °C. The insulated cabinet is a sandwich construction of wood and foam.
The inside of the cabinet is covered with aluminum/plastic insulation foil. A uniform
temperature distribution in the specimen was realized by controlling the airflow rate in
the chamber. During the tests, the temperature of the plates and the air are monitored.

The force data are measured via the MTS load cell, which is positioned between the top
plate and the actuator. The axial deformations are recorded by three external
displacement transducers (LVDT’s) placed vertically around the specimen. The radial
deformations at the middle of the specimen are registered by two circumferential
measurement systems (a string and an extensometer). The range of the axial LVDT’s is ±
20 mm. The string and the extensometer have a range of ± 150 mm and ± 3.75 mm,
respectively. The purpose of the extensometer is to enable accurate radial deformation
measurements to be made in the initial stages of the test. When the extensometer is out of
range, the radial deformations are measured by the string. The measurement systems are
connected to a PC-based data acquisition system, which produces a single ASCII output
file for each test. The measured data are captured at sampling rates ranging from 1 to
1000 Hz.

160
Chapter 7 Monotonic Uniaxial Tension and Compression Tests

7.2.2 Monotonic Uniaxial Tension Test


The tension test set-up consists of a closed temperature cabinet with a 50 kN hydraulic
actuator inside. The actuator is connected rigidly to the bottom of the temperature
chamber. In the temperature cabinet the specimen is placed in a rigid framework that can
resists the high forces that occur during the test without deforming. In the test setup, the
specimen is placed between three hinges to ensure that the specimen is subjected to pure
uniaxial tension. Two hinges are placed above and one under the specimen to prevent
bending moments to occur in the specimen. The specimens are glued to the top and
bottom end caps using a 2-component fast curing adhesive, X60. In this case, the radial
deformation will cause stress concentrations near the specimen ends. In order to provide
confinement at the ends of the specimen, PVC rings were glued around the specimen
ends and the caps to prevent specimens from cracking near the ends.

Hinge
Load cell

LVDT’s

Radial
extensometer
Plastic ring

Hinge Hydraulic
Actuator

Figure 7-2 Tension test set-up (left) and a close-up of a specimen inside the temperature
cabinet (right)

The force is measured by means of a load cell, which is positioned between the two
hinges above the specimen. An internal 407 MTS controller is used to impose the
required controlled deformation rate. The axial deformation is registered by means of
three displacement transducers (LVDT’s). These LVDT’s are fixed in an aluminium ring
which is placed around the steel cap at the bottom of the specimen. On top of the
specimen, the three LVDT’s were positioned such that they touch a second aluminium
ring, which is placed around the steel cap on top of the specimen. To obtain an accurate
axial displacement curve, the LVDT’s with a range of ± 1 mm are used for measurements
at low temperatures (below 20 °C and high strain rates at 20 °C) and the LVDT’s with a
range of ± 5 mm are used for high temperatures (above 20 °C and low strain rates at 20
°C). The measurement systems are connected to a PC-based data acquisition system,
which produces a single ASCII output file for each test. Moreover, an oscilloscope is
used as a backup for the measurements. A close-up of the tension test set-up with an
instrumented specimen is shown in Figure 7-2.

161
Chapter 7 Monotonic Uniaxial Tension and Compression Tests

7.3 Test Condition


7.3.1 Central Composition Rotatable Design
To investigate the influence of temperature and strain rate, the tests should be conducted
at several temperatures and strain rates with a number of repetitions based on a traditional
experimental design. If the tests are done at 3 temperatures and 4 strain rates and repeated
4 times for each combination, at least 48 specimens are needed for MUC or MUT test.
Therefore, it was decided to use “the central composite rotatable design technique”
[Robinson 2000] for the design of the test conditions of the MUC and MUT tests. The
procedure is quite simple and fully described by Jansen [2002], Medani [2006] and
Muraya [2007], as follows:
1) Determination of the range of the independent variables
In this study, the independent variables are the strain rate denoted by ( εɺ ) in the range of
εɺmin to εɺmax and the temperature denoted by (T) in the range of Tmin to Tmax. The
dependent variables are the compressive and the tensile strength.
Table 7-1 Coded test condition
test εɺcoded Tcoded
1 −1 −1
2 +1 −1
3 −1 +1
4 +1 +1
5 −Ψ 0
6 +Ψ 0
7 0 −Ψ
8 0 +Ψ
9~13 0 0

8

3 4
+1

Tcoded 0 5 9-13 6
(0, 0)

−1
1 2
−Ψ
7
−Ψ −1 0 +1 +Ψ
εɺcoded
Figure 7-3 Coded test conditions for the central composite rotatable design

162
Chapter 7 Monotonic Uniaxial Tension and Compression Tests

2) Each condition of the experiment to be conducted is expressed in coded terms in


Table 7-1 and graphically in Figure 7-3. Figure 7-3 shows that 5 tests are repeated in
the centre of the range of independent variables (test 9~13) and 8 tests are at the
rotatable positions around this centre.
3) Tests 1~8 are called star points and tests 9~13 are called centre points. Any value can
be chosen for Ψ. If the value of Ψ is set at 2 , there may be any number of replicated
points in the centre. However, it is recommended by Diamond [2001] to use 5 central
points with 2 independent variables.
4) To scale the coded terms of Table 7-1 into values within the range of variables, Ψ is
set equal to half the range of the variables and scaling factors for the tests follow from:

1
2 × εɺscaling = × ( εɺmax − εɺmin ) (7-1)
2
1
2 × Tscaling = × (Tmax − Tmin ) (7-2)
2
Where: εɺscaling : scaling factor for the strain rate;
Tscaling : scaling factor for the temperature;
εɺmin : minimum strain rate, %/s;
εɺmax : maximum strain rate, %/s;
Tmin : minimum test temperature, °C;
Tmax : maximum test temperature, °C;

Finally the experimental values of the strain rate εɺ and temperature T follow from:
1
εɺscaling = εɺcoded × εɺscaling + × ( εɺmax − εɺmin ) (7-3)
2
1
Tscaling = Tcoded × Tscaling + × (Tmax − Tmin ) (7-4)
2

7.3.2 Test Conditions for MUC and MUT Tests


The MUC and MUT tests are performed under displacement controlled conditions by
applying a monotonic increasing axial displacement (constant deformation rate) until
complete annihilation of the strength. The chosen range of strain rate and temperature are
0.001 ~ 4 %/s and 5 ~ 35 ºC, respectively. According to the central composite design, the
monotonic tension test and compression test are conducted at the target test conditions
shown in Figure 7-4 and tabulated in Table 7-2. Two extra tests which are not part of the
central composite design are added at 4 and 38ºC to check the accuracy of the
experimental design for failure test conditions outside the range of the experimental
design.

163
Chapter 7 Monotonic Uniaxial Tension and Compression Tests

Strain rate [% /s]


3

0
-5 5 15 25 35 45
Tem pe ra ture [ºC]

Figure 7-4 Test conditions of the MUC and MUT test

Table 7-2 Test conditions of the MUC and MUT test


MUC test MUT test
Temperature strain rate Temperature strain rate
[ºC] [%/s] [ºC] [%/s]
4 1 5 0.01, 0.05, 2
5 2.001 9.4 0.05, 0.587, 3.416
9.4 0.587, 3.414 12.9 0.587, 2
20 0.01, 2 (repeat 5 times), 4 20 0.01, 2 (repeat 5 times), 4
30.6 0.587, 3.414 30.6 0.587, 3.416
35 2.001 35 2
38 3 38 3

7.4 Test Procedure


7.4.1 Test Procedure for MUC Tests
 The specimen is placed between the two metal-soap-rubber-soap sandwiches.
 The specimen with friction reduction system is placed between the top and bottom
loading plates in the temperature cabinet and fitted with the radial measurement
systems. One must ensure that all three LVDT’s, string and extensometer are in the
measurement range.
 Then the temperature cabinet is closed in order to allow the temperature of top and
bottom plate to reach the target value. A minimum period of 3 hours is used for
temperature stabilization.
 Before starting the test, a small pre-load (0.3kN) is applied on the specimen to
prevent undesirable moments that can lead to erroneous results [Erkens and Poot,
2000].

164
Chapter 7 Monotonic Uniaxial Tension and Compression Tests

 In the displacement-controlled mode, the load is applied such that a constant


displacement rate is achieved, until the specimen breaks. The reported displacement
rate is based on the average value of the three external axial LVDT’s and is fed
back to the system.

7.4.2 Test Procedure for MUT Tests


 Two steel caps are glued to the top and bottom surface of the specimen. The steel
caps are 30 mm thick and have a diameter of 80 mm. The caps have to be glued
parallel to each other to ensure a state of pure tension during the test. When the caps
are not parallel, bending moments are introduced during the test which is
unacceptable. The set up for gluing the caps parallel on the specimen is shown in
Figure 7-5.
 Special PVC rings (thickness 8 mm) are glued around the specimen at the cap to
prevent the specimen failure between the cap and the specimen. The glue is a two
component fast curing adhesive, X60, consisting of a liquid component (B) and a
powder component (A), which are mixed at the weight ratio 1:1. After gluing, the
specimen was kept in the set-up for ten minutes.

Figure 7-5 The gluing mould for MUT test


 Then the specimen is connected to the upper and lower hinges; three LVDT’s
around the steel caps and extensometer in the middle of the specimen are fixed to
the specimen.
 The temperature cabinet is closed in order to allow the temperature of top and
bottom caps to reach the target value. A minimum of 3 hours is used for
temperature is stabilization.
 In the displacement-controlled mode, a load is applied such that a constant
displacement rate is achieved till the specimen breaks. This displacement rate is
based on the average value of the three external axial LVDT’s and is fed back to the
system.

165
Chapter 7 Monotonic Uniaxial Tension and Compression Tests

7.5 Data Processing


7.5.1 Stress and Strain
The first step of data processing is to calculate the magnitude of the strains and stresses to
be determined during the MUC and MUT tests. They include vertical stress, axial strain,
radial strain and volumetric strains and are calculated with Equations (7-5), (7-6), (7-7)
and (7-8).
Load + sample weight
vertical stress σ = (7-5)
initial cross sec tion area
change in height ∆h
axial strain ε a = (7-6)
original height h0
change in radius ∆r
radial strain ε r = (7-7)
original radius r0
volumetric strain ε v = 2ε r + ε a (7-8)

During the MUT test and the beginning of the MUC test, the radial deformation, ∆r, was
measured using an extensometer as shown in Figure 7-6 and calculated using Equation
(7-9) (MTS manual).

r
li =initial cord length, mm
lf =final cord length, mm
Rf r =radius of roller, mm
∆r =radial deformation, mm
Ri Ri =initial radius of specimen, mm
θi θf li lf Rf =final radius of specimen, mm
θi =angle subtended by li, in radians
θf =angle subtended by lf, in radians
∆r

End roller of extensometer Chain

Figure 7-6 Schematic of the measurement of radial deformation using an extensometer

∆l
∆r = R f − Ri =
 θ   θi   θi  (7-9)
2 sin  i  + π − 2  cos  2  
 2     
∆l = l f − li (7-10)

166
Chapter 7 Monotonic Uniaxial Tension and Compression Tests

lc
θi = 2π − (7-11)
( Ri + r )
Where: ∆l : change in cord length, extensometer output, mm;
lc : chain length, remains constant, mm;

In the MUC test, the radial deformation is much larger than that in the MUT test and
beyond the range of the extensometer. In that case the radial deformation is measured
using two potentiometers and a string as illustrated in Figure 7-7 and calculated
iteratively from Equation 7-12 (Erkens and Poot 2000).

  ∆ h 
2    ∆h 
2 
∆L − 2  L1 + 
2
+ ∆R 2  + L −  L2
+ + ∆R 2 
−L
  2   1  2  2   2 (7-12)
∆R =    

Where: ∆R : radial deformation, mm;


∆L : change in length of the string, mm;
∆h : change in height of the specimen, mm;
L1, L2 : distance between potentiometer and the middle of specimen,
mm

∆R

∆h

Initial position of the string

Potentiometer Final position of the string ∆h Potentiometer


2

L1 L2

Figure 7-7 Schematic of the measurement of radial deformation using potentiometers and
string

Figure 7-8 gives an example of the radial strains measured by the extensometer and string.
It is clear that the signals from extensometer and string are not exactly consistent. A
correction was made to combine these two signals. The range of the extensometer is
much smaller than that of the string. So the signal from the extensometer is more accurate
when the radial deformation is in the small range. Therefore, in the beginning of the test,
the radial strain was calculated from the signal of the extensometer.

167
Chapter 7 Monotonic Uniaxial Tension and Compression Tests

With an increasing vertical load, the radial deformation becomes larger as well and will
be outside the range of the extensometer at a certain moment (point A). From this
moment on, the strain was calculated from the signal of the string. After point A, the
signal of the string was shifted to this point.

C-1-5@20ºC, 2.5mm/s
0.1
1- Signal from extensometer
0.08 2- Signal from string
3- Corrected signal
Strain [m/m]

0.04
0.06 2
2

0.04 0.03 3
3

A
1
0.02 A
1 0.02
2.5 2.7 2.9 3.1 3.3
0
0 2 4 6 8 10
Time [sec]

Figure 7-8 Correction of the radial strain measured by extensometer and string

7.5.2 Strain Rate, Maximum Stress, Tangent Stiffness and Onset of


Dilation
Based on the strains and stresses mentioned above, the strain rate, elastic modulus,
maximum stress, failure strain and point of onset of dilation can be determined. Figure 7-
9 shows an illustration of how these parameters are determined.

C-2-11@30ºC, 4mm/s
-10 0.0025

Etan
-8 0.002
P
σdil
Volume strain [m/m]

-6 0.0015
Stress [MPa]

-4 Strain rate 0.001


Calculation Volume strain
-2 0.0005
Stress

0 0
0 -0.005 -0.01 -0.015 -0.02 -0.025 -0.03
2 -0.0005
D
4 -0.001
Axial strain [m /m ]

Figure 7-9 Calculation of strain rate, tangent stiffness, maximum stress, failure strain and
onset of dilation from a MUC test

168
Chapter 7 Monotonic Uniaxial Tension and Compression Tests

The peak value of the stress curve (point P) is defined as the strength and the
corresponding axial and radial strain are the failure strain in vertical and horizontal
direction at the maximum stress. The strain rate up to the maximum stress was chosen as
the actual strain rate. The slope of the linear part of the stress-strain curve is used to
determine the tangent stiffness, Etan. In the beginning of the MUC test, the axial strain is
larger than the radial strain, which leads to a reduction in volume strain. The onset of
dilation is the point at which the volume starts to increase due to internal crack growth.
This point is determined by means of the axial strain versus volume strain plot, in which
the lowest point (point D) is defined as the onset of dilation.

7.6 Test Results


7.6.1 MUC Test Results
Figure 7-10 shows the failed specimens after the MUC test. At high temperature and low
strain rate, the specimen deforms uniformly and does not completely fracture. But at low
temperature and high strain rate, the fracture plane lies approximately on one of the
diagonals (see Figure 7-10b), which indicates there is still some effect of shear stresses at
the loading platens. At the end the cylindrical specimen is completely split into two main
parts.

(a) C-2-1@ 30.6 ºC, 0.594 %/s; (b) C-1-10@ 9.4 ºC, 3.416 %/s
Figure 7-10 Failed specimens after being tested at different test conditions in the MUC
test

During testing, the force, axial and radial deformation were recorded. The average axial
deformation was used to calculate the axial strain. This was done because all three
LVDT's did not measure the same displacement rate (see Figure 7-11). This indicates that
in spite of all precautions still some bending occurred.

169
Chapter 7 Monotonic Uniaxial Tension and Compression Tests

C-2-1@30C, 0.594%/s
-12 -10

-10
-8
Displacement [mm]
-8

Force [kN]
-6
-6
-4
-4

LVDT1 LVDT2 -2
-2 LVDT3 Average
Force
0 0
0 5 10 15
Time [s]

Figure 7-11 Measured axial displacements and force

The vertical displacement measured by the LVDT’s includes two parts, the deformation
of the specimen and the deformation of the friction reduction system. In order to remove
the influence of the friction reduction system, Pramesti performed the calibration tests to
determine the relationship between the compressive stress and the deformation of the
friction reduction system [Pramesti, 2013]. During testing, the compressive force was
applied to a steel cylinder with the friction reduction system. Three LVDTs with a range
of 2 mm were used to measure the deformation. The characteristics of the material and
the test conditions are shown in Table 7-3.

Table 7-3 The characteristics of the material and test conditions [Pramesti, 2013]
Specimen Descriptor Test condition
1. Force control of 0.02 kN/s
E = 210,000 MPa 2. T = 20 °C
Steel cylinder H = 130 mm 3. Preload = 0.15 kN
∅ = 65 mm 4. Max Force = 60 kN
5. 2 hour precondition before start of the test

Figure 7-12 shows the relation between compressive stress and the measured
displacement. In order to determine to what extent the friction reduction deformation
influences the overall deformation, a mathematical model relating displacement to stress
has been formulated. From Figure 7-12 it can be seen that there is a simple
exponential/logarithmic relation between displacement measured by the LVDTs and the
stress. It is assumed that there is a maximum displacement when the friction reduction
deformation reaches its maximum value. By using the solver function in Microsoft Excel,
the constants for the relation between displacement and stress were obtained. The model
is expressed in Equation 7-13 and the comparison between measured and predicted
values is presented in Figure 7-12.

170
Chapter 7 Monotonic Uniaxial Tension and Compression Tests

0.14

0.12

Displacement (mm)
0.1

0.08

0.06

0.04

0.02 Measuerement Model

0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15

Stress (MPa)

Figure 7-12. Measured displacement vs. compressive stress [Pramesti, 2013]

−0.5998σ c + 0.6295σ c 0 ≤ σ c ≤ 15MPa


0.9869
∆Lm = ∆Ls + ∆L f =  (7-13)
0.118 σ c > 15MPa

Where: ∆Lm : measured total deformation, mm;


∆Ls : deformation of the steel cylinder, mm;
∆Lf : deformation of the friction reduction system and frame, mm;
σc : compressive stress, MPa

With the height and stiffness of the steel cylinder (see Table 7-3), the deformation of the
steel cylinder can be calculated by Equation 7-13.

Hs 130
∆Ls = H sε s = σc = σ c = 6.19 ×10−4 σ c (7-14)
Es 210000

Where: εs : strain of the steel cylinder, m/m;


Hs : height of the steel cylinder, mm;
Es : stiffness of the steel cylinder, MPa.

By substituting Equation 7-14 into Equation 7-13, the deformation of the friction
reduction system can be expressed as follows:

−0.6004σ c + 0.6295σ c 0 ≤ σ c ≤ 15MPa


0.9869

∆L f = ∆Lm − ∆Ls =  (7-15)


0.118 − 6.19 × 10 σ c σ c > 15MPa
−4

Therefore, when testing the asphalt specimen, the deformation of the specimen can be
simply calculated by the Equation 7-16.

171
Chapter 7 Monotonic Uniaxial Tension and Compression Tests

∆Lm − ( −0.6004σ c + 0.6295σ c0.9869 ) 0 ≤ σ c ≤ 15MPa



∆La = ∆Lm − ∆L f =  (7-16)
∆Lm − ( 0.118 − 6.19 × 10 σ c ) σ c > 15MPa
−4

Where: ∆La : deformation of the asphalt specimen, mm.

Equation 7-16 then can be used to calculate the corrected value of the displacement.
Figure 7-13 gives the comparison of the axial strain before and after correction. It can be
seen that the influence of the friction reduction system is significant before the
compressive stress reaches the peak value.

C-2-8@4°C,1%/s
-35
Compressive stress [MPa]

-30

-25

-20

-15

-10
axial strain
-5
Corrected axial strain
0
0 -0.005 -0.01 -0.015 -0.02
Axial strain [m/m]

Figure 7-13 Axial strain-stress relation at T = 4°C and εɺ =1 %/s before and after
correction

The stress-strain curves at the different test conditions are presented in Figure 7-14 and
Figure 7-15. During the test, the stress-strain curve contains two parts, loading and
unloading. The peak point and the slope of the loading part represent the strength and the
stiffness of the material. It can be seen that the behavior of an asphalt mix depends on
temperature and strain rate. Based on the stress-strain curve, the compressive strength fc,
tangential stiffness Etan,c, compressive failure strain in axial direction εa,c and in radial
direction εr,c were determined. The results are given in Table 7-4.

The friction reduction system influences the tangential stiffness Etan,c, compressive failure
strain in axial direction εa,c. Table 7-5 gives the values of Etan,c and εa,c before and after
the correction. In general the tangential stiffness increases and the compressive failure
strain in axial direction decreases after the correction.

172
Chapter 7 Monotonic Uniaxial Tension and Compression Tests

-16
σ [MPa] C-2-4, 20°C, ε=4 %/s
C-2-12, 20°C, ε=2.005 %/s
-14 C-1-11, 20°C, ε=2.005 %/s
C-1-8, 20°C, ε=2.005 %/s
-12 C-1-3, 20°C, ε=2.005 %/s
C-1-4, 20°C, ε=0.01 %/s
-10

-8

-6

-4

-2

0
-0.15 -0.1 -0.05 0 0.05 0.1 0.15 0.2
εaxial [m/m] εradial [m/m]

Figure 7-14 Compressive stress versus axial and radial strain at 20 ºC

-40
4°C, ε=1 %/s
σ [MPa] 5°C, ε=2.005 %/s
-35 9.4°C, ε=3.416 %/s
9.4°C, ε=0.594 %/s
-30 30.6°C, ε=3.416 %/s
30.6°C, ε=0.594 %/s
-25 35°C, ε=2.005 %/s
38°C, ε=3 %/s
-20

-15

-10

-5

0
-0.15 -0.1 -0.05 0 0.05 0.1 0.15 0.2
εaxial [m/m] εradial [m/m]

Figure 7-15 Compressive stress versus axial and radial strain at other temperatures

173
Chapter 7 Monotonic Uniaxial Tension and Compression Tests

Table 7-4 Overview of the MUC test results


Sample Temp. Target strain Real strain Etan,c εr,c
fc [MPa] εa,c [%]
code [ºC] rate [%/s] rate [%/s] [MPa] [%]
C-2-8 4 1 1.047 -30.94 9243 -0.666 0.231
C-2-9 5 2.005 2.009 -36.45 8292 -0.598 0.169
C-2-5 9.4 0.594 0.625 -22.27 6239 -0.723 0.288
C-1-10 9.4 3.416 3.364 -35.02 14390 -0.538 0.222
C-1-4 20 0.01 0.0102 -2.22 474 -2.294 1.400
C-2-2 20 2.005 2.002 -14.08 4836 -0.710 0.396
C-2-12 20 2.005 2.091 -11.83 4595 -0.912 0.400
C-1-11 20 2.005 2.037 -11.06 3462 -1.011 0.000
C-1-8 20 2.005 2.011 -11.71 4249 -0.828 0.411
C-1-3 20 2.005 2.02 -12.49 5826 -0.894 0.193
C-2-4 20 4 4.189 -13.86 6549 -0.619 0.265
C-2-1 30.6 0.594 0.620 -3.39 596 -1.985 1.047
C-2-11 30.6 3.414 3.582 -6.45 3079 -1.280 0.705
C-1-9 35 2.005 2.051 -3.63 722 -1.658 1.060
C-1-6 38 3 3.025 -2.75 719 -2.196 1.370

Table 7-5 Influence of the friction reduction system on tangential stiffness Etan,c and
compressive failure strain in axial direction εa,c
strain Corrected Corrected
Sample Temp. Etan,c εa,c
rate Etan,c Ratio εa,c Ratio
code [ºC] [MPa] [%]
[%/s] [MPa] [%]
C-2-8 4 1.05 8579 9243 1.08 -0.745 -0.666 0.89
C-2-9 5 2.01 8168 8292 1.02 -0.672 -0.598 0.89
C-2-5 9.4 0.63 5287 6239 1.18 -0.806 -0.723 0.90
C-1-10 9.4 3.36 9475 14390 1.52 -0.612 -0.538 0.88
C-1-4 20 0.01 424 474 1.12 -2.333 -2.294 0.98
C-2-2 20 2.00 4639 4836 1.04 -0.797 -0.710 0.89
C-2-12 20 2.09 3818 4595 1.20 -0.997 -0.912 0.91
C-1-11 20 2.04 3406 3462 1.02 -1.094 -1.011 0.92
C-1-8 20 2.01 4168 4249 1.02 -0.910 -0.828 0.91
C-1-3 20 2.02 3810 5826 1.53 -0.829 -0.894 1.08
C-2-4 20 4.19 4066 6549 1.61 -0.705 -0.619 0.88
C-2-1 30.6 0.62 550 596 1.08 -1.849 -1.985 1.07
C-2-11 30.6 3.58 2411 3079 1.28 -1.351 -1.280 0.95
C-1-9 35 2.05 712 722 1.01 -1.711 -1.658 0.97
C-1-6 38 3.03 654 719 1.10 -2.240 -2.196 0.98
Ratio = (the value after correction)/(the value before correction)

174
Chapter 7 Monotonic Uniaxial Tension and Compression Tests

7.6.2 MUT Test Results


Unlike in the compression tests, failure in a tension test is localized and highly depends
on a local defect of the specimen. Therefore the failure plane occurs randomly on the
sample; this is shown in Figure 7-16. A compression specimen fails in a kind of overall
split cracking pattern, a tension specimen breaks in two and the two halves are mostly
undamaged. In a tension test, the end effects do not influence the observed fracturing
response if the crack occurs far enough from the end caps.

Figure 7-16 Failed samples after being subjected to the MUT test

7
σ [MPa]
6

4
5°C, 0.05 %/s
5°C, 0.01 %/s 3
9.4°C, 3.414 %/s
9.4°C, 0.587 %/s 2
9.4°C, 0.05 %/s
12.9°C, 2 %/s 1
12.9°C, 0.587 %/s
0
0.0012 0.001 0.0008 0.0006 0.0004 0.0002 0 -0.0002 -0.0004 -0.0006
εaxial [m /m ] εradial [m /m ]

Figure 7-17 Tensile stress versus axial and radial strain at low temperature

The average value of the three LVDT’s was also used to determine the strain rate. In
Figure 7-17 and Figure 7-18, the tensile stress from the MUT test is plotted versus the
axial and radial strain at different test conditions. Similar to the MUC test results, the
material behavior of an asphalt mixture in tension also strongly depends on the strain rate
and temperature. The difference is that at low temperature or high strain rate, once the
cracks initiate, crack propagation happens in a very short time. That is the reason why in
Figure 7-17 the stresses suddenly reduce to zero after reaching the peak value and the

175
Chapter 7 Monotonic Uniaxial Tension and Compression Tests

unloading part of the stress-strain curve is hard to see. Table 7-6 shows the target and
actual tests conditions and the results of the MUT tests, namely tensile strength (ft),
tangential stiffness modulus in tension (Etan,t), tensile failure strain in axial direction εa,t
and in radial direction εr,t.
8
30.6°C, 0.587 %/s
σ [MPa]
30.6°C, 3.416 %/s 7
35°C, 2.005 %/s
38°C, 3 %/s 6
20°C, 0.01 %/s
5
20°C, 2.005 %/s
20°C, 4 %/s 4

0
0.015 0.01 0.005 0 -0.005 -0.01
εaxial [m /m ] εradial [m /m ]

Figure 7-18 Tensile stress versus axial and radial strain at high temperature

Table 7-6 Overview of the MUT test results


Sample Temp. Target strain Real strain Etan,t
ft [MPa] εt,v [%] εt,h [%]
code [ºC] rate [%/s] rate [%/s] [MPa]
C-2-3 5 0.01 0.010 4.51 9990 0.108 -0.038
C-8-4 5 0.05 0.056 5.75 15316 0.055 -0.015
C-6-5 5 2.005 1.270 6.05 19478 0.045 -0.011
C-6-9 5 2.005 1.274 5.97 16036 0.033 -0.007
C-8-7 9.4 0.05 0.051 5.17 10453 0.100 -0.031
C-8-10 9.4 0.587 0.535 6.29 15361 0.079 -0.015
C-8-5 9.4 3.414 1.148 5.89 17130 0.062 -0.012
C-7-7 9.4 3.414 1.940 6.00 12494 0.071 -0.031
C-6-7 9.4 3.414 1.970 6.36 17865 0.049 -0.023
C-8-6 12.9 0.587 0.620 5.67 12474 0.080 -0.019
C-8-9 12.9 2 1.180 6.02 14793 0.056 -0.015
C-6-1 20 0.01 0.010 1.08 405 0.512 -0.233
C-8-2 20 1 1.560 6.55 8796 0.180 -0.068
C-7-1 20 2.005 1.740 6.20 10213 0.111 -0.064
C-6-2 20 2.005 2.360 6.51 9145 0.104 -0.038
C-7-10 20 2.005 2.127 6.46 7665 0.295 -0.124
C-6-4 20 2.005 2.390 6.76 9392 0.221 -0.108
C-6-3 20 4 3.580 7.43 8739 0.178 -0.069
C-8-1 30.6 0.587 1.289 1.35 2068 0.619 -0.191
C-6-8 30.6 3.416 3.856 3.23 4189 0.378 -0.158
C-6-6 35 2.005 3.350 1.68 2193 0.589 -0.117
C-7-2 38 3 4.370 1.14 956 0.598 -0.238

176
Chapter 7 Monotonic Uniaxial Tension and Compression Tests

7.7 The Unified Model


As a viscoelastic material, material properties of asphalt mixtures such as strength,
tangential stiffness modulus and failure strain are related to the temperature and time
derivative variables. The unified model, proposed by Medani and Huurman [Medani,
2005] was used to establish the relationships between material properties and the test
conditions; the model is shown in Equation 7-17. Based on the Time-Temperature
superposition principle, the model allows shifting the data obtained at various
temperatures with respect to strain rate to a selected reference temperature. It is believed
that when the reduced strain rate converges to 0 or infinite, the material properties tend to
increase or decrease to a limit value [Pellinen and Witczak, 2002] [Scarpas et al., 1997]
[Erkens, 2002].

P = Ph + ( Pl − Ph ) S (7-17)
γ
 u  
Where: S = exp  −  r   ; ur = u βT ; βT = exp  −Ts (T − T0 )
  u0  
 
P : a material property e.g. compressive strength, tensile
strength, MPa; compressive strength, MPa;
Pl : value of P when u→0;
Ph : value of P when u→∞;
u : time derivative variable e.g. strain rate, 1/s;
ur : reduced time derivative variable;
u0 : reference value of time derivative variable u;
βT : temperature susceptibility function;
T : temperature, K;
T0 : reference temperature, K;
Ts : model parameter, 1/K;
γ : model parameter

The temperature susceptibility factor, βT, can be determined at a certain reference


temperature from the stiffness tests mentioned in Chapter 5. The theoretical minimum
value, Pl of the properties derived from the MUC and MUT tests was set to zero since
little or no strength and stiffness can be expected at high temperatures without
confinement. The other model parameters Ph, u0 and γ can be obtained by minimizing the
differences between the measured and fitted values for all material properties. This
procedure can simply be completed by means of the Solver function in an Excel
spreadsheet.

By means of the unified model parameters, the master curves at the reference temperature
of 10 ºC (283.15 K) were constructed for the strength, tangential stiffness, axial and
radial strain at failure. The results are shown in Figure 7-19, Figure 7-20, Figure 7-21 and
Figure 7-22. The model parameters for each material property are presented in Table 7-7.

177
Chapter 7 Monotonic Uniaxial Tension and Compression Tests

100

fc or ft [MPa]
10

fc ft

0.1
1E-06 0.00001 0.0001 0.001 0.01 0.1 1
Reduced strain rate [1/s]

Figure 7-19 Master curves of the compressive and tensile strength at the reference
temperature of 10 ºC

100000
Etan,c or Etan,t [MPa]

10000

1000
Etan,c
Etan,t

100
1E-06 0.00001 0.0001 0.001 0.01 0.1 1
Reduced strain rate [1/s]

Figure 7-20 Master curves of the compressive and tensile tangent stiffness at the
reference temperature of 10 ºC
0.1

εa,c εa,t
εa,c or εa,t [m/m]

0.01

0.001

0.0001
1E-07 1E-06 1E-05 0.0001 0.001 0.01 0.1 1
Reduced strain rate [1/s]

Figure 7-21 Master curves of the vertical compressive and tensile failure strain at the
reference temperature of 10 ºC

178
Chapter 7 Monotonic Uniaxial Tension and Compression Tests

0.1
εr,c εr,t

0.01

εr,c or εr,t [m/m]


0.001

0.0001

0.00001
0.0000001 0.00001 0.001 0.1 10
Reduced strain rate [1/s]

Figure 7-22 Master curves of the horizontal compressive and tensile failure strain at
the reference temperature of 10 ºC

Table 7-7 Unified Model parameters for the MUC and MUT tests
Model
T0 Ts Pl Ph u0 γ R2
parameters
fc 283.15 0.269 0 40.5 2.0×10-2 0.362 0.98
-3
Ec 283.15 0.269 0 9000 2.9×10 0.480 0.85
εa,c 283.15 0.269 0 3.8×10-2 9.9×10-7 -0.179 0.94
εr,c 283.15 0.269 0 2.0×10-2 1.2×10-4 -0.354 0.92
ft 283.15 0.269 0 6.1 2.2×10-4 0.694 0.95
-3
Et 283.15 0.269 0 21000 3.9×10 0.598 0.83
-3 -5
εa,t 283.15 0.269 0 6.9×10 1.6×10 -0.357 0.82
εr,t 283.15 0.269 0 2.3×10-3 1.3×10-4 -0.561 0.79

The figures show that the model provides a reasonable description of the test results.
However, the graph showing the tension test results exhibits quite some scatter. This
might be related to the mechanism of the tensile failure which was discussed in section
7.6.2. By comparing the results from MUC and MUT tests, it can be seen that the
compressive strength and failure strain are higher than those obtained in the tension test.
The reason for this is that in compression the aggregate skeleton efficiently carries the
compression load and the deformation distributes quite uniformly throughout the
specimen, while for the tension test, the tensile strength mainly depends on the cohesion
and adhesion of the bitumen and the bituminous mortar. Conversely, the tangential
stiffness in tension is higher than that in compression at high reduced strain rate.

For all material properties, the slopes of the curves in tension are steeper than those in
compression. This indicates that the material properties in tension are more sensitive to
the temperature and strain rate.

179
Chapter 7 Monotonic Uniaxial Tension and Compression Tests

7.8 Prediction of the Unified Model Parameters


The test procedures described in this chapter, clearly show that in order to obtain reliable
results from the MUC and MUT test, many aspects should be carefully taken into account,
such as specimen preparation, test setup configuration, data analysis, etc.. Because of
their complexity (especially of the tension test) it is not likely that these tests will be done
on a routine basis in e.g. contractor’s laboratories. Therefore, it is necessary to find a
simple way to predict tensile and compressive strength from the material properties, such
as mixture composition without the need to conduct the MUC and MUT test. It should be
noted that such procedures are common practice in pavement engineering. They exist for
estimating the stiffness and fatigue resistance of asphalt mixtures.
It is reasonable to assume that the shape of the unified model depends on the aggregate
gradation, shape of the aggregate, mix composition, binder characteristics and level of
compaction. This then should offer possibilities to develop a simplified unified model,
which allows the compressive and tensile strength to be directly calculated from the
material properties and mixture characteristics.

In previous research, MUC and MUT tests have been performed to measure the
compressive and tensile strength for a number of road materials [Jansen, 2002] [Muraya,
2007] [Erkens, 2002] [Pungky, 2012]. The test results of 7 different kinds of asphalt
mixtures were collected in this Section to establish a relationship between the unified
model parameters and the material properties. Hereafter some general information of the
involved mixtures is given:

• DAC 0/8: Dense asphalt concrete with conventional bitumen (40/60 pen) used in this
research. Detailed information about this mixture was shown in Chapter 4.
• EME 0/14: Enrobé á Modele Elevé, a bituminous base course material with “special”
hard binder (10/20 pen). The grain size is between 0 and 14 mm. Details about this
mixture can be found in [Jansen, 2002].
• PAC 0/16: Porous asphalt concrete with conventional bitumen (70/100 pen). The
grain size is between 0 and 16 mm. Details of this mixture are given in [Muraya,
2007].
• DAC 0/16: Dense asphalt concrete with conventional bitumen (40/60 pen). The grain
size is between 0 and 16 mm. Also details of this mixture can be found in [Muraya,
2007].
• SMA 0/11: Stone mastic asphalt mixture with conventional bitumen (70/100 pen).
The grain size is between 0 and 16 mm. This mixture was tested by Muraya as well
[Muraya, 2007].
• ACRe 0/4: A kind of sand mixture with bitumen (45/60 pen). The grain size is
between 0 and 4 mm. This fine grained mixture was used by Erkens for development
of the ACRe model [Erkens, 2002].
• GAC 0/32: Gravel asphalt concrete with conventional bitumen (40/60 pen). The grain
size is between 0 and 32 mm. This mixture was tested by Pramesti [Pramesti, 2012].

180
Chapter 7 Monotonic Uniaxial Tension and Compression Tests

Table 7-8 Material properties of various asphalt mixtures


Gradation E [MPa] Va [%] Vb [%] Cc
DAC 0/8 5200 3.0 14.9 3.25
EME 0/14 5700 3.4 12.2 0.73
PAC 0/16 2500 20.0 8.2 12.25
DAC 0/16 6900 2.7 12.9 20.48
SMA 0/11 2800 5.2 14.5 0.95
ACRe 0/4 3543 2.6 19.3 0.99
GAC 4700 4.4 8.9 1.60

The strength of an asphalt mixture is not only influenced by temperature and strain rate
but also by the properties by the bitumen and the volumetric composition of the mixture.
Based on work presented in [Molenaar, 1994], four material properties being stiffness of
the mixture E, volume percentage of the air void Va, volume percentage of the bitumen
Vb and coefficient of curvature of the gradation curve Cc were taken into account. The
material properties of all the asphalt mixtures are collected in Table 7-8. The dynamic
stiffness is a very important property, since it reflects the characteristics of the bitumen,
the aggregate skeleton and volumetric composition of the mixture. The dynamic stiffness
varies with the frequency and the temperature. In previous studies, the stiffness was
measured using different tests, different waveforms and loading amplitudes. To obtain
comparable results from different mixtures, it is assumed that all stiffness tests were
conducted using a triangle waveform. The strain rate was computed as follows from the
frequency and strain amplitude.

ε
εɺ = 100 ⋅ (7-18)
t
Where: εɺ : approximate strain rate, %;
ε : strain amplitude, m/m;
t : time duration when the strain level changes from 0 to peak
value in a cycle, s;

In Table 7-8, E represents the dynamic stiffness measured at the strain rate of 0.1 %/s and
a temperature of 20 °C. The coefficient of curvature of the gradation curve, Cc is
calculated by:

Cc =
( D30 )
(7-19)
D10 ⋅ D60

Where: D30 : sieve diameter through which 30 % of the aggregates is


passing, mm;
D10 : sieve diameter through which 10 % of the aggregates is
passing, mm;
D60 : sieve diameter through which 60 % of the aggregates is
passing, mm.

181
Chapter 7 Monotonic Uniaxial Tension and Compression Tests

With the same procedure as mentioned in Section 7.7, the unified model parameters for
the compressive and tensile strength were determined at the reference temperature of 10
°C. The model parameters for compressive and tensile strength are shown in Table 7-9.

Table 7-9 Model parameters for various asphalt mixtures


Mixture type T0 Ts Pl Ph u0 γ R2
DAC 0/8 283.15 0.27 0 48.6 4.4×10-2 0.350 0.84
EME 0/14 283.15 0.27 0 37.1 5.1×10-2 0.279 0.84
PAC 0/16 283.15 0.31 0 18.8 8.6×10-2 0.368 0.90
Compressive
DAC 0/16 283.15 0.31 0 61.3 9.7×10-2 0.315 0.77
strength
SMA 0/11 283.15 0.33 0 35.1 1.4×10-1 0.320 0.92
ACRe 0/4 283.15 0.30 0 60.6 2.9×10-1 0.304 0.94
GAC 0/32 283.15 0.25 0 50.6 5.4×10-2 0.316 0.77
DAC 0/8 283.15 0.27 0 6.0 2.3×10-4 0.686 0.93
EME 0/14 283.15 0.27 0 6.0 1.2×10-4 0.400 0.85
PAC 0/16 283.15 0.31 0 2.0 9.8×10-4 0.685 0.89
Tensile
DAC 0/16 283.15 0.31 0 5.8 2.5×10-4 0.569 0.91
strength
SMA 0/11 283.15 0.33 0 3.9 1.2×10-4 0.538 0.94
ACRe 0/4 283.15 0.30 0 6.1 7.4×10-4 0.529 0.96
GAC 0/32 283.15 0.25 0 4.9 3.1×10-4 0.551 0.86

Figure 7-23 summarizes the predicted and measured values for compressive strength and
tensile strength. It is clear that the unified model as proposed for the different mixtures is
capable of providing good estimates for the compressive and tensile strength. Most of the
data points lie in the range of ± 20%. The correlation coefficient R2 of each data set listed
in Table 7-9 is higher than 0.75.

Unified model
60 Line of
DAC0/8 EME 0/14 +20%
equality
50 PAC 0/16 DAC 0/16
SMA 0/11 ACRe 0/4
Predicted fc [MPa]

40 GAC 0/32 -20%

30 R2=0.986

20

10

0
0 10 20 30 40 50 60
Measured fc [MPa]

(a) Compressive strength

182
Chapter 7 Monotonic Uniaxial Tension and Compression Tests

Unified model
10
+20% Line of
R2=0.949 equality
8

Predicted ft [MPa]
-20%
6

4
DAC0/8 EME 0/14
2 PAC 0/16 DAC 0/16
SMA 0/11 ACRe 0/4
GAC 0/32
0
0 2 4 6 8 10
Measured ft [MPa]

(b) Tensile strength


Figure 7-23 Comparison between prediction of the unified model and measurement of
compressive strength (a) and tensile strength (b)

As mentioned in Section 7.7, the parameter T0 and Ts relates to the temperature


susceptibility and the parameters Pl for various mixtures are always equal to zero. So it is
believed that the parameter Ph, u0 and γ are influenced by the material properties. Figure
7-24, Figure 7-25 and Figure 7-26 show the relationships between the selected material
properties and the model parameters Ph, u0 and γ. Similar results were also found for the
model parameters of the tensile strength. From Figure 7-24, it can be seen that the
parameter Ph highly depends on the stiffness of the mixture, the properties of the bitumen
and the volumetric composition of the mixture. In general the value of Ph increases with
the mixture stiffness E, bitumen content Vb and decreases with void content Va. These
observed results are logical. Based on work presented by Molenaar [Molenaar, 1994],
also the degree of filling of the voids in the aggregate skeleton by bitumen was
considered to be an important parameter. By taking these parameters into account, the
following relationship was developed for Ph.

a3
 Vb 
Ph = a1 ⋅ E 
a2
 (7-20)
 Vb + Va 

Where: Ph : value of P when strain rate εɺ → ∞ ;


E : stiffness at the temperature of 20 °C and the strain rate of 0.1
%/s, MPa;
Vb : volume content of the bitumen, %;
Va : volume content of air void, %;
a1, a2 and a3 : model parameters.

183
Chapter 7 Monotonic Uniaxial Tension and Compression Tests

70 70
60 60
50
50
40
Ph

40

Ph
30 30
20 20
10
10
0
0
0 2000 4000 6000 8000
5 10 15 20
E [MPa] V b [%]

70 70
60 60
50 50
40 40
Ph

Ph
30 30
20 20
10 10
0 0
0 5 10 15 20 25 0 5 10 15 20 25
Va [%] Cc

Figure 7-24 Relations between Ph and the material properties for compressive strength

0.35 0.35
0.3 0.3
0.25 0.25
0.2 0.2
u0

u0

0.15 0.15
0.1 0.1
0.05 0.05
0 0
0 2000 4000 6000 8000 5 10 15 20
V b [%]
E [MPa]

0.35 0.35

0.3 0.3
0.25 0.25
0.2 0.2
u0
u0

0.15 0.15
0.1 0.1
0.05 0.05
0 0
0 5 10 15 20 25 0 5 10 15 20 25
Va [%] Cc

Figure 7-25 Relations between u0 and the material properties for compressive strength

184
Chapter 7 Monotonic Uniaxial Tension and Compression Tests

0.4 0.4
0.35 0.35
0.3 0.3
0.25 0.25

γ
0.2 0.2
γ

0.15 0.15
0.1 0.1
0.05 0.05
0 0
0 2000 4000 6000 8000 5 10 15 20
V b [%]
E [MPa]
0.4 0.4
0.35 0.35
0.3 0.3
0.25 0.25
γ

0.2

γ
0.2
0.15
0.15
0.1
0.1
0.05
0.05
0
0
0 5 10 15 20 25
0 5 10 15 20 25
Va [%] Cc

Figure 7-26 Relations between γ and the material properties for compressive strength

The parameters in this function can be simply determined by means of the Solver
function in Excel. The following equations give the functions of the parameter Ph for
compressive and tensile strength.
0.623
 Vb 
For compressive strength: Ph = 1.755 ⋅ E 0.402
  (7-21)
 Vb + Va 
0.849
 Vb 
For tensile strength: Ph = 0.505 ⋅ E 0.308
  (7-22)
 Vb + Va 

Table 7-10 gives the comparison between the measured and predicted parameter Ph. The
mean percent relative error (MPRE) of the prediction value is calculated by means of
Equation 7-23. Except for the compressive strength data points of the mixtures ACRe 0/4
and EME 0/14, the relative errors are all less than 20 %.

n Ppi − Pmi

i =1 Pmi (7-23)
MPRE = ×100%
n

Where: Pmi : measured data;


Ppi : predicted data;
n : number of test data;

185
Chapter 7 Monotonic Uniaxial Tension and Compression Tests

Table 7-10 Predicted model parameters Ph for various asphalt mixture


Compression tension
Mixture type
Measured Predicted MPRE Measured Predicted MPRE
DAC 0/8 48.6 48.6 0.0% 6.02 6.02 0.0%
EME 0/14 37.1 48.6 31.2% 5.99 5.89 1.6%
PAC 0/16 18.8 18.9 0.0% 1.97 1.97 0.0%
DAC 0/16 61.3 54.3 11.5% 5.80 6.54 12.7%
SMA 0/11 35.1 35.1 0.1% 3.86 4.48 16.3%
ACRe 0/4 60.6 43.2 28.7% 6.15 5.62 8.5%
GAC 0/32 50.6 40.7 19.6% 4.88 4.84 0.8%

Figure 7-25 and Figure 7-26 show that the change of the parameter u0 and γ with the
material properties is not significant. Nevertheless it was assumed that the parameter u0
and γ have a linear relationship with the material properties. By minimizing the MPRE
between the predicted and real value, the following equations for u0 and γ were
determined.

For the compressive strength:

u0 = ( −2.83 × 10 −5 ) ⋅ E + ( 6.81 × 10 −3 ) ⋅ Vb + ( −4.21× 10 −3 ) ⋅ Va + ( 3.73 × 10 −3 ) ⋅ Cc + 0.139 (7-24)

γ = ( −4.36 × 10 −6 ) ⋅ E + ( 4.24 × 10 −4 ) ⋅ Vb + ( 2.62 × 10 −3 ) ⋅ Va + (1.03 × 10 −3 ) ⋅ Cc + 0.311 (7-25)

For the tensile strength:

u0 = ( −3.64 × 10 −7 ) ⋅ E + ( −1.46 × 10 −5 ) ⋅ Vb + ( −4.51× 10 −5 ) ⋅ Va + ( 2.77 × 10 −5 ) ⋅ Cc + 0.0025 (7-26)

γ = ( −1.5 × 10 −5 ) ⋅ E + ( 4 × 10 −3 ) ⋅ Vb + ( 5.86 × 10 −3 ) ⋅ Va + ( 4.85 × 10 −3 ) ⋅ Cc + 0.503 (7-27)

On the basis of the calculated model parameters given by Equation 7-21, 7-22, 7-24, 7-25,
7-26 and 7-27, the unified model is simplified and the compressive and tensile strength
can be simply predicted from the material properties. All the predictions for the model
parameters are listed in Table 7-11. In Figure 7-27, the predicted values by the simplified
unified model are compared to the measured results.

186
Chapter 7 Monotonic Uniaxial Tension and Compression Tests

DAC 0/8 EME 0/14


100 100
fc or ft [MPa]

fc or ft [MPa]
10 10

1 fc (Measured) 1
fc (Measured)
ft (Measured) ft (Measured)
Prediction Prediction
0.1 0.1
1E-06 0.0001 0.01 1 100 10000 0.00001 0.001 0.1 10 1000
Reduced strain rate [1/s] Reduced strain rate [1/s]

PAC 0/16 DAC0/16


100 100

10 10

fc or ft [MPa]
fc or ft [MPa]

1
1

fc (Measured)
fc (Measured)
0.1 ft (Measured)
0.1
ft (Measured) Prediction
Prediction 0.01
0.01
1E-06 0.0001 0.01 1 100 10000
1E-07 0.00001 0.001 0.1 10 1000
Reduced strain rate [1/s] Reduced strain rate [1/s]

SMA 0/11 ACRe 0/4


100 100
fc (Measured)
ft (Measured)
10
fc or ft [MPa]

Prediction
fc or ft [MPa]

10

1
0.1 fc (Measured)
ft (Measured)
Prediction
0.01 0.1
1E-07 0.00001 0.001 0.1 10 1000 1E-07 0.00001 0.001 0.1 10 1000
Reduced strain rate [1/s] Reduced strain rate [1/s]

GAC 0/32
100
fc or ft [MPa]

10

1
fc (Measured)
ft (Measured)
Prediction
0.1
0.00001 0.001 0.1 10 1000
Reduced strain rate [1/s]

Figure 7-27 Prediction of compressive and tensile strength from the material properties

187
Chapter 7 Monotonic Uniaxial Tension and Compression Tests

Table 7-11 Calculation of the unified model parameters


Model
Prediction
parameters
T0 , Ts Determined by frequency sweep test
Pl 0
0.623
Compressive  Vb 
Ph = 1.755 ⋅ E 0.402
 
strength  Vb + Va 
Ph 0.849
Tensile  Vb 
Ph = 0.505 ⋅ E 0.308
 
strength  Vb + Va 
Compressive u0 = ( −2.83 × 10 −5 ) ⋅ E + ( 6.81 × 10 −3 ) ⋅ Vb + ( −4.21× 10 −3 ) ⋅ Va
strength + ( 3.73 × 10 −3 ) ⋅ Cc + 0.139
u0
Tensile u0 = ( −3.64 × 10 −7 ) ⋅ E + ( −1.46 × 10 −5 ) ⋅ Vb + ( −4.51 × 10 −5 ) ⋅ Va
strength + ( 2.77 × 10 −5 ) ⋅ Cc + 0.0025

Compressive γ = ( −4.36 × 10 −6 ) ⋅ E + ( 4.24 × 10 −4 ) ⋅ Vb + ( 2.62 × 10 −3 ) ⋅ Va


strength + (1.03 × 10 −3 ) ⋅ Cc + 0.311
γ
Tensile γ = ( −1.5 × 10 −5 ) ⋅ E + ( 4 × 10 −3 ) ⋅ Vb + ( 5.86 × 10 −3 ) ⋅ Va
strength + ( 4.85 × 10 −3 ) ⋅ Cc + 0.503

For the “goodness-of-fit” statistics of the simplified unified model, the statistical
parameters, mean percentage relative error (MPRE) and the correlation coefficient (R2)
are used. For all mixtures, these parameters are shown in Table 7-12 and Table 7-13,
respectively.

∑( P pi − Pmi )
n 2

R2 = 1 − i =1
(7-28)
∑( P
n
− Pm )
2
mi
i =1
Where: Ppi : predicted value;
Pmi : measured data;
Pm : average value of measured data;

From MPRE and R2 values listed in Table 7-12 and Table 7-13, it can be seen that the
predictions of the simplified unified model for most of the data sets show a good
agreement with the measured results. The prediction for the total compressive strength
data has a lower MPRE value. This implies that the model has a better prediction
capability for the compressive strength compared to the tensile strength. The same
conclusions can also be derived from the plots of predicted versus measured values, (see
Figure 7-28).

188
Chapter 7 Monotonic Uniaxial Tension and Compression Tests

Table 7-12 Mean percentage relative errors for different mixtures


DAC EME PAC DAC SMA ACRe GAC
Total
0/8 0/14 0/16 0/16 0/11 0/4 0/32
Compressive
11.2% 20.5% 7.5% 10.5% 2.7% 20.5% 20.5% 12.9%
strength
Tensile strength 11.8% 13.2% 14.7% 15.5% 19.5% 13.3% 15.0% 14.6%
Overall 13.8%
Table 7-13 Correlation coefficients for different mixtures
DAC EME PAC DAC SMA ACRe GAC
Total
0/8 0/14 0/16 0/16 0/11 0/4 0/32
Compressive
0.91 0.84 0.98 0.98 1.00 0.82 0.82 0.90
strength
Tensile strength 0.84 0.84 0.90 0.77 0.92 0.94 0.77 0.91
Overall 0.92
Simplified unified model
60 Line of
DAC0/8 EME 0/14 +20%
PAC 0/16 DAC 0/16 equality
50
SMA 0/11 ACRe 0/4
Predicted fc [MPa]

GAC 0/32 -20%


40

30

20

10

0
0 10 20 30 40 50 60
Measured fc [MPa]

(a) Compressive strength


Simplified unified model
10
DAC0/8 EME 0/14 Line of
+20%
PAC 0/16 DAC 0/16 equality
8 SMA 0/11 ACRe 0/4
Predicted ft [MPa]

GAC 0/32 -20%


6

0
0 2 4 6 8 10
Measured ft [MPa]

(b) Tensile strength

189
Chapter 7 Monotonic Uniaxial Tension and Compression Tests

60
Line of
+20%
equality
50

Predicted value [MPa]


-20%
40

30

20

10

0
0 10 20 30 40 50 60
Measured value [MPa]

(c) Compressive and tensile strength results


Figure 7-28 Comparison of predicted and measured values

From the statistical analysis discussed above, it can be seen that in the simplified
procedure, the model parameter Ph, u0 and γ can be accurately determined by the
proposed equations. It gives a good prediction of compressive and tensile strength
directly from the material properties. However, with regard to the regression analysis of
the linear equation, the number of independent variables is 4 and the number of
measurements is 7. It implies that the degree of freedom is not high enough to provide a
reliable regression analysis. Therefore, more test results are required to obtain a reliable
linear relationship between model parameter u0 and γ the material properties.

From Figure 7-25 and 7-26, it seems that the values of the parameter u0 and γ do not
differ too much for the different asphalt mixtures. To make a further simplification, it is
assumed that the parameters u0 and γ are independent of the material properties. Then the
average values of u0 and γ for all seven mixtures can be used in the unified model. The
determinations of the unified model parameters for the compressive and tensile strength
are presented in Table 7-14.

Table 7-14 Prediction of the unified model parameters (u0 and γ are constant)
T0 Ts Pl Ph u0 γ
0.623
Compressive  Vb 
0 Ph = 1.755 ⋅ E 0.402   7.87E-02 0.322
strength Determined by  Vb + Va 
stiffness test  Vb 
0.849

Tensile strength 0 Ph = 0.505 ⋅ E 0.308


  5.44E-04 0.565
 Vb + Va 

By means of the equations and values listed in Table 7-14, the predicted mastercurves of
the compressive and tensile strength are compared with the measured values in Figure 7-
29.

190
Chapter 7 Monotonic Uniaxial Tension and Compression Tests

DAC 0/8 EME 0/14


100 100

fc or ft [MPa]
fc or ft [MPa]

10 10

1 1
fc (Measured) fc (Measured)
ft (Measured) ft (Measured)
Prediction Prediction
0.1 0.1
1E-06 0.0001 0.01 1 100 10000 0.00001 0.001 0.1 10 1000
Reduced strain rate [1/s]
Reduced strain rate [1/s]
PAC 0/16 DAC0/16
100 100

10 10
fc or ft [MPa]

fc or ft [MPa]
1
1
fc (Measured)
0.1 ft (Measured)
0.1 fc (Measured)
ft (Measured) Prediction
Prediction 0.01
0.01 1E-07 0.00001 0.001 0.1 10 1000
1E-06 0.0001 0.01 1 100 10000
Reduced strain rate [1/s]
Reduced strain rate [1/s]
ACRe 0/4
SMA 0/11 100
100
fc (Measured)
ft (Measured)
Prediction
fc or ft [M Pa]

10
fc or ft [MPa]

10

1
1
0.1 fc (Measured)
ft (Measured)
Prediction
0.01 0.1
1E-07 0.00001 0.001 0.1 10 1000 1E-07 0.00001 0.001 0.1 10 1000
Reduced strain rate [1/s] Reduced strain rate [1/s]
GAC
100
fc or ft [MPa]

10

1
fc (Measured)
ft (Measured)
Prediction
0.1
0.00001 0.001 0.1 10 1000
Reduced strain rate [1/s]

Figure 7-29 Prediction of compressive and tensile strength (u0 and γ are constant)

191
Chapter 7 Monotonic Uniaxial Tension and Compression Tests

Similarly, the mean percent relative error (MPRE) and the correlation coefficients R2 are
calculated to estimate the accuracy of the prediction. The values of MPRE and R2 for all
the mixtures are presented in Table 7-15 and 7-16, respectively.

Table 7-15 Mean percentage relative errors for each mixture


DAC EME PAC DAC SMA ACRe
GAC Total
0/8 0/14 0/16 0/16 0/11 0/4
Compressive
9.4% 9.9% 26.1% 8.9% 16.3% 11.3% 32.7% 16.2%
strength
Tensile strength 15.9% 29.8% 33.2% 18.9% 50.2% 6.7% 21.3% 25.1%
Overall 20.6%

Table 7-16 Correlation coefficients for each mixture


DAC EME PAC DAC SMA ACRe
GAC Total
0/8 0/14 0/16 0/16 0/11 0/4
Compressive
0.92 0.96 0.92 0.99 0.96 0.88 0.64 0.92
strength
Tensile strength 0.73 0.54 0.86 0.77 0.69 0.98 0.72 0.86
Overall 0.94

When it is assumed that the parameters u0 and γ are constant, the predictions have a
higher MPRE value and a lower R2 value compared to the linear relations, especially for
the tensile strength. The MPRE value for the tensile strength prediction increases from
14.6% to 25.1 %. With regard to the correlation coefficients, in most cases the R2 values
are larger than 0.7, except for the mixture EME0/14 and SMA0/11. The same conclusions
also can be derived from the comparison of prediction and measurement results (see
Figure 7-30). It indicates that when taking the average value for the parameter u0 and γ,
the accuracy of the unified model becomes lower, but it still gives acceptable predictions
for both compressive and tensile strength.

60 Line of
DAC0/8 EME 0/14 PAC 0/16 +20%
DAC 0/16 SMA 0/11 ACRe 0/4 equality
50
GAC
-20%
Predicted fc [MPa]

40

30

20

10

0
0 10 20 30 40 50 60
Measured fc [MPa]

(a) Compressive strength

192
Chapter 7 Monotonic Uniaxial Tension and Compression Tests

10
DAC0/8 EME 0/14 PAC 0/16 Line of
+20%
DAC 0/16 SMA 0/11 ACRe 0/4 equality
8 GAC
-20%

Predicted ft [MPa]
6

0
0 2 4 6 8 10
Measured ft [MPa]

(b) Tensile strength

Figure 7-30 Comparison of predicted and measured values

When it is assumed that all data sets are normally distributed around the line of equality,
the probability density function of the MPRE can be expressed as follows.
2
1  x− µ 
− 
f ( x; µ; σ 2 ) =
1 2 σ 

e (29)
σ 2π

∑( P
n
− Pm )
2
mi
(30)
σ= i =1

n −1

Where: σ : sample standard deviation;


Pm : average value of measured data;
Pmi : measured data;
n : number of test data;
x : a variable of MPRE, %;
µ : expected value of MPRE, in this study µ=0;

The normal distributions of each data set are plotted in Figure 7-31 and the confidence
intervals at 80 % and 95 % confidence levels are presented in Table 7-17.

193
Chapter 7 Monotonic Uniaxial Tension and Compression Tests

3
f ( x; µ;σ 2 ) Com pression_linear relation
Tension_linear relation
2.5
Com pression_average value
Tension_average value
2

1.5

0.5

0
-1.5 -1 -0.5 0 0.5 1 1.5

Figure 7-31 Normal distribution for compressive and tensile strength

Confidence intervals are normally used to indicate the reliability of a prediction. At the
same confidence level, the confidence interval of MPRE of the compressive strength is
smaller than that of tensile strength. The reason is that, as mentioned before, the
specimen in a MUC test is more or less uniformly compressed (if a good friction
reduction system and a reliable temperature cabinet are used). This leads to a rather high
reproducibility of the experimental results, while for the uniaxial tension test, the failure
is localized and the failure plane highly depends on a local defect in the sample. This
leads to a lower reproducibility of the experimental results. When using the linear relation
to determine the model parameters, with a probability of 80% of the confidence intervals,
MPRE of the predicted compressive and tensile strength are within ± 21.1% and ± 24.7%,
respectively. In terms of the determination of the parameters u0 and γ, when taking
average values, the variation of the unified model becomes larger compared to the case
when using the linear relation, especially for the tensile strength.

Table 7-17 80% and 95% level of confidence intervals of MPRE


Equation for
Compressive strength Tensile strength
u0 and γ
σ1 80% l.o.c.2 95% l.o.c. σ 80% l.o.c. 95% l.o.c.
Linear
16.4% ± 21.1% ± 32.2% 19.3% ± 24.7% ± 37.8%
relation
Average
20.6% ± 26.4% ± 40.4% 36.3% ±46.5% ±71.1%
value
σ1: sample standard deviation
l.o.c.2: level of confidence

194
Chapter 7 Monotonic Uniaxial Tension and Compression Tests

7.9 Conclusions
Based on the results discussed in this chapter, the following conclusions and
recommendations can be made.
1. To make an efficient test plan, “the central composite rotatable design techniques”
was used to determine the combination of the temperature and strain rate. For the
MUC test, this method significantly reduces the number of tests and the results
measured at various test conditions are able to describe a reliable relationship
between compressive strength and test conditions. For the MUT test, it is different,
because the tensile strength is very sensitive to the strain rate at low temperature.
So it is difficult to select an appropriate range of the test conditions. The data
points obtained from the central composite rotatable design normally do not
distribute uniformly in a wide range, which can not give a full description of the
tensile strength evolution. Therefore from the author's viewpoint, before using
the central composite rotatable design methods, some preliminary tests are needed
to determine an appropriate range of test conditions.
2. The tensile and compressive strength are, like the mixture stiffness highly
dependent on the temperature and loading rate. When the loading rate or
temperature becomes very high or low, the material properties tend to reach a
threshold value. This behavior of asphalt mixtures can successfully be described
with the unified model.
3. The compressive strength and strain at failure are higher than those in tension,
while the tangential stiffness in tension is higher than that in compression at high
reduced strain rate. For all the material properties, the slopes of the curves of
tensile strength and failure strain vs. reduced strain rate are normally larger than
those in compression. This indicates that the material properties in tension are
more sensitive to the temperature and strain rate.
4. The model parameters of the unified model depend on material properties, such as
aggregate gradation, mixture composition and binder characteristics. Based on the
collected compression and tension test results of 7 different types of asphalt
mixtures, the unified model was simplified. Equations to predict the model
parameters Ph, u0 and γ were developed for compressive and tensile strength. With
the temperature susceptibility function obtained from the stiffness test and the
prediction equations of the unified model parameters (see Table 7-11), the
compressive and tensile strength at a certain strain rate and temperature can
directly be calculated from the stiffness of the mixture, volume content of the
bitumen, volume content of air voids and the coefficient of curvature of the
aggregate gradation. Based on a statistical analysis, it is concluded that the
predictions made by means of the simplified unified model show a good
agreement with the measured value; this is especially the case for the compressive
strength predictions.
5. For further simplification, the parameter u0 and γ can be considered constant (see
Table 7-14). Although the variation of the unified model becomes larger, the
predictions of the compressive and tensile strength are still quite acceptable.

195
Chapter 7 Monotonic Uniaxial Tension and Compression Tests

References
Diamond, W.J, “Practical Experiment Design for Engineers and Scientists,” Wiley, New
York, U.S.A, 2001.

Erkens, S.M.J.G. and Poot, M.R., The Uniaxial Compression Test Asphalt Concrete
Response (Acre), Report 7-98-117-4, Delft University of Technology, the Netherlands,
2000.

Erkens, S.M.J.G., Asphalt Concrete Response (ACRe)-Determination, Modelling and


Prediction, PhD Thesis, Delft University of Technology, The Netherlands, 2002.

Jacobs, M.M.J., Crack Growth in Asphalt Mixes, PhD Thesis, Delft University of
Technology, the Netherlands, 1995.

Jansen, J.P.M. Characterising EME Enrobé à Modele Elevé, Master thesis, Civil
Engineering and Geosciences, Delft University of Technology, 2002.

Jenkins, K.J., Mix Design Considerations for Cold and Half-Warm Bituminous Mixes
with Emphasis on Foamed Bitumen, PhD Thesis, University of Stellenbosch, South
Africa, 2000.

Molenaar, A.A.A. Road Material, Part III, Asphalt Materials, lecture notes e52 (CT4850),
Delft University of Technology, The Netherlands, 1994.

Muraya, P.M., Permanent Deformation of Asphalt Mixtures, PhD Thesis, Delft


University of Technology, the Netherlands, 2007.

Pellinen, T.K. and Witczak M.W., “Stress Dependent Master Curve Construction for
Dynamic (Complex) Modulus,” Annual Meeting Association of Asphalt Paving
Technologists, Colorado Springs, Colorado, USA, March 2002.

Pramesti, F. P., Molenaar, A. A. A., and van de Ven, M. F. C., Fatigue Cracking of
Gravel Asphalt Concrete; Cumulative Damage Determination, Proceedings of the 7th
International RILEM Conference on Cracking in Pavements, Delft, The Netherlands,
June 20–22, 2012.

Pramesti, F. P. and Poot M.R., Special report; Correction on compression test result due
to friction reduction test set-up (frame test+friction reduction system), Delft University of
Technology, 2013.

Robinson, G.K. Practical Strategies for Experimenting, Sussex, England, 2000.

Scarpas, A., Gurp, C.A.M.P. van, Al-Khoury, R.I.N. and Erkens, S.M.J.G., “Finite
Element Simulation of Damage Development in Asphalt Concrete Pavements,” 8th
International Conference on Asphalt Concrete Pavements, Seattle, Washington, U.S.A.,
1997.

196
Chapter 8 Yield Surface and Fatigue Tests

Chapter 8 Yield Surface and Fatigue Tests

8.1 Introduction
In this chapter, the yield surface concept is applied to analyze different fatigue test results.
By means of the hierarchical single surface (HISS) criterion proposed by Desai [1986],
the yield surface at a certain stress state can be determined by using the unified models
given in Chapter 7. The distance between the stress state at the critical location and the
corresponding yield surface is an indication of the safety margin to failure and it is
considered to be related to the number of load repetitions until fatigue damage. A new
parameter R∆ is introduced as safety factor. By combining the fatigue results described in
Chapter 5 and the unified models given in Chapter 7, a new fatigue analysis approach
was developed to reduce or exclude the influence of the test type and specimen size on
the fatigue results.

8.2 Material Model Concept

The response of a material is normally described by a material model consisting of a


yield surface and constitutive relations. The yield surface is the collection of all stress
combinations that cause the transition from one type of response to the other. The
constitutive relation for each separated response describes the relation between stress and
strain. Figure 8-1 shows a simple example of the elastic-plastic material model.

σ
fy

εe εp ε
Figure 8-1 Relation between stress and strain for the elastic-plastic model

In this model, the material is subjected to a uniaxial load. When the applied stress is
below its yield strength, the material behavior is purely linear elastic and governed by
Hooke’s law. When the yield strength is reached, the material behaves ideally plastic
associated with the development of irreversible strain. The deformation increases
continuously at the same stress level. It means that the constitutive relation has changed.
The constitutive equations for these two parts are expressed as follows [Hill, 1950]
[Erkens, 2002]:

197
Chapter 8 Yield Surface and Fatigue Tests

σ = E ⋅ εe If σ < fy (Elasticity) (8-1)


σ = E ⋅(ε − εp ) If σ = fy (Plasticity) (8-2)
ε = εe + ε p (8-3)
∂f ( σ )
εɺ p = λ (8-4)
∂σ

Where: σ : applied stress, [MPa];


E : elastic modulus, [MPa];
ε : total strain, [m/m];
εp : plastic strain, [m/m];
εe : elastic strain, [m/m];
εɺ p : plastic strain rate, [s-1];
f(σ) : yield function of stress and strain;
λ : plastic multiplier, λ=0 for elasticity, λ>0 for plasticity.

In the axial case, the yield surface is a point (fy, εe), which separates the responses into an
elastic part and a plastic part. In Figure 8-2, the yield surfaces are represented as a ellipse
and a surface in a two-dimensional and a three-dimensional principal stress space,
respectively [Zienkiewicz, O.C., 1999]. If the yield function f(σ) is negative, the stress
state is within the circle or the surface and the response is elastic. If the yield function is
equal to zero, the stress state is positioned on the circle or the surface and the response is
plastic. States of stress outside the yield surface do not exist.

σ1 σ1

σ2 σ2
σ3

(a) (b)
Figure 8-2 Yield surface in the two-dimensional (a) and three-dimensional (b) principal
stress space

Asphalt concrete is a complicated material exhibiting both visco-elastic and visco-plastic


behavior. Under uniaxial monotonic loading to failure in the displacement controlled
mode, three phases can be observed in the stress-strain curve being the elastic, hardening
and softening phase (see Figure 8-3 [Erkens, 2000]). In the beginning, the material
behaves linear-elastic (elastic part). At point 1, the non-linear behavior starts and the
slope of the curve becomes smaller, but the load carrying capacity is still increasing

198
Chapter 8 Yield Surface and Fatigue Tests

(hardening part). The nonlinear deformations are basically plastic, in other words upon
unloading only the portion of elastic strain εe can be recovered from the total strain. The
material is weakened by internal microcracks. After reaching the peak value (point 2), the
stress decreases and the softening part starts. Microcracking continues and macrocracks
will develop.

Hardening
2
Stress

1
Softening
Elastic

Strain

Figure 8-3 Material response of asphalt concrete under uniaxial loading [Erkens, 2002]

In the three-dimensional situation, the material model describing the response of asphalt
concrete was developed in the framework of the ACRe project [Scarpas, 1997] [Erkens,
2002] [Liu, 2003]. The hierarchical single surface (HISS) criterion proposed by Desai
[1986] was utilized as the response function of the yield surface. In the stress invariant
space it is defined as Equation 8-5:

  I − R n  I1 − R  
2

 −α  1
 +γ   
J2   Pa   Pa   (8-5)
f = − =0
(1 − β cos ( 3θ ) )
2
P a

I1 = σ xx + σ yy + σ zz (8-6)
1
J 2 = (σ 1 − σ 2 ) + (σ 2 − σ 3 ) + (σ 1 − σ 3 ) 
2 2 2

6 
(8-7)
= (σ xx − σ yy ) + (σ yy − σ zz ) + (σ zz − σ xx )  + τ xy2 + τ yz2 + τ zx2
1 2 2 2

6 
J 3 = (σ xx − p ) (σ yy − p ) (σ zz − p ) + 2τ xyτ yzτ zx
(8-8)
− (σ xx − p )τ yz2 − (σ yy − p )τ zx2 − (σ zz − p )τ xy2
3 3 J3
cos ( 3θ ) = ⋅ 3 (8-9)
2
( J2 )2
σ xx + σ yy + σ zz
p= (8-10)
3

199
Chapter 8 Yield Surface and Fatigue Tests

Where: σi : i-th principal deviator stress, [MPa];


I1 : the first stress invariant;
J2 : the second deviatoric stress invariant;
J3 : the third deviatoric stress invariant;
p : the isotropic stress, [MPa];
θ : lode angle;
Pa : atmospheric pressure, 0.1 MPa;
α, β, γ, n, R : model parameters.

Desai's response surface exhibits a spindle shape in the (I1, √J2, lode θ) plane (see Figure
8-4). The size, shape and position of the yield surface are controlled by the model
parameters α, β, γ, n and R, which are related to the properties of the material and can be
directly determined by means of laboratory tests [Erkens et al., 2002].

Figure 8-4 Schematic of Desai response surface in the (I1, √J2, lode θ) plane
[Erkens, 2002]
Stress

Strain
α=C 0<α<C α=0

Figure 8-5 Influence of model parameter α on the response surface in I1-√J2 space
[Erkens, 2002] [Medani, 2006]

By plotting the model in the I1-√J2 space, the yield surface changes from 3D to 2D.
Figure 8-5 shows the influence of the model parameter α on the yield surface in the I1-√J2
space. It can be seen that the parameter α determines the size of the response surface. The
size of the surface increases with decreasing α value. So the hardening response of the

200
Chapter 8 Yield Surface and Fatigue Tests

material is controlled by the parameter α. In the elastic part, the parameter α has a
constant value. Once the hardening response occurs, the value of α starts to decrease. At
the peak stress, the parameter α reduces to zero and the yield surface reduces to a straight
line in the I1-√J2 space.

The model parameter γ controls the slope of the yield surface in the I1-√J2 space, which
increases with increasing γ (see Figure 8-6a). The parameter γ is stress-state independent,
but it can vary as a function of temperature and loading rate. The parameter n determines
the apex of the surface in the I1-√J2 plane. The apex is defined as the point on the
response surface where the tangent is a horizontal line (Figure 8-6b), indicating a fully
deviatoric state of stress. It expresses the state of stress after which the material starts to
dilate [Scarpas et al., 1997] [Liu, 2003]. The parameter R represents the triaxial strength
of the material in tension, which is an indication of the cohesion. For cohesionless
materials R=0. For increasing R values, the response surface moves on the positive I1 axis
(Figure 8-6c).

(a) (b)

(c)
Figure 8-6 Influence of the model parameters on the response surface in the I1-√J2 space
[Medani, 2006]

The parameter β determines the shape of the model on the octahedral-plane. For β=0, it is
a circle and with increasing β it becomes triangular. To simplify the model, it is assumed
that the cross section on the octahedral-plane is a circle (β=0). Then the response surface
reduces from Equation 8-5 to:

201
Chapter 8 Yield Surface and Fatigue Tests

J2   I − R n  I1 − R  
2

f = 2 −  −α  1
 +γ   =0 (8-11)
Pa   Pa   Pa  
 

8.3 Determination of Model Parameters

As described in the previous Section, the parameters α and γ are controlled by the
hardening and softening response, respectively. The parameter R is the tensile strength of
the material when the deviatoric stress is zero. The parameter n determines the shape of
the yield surface. In this Section, the model parameters in Equation 8-11 will be
determined using the results obtained from the MUC and MUT test presented in Chapter
7. Thus the model parameters can be expressed as function of the strain rate εɺ and
temperature T. The methodology was originally presented by many researchers [Scarpas,
1997] [Erkens, 2002] [Liu, 2003].

In the case of the uniaxial states of stress, the expression of the yield surface can be
simplified as follows:
σ1 = σ ,σ 2 = σ 3 = 0 ⇒
I1 = σ1 + σ 2 + σ 3 = σ (8-12)
1 1
J 2 =  (σ 1 − σ 2 ) + (σ 2 − σ 3 ) + (σ 1 − σ 3 )  = σ 2
2 2 2
(8-13)
6   3
Substitution in Equation 8-11:
σ2   σ − R n σ − R  
2

=  −α   +γ    (8-14)
3Pa2   Pa   Pa  
 

8.3.1 Model Parameters R and γ

At the peak stress, the hardening parameter α is zero and the expression for the yield
surface is simplified as follows:

1
f c2 = γ ( − f c − R )
2
(8-15)
3
1 2
f t = γ ( ft − R )
2
(8-16)
3

Then the solutions for the parameter R and γ are obtained by substituting Equation 8-16
in Equation 8-15:
2 fc ⋅ ft
R= (8-17)
fc − ft
(f − ft )
2

γ =
c
(8-18)
3( f + ft )
2
c

202
Chapter 8 Yield Surface and Fatigue Tests

The parameters R and γ are the intercept and slope of the yield surface in the I1-√J2 plane
(see Figure 8-7). Since the uniaxial tensile strength ft and the uniaxial compressive
strength fc relate to the strain rate and temperature and can be calculated by the unified
models of compressive and tensile strength determined in chapter 7, the parameters R and
γ can also be expressed as functions of the strain rate and temperature (see Figure 8-8 and
Figure 8-9).

0.001_0°C 0.01_0°C 0.1_0°C 0.001_20°C 0.01_20°C


0.1_20°C 0.001_40°C 0.01_40°C 0.1_40°C
25

20
γ
√J 2 [MPa]

15

10

0
-50 -40 -30 -20 -10 0 10 20 30
I1 [MPa] R

Figure 8-7 Yield surface at the peak stress at different strain rates and temperatures

30

25 0°C

20 10°C
R [MPa]

20°C
15

30°C
10

40°C
5

0
0.0001 0.001 0.01 0.1 1
Strain rate [1/s]

Figure 8-8 The parameter R as a function of strain rate and temperature

203
Chapter 8 Yield Surface and Fatigue Tests

0.3

0°C
0.25
10°C
0.2
γ 20°C
0.15
30°C
0.1
40°C
0.05

0
0.0001 0.001 0.01 0.1 1
Strain rate [1/s]

Figure 8-9 The parameter γ as a function of strain rate and temperature

The value of the parameter γ also changes during softening. At the stress peak, the
parameter γ has a maximum value. After the initiation of the degradation response, the
evolution of γ can be expressed as a decaying function of equivalent post fracture strain
ξpf, the strain rate and temperature (Equation 8-19). The effective post-fracture plastic
strain is defined based on the post-fracture plastic strain increment ξpf. The proposed
relationship between γ and the effective post fracture plastic strain ξpf (Figure 8-10) is
given by:

0
Figure 8-10 Evolution of the parameter γ in the softening response [Medani, 2006]

γ = γ min + ( γ max − γ min ) exp −η1 (ξ pf ( )


η2
) (8-19)

γ min = γ min1 ( εɺ ⋅ β T )
γ min 2
+ γ min 3 (8-20)

204
Chapter 8 Yield Surface and Fatigue Tests

(f − ft )
2

γ max =
c
(8-21)
3( f + ft )
2
c

η1 = η11 ( εɺ ⋅ β T )
η12
(8-22)
β T = exp  −Ts (T − T0 ) 

Where: γmin : value of γ at the point of complete annihilation;


γmax : value of γ at the point of peak stress;
γmin1, γmin2, γmin3 : material constants;
η1, η2, η11, η12 : material constants;
ξ pf : effective post fracture plastic strain;
εɺ : strain rate;
βT : temperature susceptibility function;
T0 : reference temperature;
Ts : regression constant.

8.3.2 Model Parameter n

The model parameter n is related to the dilation in the specimen. Dilation is the increase
in volume that results from the opening of internal cracks. At the beginning of a
compression test, the axial strain is larger than the radial strain, which leads to a decrease
in volume. This is caused by the closure of initial flaws or the reorientation of grains.
Once dilation starts, the volume starts to increase. Therefore, the onset of dilation is
related to the point at which the volumetric strain changes from decreasing to increasing
[Erkens, 2002] [Liu, 2003]. This point can be determined by means of the axial versus
volumetric strain plot and the axial strain versus axial stress plot, as described in Chapter
7. During this dilation phase, whole groups of grains slide past each other [Sitters, 1998].

2 (8-23)
n=
 σ dil
2 
1 − 
 3γ (σ − R )2 
 dil 

The compressive stress at which dilation starts (σdil) is expressed as a function of strain
rate and temperature and is described using the unified model. In addition, the model
parameters γ and R are expressed as a function of strain rate and temperature. Thus, the
model parameter n can be expressed as a function of strain rate and temperature.

8.3.3 Model Parameter α

The model hardening parameters can be determined directly from the expression of the
response surface for each state of stress (Equation 8-24).

205
Chapter 8 Yield Surface and Fatigue Tests

2
J I −R
− 22 + γ  1 
Pa  Pa 
α= n
(8-24)
 I1 − R 
 
 Pa 

Based on Equation 8-24, values for α for all states of stress throughout the stress strain
curves can be found. As a hardening parameter, the value of parameter α varies between
the onset of non-linearity and peak strength. The evolution of the parameter α during the
hardening response is shown in Figure 8-11. The value of α evaluated at the stress where
the plastic behavior (σ=σplas) starts is given by:

σ plas σ − R 
2 2

− 2
+ γ  plas 
3Pa  Pa 
α0 = n
(8-25)
 σ plas − R 
 
 Pa 

Like the strength, σplas also depends on the strain rate and temperature and this
dependency is described using the unified model. In addition, the model parameter R is
expressed as a function of strain rate and temperature too. Thus, the parameter α0 also
depends on the strain rate and temperature.

0
Figure 8-11 Evolution of the parameter α in the hardening response [Medani, 2006]

The degradation of α from α0 at the onset of plasticity to zero at the peak stress can be
expressed as a decaying function of the effective plastic strain, strain rate and temperature
(see Figure 8-11). The proposed relationship is given by Equation 8-26:

206
Chapter 8 Yield Surface and Fatigue Tests

α0
α= (ξ − ξ ) exp ( −κ α ξ ) (8-26)
ξ lim lim
ξ lim = ξ lim1 ( εɺ ⋅ β T )
ξ lim 2
(8-27)

( (
κ α = κ α 1 1 − exp − ( εɺ ⋅ β T )
κα 2
)) (8-28)

Where: ξ : value of the effective plastic strain;


ξlim : value of the effective plastic strain at peak stress;
ξlim1, ξlim2 : material constants;
κα, κα1, κα2, : material constants that control the rate of hardening;

8.3.4 Determination of Yield Surface

The value of the parameter α is equal to zero when the stress is at the peak value. The
yield surface reduces to a straight line in the I1-√J2 space and is expressed as follows:
2
J2 I −R
=γ  1  ⇒ J 2 = γ ( I1 − R )
2
2
(8-29)
Pa  Pa 

The yield surface can be determined based on the results of the MUC and MUT test. In
the uniaxial test, the stress is only applied in one direction. As mentioned above, the
compressive and tensile strength can be determined by means of the unified model at a
given strain rate and temperature, then the parameter γ and R are calculated by Equation
8-17 and 8-18. The yield surface in the stress invariant space can be directly determined
by the two points on the yield surface representing the tensile and compression strength
[Ning, 2013a].

 1   1 2
In stress invariant:  I1 = f c , J 2 = f c2  and  I1 = ft , J 2 = ft  (8-30)
 3   3 
  

where
fc = compressive strength, MPa;
ft = tensile strength, MPa.

The strain-field however is three dimensional and this has to be taken into account when
defining the yield surface in the strain invariant space. The parameter γ and R are
determined by the strain states at the maximum stress in both compression and tension.
For simplicity, the used asphalt mixture was assumed to be laterally isotropic, and
therefore the values of the strain in two horizontal directions ε2 and ε3 were considered to
be equivalent and both of them are equal to εc,h. The yield surface line also can be
determined from the tensile and compressive failure strain, expressed by Equation 8-31
[Ning, 2013b].

207
Chapter 8 Yield Surface and Fatigue Tests

 1 2 
In strain invariant:  I1 = εc ,v + 2εc ,h , J 2 =
 3
( εc ,v − εc ,h ) 

 
 1 2 
and  I1 = εt ,v + 2εt ,h , J 2
 3
(εt ,v − εt ,h ) 

(8-31)
 

where
εc,v = vertical compressive strain when stress is at the peak, µm/m;
εc,h = horizontal compressive strain when stress is at the peak, µm/m;
εt,v = vertical tensile strain when stress is at the peak, µm/m;
εt,h = horizontal tensile strain when stress is at the peak, µm/m.

Figure 8-12 shows an example of the yield surface in the stress invariant space. At a
certain stress state, the stress state point (I1, √J2) is related to the magnitude of the applied
load. The corresponding yield surface represents the maximum load carrying capacity of
the material. To represent the safety margin to failure of the material at a certain stress
state, a new parameter, safety factor R∆ is defined using Equation 8-32 [Ning, 2013b].
This approach is also valid in the strain invariant space. When there is no applied stress,
the stress state point is at the original point (0, 0), the value of the safety factor R∆ is
equal to zero; when the applied stress reaches the strength of the material, the stress state
point is on the yield surface and the value of the safety factor R∆ is equal to 1.

∆i
R∆ = (8-32)
∆ tot

Where: ∆i : distance between original point and stress state point;


∆tot : distance between a data point and the yield surface along
the line through original point and stress state point.

4
√J 2 [MPa]

( ft , 1 )
( fc , 1 2 ) ft 2
f c 3 3
3

2
( I1 , J2 )
1
∆i ∆tot
0
-10 -8 -6 -4 -2 0 2 4 6
I1 [MPa]

Figure 8-12 Yield surface determined by uniaxial tension and compression test results

208
Chapter 8 Yield Surface and Fatigue Tests

8.4 Critical Location

To obtain the yield surface at the critical location, the stress-strain distributions of the
specimen should be taken into account. As discussed in Chapter 5, the stress-strain field
is uniform in the UT/C test and the critical stress is reached in the complete cross area.
The cracks might occur anywhere inside the specimen. For the inhomogeneous tests, the
maximum stress only occurs in the critical location of the specimen. For example, in the
4PB test, the maximum strain occurs at the surface in the midsection of the beam.

(a) Horizontal stress (b) Vertical stress

(c) Horizontal strain (d) Vertical strain


Figure 8-13 Stress and strain distributions of IT specimen

For the indirect tensile fatigue test, the stress state of the specimen is more complicated.
A repeated compressive loading is applied along the vertical diameter of a cylindrical
specimen. The vertical load produces both a vertical compressive stress and a horizontal
tensile stress at any point in the tested specimen. The magnitude of the compressive and
tensile stresses change along the diameter. The centre of the specimen with a maximum
horizontal tensile stress is generally supposed to be the critical location. Figure 8-13
illustrates the horizontal and vertical stress distribution simulated by the finite element
(FE) program ABAQUS. It is clear that the horizontal tensile stress reaches the maximum
value at the centre and the maximum horizontal strain occurs at the place between the

209
Chapter 8 Yield Surface and Fatigue Tests

centre and the edge of the specimen, because the horizontal strain is not only caused by
the horizontal tensile stress, but also by the vertical compressive stress.

Assuming that the specimen is homogeneous, isotropic and behaves linear elastic, the
solutions of the horizontal stress, vertical stress and horizontal strain along the vertical
diameter are derived by Hondros [1959].

  y2    
 1 − 2  sin 2θ y2
 1 + 
σx ( y) =
2P   R  − arctan  R 2 tanθ  (8-33)
πad   y2 y4   y2 
 1 − 2 2 cos 2θ + 4   1− 2 
  R R   R  
  y2    
  − y2
2 
1 sin 2θ +
 1 
σy ( y) = −
2P   R  + arctan  R 2 tan θ  
 (8-34)
πad  y  
2 4 2
y  y
 1 − 2 2 cos 2θ + 4   1− 2 
  R R   R  
1
εx ( y ) = σ x ( y ) − νσ y ( y )  (8-35)
E

Where: P : applied load, N;


a : loading strip width, mm;
d : thickness of specimen, mm;
R : specimen radius, mm;
θ : half the top angle of the triangle between loading strip
and specimen centre;
y : distance to the centre of the specimen, mm;
E : Young’s modulus, MPa;
v : Poisson’s ratio

Figure 8-14 shows the comparison of the stress-strain distribution calculated by the
Hondros' equations and the FE model. The vertical and horizontal stresses obtained by
these two methods are very close. The maximum horizontal strain occurs at the locations
y = ± 36 mm from the centre of the specimen and the value calculated by ABAQUS is a
little smaller than the results calculated by Equation 8-35. A relatively large area (from y
= 36 to y = -36) is created in which the stress-strain field is more or less uniform. This
might be the reason why the size effect on the fatigue life is not significant for IT test.

210
Chapter 8 Yield Surface and Fatigue Tests

60 IT specimen
50
40
horizontal σx_Eqn. 30

Vertical diameter [mm]


vertical σy_Eqn. 20
horizontal σx_Abaqus
10
vertical σy_Abaqus
0
-8 -7 -6 -5 -4 -3 -2 -1 -10 0 1 2
-20
-30
-40
-50
-60
Stress [MPa]

(a)

60
IT specimen
50
40
horizontal εx_Eqn. Point A
30 (0, 36)
vertical diameter [mm]

horizontal εx_Abaqus 20
10
0
-250 -200 -150 -100 -50 -10 0 50 100
Point O
-20
(0, 0)
-30
-40
-50
-60
Strain [µm /m ]

(b)
Figure 8-14 Stress (a) and strain (b) distributions along vertical diameter of a 100 mm
diameter IT specimen

The stress-strain state at the locations of y = 0 (point O), y = 36 (point A) are analysed for
the IT specimen. The horizontal and vertical stress and the horizontal strain at each point
are presented in Table 8-1. It is clear that the horizontal strain at point A is around 27 %
higher than that at the centre point. Therefore in this chapter the point O and point A are
considered as the critical locations for the IT test.

211
Chapter 8 Yield Surface and Fatigue Tests

Table 8-1 Stress and strain at different locations for size 1.0 and 1.5
Point O Point A
Specimen Force level
Sample code [N] σxx σyy εxx σxx σyy εxx
size
MPa MPa µm/m MPa MPa µm/m
I-4-3 2218 0.45 -1.38 50 0.28 -3.09 64
I-3-1 2428 0.48 -1.46 54 0.30 -3.27 68
I-1-4 2819 0.57 -1.73 61 0.36 -3.88 78
Size1.0 I-1-9 3320 0.70 -2.13 77 0.44 -4.77 99
I-1-5 4213 0.88 -2.68 98 0.55 -6.01 124
I-1-7 5222 1.04 -3.15 115 0.65 -7.07 146
I-1-6 6108 1.26 -3.82 142 0.79 -8.57 181
I-3-4 6020 0.54 -1.66 60 0.32 -3.68 74
I-2-6 6518 0.60 -1.82 60 0.35 -4.03 74
I-4-4 7011 0.63 -1.93 76 0.37 -4.27 95
Size1.5 I-3-6 7513 0.68 -2.07 72 0.40 -4.58 90
I-2-7 8505 0.82 -2.48 87 0.48 -5.51 108
I-4-7 9491 0.86 -2.63 90 0.51 -5.84 111
I-4-6 9969 0.92 -2.79 95 0.54 -6.19 118

8.5 Determination of Yield Surface for Fatigue Test

As described in Section 8.3.4, the yield surface determined by the tension and
compression test results is a function of strain rate and temperature. During the UT/C and
4PB fatigue test, the specimen is alternately subjected to a tension and compression load.
Since the crack damage is mainly caused by tension, the tension loading is considered as
the critical load. During a fatigue test, the specimen is tested under a sine or haversine
cyclic loading, so the strain rate is not a constant but follows a cosine. To simplify the
calculation of the strain rate, the full sine signal is converted into a triangle signal with
the same maximum strain level and time duration (see Figure 8-15). The strain rate of this
alternating triangular signal is considered to be an acceptable approximation, computed
as follows:

ε0
εɺ = (8-36)
t

Where: ε0 : amplitude of the applied strain, m/m;


t : time duration when the strain changes from 0 to peak
value, s;

212
Chapter 8 Yield Surface and Fatigue Tests

200
150
100
50 εɺ εɺ

strain
0
-50 0 2 4 6 8 10 12

-100
-150
-200
Time [s]
Figure 8-15 Conversion from a sine signal to a triangle signal [Erkens, 2002]

Knowing the tensile and compressive strength as a function of temperature and strain rate,
the yield surfaces which are applicable for the stress or strain controlled mode fatigue test
can be determined by Equation 8-30 or 8-31. Figure 8-16 and 8-17 give the initial stress
states and their yield surfaces at different force levels in the IT fatigue test. It can be seen
that at the same location the data points determined by the initial stress states are aligned
and the straight line also goes through the origin (0, 0). With increasing load, the data
points of the initial stress states increase along this path. Based on the plots of the stress
state point and its yield surface line, the safety indicator R∆ can be calculated by Equation
8-32.

IT-size1.0@point O
16
14

12
10
ΓJ2 [MPa]

6
4
2.2kN 2.8kN ∆tot
3.3kN 4.2kN 2 ∆i
5.2kN 6.1kN
0
-30 -25 -20 -15 -10 -5 0 5 10
I1 [MPa]

Figure 8-16 Stress state and yield surface for each force level for IT_size 1.0 at point O

213
Chapter 8 Yield Surface and Fatigue Tests

IT_size1.0@point A
16
14

12

ΓJ2 [MPa] 10

8
6

2.2kN 2.8kN ∆tot 4


3.3kN 4.2kN
5.2kN 6.1kN
2 ∆i
0
-30 -25 -20 -15 -10 -5 0 5 10
I1 [MPa]

Figure 8-17 Yield surface lines for each force level for IT_size 1.0 at point A

The safety factor R∆ at the different locations of the IT specimen are compared in Figure
8-18 and Figure 8-19. The values of R∆ at point O and point A for the IT specimens with
the different sizes are presented in Table 8-2. It is clear that the R∆ value at point A is
around 35% larger than that at the centre point O at various force levels. It indicates that
although the maximum horizontal stress occurs at the centre of the specimen, point A
seems to be the weakest point for the IT fatigue test. In the fatigue analysis, therefore the
stress state at the point A should be taken into account instead of the stress state at the
centre point.

0.50
IT-size1@point O
0.40 IT-size1@point A

0.30
R∆

0.20

0.10

0.00
2.2kN 2.4kN 2.8kN 3.3kN 4.2kN 5.2kN 6.1kN
Force level

Figure 8-18 Comparison of the safety factor R∆ at point O and point A for the IT
specimen with size 1.0

214
Chapter 8 Yield Surface and Fatigue Tests

0.5
IT-size1.5@point O
0.4 IT-size1.5@point A

0.3

R∆
0.2

0.1

0
6.0kN 6.5kN 7.0kN 7.5kN 8.5kN 9.5kN 10.0kN
Force level

Figure 8-19 Comparison of the safety factor R∆ at point O and point A for the IT
specimen with size 1.5

Table 8-2 Stress and strain at different locations for size 1.0 and size 1.5
Specimen sample Point O Point A R ∆ ( A ) − R∆ ( O ) Fatigue
F [N]
size code εɺ [%/s] R∆ εɺ [%/s] R∆ R∆ ( O ) life Nf
I-4-3 2218 0.10 0.13 0.13 0.20 48.0% 1.8×106
I-3-1 2428 0.11 0.14 0.14 0.21 46.5% 7.4×105
I-1-4 2819 0.12 0.16 0.16 0.24 45.7% 1.0×106
Size 1.0 I-1-9 3320 0.16 0.19 0.20 0.27 42.9% 4.4×105
I-1-5 4213 0.20 0.23 0.25 0.33 40.3% 1.4×105
I-1-7 5222 0.23 0.27 0.29 0.37 38.6% 8.4×104
I-1-6 6108 0.28 0.31 0.36 0.43 36.5% 3.0×104
I-3-4 6020 0.12 0.16 0.15 0.23 45.2% 7.7×105
I-2-6 6518 0.12 0.17 0.15 0.24 41.7% 6.9×105
I-4-4 7011 0.15 0.17 0.19 0.25 45.9% 4.5×105
Size 1.5 I-3-6 7513 0.14 0.19 0.18 0.27 41.7% 3.4×105
I-2-7 8505 0.17 0.22 0.22 0.31 39.7% 1.6×105
I-4-7 9491 0.18 0.23 0.22 0.32 38.4% 1.3×105
I-4-6 9969 0.19 0.24 0.24 0.33 37.7% 9.4×104

8.6 Relationship between R∆ and Fatigue Life

Based on the analysis of the stress or strain states, the yield surface at various loading
levels in the fatigue test can be determined by Equation 8-30 and 8-31. Table 8-3 shows
the stress states and the corresponding yield surfaces for the stress-controlled fatigue tests
discussed in Chapter 5. The safety factor R∆ is determined from the plot of the stress
states and the yield surfaces in the I1-√J2 space, as shown in Figure 8-20, Figure 8-21,
Figure 8-22 and Figure 8-23.

215
Chapter 8 Yield Surface and Fatigue Tests

Table 8-3 Stress states and corresponding yield surfaces at the critical locations for the
stress-controlled fatigue tests at 5 ºC
Stress state Yield surface
Test Sample
σ1 σ2 fc ft R∆ Nf
type code R γ
[MPa] [MPa] [MPa] [MPa]
C-15-5 1.24 0 -23.5 6.23 17.0 0.112 0.204 1.5×106
C-16-9 1.28 0 -24.0 6.23 16.8 0.115 0.216 1.1×106
C-16-2 1.41 0 -24.2 6.23 16.8 0.116 0.234 8.4×105
UT/C-
C-15-7 1.49 0 -24.6 6.23 16.7 0.118 0.247 4.7×105
size1.0
C-15-8 1.55 0 -25.2 6.23 16.6 0.121 0.260 2.9×105
C-15-3 1.59 0 -25.7 6.23 16.5 0.124 0.281 1.6×105
C-15-10 1.82 0 -27.3 6.23 16.2 0.132 0.301 8.3×104
B-11-4 1.61 0 -25.3 6.23 16.6 0.122 0.362 2.4×106
B-11-1 1.82 0 -25.8 6.23 16.4 0.125 0.429 1.7×106
B-11-2 1.82 0 -25.7 6.23 16.5 0.124 0.432 1.9×106
4PB-
B-7-3 1.93 0 -26.1 6.23 16.4 0.126 0.473 9.1×105
size1.0
B-7-4 2.04 0 -26.0 6.23 16.4 0.125 0.511 7.3×105
B-7-1 2.13 0 -26.1 6.23 16.4 0.126 0.543 6.0×105
B-16-3 2.32 0 -26.7 6.23 16.3 0.129 0.624 5.8×105
I-4-3 0.28 -3.09 -18.5 6.23 18.8 0.082 0.199 1.8×106
I-3-1 0.30 -3.27 -18.9 6.23 18.6 0.085 0.206 1.0×106
I-1-4 0.36 -3.88 -19.7 6.23 18.2 0.090 0.236 7.4×105
IT-
I-1-9 0.44 -4.77 -21.3 6.23 17.6 0.100 0.273 4.4×105
size1.0
I-1-5 0.55 -6.01 -22.9 6.23 17.1 0.109 0.325 1.5×105
I-1-7 0.65 -7.07 -24.0 6.23 16.8 0.115 0.368 8.4×104
I-1-6 0.79 -8.57 -25.5 6.23 16.5 0.123 0.426 3.0×104
I-3-4 0.32 -3.68 -19.3 6.23 19.0 0.079 0.226 7.7×105
I-2-6 0.35 -4.03 -19.9 6.23 19.0 0.079 0.242 6.9×105
I-4-4 0.37 -4.27 -20.3 6.23 18.3 0.089 0.253 4.5×105
IT-
I-3-6 0.40 -4.58 -20.7 6.23 18.4 0.087 0.266 3.4×105
size1.5
I-2-7 0.48 -5.51 -22.0 6.23 17.9 0.095 0.305 1.6×105
I-4-7 0.51 -5.84 -22.4 6.23 17.8 0.096 0.319 1.3×105
I-4-6 0.54 -6.19 -22.8 6.23 17.7 0.098 0.333 9.4×104

Due to the existence of a vertical compressive stress at the critical location, the value of I1
for the IT fatigue test is negative and the differences of the stress states between the
different load levels are relatively large compared to the other two types of fatigue tests.

216
Chapter 8 Yield Surface and Fatigue Tests

UT/C-size1@5°C_ σ control 1.2


18

√ J2 [MPa]
1
16
0.8
14
0.6
12 1 1.5 2 2.5

√ J2 [MPa]
10 I1 [MPa]

8
6
C-15-5 C-16-9
C-16-2 C-15-7 4
C-15-8 C-15-3
2
C-15-10
0
-30 -20 -10 0 10 20
I 1 [MPa]

Figure 8-20 Stress states and yield surfaces for the UT/C test in stress controlled mode

4PB-size1@5°C_ σ control 1.6


18

ΓJ2 [MPa]
1.4
16 1.2
1
14
0.8
12 1 1.5 2 2.5 3
ΓJ2 [MPa]

I1 [MPa]
10
8
B-11-4 B-11-1 6
B-11-2 B-7-3 4
B-7-4 B-7-1
2
B-16-3
0
-30 -20 -10 0 10 20
I1 [MPa]

Figure 8-21 Stress states and yield surfaces for the 4PB test in stress controlled mode

IT_size1.0@5°C_ σ control
16

14

12

10
ΓJ2 [MPa]

6
I-4-3 I-1-4
4
I-1-9 I-1-5
I-1-7 I-1-6 2
I-3-1
0
-30 -25 -20 -15 -10 -5 0 5 10
I1 [MPa]

Figure 8-22 Stress states and yield surfaces for the IT specimen size 1.0 in stress
controlled mode

217
Chapter 8 Yield Surface and Fatigue Tests

IT_size1.5@5°C_ σ control
16

14

12

10

ΓJ2 [MPa]
8

6
I-3-4 I-2-6
I-4-4 I-3-6 4
I-2-7 I-4-7
2
I-4-6
0
-30 -20 -10 0 10
I 1 [MPa]

Figure 8-23 Stress states and yield surfaces for the IT specimen size 1.5 in stress
controlled mode

Similarly, the safety factor R∆ at a certain strain level for the strain-controlled fatigue
tests can be calculated from the strain state and the corresponding yield surface.
Examples of the I1-√J2 plots for the UT/C test with specimen size 1.0 at 20 ºC and 5 ºC
are shown in Figure 8-24 and 8-25, respectively. At 20 ºC the yield surface lines at
different strain levels show significant differences, because the unified model of the
failure strain at 20 ºC is more sensitive to the strain rate than at the relatively low or high
temperatures (T ≤ 5 ºC or T ≥ 35 ºC). Due to the Poisson effect, the strain at the critical
location for the UT/C and 4PB test changes in three dimensions under the axial loading.
So in the determination of the yield surface, the failure strains in three mutually
perpendicular directions are taken into account. The values of R∆ and the fatigue lives for
some strain controlled fatigue tests are listed in Table 8-4. ε1 means the initial tensile
strain along the loading direction, ε2 and ε3 are the strain in the other two directions
perpendicular to the strain ε1. For simplicity purposes, ε3 is considered equivalent to ε2 in
magnitude.
UT/C-size1@20ºC 200
15000
√J 2 [µm/m]

150
100
12000 50
0
√J2 [µm/m]

9000 0 50 100 150


I1 [µm/m]

6000
68µε 84µε
110µε 131µε
3000 155µε 172µε
199µε

0
-2000 -1000 0 1000 2000 3000
I1 [µm/m]

Figure 8-24 Strain states and yield surfaces for the UT/C test in strain controlled mode at
20 ºC

218
Chapter 8 Yield Surface and Fatigue Tests

Table 8-4 Strain states and corresponding yield surfaces at the critical locations for the
strain-controlled fatigue tests
Strain state Yield surface
Test type εt,v εt,h εc,v εc,h R∆ Nf,50
ε1 [µm/m] ε2 [µm/m]
[µm/m] [µm/m] [µm/m] [µm/m]
88 -26 3105 -1208 -13047 6140 0.043 1.6×106
117 -35 2569 -1071 -12507 5695 0.079 5.4×105
130 -39 2384 -1020 -12304 5531 0.101 3.0×105
UT/C-
136 -41 2310 -999 -12221 5465 0.111 2.8×105
size0.5
138 -41 2285 -992 -12191 5441 0.116 2.3×105
@20 ºC
146 -44 2196 -966 -12085 5357 0.132 1.3×105
168 -50 1996 -906 -11835 5162 0.179 9.0×104
186 -56 1862 -863 -11654 5022 0.288 5.6×104
68 -18 3899 -1396 -13785 6772 0.022 6.0×106
84 -25 3195 -1230 -13133 6213 0.040 1.9×106
UT/C- 110 -33 2666 -1097 -12610 5778 0.070 5.5×105
size1.0 131 -39 2365 -1015 -12284 5515 0.103 2.8×105
@20 ºC 155 -47 2109 -940 -11979 5274 0.149 1.1×105
172 -52 1962 -895 -11790 5126 0.187 7.2×104
199 -60 1780 -835 -11537 4933 0.255 3.2×104
68 -21 3642 -1336 -13550 6568 0.027 4.3×106
89 -27 3075 -1200 -13018 6116 0.046 1.3×106
UT/C-
115 -34 2593 -1077 -12533 5716 0.080 4.1×105
size1.5
133 -40 2351 -1011 -12267 5502 0.110 1.9×105
@20 ºC
161 -48 2055 -924 -11911 5220 0.171 7.0×104
182 -55 1888 -871 -11690 5050 0.225 4.0×104
69 -15 637 -183 -7433 2390 0.123 2.7×106
87 -19 624 -160 -7194 2280 0.149 1.5×106
UT/C- 94 -21 620 -153 -7114 2244 0.159 6.6×105
size1.0 105 -23 615 -143 -7003 2196 0.175 2.8×105
@5 ºC 118 -26 611 -134 -6893 2148 0.193 1.6×105
124 -27 609 -130 -6844 2127 0.201 1.6×105
136 -30 606 -123 -6754 2090 0.219 8.3×104
100 -30 2843 -1142 -12791 5927 0.045 5.7×106
119 -36 2531 -1061 -12466 5662 0.081 1.3×106
4PB-
140 -42 2266 -987 -12169 5424 0.116 8.5×105
size0.5
161 -48 2060 -926 -11918 5226 0.158 3.8×105
@20 ºC
179 -54 1909 -878 -11719 5072 0.202 2.2×105
201 -60 1770 -832 -11522 4922 0.258 2.1×105
87 -26 3134 -1215 -13075 6164 0.041 4.0×106
105 -31 2762 -1122 -12709 5860 0.061 1.9×106
4PB-
124 -37 2458 -1041 -12387 5598 0.089 6.6×105
size1.0
142 -43 2240 -979 -12138 5399 0.120 4.2×105
@20 ºC
161 -48 2057 -924 -11913 5222 0.159 3.1×105
180 -54 1908 -878 -11718 5071 0.203 1.9×105
74 -22 3478 -1298 -13398 6438 0.029 7.9×106
4PB- 88 -26 3100 -1207 -13043 6137 0.042 1.7×106
size1.5 93 -28 2995 -1181 -12941 6052 0.047 1.6×106
@20 ºC 103 -31 2801 -1132 -12748 5892 0.058 2.5×106
119 -36 2535 -1062 -12471 5666 0.080 5.8×105
89 -20 623 -158 -7170 2269 0.152 2.0×106
105 -23 615 -143 -7009 2198 0.174 1.6×106
4PB-
109 -24 613 -140 -6965 2179 0.181 1.3×106
size1.0
116 -26 611 -135 -6908 2155 0.190 3.4×105
@5 ºC
124 -27 609 -130 -6843 2127 0.202 4.7×105
135 -30 606 -123 -6761 2092 0.217 4.1×105

219
Chapter 8 Yield Surface and Fatigue Tests

UT/C-size1@5ºC
6000 150

√J2 [µm/m]
100
5000 50

0
4000
√J2 [µm/m] 0 50 100
I1 [µm/m ]
3000

2000
69µε 87µε
94µε 105µε 1000
118µε 124µε
136µε
0
-3000 -2000 -1000 0 1000 2000
I1 [µm/m]

Figure 8-25 Strain states and yield surfaces for the UT/C test in strain controlled mode at
5 ºC

For the fatigue test, each stress or strain level corresponds to a R∆ value and also a fatigue
life. Assuming that 50% stiffness reduction corresponds to a certain amount of damage, it
is possible to find a relationship between R∆ and the fatigue life.

1. Relation in double logarithm coordinates

The plots of R∆ versus the fatigue life for the strain-controlled and stress-controlled tests
are presented in Figure 8-26 and Figure 8-27. It can be seen that a straight line is found
on a double logarithm scale similar to the traditional fatigue line. Equation 8-37 shows
the relation between R∆ and fatigue life. All the coefficients for the different fatigue tests
are given in Table 8-5.

N f = r1 ⋅ R∆ r2 (8-37)

Where: Nf : fatigue life;


R∆ : safety factor, ratio of ∆i and ∆tot;
r1 and r2 : regression coefficients.

220
Chapter 8 Yield Surface and Fatigue Tests

Strain-controlled mode
1.E+08
UT/C-size0.5@20ºC UT/C-size1.0@20ºC
UT/C-size1.5@20ºC UT/C-size1.0@5ºC
4PB-size0.5@20ºC 4PB-size1.0@20ºC
1.E+07 4PB-size1.5@20ºC 4PB-size1.0@5ºC

Nf,50
1.E+06

1.E+05

1.E+04
0.01 0.1 1
R∆

Figure 8-26 Relation between R∆ and Nf,50 for the fatigue tests in strain-controlled mode

Stress-controlled mode
1.E+08
UT/C-size1.0@5ºC 4PB-size1.0@5ºC
ITT-size1.0@5ºC ITT-size1.5@5ºC
1.E+07
Nf

1.E+06

1.E+05

1.E+04
0.1 1
R∆

Figure 8-27 Relation between R∆ and Nf for all the fatigue tests in stress-controlled mode

Figure 8-26 shows that in the case of the 4PB fatigue test at 20 ºC, there is no significant
difference between the fatigue lines obtained from the specimens with different specimen
sizes. The results indicate that the size effect on the fatigue line can be excluded in this
new fatigue relation. At 20 ºC, the slopes of the fatigue lines obtained from the UT/C and
4PB tests are very close. With decreasing temperature, the slope of the fatigue line
increases. The reason is that the yield surface does not change significantly with the
strain rate at low temperature in general. Comparison with the results of the UT/C fatigue
test shows that the fatigue lines of the 4PB test are still higher, because the safety factor
R∆ represents the safety level at the critical location, being at the surface in the middle
part of the beam. The amount of damage in the rest of the beam is not as much as at the
critical location. Therefore the fatigue life obtained from the 4PB test is longer than from
the UT/C test when the value of R∆ is the same.

221
Chapter 8 Yield Surface and Fatigue Tests

Table 8-5 Regression coefficients of Nf-R∆ curve determined by the different fatigue tests
Test type r1 r2 R2
UT/C_size0.5@20 oC 3065 -1.93 0.93
UT/C_size1.0@20 oC 2045 -2.11 0.99
UT/C_size1.5@20 oC 2137 -1.99 0.99
Controlled-strain UT/C_size1.0@5 oC 5.22 -6.37 0.97
mode 4PB_size0.5@20 oC 3244 -2.17 0.99
4PB_size1.0@20 oC 3866 -2.02 0.97
4PB_size1.5@20 oC 1432 -2.41 0.78
4PB_size1.0@5 oC 101.5 -5.32 0.74
UT/C_size1.0@5 oC 14.34 -7.37 0.89
Controlled-stress 4PB_size1.0@5 oC 1.1×106 -3.11 0.89
mode IT_size1.0@5 oC 523 -4.99 0.99
IT_size1.5@5 oC 334 -5.12 0.99

With regard to the stress-controlled test, the fatigue life in the 4PB test is also longer than
those in the UT/C and IT test. The coefficients r1 (intercept) and r2 (slope) of the fatigue
lines are not comparable for the different fatigue tests (see Table 8-5). However, the data
points themselves show that UT/C fatigue data are coinciding rather well with the IT
fatigue data. The reason is that the FE modeling showed that the distributions of the
tensile stress and strain are relatively uniform in the middle part along the diametrical
direction. This critical area is around 70% of the whole cross area. Furthermore it is clear
that there is no difference between the fatigue lives of the two IT fatigue specimen sizes.

2. Limit value of the safty factor R∆

Figure 8-28 shows the plots of the fatigue life vs. the safety factor R∆ in the normal
coordinate system for both strain-controlled and stress-controlled tests. In general, Nf,50
increases with decreasing R∆ value. When the R∆ is lower than a certain value, the fatigue
life Nf tends to be inifinite. Taking into account the shape and the trend of the data points,
the following regression equation can be used to simulate the R-N (initial stress ratio
versus fatigue life) lines:

( )
b
− aN f
R∆ = R0 − R1 1 − e (8-38)

Where: R0, R1, a and b : regression coefficients.

222
Chapter 8 Yield Surface and Fatigue Tests

Strain-controlled mode
0.3
UT/C-size0.5@20ºC UT/C-size1.0@20ºC
0.25 UT/C-size1.5@20ºC UT/C-size1.0@5ºC
4PB-size0.5@20ºC 4PB-size1.0@20ºC
4PB-size1.5@20ºC 4PB-size1.0@5ºC
R∆ 0.2

0.15

0.1

0.05

0
0 2000000 4000000 6000000 8000000 10000000 12000000
Nf,50

(a) Strain-controlled mode

Stress-controlled mode
0.8
UT/C-size1.0@5°C 4PB-size1.0@5°C

IT-size1.0@5°C IT-size1.5@5°C
0.6
R∆

0.4

0.2

0
0 500000 1000000 1500000 2000000 2500000 3000000
Nf

(b) Stress-controlled mode


Figure 8-28 Initial stress ratio versus fatigue life Nf,50

The correlation coefficients R2 of the equations for these three test types are all
reasonable, as the values of all the specimen types are larger than 0.8. Therefore, it is
believed that Equation 8-38 is capable of providing a fair prediction for the R-N lines.
When the fatigue life becomes infinite, the safety factor R∆ tends to a limit value Rlimit.

Rlim it = R0 − R1 (8-39)

It indicates that the specimen will not have any fatigue damage under the loading
condition Rinitial = Rlimit, which relates to the fatigue endurance limit εlimit.
223
Chapter 8 Yield Surface and Fatigue Tests

Table 8-6 Regression coefficients and equations of three kinds of binder


Regression coefficients
Specimen type Rlimit R2
R0 R1 a b
UTC Size0.5_ε_20°C 0.173 0.140 1.60E-06 0.66 0.0332 0.809
UTC Size1.0_ε_20°C 0.319 0.282 3.06E-06 0.460 0.0376 0.904
UTC Size1.5_ε_20°C 0.419 0.393 2.86E-06 0.280 0.0261 0.835
UTC Size1.0_ε_5°C 0.524 0.423 1.33E-07 0.058 0.1011 0.887
UTC Size1.0_σ_5°C 0.412 0.212 2.10E-06 0.428 0.2000 0.901
4PB Size0.5_ε_20°C 0.258 0.213 1.86E-06 0.564 0.0454 0.870
4PB Size1.0_ε_20°C 0.921 0.881 2.42E-06 0.138 0.0404 0.873
4PB Size1.5_ε_20°C 0.895 0.586 2.07E-06 0.135 0.0394 0.900
4PB Size1.0_ε_5°C 0.559 0.416 2.20E-07 0.066 0.1428 0.837
4PB Size1.0_σ_5°C 0.911 0.687 1.65E-07 0.250 0.2238 0.796
IT Size1.0_σ_5°C 0.725 0.521 2.01E-06 0.198 0.2047 0.870
IT Size1.5_σ_5°C 0.727 0.510 3.35E-06 0.197 0.2171 0.920

UTC Size0.5_ε_20°C
UTC Size1.0_ε_20°C
UTC Size1.5_ε_20°C
UTC Size1.0_ε_5°C
UTC Size1.0_σ_5°C
4PB Size0.5_ε_20°C
4PB Size1.0_ε_20°C
4PB Size1.5_ε_20°C
4PB Size1.0_ε_5°C
4PB Size1.0_σ_5°C
IT Size1.0_σ_5°C
IT Size1.5_σ_5°C

0 0.05 0.1 0.15 0.2 0.25


Rlimit

Figure 8-29 The values of Rlimit in all different cases

From Figure 8-29, it can be seen that the Rlimit value does not change too much with the
specimen size and test type but is strongly influenced by the temperature and loading
mode. For all the UT/C, 4PB and IT fatigue tests, the value of Rlimit obtained at the low
temperature and stress controlled mode is larger than that obtained from the high
temperature and strain controlled mode, respectively.

224
Chapter 8 Yield Surface and Fatigue Tests

8.7 Conclusions
Based on the results presented in this chapter, the following main findings and
conclusions are summarized below:

1. By means of the unified models described in Chapter 7, the yield surface based on
the Desai yield function can be determined when the strain rate and the
temperature are known. At the peak stress, the yield surface reduces to a straight
line in the I1-√J2 space.
2. A new parameter, called the safety factor R∆ is introduced as an indicator of the
“safety against failure” at a certain stress-strain state. When there is no applied
stress, the material is absolutely safe, the value of the safety factor R∆ is equal to
zero; when the applied stress reaches the strength of the material, the material
starts to yield, the value of the safety factor R∆ is equal to 1.
3. The critical locations of the cylinder and beam specimen can be easily determined
for the UT/C and 4PB fatigue tests, respectively. During the IT fatigue test, the
specimen is subjected to both horizontal and vertical stresses along the vertical
diameter. Based on an FE model, the maximum horizontal tensile stress occurs in
the center point (y=0) and the maximum horizontal strain is found at the locations
with y = ± 36 R (R is the radius). The R∆ value at the points y = ± 36 R is around
50 50
35% larger than at the centre point at various force levels. Therefore these two
points seem to be the weakest points instead of the centre point for the IT fatigue
test.
4. By converting the sine wave to a triangle wave, the strain rate can be calculated at
a given strain amplitude and frequency. For the fatigue test, each load level
corresponds to a safety factor R∆ and a fatigue life. In the stress-controlled mode,
the value of R∆ is calculated by the stress state and the yield surface determined
by the failure stress. In the strain-controlled mode, the value of R∆ is calculated by
the strain state and the failure strains.
5. Similar to the traditional fatigue line, a straight line is found by plotting R∆ at the
critical location and the fatigue life on a log-log scale. In contrast to the traditional
fatigue line, the size effect on the 4PB and IT fatigue results can be ignored by
using this new fatigue relation. For the stress-controlled mode, the fatigue lines
obtained from the UT/C shows good agreement with the IT fatigue results.
However, the 4PB test shows a longer fatigue life when the safety factor at the
critical points are the same. The differences of the fatigue relations caused by
temperature and loading mode still exist in this new fatigue method.
6. The relationship between R∆ and fatigue life in the normal coordinate can also be
( )
b
expressed as the function R∆ = R0 − R1 1 − e − aN f
. When the safety factor R∆
reduces to a limit value Rlimit, the fatigue life becomes infinite. The Rlimit is
independent of specimen size and test type but is strongly influenced by the
temperature and loading mode. Rlimit obtained at the low temperature and stress
controlled mode is larger than that obtained from the high temperature and strain
controlled mode, respectively.

225
Chapter 8 Yield Surface and Fatigue Tests

References

Desai, C.S., Somasundaram, S. and Frantziskonis, G., a Hierarchical approach for


Constitutive Modelling of Geologic Materials, International Journal of Numerical and
Analytical Methods in Geomechanics, 1986; 10(3): 225-257.

Erkens, S.M.J.G., Liu, X., Scarpas, A. and Molenaar, A.A.A., 3D Finite Element Model
for Asphalt Concrete Response Simulation, Paper presented at the 2nd International
Symposium on 3D Finite Element in Pavement Engineering, Charleston, West-Virginia,
USA, 2000.

Erkens, S.M.J.G., Asphalt Concrete Response - Determination. Modelling and Prediction.


PhD Thesis. Delft University of Technology, the Netherlands, 2002.

Hill, R., The Mathematical Theory of Plasticity. Oxford University Press, London, U.K.,
1950.

Hondros, G., Evaluation of Poisson’s Ratio and the Modulus of Materials of a Low
Tensile Resistance by the Brazilian (Indirect Tensile) Test with Particular Reference to
Concrete. Austr. J. Appl. Sci.. 1959; 10(3): 243-268.

Liu, X., Numerical Modelling of Porous Media Response Under Static and Dynamic
Load Conditions. PhD Thesis. Delft University of Technology, the Netherlands, 2003.

Li, N., Molenaar, A.A.A., van de Ven, M.F.C., Wu S., Application of a New Fatigue
Analysis Approach on laboratory fatigue tests. The 5th European Asphalt Technology
Association (EATA) conference. Braunschweig, Germany. June 3-5, 2013a.

Li, Ning, Molenaar, A.A.A., Pronk A.C., van de Ven, M.F.C., Characterization of
Fatigue Performance of Asphalt Mixture Using a New Fatigue Analysis Approach.
Construction and Building Materials, 2013b, Vol. 45, pp. 45-52.

Medani, T.O., Huurman, M., Superposition Principle to Determine Properties of


Bituminous Mixtures in the Time–Temperature Domain. Proc. of the 7th International
Conference on the Bearing Capacity of Roads. Railways and Airfields. Norway, 2005.

Scarpas, A., Gurp, C.A.P.M., van, Al-Khoury, R.I.N., Erkens, S.M.J.G., Finite Elements
Simulation of Damage Development in Asphalt Concrete Pavements. 8th International
Conference on Asphalt Pavements. Seattle, U.S.A., 1997.

Sitters, C.M.W., Material Models for Soil and Rock, Delft University of Technology,
Faculty of Civil Engineering, the Netherlands, 1998.

Zienkiewicz, O.C., Chan, A.H.C., Pastor, M., Schreer, B.A., Shiomi, T., Computational
Geomechanics, with Special Reference to Earthquake Engineering. Wiley& Sons, 1999.

226
Chapter 9 Conclusions and Recommendations

Chapter 9 Conclusions and


Recommendations
The main goal of this research is to investigate the influence of the test type and specimen
size on laboratory fatigue test results and to find a way to harmonize the fatigue results
obtained from different tests. In Chapter 5, the results of three types of fatigue tests were
discussed. For each test, specimens with different dimensions were used. The detailed
analysis approaches used in this research were described in Chapter 6 to Chapter 8.

In this chapter, the important findings of each Chapter are summarized. Section 9.1 gives
the generalized conclusions and Section 9.2 presents the recommendations for future
work.

9.1 Conclusions
9.1.1 Conclusions Related to Literature Review
• Since fatigue cracking due to repeated traffic loading is one of the main failure
modes of an asphalt pavement, the fatigue characteristics of asphalt concrete are
important for the thickness design of asphalt pavements.
• In pavement design methods, normally the fatigue laws are use obtained from
laboratory fatigue testing. Various fatigue test devices are currently used to evaluate
the fatigue performance of asphalt concrete. Even at the same test conditions, the
results from the different tests are difficult to compare to each other. It is therefore
important to understand the effect of the test type and specimen size on fatigue
results.
• Fatigue testing is normally done in a one-dimensional or two-dimensional stress
state. In reality however pavement materials are subjected to a three-dimensional
stress state. This makes it necessary to represent the failure condition obtained by
means of laboratory testing in a three-dimensional space.

9.1.2 Conclusions Related to Fatigue Test Equipment


• The test configuration of the UT/C test is relatively simple and the results are hardly
influenced by the test set up. Due to the uniform stress and strain state, data
processing is simple. However, to prevent bending moments, there are high
demands for the used specimen and test setup. The upper and lower steel platens
should be parallel to each other and perpendicular to the loading direction. The
cylindrical specimen should be polished and its top and bottom ends should be
exactly parallel to each other and perpendicular to the axis of the cylinder. This asks
for very careful preparation of specimens and equipment.
• Compared to the UT/C and IT test, the test results measured by the 4PB test are
highly influenced by the test equipment, such as the clamping force, distance
between the clamps, structural strength of the whole set up, etc. To avoid high shear

227
Chapter 9 Conclusions and Recommendations

and tensile stress concentrations near the supports, the four clamps should be
allowed to rotate freely and horizontal translation and the clamping force should be
as low as possible. Furthermore, shear deflection of the beam specimen and
deformation of the frame should be taken into account in the calculation of the
stiffness of the beam specimen.
• The IT test is simple and easy to perform. The specimen can be easily prepared in
the laboratory or cored from the field. However the stress-strain distributions in the
specimen are complex. For this test, it is impossible to apply a sinusoidal loading to
the specimen; a haversine or half sine shaped load pulse need to be applied, which
induces permanent deformation during testing even at a low temperature.

9.1.3 Conclusions Related to Stiffness and Fatigue Results


• The complex stiffness values measured by the UT/C and 4PB test are similar and
are not influenced by the specimen size. A lower stiffness is obtained from the IT
test and the measured IT stiffness is slightly influenced by the specimen size. These
differences are most probably caused by the difference in the mode of loading and
state of stress.
• The fatigue life NR determined by the dissipated energy ratio is smaller than the
traditional fatigue life. The difference between NR and Nf measured by the UT/C
test is smaller compared to those from the 4PB and IT test. It indicates that the
crack propagation is much shorter during the UT/C fatigue test.
• The effect of specimen size is negligible for the UT/C and IT fatigue results, but for
the 4PB fatigue test the influence of the specimen size is significant. The smaller
specimen size with a lower height has a longer fatigue life. The reason is that the
tensile stress-strain fields inside the UT/C and IT specimen are relatively uniform
compared to the 4PB specimen.
• With regard to the influence of the test type, at the same test condition and loading
mode, the fatigue life from the 4PB test is the longest and from the IT test it is the
shortest. The differences might be explained by the stress-strain distributions of the
specimen in the different fatigue tests. Apart from that, failure of the IT is also
partly caused by permanent deformation.
• With regard to the NR-Wfat relation, the effect of specimen size on the 4PB test is
obvious but not for the UT/C and IT test. The NR-Wfat line is independent of the
temperature but influenced by the loading mode. The slope of the curves in the
stress-controlled mode is a little steeper than in the strain-controlled mode. The total
dissipated energy measured by the IT test is lower compared to the UT/C and 4PB
tests at the same fatigue life NR. This difference is caused by the irreversible
deformation, biaxial stress state at the center of the specimen and the smaller phase
angle measured in the IT test. By taking into account the dissipated energy in
vertical direction, the corrected NR-Wfat lines from the IT test is much closed to the
results from the other tests.

9.1.4 Conclusions Related to Partial Healing Model


• The PH model provides a good simulation of the evolution of the complex modulus
and phase angle for a unit volume in a strain-controlled fatigue test. The results

228
Chapter 9 Conclusions and Recommendations

from the UT/C fatigue test can be directly used for the determination of model
parameters.
• The model parameters can be expressed as functions of the applied strain level and
are not significantly influenced by the specimen size. The parameter δα1 (reversible
damage in the loss modulus) is always nil. The parameter δα2 (reversible damage in
the storage modulus) becomes smaller with a decrease in temperature. The
parameter δγ2 (irreversible damage in storage modulus) shows an opposite tendency.
This indicates that in the UT/C test, the healing ability of the asphalt mixture
decreases at a lower temperature and the more irreversible damage happens at a
higher strain level or at a lower testing temperature.
• For the 4PB test, the strain is not distributed uniformly throughout the beam
specimen. The functions of the PH model parameters are used to calculate the local
stiffness for different parts of the specimen. In this way the weighted overall
stiffness can be calculated by a weighing procedure. For the specimen size 1.0, the
functions of the PH model parameters obtained from the UT/C and 4PB test are
similar to each other. It indicates that the 4PB test results can be predicted by means
of the PH model parameters determined from the UT/C test results.
• By means of the PH model, the local stiffness at the surface in the midsection of the
beam is comparable with the UT/C fatigue test results when the pure bending strain
on the beam surface is equal to the strain applied to the cylinder. The fatigue line
based on the local surface stiffness of the beam shows a good agreement with the
results obtained from the UT/C fatigue test.
• The trends of the values of the parameter δγ1 and δγ2 indicate the existence of an
endurance limit. The predicted endurance limit obtained by the UT/C test is not
influenced by the specimen size and temperature. The range of the endurance limit
is between 66 and 69 µm/m. For the 4PB test, the predicted endurance limit for a
unit volume is influenced by the specimen size. A larger specimen has a lower
endurance limit.

9.1.5 Conclusions Related to Monotonic Test Results


• The tensile and compressive strength are dependent on the temperature and loading
rate. When the loading rate or temperature becomes very high or low, the material
properties tend to reach a threshold value. This behavior of asphalt mixtures can
successfully be described by the unified model.
• The model parameters of the unified model were expressed as functions of the
material properties including the stiffness of the mixture, volume content of the
bitumen, volume content of air voids and the coefficient of curvature of the
aggregate gradation. With the temperature susceptibility function obtained from the
stiffness test and the functions of the unified model parameters, the compressive
and tensile strength at a certain strain rate and temperature can be predicted directly.
• For further simplification, the parameters u0 and γ can be simply considered to be
constant. Although the variation of the unified model becomes larger, the
predictions of the compressive and tensile strength are quite acceptable.

229
Chapter 9 Conclusions and Recommendations

9.1.6 Conclusions Related to Yield Surface Approach


• By means of the unified models discussed above, the yield surface based on the
Desai yield function could be determined at a given strain rate and temperature. At
the peak stress, the yield surface reduces to a straight line in the I1-√J2 space.
• The safety factor R∆ is introduced as an indicator of the “safety against failure”. A
smaller R∆ value means that the material is safer. The range of R∆ value is from 0 to
1.
• For the IT test, based on the FE model, The R∆ value at the points y = ± 36 R is
50
around 35% larger than that at the centre point y=0. Therefore these two points
were considered as the higher loaded places instead of the centre point for the IT
fatigue test.
• By converting the sine wave to a triangle wave, the strain rate takes a constant value
and can be calculated by the given strain amplitude and frequency. For the fatigue
test, each loading level corresponds to a safety factor R∆ and a fatigue life. A
straight line is found by plotting R∆ at the critical location and the fatigue life on a
log-log scale. Compared to the traditional fatigue line, the size effect on the 4PB
and IT fatigue results can be ignored by using this new fatigue relation. For the
stress-controlled mode, the fatigue lines obtained from the UT/C test show a good
agreement with the IT fatigue results. Due to the inhomogeneity of the stress-strain
distribution, the fatigue life from the 4PB test is higher than those obtained from the
other fatigue tests when the safety factor at the critical point is the same. The
differences of the fatigue relations caused by the temperature still exist in this new
fatigue method.
• The relationship between R∆ and the fatigue life in the normal coordinate also can
( ) . When the safety factor R∆
b
be expressed as the function of R∆ = R0 − R1 1 − e − aN f

reduces to a limit value Rlimit, the fatigue life becomes infinite. The Rlimit value does
not change too much with the specimen size and test type but is strongly influenced
by the temperature and loading mode. For all the UT/C, 4PB and IT fatigue tests,
the value of Rlimit obtained at the low temperature and stress controlled mode is
larger than that obtained from the high temperature and strain controlled mode,
respectively.

9.2 Recommendations
9.2.1 Recommendations Related to Experimental Work
• To make an efficient test plan, the “central composite rotatable design technique”
was used to determine the combination of the temperature and strain rate. This test
design method significantly reduces the number of tests. However, the tensile
strength is very sensitive to the strain rate at low temperatures. It is difficult to
select an appropriate range of test conditions, which can give a complete description
of the tensile strength evolution. Therefore, before using the central composite
rotatable design methods, some preliminary tests are needed to determine a
appropriate range of the test conditions.

230
Chapter 9 Conclusions and Recommendations

• Based on the calibration of the 4PB test set up, it was found that the frame stiffness
is only 4 times higher than the stiffness of the aluminum beam. For the large
specimen or the test at low temperature, the deformation of the test frame
significantly influences the test results because of the relatively high applied force.
In future research, the 4PB set up should be strengthened to resist this deformation
and to allow applying fairly high load levels when tests need to be performed at low
temperatures or when using a large specimen.
• After modification of the 4PB equipment, the two outer clamps are still controlled
by the springs so that the outer clamping force varies with the applied force. It is
therefore recommended to further improve the outer clamps to keep the outer
clamping force constant.
• Since the fatigue test is time-consuming, only one asphalt mixture type was used in
this research. It is recommended that different asphalt mixture types should be
tested to verify the conclusions based on the results of the DAC 0/8 mixture.

9.2.2 Recommendations Related to Partial Healing Model


• At present, the Partial Healing model is only valid for the fatigue test in strain-
controlled mode. It is proposed to develop the solutions for the stress-controlled test
in the near future.
• By taking into account the stress-strain distribution and the specimen shape, the
Partial Healing model is also applicable for the fatigue results from the two-point
bending or three-point bending tests.

9.2.3 Recommendations Related to Yield Surface Approach


• The model parameters of the unified model depend on material properties. Based on
the collected compression and tension test results of 7 different asphalt mixture
types, the model parameters were expressed as function of the stiffness of the
mixture, volume content of bitumen, volume content of air voids and the coefficient
of curvature of the aggregate gradation. In the future, the monotonic test results
from more asphalt mixtures are needed to improve those functions of the model
parameters.
• The relation between R∆ and fatigue life obtained from the 4PB test shows a big
difference compared to the other two tests, because the safety factor R∆ of the 4PB
specimen simply represents the stress state at the most severe location and is much
larger than those of the UT/C and IT test. It is recommended that for the 4PB test, a
new safety factor should be defined to adequately represent the safety condition of
the beam specimen.

231
Chapter 9 Conclusions and Recommendations

232
Appendix A Calculations for 4-Point Bending Test

Appendix A Calculations for 4-Point Bending


Test
The four-point bending test is schematically given in Figure A-1. All dimensions are in
SI units unless other specified. For the test setup used in this research, the distance
between two outer clamps is 400 mm. The distance between inner two clamps A is equal
to one third of L.

Clamping Clamping

Sample
L

Actuator

Figure A-1 Principle four-point bending test

A.1 Shape factor Z [m-1]:

6 AL2 − 8 A3
Z= (A-1)
8bh3
where: A : distance between inner clamp and outer clamp, next to each
other, [m];
L : effective length=distance between outer clamps, [m];
b : width of specimen, [m];
h : height of specimen, [m].

233
Appendix A Calculations for 4-Point Bending Test

For L= 3A (evenly spaced) the formula becomes:

46L3 L3
Z= = 0.213 ⋅ 3 (A-2)
216bh3 bh

A.2 Weighing function R{x} for mass compensation

12 L 1
 A 2 2
for x > A
 x  x   A
3 − 3  −  
 L L  L
R{x} =  (A-3)
12 L 1
for x < A
 x 2
A  A  x 
2
 3 − 3  −  
 L  L L

Because L= 3A and in this configuration (standard) the sensor is in the middle so x= L/2,
R{x} becomes:

L 36
R{ } = 2 2
= 56.348
2  1  1 1 (A-4)
3  − 3  −  
 2  2 3

And the weighing factor at position x=A becomes (for the moving masses at x=A):

36
R{x} = 2 2
= 64.8
1 1 1 (A-5)
3  − 3  −  
 3  3 3

A.3 Equivalent masses meq[kg]

The equivalent beam mass used in the back calculation depends on the x coordinate
where the deflection (used for back calculation) is measured. Here, the sensor is in the
middle so x= L/2:

L
R 
=  4  ⋅ mbeam = 0.5785 ⋅ mbeam
2 (A-6)
meq.beam
π

The equivalent masses for other moving masses like the clamps and the sensor depend on
the x coordinate xback used for the back calculation and the coordinate xplace where the
extra mass is located.

234
Appendix A Calculations for 4-Point Bending Test

R{xback }R{ A}
M eq.mass = M place (A-7)
R{x place }R{x place }

In general you will use x = L/2 for the back calculation. So, the equivalent mass for the
clamps (and all other moving masses at x = A) is given by:

L
R 
=   ⋅ mclamps = 0.8696 ⋅ mbeam
2 (A-8)
meq.beam
R { A}

The equivalent mass of the sensor used at x=L/2 for the back calculation is given by:

R { A}
meq.sensor = ⋅ msensor = 1.15 ⋅ msensor
L (A-9)
R 
2

So, for the given configuration of the 4-point bending test, the equivalent mass is given
by:

meq = 0.5785 ⋅ msensor + 0.8696 ⋅ mclamps + 1.15 ⋅ msensor (A-10)


where: mbeam : mass of the beam [kg];
mclamps : Moving masses (except beam) between load cell
and outer supports = 1.6 [kg];
msensor : mass sensor =0.14 [kg].

A.4 Stiffness [Pa], (D=0, S=0)

Neglecting damping D (non viscous) effects in the system (system losses) and shear, the
complex stiffness E*, which consists of a real part (ERe) and an imaginary part (EIm), can
be written as:

F 
ERe = Z  cos (ϕs ) + meqω 2  (A-11)
 vb 
F 
EIm = Z  sin (ϕ s )  (A-12)
 vb 
where: F : amplitude of the force, [N];
vb : (bending) amplitude of the deformation, [m];
φs : measured phase lag [º]
meq : equivalent mass [kg] ; see (A4.4)
ω : circular frequency [rad/s]

235
Appendix A Calculations for 4-Point Bending Test

Edyn = ERe
2
+ EIm
2
(A-13)
 EIm 
ϕ mix = arctan   (A-14)
 ERe 
where: F : amplitude of the force, [N];
vb : (bending) amplitude of the deformation, [m];
φs : measured phase lag [º]
meq : equivalent mass [kg] ; see (A4.4)
ω : circular frequency [rad/s]

Substituting Equation A-2 in Equation A-11 and A-12, Equation A-13 and A-14 are
expressed as follows [N/m2]:

2
ZF v v 
Edyn = 1 + 2cos (ϕs ) b meqω 2 +  b meqω 2  (A-15)
vb F F 
 
 sin (ϕ s ) 
ϕ mix = arctan   (A-16)
 cos (ϕ s ) + vb meqω 2 
 F 

In general the equations used for this calculation are given by:

FL3
Edyn = 1 + 2cos (ϕs ) ⋅ Q + Q2 (A-17)
vb R ( x ) I
v
Q = b meqω 2 (A-18)
F
1 3
I = bh (A-19)
12
R { x} R { x}  R{x}R{ A} 
meq =
π 4
⋅ mbeam +
R { A}
⋅ mclamps + ∑  R{ y}R{ y} m y

(A-20)

A.5 Maximal tensile strain ε{x} [m/m]

h ⋅ A ⋅ R { x}
ε { x} = vb { x} (A-21)
4 L3

So for L=3A and calculating the strain at the center, at x=L/2, this gives:

L
h⋅R 
L
ε =  2  v  L  = 56.348 ⋅ h v (A-22)
b   b
2 12 L2 2 12 L2

236
Appendix A Calculations for 4-Point Bending Test

A.6 Maximum stress σ [Pa]

1
F⋅A
M 2 3F ⋅ A FL
σ= = = = 2 (A-23)
W 1 b ⋅ h2 bh2 bh
6
where: M : bending moment of the beam [Nm];
W : resistance of the cross section against bending [m3];

A.7 Energy dissipation of beam W [J]

Total consumed energy for the beam (so including system loss) for one cycle is
calculated by:

Wbeam ,tot = π ⋅ sin (ϕ s ) vb ⋅ F (A-24)

And the total dissipated (viscous) energy for the beam for one cycle is calculated by:

R { x}
Wbeam ,visc = π ⋅ sin (ϕ mix ) vb2 ⋅ I ⋅ (A-25)
L3

In this set-up R{x}=R{L/2}= 56.348, so Equation A-25 becomes:

56.348 vb2 ⋅ b ⋅ h 3
Wbeam ,visc = π ⋅ sin (ϕ s ) vb2 ⋅ b ⋅ h3 = 4.7π sin ( ϕ s ) ⋅ (A-26)
12 L3 L3

A.8 Energy dissipation per unit of volume W [J/m3]

And the total dissipated (viscous) energy per unit m3 for one cycle is calculated by:

Wvisc , per m3 = π ⋅ ε ⋅ σ ⋅ sin (ϕ s ) (A-27)

The cumulated dissipated energy per cycle up to cycle n:

n
Wsum ,visc , per m3 = ∑ ∆Wvisc , per m3 ,i (A-28)
i =1

If the measurements are taken at intervals Ni we use Simpson for numerical integration:

∑ i=1 ∆Ni
m n
n= N =
Wsum ,visc , per m3 = ∫ Wvisc , per m3 ⋅ dn ≈ ∑  ∆Wvisc , per m3 ,i ⋅ ∆N i  (A-29)
i =1
i =1
 

237
Appendix A Calculations for 4-Point Bending Test

238
Appendix B Determination for Prony Series Model Parameters

Appendix B Determination for Prony Series


Model Parameters
In the Abaqus software, the visco-elasticity of the asphalt mixture is characterized by the
Prony series representation. The time domain stiffness relation for this model is given in
Equation B-1.

n
G ( t ) = Gr + ∑ Gi ⋅ e −t /τ i (B-1)
i =1
where: Gr : rubbery shear modulus, [MPa];
Gi : shear stiffness, [MPa];
τi : relaxation time;
n : number of Prony parameters terms.

The model parameters Gr, Gi and τi are determined by conducting the stiffness frequency
sweep tests. The parameters determination procedure is incorporated in a user friendly
graphical user interface program (GUI) in Matlab environment developed by Woldekidan
[Woldekidan, 2011]. By changing the number of Kelvin-voigt elements, optimization can
be performed until a satisfactory fit is obtained. The procedure for the parameter
determination is summarized as follows:

1. Import the master curve data, frequency (rad/s), complex modulus (MPa) and
phase angle (degree) to the Matlab GUI program; in a comma delimited file.
2. In the program interface, fill in the number of Kelvin-voigt elements.
3. Run the optimization to perform Prony series fitting to the imported master curve
data.
4. Change the number of Kelvin-voigt terms and re-run the optimization till
satisfactory fit is obtained. The weighting factor for complex stiffness, phase
angle, loss and storage modulus can also be altered from the program interface
helping to obtain better fits.
5. The output data are exported to an excel file in the working direction.
The number of Prony parameter terms is determined based on the quality of the model
description. In this study, the number of the Prony terms is 12. Figure B-1 illustrates the
obtained regression fit for the specimen UT/C-size 1.0 at a reference temperature of
20°C. Similar qualities of fits were also obtained for the other specimens.

The generalized Prony series model parameters for the different specimen types are given
in Table B-1 and Table B-2. These data were used as the input data for the material
properties in the ABAQUS software.

239
Appendix B Determination for Prony Series Model Parameters

Figure B-1 Matlab GUI for model parameter determination

Table B-1 Prony series model parameters for the UT/C specimen at Tref = 20 oC (12
terms)
n Size0.5 Size1.0 Size1.5
τi Gi τi Gi τi Gi
-6
1 1.2×10 1.8×10-1 1.2×10 -6
1.7×10-1 1.2×10 -6
1.3×10-1
-6
2 7.3×10 5.9×10-2 7.3×10 -6
5.8E-02 7.3×10 -6
1.0×10-1
3 4.3×10-5 1.2×10-1 4.3×10 -5
1.2×10-1 4.3×10-5 6.0×10-2
-4
4 2.6×10 1.0×10-1 2.6×10 -4
1.1×10-1 2.6×10 -4
1.1×10-1
5 1.5×10-3 1.2×10-1 1.5×10-3 1.3×10-1 1.5×10-3 9.2×10-2
-3
6 8.9×10 1.2×10-1 8.9×10 -3
1.3×10-1 8.9×10 -3
1.7×10-1
7 5.3×10-2 1.2×10-1 5.3×10-2 1.3×10-1 5.3×10-2 1.4×10-1
8 3.1×10-1 9.8×10-2 3.1×10-1 9.7×10-2 3.1×10-1 1.3×10-1
9 1.8 5.4×10-2 1.8 5.0×10-2 1.8 5.4×10-2
10 11 1.6×10-2 11 1.3×10-2 11 1.5×10-2
11 64 2.1×10-3 64 1.6×10-3 64 2.3×10-3
12 3.8×102 5.0×10-4 3.8×102 3.1×10-4 3.8×102 7.4×10-4
Gr [MPa] 28367 28127 28627

240
Appendix B Determination for Prony Series Model Parameters

Table B-2 Prony series model parameters for the 4PB specimen at Tref = 20 oC (12 terms)
n Size0.5 Size1.0 Size1.5
τi Gi τi Gi τi Gi
-6
1 1.0×10 1.5×10-1 1.0×10 -6
1.3×10-1 1.0×10 -6
1.6×10-1
2 5.0×10-6 3.6×10-2 5.0×10-6 3.5×10-2 5.0×10-6 3.6×10-2
3 2.5×10-5 1.0×10-1 2.5×10-5 9.3×10-2 2.5×10-5 1.0×10-1
-4
4 1.2×10 8.0×10-2 1.2×10 -4
7.6×10-2 1.2×10 -4
7.9×10-2
5 5.9×10-4 1.0×10-1 5.9×10-4 1.0×10-1 5.9×10-4 1.0×10-1
-3
6 2.9×10 1.1×10-1 2.9×10 -3
1.1×10-1 2.9×10 -3
1.1×10-1
7 1.4×10-2 1.1×10-1 1.4×10-2 1.3×10-1 1.4×10-2 1.2×10-1
8 6.9×10-2 1.1×10-1 6.9×10-2 1.3×10-1 6.9×10-2 1.3×10-1
-1
9 3.4×10 9.2×10-2 3.4×10 -1
1.0×10-1 3.4×10 -1
9.2×10-2
10 1.6 6.4×10-2 1.6 6.1×10-2 1.6 4.9×10-2
11 8.0 2.7×10-2 8.0 1.7×10-2 8.0 1.3×10-2
12 39 1.4×10-2 39 1.3×10-2 39 1.1×10-2
Gr [MPa] 25032 22400 24982

Due to the big variation of the measured phase angle, it was decide not to calculate the
prony model parameters for the IT test results. In the simulation of the IT specimen, the
material was simply defined as an elastic material.

References
Woldekidan, M.F., Mixture Performance Optimization: Laboratory Investigation and
LOT Performance Computations. Report 7-12-186-2, Delft University of Technology,
Delft, the Netherlands, 2012.

Woldekidan, M.F., Response Modelling of Bitumen, Bitumenous Mastic and Mortar,


PhD Thesis, Delft University of Technology, The Netherlands, 2011.

241
Appendix B Determination for Prony Series Model Parameters

242
Curriculum Vitae
Personnel Information

Name: Ning Li (李宁)


Gender: Male
Date of Birth: July 27, 1983
Place of Birth: Tianmen, Hubei Province, P.R. China
Email: Ning.Li@tudelft.nl; lining0618@gmail.com

Education Experience

Oct. 2008 – Nov. 2013 PhD student in Section of Road and Railway Engineering,
Faculty of Civil Engineering and Geosciences,
Delft University of Technology, the Netherlands.

Sep. 2006 – Jun. 2008 Master student in Material Science,


School of Materials Science and Engineering,
Wuhan University of Technology, Wuhan, P.R.China.

Sep. 2002 – Jun. 2006 Bachelor student in Material Science and Engineering,
School of Materials Science and Engineering,
Wuhan University of Technology, Wuhan, P.R.China.

243
Publications
1. Li, N., Molenaar A.A.A., van de Ven, M.F.C., and Wu, S. Estimation of the Fatigue
Endurance Limit of HMAC for Perpetual Pavements. Journal of Wuhan University
of Technology-Mater. Sci. Ed. Vol.25 No.4, 2010, pp 645-649.
2. Li, N., Molenaar, A.A.A., van de Ven, M.F.C., and Wu S., Characterization of
Fatigue Performance of Asphalt Mixture Using a New Fatigue Analysis Approach.
Construction and Building Materials. Volume 45, 2013, pp 45–52.
3. Li, Ning, Molenaar, A.A.A., Pronk A.C., van de Ven, M.F.C. and Wu, S.,
Comparison of Uniaxial and Four Point Bending Fatigue Tests for asphalt mixtures.
Journal of Transportation Research Record (TRR), 2013. (Accepted)
4. Qiu J., Li N., Pramesti F.P., Van de Ven M.F.C. and Molenaar A.A.A. Evaluating
Laboratory Compaction of Asphalt Mixtures using the Shear Box Compactor.
Journal of testing and evaluation. 2012; 40 (5): 844-852.
5. Li, Ning, Molenaar, A.A.A., Pronk A.C. and van de Ven, M.F.C., Effect of
Specimen Size on Fatigue Behavior of Asphalt Mixture in Laboratory Fatigue Tests,
Proceedings of the 7th RILEM International Conference on Cracking in Pavements,
Delft, the Netherlands, pp 827-836, 2012.
6. Li, Ning, Molenaar, A.A.A., Pronk A.C., van de Ven, M.F.C., Investigation into the
Size effect on Four Point Bending Fatigue Tests, Proceedings of the 3rd Workshop
on 4PB, Davis, California, September 17-18 2012.
7. Li, Ning, Molenaar, A.A.A., Prediction of tensile strength of asphalt concrete,
Proceedings of the 2nd International Conference on Sustainable Construction
Materials (SusCoM2012). Wuhan, China, October 18-22 2012.
8. Li, N., Molenaar, A.A.A. and van de Ven, M.F.C., Application of a New Fatigue
Analysis Approach on laboratory fatigue tests. 5th European Asphalt Technology
Association (EATA) conference. Braunschweig, Germany. June 3-5 2013.
9. A.A.A. Molenaar, Ning Li, Pungky Pramesti, Fatigue Characterization of Asphalt
Mixtures for Designing Long Life Pavements, 15th AAPA International Flexible
Pavements Conference, Brisbane, Australia, September 22-25, 2013.

244

You might also like