Download as pdf or txt
Download as pdf or txt
You are on page 1of 228

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/280564830

Asphalt Concrete Response (ACRe) - determination, modelling and prediction

Thesis · October 2002

CITATIONS READS

54 818

1 author:

Sandra Erkens
Delft University of Technology
95 PUBLICATIONS 1,318 CITATIONS

SEE PROFILE

All content following this page was uploaded by Sandra Erkens on 30 July 2015.

The user has requested enhancement of the downloaded file.


ASPHALT CONCRETE RESPONSE (ACRe)
-DETERMINATION, MODELLING AND PREDICTION-
HET GEDRAG VAN ASFALT (ACRe)
-BEPALEN, MODELEREN EN VOORSPELLEN-

PROEFSCHRIFT

ter verkrijging van de graad van doctor


aan de Technische Universiteit Delft
op gezag van de Rector Magnificus prof. dr. ir. J.T. Fokkema,
voorzitter van het College voor Promoties aangewezen
in het openbaar te verdedigen op maandag 7 oktober 2002 te 13:30 uur

door

Sandra Maria Johanna Grada ERKENS

civiel ingenieur
geboren te Nijmegen
Dit proefschrift is goedgekeurd door de promotoren:

Prof. dr. ir. A.A.A. Molenaar,


Prof. dr.ir. J. Blaauwendraad

Samenstelling Promotiecommissie:

Rector Magnificus, Technische Universiteit Delft, voorzitter


Prof.dr.ir. A.A.A. Molenaar, Technische Universiteit Delft, promotor
Prof.dr.ir. J. Blaauwendraad, Technische Universiteit Delft, promotor
Prof. dr. ir. A. Bakker, Technische Universiteit Delft
Prof. C.S. Desai, BSc., MSc., Phd., University of Arizona
A. Scarpas, BSc., MSc., Technische Universiteit Delft
Dr. B.B. Koenders, Shell Global Solutions

Prof. dr. ir. C. Esveld Technische Universiteit Delft, reserve lid

Published and distributed by: DUP Science

DUP Science is an imprint of


Delft University Press
P.O. Box 98
2600 MG Delft
The Netherlands
Telephone: +31 15 2785678
Telefax: +31 15 2785706
E-mail: DUP@Library.TUDelft.NL

ISBN 90-407-2326-5

Keywords: Asphalt Concrete, experiments, characterisation, material modelling

© 2002 by Sandra Erkens

All rights reserved. No part of the material protected by this copyright may be reproduced or
utilized in any form or by any means, electronic or mechanical, including photocopying,
recording or by any information storage and retrieval system, without written consent from the
publisher: Delft University Press.

Printed in the Netherlands


ACKNOWLEDGEMENTS
“The important thing to recognize is that it takes a team, and the team ought to get
credit for the wins and the losses. Successes have many fathers, failures have none.”
- Philip Caldwell -

The research presented in this thesis was carried out at the Road & Railway Research
Laboratory of the department of Civil Engineering at the Delft University of
Technology. The project itself was a co-operation between this group and the Structural
Mechanics Section of the same department. I want to thank both groups for providing
the facilities necessary to carry out this poject.
In particular I want to thank Tom Scarpas, who created a group spirit that is necessary
for the interaction between numerics and experiments. I enjoyed all our over-coffee
discussions about the project and many other things. Your encouragement at difficult
times has been important to me and I’m glad that we’ll continue to work together, but
remember: when I’m good I’m very, very good, but when I’m bad…I’m better!
From this place I want to thank all my colleagues from both the road engineering and
structural mechanics group for their support. It was a great advantage to be working
with people who could (and did!) contribute to my work.
For the experimental part of the project, I really appreciated the contribution, in ideas
and facilities as well as actual assistance, of Hans Kalf, Jan Willem Bientjes, Jan
Moraal, Abdol Miradi, Peter Galjaard, Rien Huurman, Jacob Groenendijk. Special
thanks are due to Gerard Galjé, who coordinated the specimen production and
developed an ingeneous system to remove the split mould from the tension specimens
and of course Marco Poot whose contribution was essential to my work.
Also the support of the laboratory secretariat in organising anything from meetings to
trips and the printing of reports has saved me a lot of time and worries. For that reason I
would like to thank all people that worked there over the years and in particular
Jacqueline Barnhoorn and Sonja van den Bos.
I also want to thank the “computational guru’s”, Frank Custers, Dung de Ha, Xueyan
Liu, Yi Zhao and especially Cor Kasbergen for their help with respect to the numerical
aspects of the project, concerning both the hard- and the software.
Finally, I want to take this opportunity to extend a whole hearted “thank you” to our
sponsors, the Technology Foundation STW, Applied Science Division of NWO via the
Technology Programme of the Ministry of Economic Affairs, Shell Global Solutions
(formerly the Royal Shell Laboratories in Amsterdam) and the Road and Hydraulic
Engineering Division of the Ministry of Transportation, this project would not have
been possible without their financial support. Apart from their financial contribution, I
appreciated the interest in the project itself that these sponsors showed by providing
members for the STW-user group. The six monthly meetings we had were a pleasure
because of the interest and support provided by the individual members.

Sandra Erkens,
17 juni 2002

Acknowledgements v
()
(vi) Acknowledgements
SUMMARY
“Perfection is achieved, not when there is nothing more to add,
but when there is nothing left to take away”
-Antoine de Saint Exupery (1900-1944)-

In this dissertation the development of a realistic model to describe asphalt concrete


reponse (ACRe) is described. The model expresses this response as a function of the
state of stress, the temperature and the strain rate. The ACRe project, carried out in co-
operation between the Road and Railway Research Laboratory and the Structural
Mechanics group of the department of Civil Engineering and Geosciences of the Delft
University of Technology, was initiated because the available methods are not capable
to describe many of the phenomena observed in both laboratory tests and actual
pavement structures. Examples of such elusive phenomena are surface cracking,
ravelling and combinations of damage types. It was felt that a material model that takes
into account the main influences on the response of asphalt concrete would allow the
analysis of such phenomena, adding to the understanding of the mechanisms behind
them.

The model is a plasticity model, using the flow surface proposed by Desai and co-
workers as the basis and extending and adapting it to facilitate the description of asphalt
concrete behaviour. Laboratory tests on the material with an a-priory known, uniform
state of stress are used to determine the material response and the model parameters. In
this project uniaxial compression, uniaxial tension and multiaxial four-point shear tests
are used to provide this information. These tests are not commonly used in road
engineering, therefore the set-ups and test procedures were developed in the course of
the project. This resulted in three sophisticated test-set-ups with, in case of the uniaxial
tests, elaborate instrumentation. The set-ups were used to characterise a single,
relatively homogenous asphalt mixture (the Asphalt Concrete Response (ACRe)
mixture). On the basis of this test programme data analysis procedures to determine the
model parameters were developed.

Because the tests developed in the course of this project are non-standard, the ACRe
mixture used was also tested in some standard road engineering tests. These were the
Marshall test, the Indirect Tension Test (ITT) and the Four-point Bending Fatigue
(FPBF) test. The results from these tests provide a frame of reference that shows that the
mixture is a down-scaled Dense Asphalt Concrete (DAC 0/5).

In developing the uniaxial compression test, also an effective friction reduction system
for asphalt concrete was invented. This is particularly interesting because in road
engineering often relatively flat specimens are used. In these specimens the effect of
friction is considerable, but using the ACRe friction reduction system reliable uniaxial
strength values are obtained for specimens with height to diameter ratios as low as 0.5.
During the development of the compression test it was furthermore noted that the
influence of the temperature on the response of asphalt concrete is even more
pronounced than usually expected. As a result testing of asphaltic materials without
temperature control is strongly discouraged. Once the set-up and test procedures were

Summary (vii)
developed, they were used to run a series of monotonic tests on the asphalt mixture. A
wide range of temperatures and strain rates was used, yielding responses that ranged
from 1.9 N/mm2 (3.7 kN) to 56.5 N/mm2 (111 kN). The repeatability of the test results
was very good.

During the development of the tension set-up also some spin-off information was
obtained. Because in a tension test after the peak different types of response are
interacting, it was tried to separate them in the measurements. To achieve this, a
parabolic specimen shape was used to ensure that the position of the crack was known
beforehand, without inducing the stress concentrations that are caused by notches. In
this way, both the overall response and the unloading of the undamaged parts could be
measured, allowing the overall response to be separated into crack-opening and
unloading. To produce specimens that allowed this, a special split mould was developed
for use inside the gyratory mould. Although this worked rather well, it also showed that
the compaction of asphalt concrete is extremely sensitive to the temperature. Because
the mixture temperature cannot be maintained during compaction, the period over which
effective compaction is possible is limited. This indicates that it would be profitable to
develop a method to maintain the mixture temperature until the desired level of
compaction is reached. The tesion test was also used for a series of monotonic tests on
the ACRe mixture.

On the basis of the uniaxial tests only, already four out of five model parameters could
be determined. Because the test results are available for different combinations of
temperature and strain rate, the model parameters can also be expressed as functions of
these influences. The procedures used to determine these four model parameters from
the test results and the methods used to express them as general functions of
temperature and strain rate are presented in detail. Using the additional information
from the multi-axial tests, the model parameters could be determined more accurately.
This test also provided information on the confinement sensitivity of the material. From
the combined test results it appeared that the fifth model parameter is zero for the
mixture considered in this thesis. As a result, the parameters that were initially
determined did not change. Furthermore it is shown that, even if the fifth model
parameter is not equal to zero, the procedures developed for the parameter
determination remain valid.

The model that is developed on the basis of the tests distinguishes between tension and
compression damage, mainly because tension damage does not effect the compressive
strength while compression damage weakens the material in tension. Since in
pavements tension and compression stresses alternate, this difference is very important.
The model was verified by simulating the tests used to determine the model parameters.
As expected, the simulations agreed well with these laboratory results. In the next step,
the monotonic Indirect Tensile Test was simulated. This test was used because results of
the ITT on the ACRe mixture are available, but also because in this test tension and
compression stresses occur. This allows a study of the effect of the different damage
mechanisms (in tension and compression, respectively) on the overall response as well
as the interaction between them. From these simulations it can be seen that the model
really does provide a realistic representation of asphalt concrete response if the different
damage types are used. If however only isotropic or localised damage is used, only
specific parts of the material response is captured. This can clearly be seen from the

(viii) Summary
predicted damage patterns and a comparison thereof with the response observed in
laboratory tested specimens.
Finally, an analytical approach in which the model and tests described in thesis can be
used in a simplified way is presented. Although this does not provide the same insight
as the elaborate analyses, it provides a powerful extension of the currently available
analysis methods because it introduces a strength limit, thus providing an indication of
the resistance against failure.

Summary (ix)
(x) Summary
SAMENVATTING
“Perfectie is niet bereikt wanneer er niets meer toe te voegen is,
wanneer als er niets meer weggelaten kan worden”
-Antoine de Saint Exupery (1900-1944)-

In dit proefschrift wordt beschreven hoe een realistisch materiaalmodel voor asfalt
ontwikkeld is. Het model beschrijft het gedrag van asfalt als functie van de
spanningstoestand, de temperatuur en de reksnelheid. Het project is uitgevoerd door het
Laboratorium voor Wegen & Spoorwegen in samenwerking met de Sectie Constructie
Mechanica van de Faculteit der Civiele Techniek en Geowetenschappen van de
Technische Universiteit Delft. Het werd opgestart omdat de beschikbare asfaltmodellen
niet in staat bleken om vele van de in laboratoriumproeven en wegconstructies
geobserveerde schadevormen te voorspellen of te beschrijven. Voorbeelden van zulke
niet beschrijfbare schadevormen zijn oppervlaktescheuren, rafeling en elke combinatie
van schadevormen. Het idee was dat een model dat de belangrijkste invloeden op het
gedrag van asfalt mee nam de analyse van dergelijke fenomenen mogelijk zou maken,
waardoor er inzicht kan worden verkregen in de achterliggende mechanismen.

Het ontwikkelde model is een plasticiteitsmodel, gebaseerd op de vloeifunctie die door


Desai en zijn medewerkers is ontwikkeld voor granulaire materialen. Door
aanpassingen en uitbreidingen is een model verkregen dat het gedrag van asfalt kan
beschrijven. Om de parameters in het model te bepalen zijn laboratoriumproeven nodig
waarbij het materiaal onderworpen wordt aan een vooraf bekende, uniforme
spanningsverdeling. In dit project zijn uniaxiale trek- en drukproeven en meer-assige
vierpuntsafschuifproeven gebruikt om de benodigde gegevens te verkrijgen. Omdat
deze of vergelijkbare proeven niet standaard zijn in de wegenbouw, zijn de opstellingen
en testprocedures ontwikkeld in het kader van dit project. Dit heeft drie hoogwaardige
opstellingen voor materiaalkarakterisering opgeleverd met, voor de trek- en
drukopstellingen, uitgebreide instrumentatie. De opstellingen zijn gebruikt om één
relatief homogeen asfaltmengsel (het Asphalt Concrete Response (ACRe) mengsel) te
karakteriseren. Deze proeven zijn vervolgens gebruikt om de data analyse procedures
voor het bepalen van de model parameters te ontwikkelen.

Aangezien de proeven die binnen het project ontwikkeld zijn niet standaard zijn binnen
de wegbouwkunde, is het materiaal ook onderzocht door middel van een aantal
standaardproeven. Het ging hierbij om de Marshall proef, de Indirecte Trekproef en de
vierpuntsbuigvermoeiingsproef. De resultaten van deze proeven toonden aan dat het
ACRe-mengsel als een verschaald Dicht Asfalt Beton (DAB 0/5) te beschouwen is.

Gedurende het ontwerpen van de drukopstelling is ook een wrijvingsreductiesysteem


voor drukproeven op asfalt ontwikkeld. Dit is vooral van belang omdat in de
wegenbouw, tengevolge van de gelaagde constructieopbouw, vaak vrij platte
proefstukken worden gebruikt. Het is bekend dat voor dergelijke proefstukken de
invloed van randeffecten aanzienlijk kan zijn, maar bij gebruik van het ACRe
wrijvingsreductiesysteem worden betrouwbare sterktewaarden verkregen voor hoogte-
diameter verhoudingen tot 0,5. Tijdens het ontwerpen van deze proef bleek verder dat
de invloed van de temperatuur op het gedrag van asfalt nog groter is dan normaal wordt

Samenvatting (xi)
aangenomen, zodat het beproeven van asfalt zonder een temperatuurbeheersingssysteem
ten zeerste moet worden afgeraden. Nadat de drukopstelling en de testprocedures klaar
waren, zijn deze gebruikt voor een serie proeven op het ACRe mengsel. De proeven zijn
uitgevoerd bij uiteenlopende temperaturen en reksnelheden, resulterend in sterktes die
variëren van 1.9 N/mm2 (3.7 kN) tot 56.5 N/mm2 (111kN). De proefresultaten bleken
goed reproduceerbaar.

Ook bij het ontwerpen van de trekopstelling werd afgeleide kennis verkregen. Omdat er
bij een trekproef na de piekbelasting een combinatie van gedrag in het proefstuk
optreedt, was het de bedoeling een proef te ontwikkelen waarbij dit gescheiden kon
worden gemeten. Om dit te realiseren is een parabolische proefstukvorm gekozen,
waarbij de positie van de scheur, bij benadering, vooraf vast stond zonder dat er
spanningsconcentraties optreden zoals bijvoorbeeld het geval is wanneer zaagsneden
worden gebruikt. Op deze manier kon zowel het totaalgedrag als het ontlasten van de
proefstuk helften gemeten worden. Om de parabolische proefstukvorm te kunnen
produceren, is een speciale deelbare binnenmal voor de gyratormal gemaakt. Hoewel dit
bijzonder goed bleek te werken om de gewenste vorm te krijgen, werd ook duidelijk dat
de verdichting van asfalt erg gevoelig is voor de verdichtingtemperatuur. Omdat de
temperatuur tijdens het verdichten niet meer gehandhaafd wordt, is de tijdspanne waarin
de verdichting nog toeneemt beperkt. Ook de trekopstelling is gebruikt voor een serie
monotone, verplaatsingsgestuurde proeven op het ACRe mengsel.

Op basis van de trek- en drukresultaten kunnen vier van de vijf modelparameters


worden bepaald. Omdat de proefresultaten voor verschillende temperaturen en
reksnelheden bepaald zijn, kunnen ook de modelparameters worden uitgedrukt als
functie van deze grootheden. De procedures die ontwikkeld zijn om de parameters uit
de proefresultaten te bepalen en de manier waarop algemene uitdrukkingen voor de
parameters als functie van temperatuur en reksnelheid zijn bepaald worden in detail
beschreven. Door aanvullende informatie uit de afschuifproeven kunnen de
modelparameters nauwkeuriger worden bepaald. Bovendien kan de
steundrukgevoeligheid van het materiaal op basis van gegevens uit deze proef in kaart
worden gebracht. Uit de combinatie van alle proefgegevens kan vervolgens de vijfde
modelparameter bepaald worden. Voor het ACRe blijkt deze parameter nul te zijn,
zodat alle eerder bepaalde waarden voor de andere parameters gelijk blijven. Het is
verder aangetoond dat voor mengsels waarvoor de vijfde parameter ongelijk aan nul is
de procedures om de parameters te bepalen hetzelfde blijven.

Het ontwikkelde model maakt onderscheid tussen schade ontstaan onder trekbelasting
en drukbelasting, voornamelijk omdat trekschade de druksterkte niet beïnvloedt terwijl
drukschade wel de treksterkte vermindert. Doordat in wegconstructies trek- en
drukspanningen elkaar afwisselen, is dit verschil van groot belang voor het correct
beschrijven van de optredende schade. Om het materiaalmodel te valideren zijn
allereerst een aantal van de laboratoriumproeven gesimuleerd die gebruikt zijn om de
parameters te bepalen. Zoals verwacht was in dit geval de overeenkomst tussen
simulatie en laboratoriumresultaat zeer goed. Daarna is de monotone Indirecte
Trekproef gesimuleerd. Deze proef bij uitstek geschikt om het effect van de
verschillende schademechanismen (voor trek en druk) op het totale gedrag te
onderzoeken, omdat er zowel trek- als drukspanningen in optreden. Bovendien zijn er
resultaten van deze proef op het ACRe mengsel beschikbaar. Uit deze simulaties blijkt
dat de combinatie van deze schade-mechanismen in het model inderdaad een realistisch

(xii) Samenvatting
gedrag oplevert. Als daarentegen slechts één van beide schademodellen wordt
meegenomen wordt, zoals duidelijk blijkt uit een vergelijking van de voorspelde
schadepatronen met in het laboratorium geteste proefstukken, maar een deel van de
optredende schade voorspeldt.

Tenslotte wordt er een vereenvoudigde, analytische toepassing van het in dit


proefschrift beschreven model en de bijbehorende proeven gepresenteerd. Dit geeft
vanzelfsprekend niet dezelfde hoeveelheid aan informatie en inzicht als de uitgebreidere
analyses, maar het vormt een goede aanvulling op de al beschikbare analyse methoden.
Het voornaamste voordeel is dat er een directe indicatie van de weerstand tegen
bezwijken kan worden gegeven doordat er een sterkte criterium wordt ingevoerd.

Samenvatting (xiii)
(xiv) Samenvatting
CONTENTS
“He who violates the form, also damages the content”
Herbert von Karajan (1908-1989)

1. INTRODUCTION .......................................................................................................1
1.1 FIRST STEPS ON THE ROAD ........................................................................................... 1
1.1.1 The beginning of roads ............................................................................. 1
1.1.1.1 Ancient roads........................................................................................ 1
1.1.1.2 Roads in the Roman Empire................................................................. 2
1.1.1.3 Recent road history............................................................................... 3
1.1.2 Asphalt concrete as a road paving material............................................. 3
1.1.2.1 What is asphalt concrete? ..................................................................... 3
1.1.2.2 Asphalt through the centuries ............................................................... 3
1.1.2.3 Asphalt today........................................................................................ 4
1.1.3 Road engineering research....................................................................... 4
1.1.3.1 Mixture design...................................................................................... 4
1.1.3.2 Pavement design ................................................................................... 7
1.1.3.3 Limitations of the classic approaches................................................... 8
1.2 OBJECTIVES OF THE ACRe PROJECT.............................................................................. 8
1.2.1 The material model................................................................................... 9
1.2.2 Test programme...................................................................................... 11
1.2.3 The interaction between model and test programme.............................. 12
1.3 ORGANISATION OF THIS THESIS ................................................................................. 12
1.4 REFERENCES .............................................................................................................. 12
1.4.1 Literature ................................................................................................ 12
1.4.2 Info from the World Wide Web............................................................... 13
2. MODELLING OF PAVEMENT MATERIALS & STRUCTURES ....................15
2.1 WHAT IS A GOOD MODEL?.......................................................................................... 15
2.2 PAVEMENT DESIGN METHODS .................................................................................... 15
2.3 THE ACRE PROJECT ................................................................................................... 17
2.3.1 Material parameter determination ......................................................... 17
2.3.2 The ACRe material model ...................................................................... 18
2.3.3 Model capabilities .................................................................................. 19
2.4 REFERENCES .............................................................................................................. 20
3. THE ACRE MIXTURE ............................................................................................23
3.1 SELECTION OF THE MIXTURE ..................................................................................... 23
3.2 MIXTURE COMPOSITION ............................................................................................. 24
3.2.1 Crushed rock .......................................................................................... 24
3.2.2 Bitumen................................................................................................... 25
3.2.3 Filler ....................................................................................................... 26
3.2.4 Comparing the mixture composition to that of standard mixtures......... 26
3.3 ROAD ENGINEERING TYPECASTING OF THE ACRe MIXTURE ...................................... 27
3.3.1 The Marshall test .................................................................................... 27
3.3.2 The monotonic and cyclic indirect tension test ...................................... 28
3.3.2.1 Comparing the results from the cyclic test to those of other mixtures32
3.3.2.2 Comparing the monotonic test results to those of other mixtures ...... 32
3.3.3 The four-point bending fatigue test ........................................................ 37

Contents (xv)
3.3.4 Comparing monotonic and cyclic stiffness values.................................. 40
3.4 CONCLUDING REMARKS ............................................................................................. 43
3.5 REFERENCES: ............................................................................................................. 44
4. THE UNIAXIAL COMPRESSION TEST .............................................................47
4.1 INTRODUCTION .......................................................................................................... 47
4.2 DEVELOPING THE TEST SET-UP .................................................................................. 47
4.2.1 Frame deformations and movements...................................................... 48
4.2.2 Set-up capacity ....................................................................................... 49
4.2.3 Influence of boundary friction ................................................................ 50
4.2.3.1 Effect of friction on the test results .................................................... 50
4.2.3.2 Developing an effective friction reduction system............................. 51
4.2.3.3 Assessing the effect of friction reduction ........................................... 55
4.2.4 Temperature effects ................................................................................ 60
4.3 THE UNIAXIAL COMPRESSION TESTS .......................................................................... 60
4.3.1 Specimens ............................................................................................... 60
4.3.2 Test set-up............................................................................................... 61
4.3.3 Instrumentation....................................................................................... 63
4.3.4 Test conditions........................................................................................ 64
4.3.5 Overview of the test results..................................................................... 65
4.3.6 Test results as function of temperature and loading rate....................... 69
4.3.6.1 Compressive strength ......................................................................... 70
4.3.6.2 Stiffness .............................................................................................. 74
4.3.6.3 Poisson’s ratio .................................................................................... 75
4.4 CONCLUSIONS AND RECOMMENDATIONS .................................................................. 76
4.4.1 Conclusions ............................................................................................ 76
4.4.2 Recommendations................................................................................... 77
4.5 REFERENCES ............................................................................................................... 77
5. THE UNIAXIAL TENSION TEST .........................................................................79
5.1 INTRODUCTION .......................................................................................................... 79
5.2 DEVELOPING THE UNIAXIAL TENSION TEST ............................................................... 79
5.2.1 Alignment................................................................................................ 79
5.2.2 End effects .............................................................................................. 81
5.2.3 Specimen shape ...................................................................................... 81
5.2.3.1 Localised failure in tension ................................................................ 81
5.2.3.2 A special specimen shape ................................................................... 82
5.2.3.3 Specimen production .......................................................................... 83
5.2.4 Temperature effects ................................................................................ 85
5.2.5 set-up ...................................................................................................... 86
5.3 THE UNIAXIAL TENSION TESTS................................................................................... 86
5.3.1 The specimens......................................................................................... 86
5.3.2 Tension set-up in detail .......................................................................... 88
5.3.3 Instrumentation....................................................................................... 89
5.3.3.1 Displacement transducers ................................................................... 90
5.3.3.2 Strain gauges ...................................................................................... 90
5.4 TEST PROCEDURES, CONDITIONS AND RESULTS ........................................................ 92
5.4.1 Test procedures ...................................................................................... 92
5.4.2 Test conditions........................................................................................ 93
5.4.2.1 Overview of the test conditions .......................................................... 93
5.4.2.2 Relation between strain rate and deformation rate ............................. 95
5.4.3 Test results.............................................................................................. 98
5.4.3.1 Control parameter and measuring length ........................................... 98

(xvi) Contents
5.4.3.2 The different response components.................................................. 100
5.4.3.3 Fracture energy ................................................................................. 102
5.4.3.4 Test results........................................................................................ 103
5.4.4 General expressions for the test results................................................ 107
5.4.4.1 Tensile strength ................................................................................ 107
5.4.4.2 Elasticity parameters......................................................................... 109
5.5 CONCLUSIONS AND RECOMMENDATIONS ................................................................ 112
5.5.1 Conclusions .......................................................................................... 112
5.5.2 Recommendations................................................................................. 112
5.6 REFERENCES ............................................................................................................ 113
6. THE MULTIAXIAL TEST ....................................................................................115
6.1 INTRODUCTION ........................................................................................................ 115
6.2 DEVELOPING THE TEST SET-UP ................................................................................ 115
6.2.1 Ideas behind the four-point shear test .................................................. 116
6.2.2 The existing set-up ................................................................................ 117
6.2.3 Changes to facilitate the ACRe tests .................................................... 118
6.2.3.1 Temperature control ......................................................................... 118
6.2.3.2 Vertical tension using pendulum bars .............................................. 119
6.2.3.3 Width of the shear zone .................................................................... 119
6.2.4 Connection of the horizontal actuator.................................................. 121
6.3 THE MULTIAXIAL TESTS ........................................................................................... 121
6.3.1 Specimens ............................................................................................. 121
6.3.1.1 Specimen production ........................................................................ 121
6.3.1.2 Specimen preparation ....................................................................... 123
6.3.2 ACRe four-point shear test set-up in detail .......................................... 124
6.3.3 Test procedures and conditions............................................................ 126
6.3.3.1 Test procedures................................................................................. 126
6.3.3.2 Test conditions.................................................................................. 127
6.3.3.3 Force or displacement control?......................................................... 129
6.3.4 Test results............................................................................................ 130
6.3.5 General relations for the test results .................................................... 134
6.4 CONCLUSIONS AND RECOMMENDATIONS ................................................................ 136
6.4.1 Conclusions .......................................................................................... 136
6.4.2 Recommendations................................................................................. 136
6.5 REFERENCES ............................................................................................................ 136
7. THE ACRE MATERIAL MODEL .......................................................................137
7.1 THE MODEL .............................................................................................................. 137
7.1.1 The flow surface ................................................................................... 138
7.1.2 Model parameters................................................................................. 139
7.1.2.1 Influence of α ................................................................................... 139
7.1.2.2 Influence of γ .................................................................................... 141
7.1.2.3 Influence of n.................................................................................... 142
7.1.2.4 Influence of β.................................................................................... 142
7.1.2.5 Influence of R ................................................................................... 143
7.1.3 Constitutive framework ........................................................................ 144
7.1.3.1 Hardening response .......................................................................... 145
7.1.3.2 Response degradation ....................................................................... 145
7.1.3.2.1 Isotropic softening via γ degradation ......................................... 145
7.1.3.2.2 Cracking....................................................................................... 146
7.2 MODEL PARAMETER DETERMINATION USING ONLY UNIAXIAL TEST RESULTS ........ 148

Contents (xvii)
7.2.1 Flow surface for uniaxial states of stress ............................................. 148
7.2.2 Model parameter β ............................................................................... 149
7.2.3 Model parameters R and γ.................................................................... 149
7.2.4 Model parameter n ............................................................................... 152
7.2.5 Hardening response parameter determination..................................... 156
7.2.5.1 Parameter α ...................................................................................... 156
7.2.5.2 Hardening in uniaxial tension........................................................... 160
7.2.6 Softening parameter determination ...................................................... 162
7.2.6.1 Isotropic softening via γ degradation................................................ 162
7.2.6.2 Tension softening ............................................................................. 164
7.2.7 Overview of the model relations and parameters................................. 166
7.3 MODEL PARAMETER DETERMINATION FROM UNIAXIAL & MULTIAXIAL TESTS ...... 170
7.3.1 Relation between β, γ and R ................................................................. 170
7.3.2 Consequences for the other model parameters .................................... 174
7.4 CONCLUSIONS AND RECOMMENDATIONS ................................................................ 175
7.5 REFERENCES ............................................................................................................ 175
8. MODEL VERIFICATION AND APPLICATION ..............................................177
8.1 INTRODUCTION ........................................................................................................ 177
8.2 DAMAGE DEFINITION ............................................................................................... 177
8.3 MODEL VERIFICATION: SIMULATIONS OF THE ACRe TESTS...................................... 178
8.3.1 Simulations of the uniaxial compression test ....................................... 178
8.3.2 Simulation of the uniaxial tension tests ................................................ 181
8.4 MODEL APPLICATION: SIMULATION OF THE INDIRECT TENSION TEST (ITT) ........... 182
8.4.1 Response simulation via the Hoffman criterion ................................... 183
8.4.2 Response simulation via the Desai criterion ........................................ 186
8.4.3 Response simulation via the ACRe constitutive model......................... 188
8.4.4 Comparison of the simulation results with lab tests............................. 190
8.5 A SIMPLIFIED MODEL APPLICATION ......................................................................... 190
8.5.1 Testing .................................................................................................. 191
8.5.2 Application ........................................................................................... 192
8.5.2.1 Evaluating monotonic test results..................................................... 192
8.5.2.2 Evaluating cyclic test results ............................................................ 195
8.6 CONCLUDING REMARKS ........................................................................................... 197
8.7 REFERENCES ............................................................................................................ 198
9. CONCLUSIONS AND RECOMMENDATIONS ................................................199
9.1 INTRODUCTION ........................................................................................................ 199
9.2 GENERAL CONCLUSIONS AND RECOMMENDATIONS ................................................ 199
9.2.1 General conclusions ............................................................................. 199
9.2.2 General recommendations.................................................................... 200
9.3 SUB CONCLUSIONS AND RECOMMENDATIONS ......................................................... 200
9.3.1 Sub conclusions .................................................................................... 200
9.3.2 Sub recommendations........................................................................... 201

(xviii) Contents
1. INTRODUCTION
“Let’s start at the very beginning, a very good place to start”
Equation Section 1
“Do-Re-Mi”, from the soundtrack of “the Sound of Music”

1.1 FIRST STEPS ON THE ROAD


Road Engineering research is aimed at a better understanding of the behaviour of
pavement structures and, via this understanding, at improving the design of both
pavement structures and materials. Pavement response is the result of the interaction
between subgrade, subbase and asphalt layers due to the applied loads and the climatic
conditions. To truly understand this complicated interaction, first of all knowledge of
the separate components is needed. For this reason, this dissertation deals with the
representation of asphalt concrete, the most popular road paving material in the
Netherlands and much of Europe. In the first part of this chapter, some historical
background information on both roads and asphalt is given. It provides an introduction
to the objectives of this dissertation, which are presented in the remainder of this
chapter. The information and the pictures used in the first part of this chapter is mainly
obtained from the following sources: Hindley (1971), the Asphalt Institute (1965),
Brzesowsky and Nijsen (1988), Lee (1953), ArtsEdNet (1999), Auburn University
Alabama (1999), Teaching Ideas (1999) and Via Domitia (1999).

1.1.1 THE BEGINNING OF ROADS


Travelling is an integral part of human history, from the perpetual trek of nomadic tribes
via the trade-caravans that crossed the continents to the current day traffic jams. The
way in which we travel has changed but the activity itself remained. What also
remained is the tendency to travel along familiar routes, so you know what to expect
and how to orientate yourself. That is the way in which the first “roads” developed:
well-trodden paths with (natural) landmarks at strategic points. Even today we
sometimes witness the development of such “natural routes”, when popular short cuts
gradually become recognisable paths.

1.1.1.1 Ancient roads


Many ancient routes were established either for trading or for government purposes.
Both trade and management of large areas require reliable, fast and safe routes for the
transportation of goods, information or troops. Well known examples are the silk and
salt routes and the extensive road systems developed by the Egyptians, Inca’s, Persians
and Romans. The Mesopotamian king Pileser I, who had a segment of road engineers
(“ummani”) in his army, presents an eloquent example of the importance of roads. His
engineers built a road through the mountains, which he used to pursue his enemies.
Once the area was conquered, it was also linked to the heart of the empire by good
roads, which made it easier to remain in control.

The roads changed with the changes in human lifestyle and especially the kind of
transport that was used. Roads that were used solely by pedestrians usually followed the

Introduction 1
contours of the terrain, keeping to high ground. If beasts of burden were used, the roads
had to be broader and more level, so they were positioned lower in the terrain. To keep
these lower roads passable in the wet season, special constructions to provide sufficient
drainage and protection for the road surface were often required. In a community where
carriages were used, the main roads were usually paved. Carriages require a smooth
surface, yet they also easily damage unpaved road surfaces. Therefore, heavily
trafficked routes required special protection to remain passable.

1.1.1.2 Roads in the Roman Empire


The most impressive of the ancient road builders were the Romans, who built a road
system that spanned approximately 53000 miles (≈85000 km), which consisted of
anything between well-paved main roads and simple tracks. Yet it is not the size of their
road system, but its quality that sets them apart. They not only created a hard surface,
but took care of the foundation and drainage as well. This made it possible to use heavy
carriages on the Roman roads.

Figure 1.1: Roman main roads, (above) a photo of the Via Domitia (1999) and
(below) a drawing of the typical, multi-layer construction of such a road
A typical Roman main road consisted of three layers (Figure 1.1), first a foundation of
rocks and chalk that was compacted to a water-resistant mass with a stone roller. Next
came a layer of granular material, usually sand or gravel with the fine materials on top
of the coarse and finally a surface layer (or “metalling”) of either gravel, flint, pebbles
or slabs of stone usually bonded by cement or chalk. The result was a durable and

2 Chapter 1
waterproof road construction. This layered approach was standardised throughout the
empire, but they used locally available materials. In marshy areas, wooden mattresses
were placed underneath the road to prevent it from sinking.

1.1.1.3 Recent road history


The European roads were not seriously maintained after the fall of Rome and they
slowly fell in disrepair. It took until the 16th century, when a centralised road
administration was started in France, before the interest in good quality roads was
revived. Real developments did not occur until les departement des Ponts et Chaussées
was founded in 1716. In the 18th century, as the increase in traffic number and weight
set new demands for road constructions Pierre Trésauguet re-established sound practices
for road building based on a well drained, strong foundation and a surface layer that was
impervious to water. When automobiles became available to the public at the end of the
19th century, the demand for level, mud and dust free roads increased. This boosted the
popularity of asphaltic road paving materials, which eventually led to the invention of
synthetic asphalt. As a result most of the main roads in the industrialised countries are
nowadays paved with asphalt concrete.

1.1.2 ASPHALT CONCRETE AS A ROAD PAVING MATERIAL

1.1.2.1 What is asphalt concrete?


Asphalt concrete is a mixture of gravel, sand and filler, bonded by asphalt. Asphalt is a
plastic, sticky substance that is a natural constituent of crude petroleum. Petroleum
asphalt consists mainly of bitumen, which is a hydrocarbon material that is soluble in
disulfide (CS2). In Dutch, petroleum asphalt is usually referred to as bitumen, while the
mixture of stones, sand and filler with asphalt is called asphalt instead of asphalt
concrete. In this thesis the binder will be called asphalt and the mixture will be called
asphalt concrete.

Unlike cement, the binder used in “ordinary” concrete, asphalt is a viscous material.
This means that its characteristics fit somewhere between solids and fluids. Just how
fluid it is depends on the temperature, at higher temperatures it is a real, water-like, fluid
while at low temperatures it is solid. This characteristic is used to produce asphalt
concrete, for which asphalt and aggregates (sand, rock and finer particles) are heated to
about 160oC and then mixed. At that temperature the binder is very fluid and mixes
easily with the other materials. While warm, the mixture is kneaded to get a compact
mixture. When this mixture cools down, a strong composite material is left.

Another effect of the viscous nature of asphalt, and the asphalt concrete in which it is
the binder, is that the latter material is relatively flexible: it can follow deformations
without cracking. This makes it an ideal road paving material since it can follow
deformations of the subgrade without excessive damage. Furthermore, the material is
rather impervious to salt and acids.

1.1.2.2 Asphalt through the centuries


Asphalt is not only a constituent of petroleum, it can also be found in nature, either as
soft mortars or as black, glass-like rock. Soft asphalt, found for example in the Trinidad
Lake depository, Bermudez Lake (Venezuala) and the tar sands in Canada, became
known as “natural asphalt”. Since asphalt is a strong, natural cement and extremely
waterproof, the material has been used throughout the centuries in the shipbuilding

Introduction 3
industry (Sumeria, 6000 B.C.), as a waterproofing material (the Egyptians, 2600 B.C.)
and to pave roads. An example of this latter application is the Royal road, from Susa to
Sardis, which the Persians paved over its full length of 2600 km.

Like many engineering materials and techniques, asphalt was forgotten after the fall of
the Roman Empire, when Europe entered the Dark Ages. In the 16th century the interest
in asphalt was rekindled, resulting in research into its characteristics and origin. In the
early 19th century, asphalt was rediscovered as a paving material. In 1838 d’Erinas used
stamped down asphalt to pave sidewalks in Paris and Philadelphia. In 1869
Threedneedle Street in London was paved with it and in 1873 the Kalverstraat in
Amsterdam became the first Dutch road with a top layer of asphalt. The Belgian
chemist de Smedt invented sheet asphalt, a mixture of natural asphalt with crushed rock
or sand, in 1876. The increased use of automobiles led to a rapid growth of the use of
asphaltic pavement materials and because of the expenses, people started looking for an
alternative for natural asphalt. This led to the invention of synthetic asphalt, which is
asphalt obtained from the distillation of crude petroleum.

1.1.2.3 Asphalt today


Nearly all current-day asphalt is obtained from the refinery of crude petroleum. Crude
petroleum is a mixture of many chemical compounds. The exact number of compounds
and the amount of each of them varies. Distillation is used to separate these compounds.
This results in, among others, paraffin oil, petrol and asphalt. By increasing or
decreasing the amount of volatile components that is removed by distillation the asphalt
that is obtained is either harder or softer. Asphalt obtained in this way is so-called
straight-run asphalt. Asphalt can also be obtained by leading air through crude oil from
which the paraffin oil and petrol components were removed by distillation. The oil
reacts with the oxygen in the air, which changes the characteristics.

Today, asphalt is used for many applications such as road pavements, runways,
sidewalks, parking lots and dike revetment. This thesis deals only with hot-mix asphalt,
the most common type of road paving material.

1.1.3 ROAD ENGINEERING RESEARCH

1.1.3.1 Mixture design


The development of petroleum asphalt resulted in an explosive increase of activities,
both in the area of road construction and in that of road engineering techniques. In order
to get good quality roads, the material was also the subject of extensive research
projects. In the 1920’s this resulted in the empirical Hubbard-Field test, which gave an
indication of the stability of a mixture based on a punching-shear test. The objective of
this test as well as the more elaborate mixture design procedures that will be discussed
hereafter is to establish the optimum binder content for a given aggregate distribution.
This optimum binder content is the amount of binder that results in the best composite
material. The basic idea is to get a load carrying skeleton of aggregates that is kept
together by the binder. If too little binder is used, the bond between the aggregates is too
weak and the material will be reduced to loose stones when it is subjected to loads. Yet
if too much binder is used, the aggregates will flow in the binder instead of forming a
skeleton and the load will be carried by the viscous binder, which results in large plastic
deformations.

4 Chapter 1
Too little binder optimal binder content too much binder

Subjected to loads the material will:


fall apart deform, but hold together deform easily

(A) (B) (C)


Figure 1.2: Mixture design procedures aim to establish the optimum binder content
for a given aggregate gradation ………………………………
In the 1930’s Francis Hveem (Hveem and Davis 1950) developed a mixture design
procedure in which a compacted mixture was tested to establish both the strength of the
bond between the aggregates and the existence of a stable aggregate skeleton. All
mixture design starts with an aggregate gradation, which is usually described by
minimum and maximum percentages of the material that pass through a series of sieves.
The actual gradation falls within these limits, but that still leaves room for variation.
The Hveem mixture design procedure starts at the point that the actual gradation is
known. Then the optimum binder content for that gradation has to be determined.
Hveem estimates the optimum binder content based on the surface area of the aggregate
particles and their absorption characteristics. The former is determined on the basis of
standard surface factors related to the sieve analysis, but the latter is found from
absorption tests. Based on this information, the required binder content is determined
from nomographs.

In the next step a series of specimens, compacted using a special mechanical kneading
compactor, with binder contents that vary around this estimated optimum value are
subjected to tests in order to arrive at the actual optimum value. The tests Hveem used
to judge the fitness of the compacted mixtures were the stabilometer, cohesiometer and
the swell test. The stabilometer is a predecessor of the triaxial test and it was used to
determine the stability of a mixture by measuring the radial expansion due to an axially
applied load. Naturally, an over-filled mixture (Figure 1.2C) would show relatively
large deformations and thus be judged unstable. The results from this test were
expressed in a relative stabilometer value, where a true liquid was considered to have
zero relative stability (lateral pressure equal to vertical pressure) while a non-deforming
solid was the end of the range (radial deformation of zero). To account for the influence
of height versus diameter ratio’s, Hveem established correction curves for specimens
with non-standard heights.

Introduction 5
Figure 1.3: The Hveem stabilometer, an early triaxial testing apparatus
The second test Hveem used, the cohesiometer test, was basically a force controlled
bending test. By dropping a controlled quantity of a material with a known weight per
time unit in a container, the applied load steadily increased. The force necessary to
arrive at a standard displacement of the loading arm is recorded as the cohesiometer
value. If the bond between the aggregate particles is weak due to a lack of binder, the
material will perform badly in this experiment. Finally, the swell test was be used to
determine the sensitivity of the mixture to water. It measures the permeability and
increase in volume due to absorption of water (swell). Hveem advised to aim for the
maximum binder content that met the stability criteria and had no less than 4% air
voids, to avoid the risk of (locally) over-filled material in the actual construction.

At the beginning of World War II, roads for war activities were rapidly established.
Also, many runways were built in a short time. This required fast design procedures for
heavy-duty pavements and these were developed by Bruce Marshall (Asphalt Institute
1965). In this design procedure the phase of determining the estimated optimum binder
content is skipped. A series of specimens with different binder contents are prepared.
These binder contents are usually selected on the basis of past experience. At the end,
there should be at least two specimens with a higher and two with a lower binder
content than the one that turns out to be the optimum value. If this is not the case, the
procedure should be repeated with additional binder contents. The specimens are
compacted by drop-hammer procedures, using simple, standardised equipment. The
compacted specimens are analysed and tested in the Marshall test device (Figure 1.4).
The specimens are placed in the bottom half of a split-breaking head, after which the top
of the head is positioned against the top of the specimen. In this configuration the top
and bottom of the head do not touch each other. During the tests the gap between them
is diminished by a constant rate of 2” per minute (0.85 mm/s), as a result the cylindrical
specimens are loaded perpendicular to their cylinder axis. During the test the applied
head displacement and the corresponding load are measured. The maximum force is
designated the Marshall stability and the displacement at the moment that this maximum
occurs is the Marshall flow. The Marshall quotient is the division of the stability by the
flow. The test is performed on specimens that were stored at 60oC, and the test is
performed rapidly, so it is assumed that a temperature controlled cabinet is not
necessary.

6 Chapter 1
Figure 1.4: The Marshall test device
Both mixture design procedures are basically meant to compare a series of binder
contents to find the optimum one for a specific aggregate gradation. Of the two
methods, Hveems procedure is more elaborate and, because it addresses both the risk of
too little and too much binder with a separate test, more transparent. The Marshall
method tries to determine the optimum binder content on the basis of a single test. By
determining two parameters (stability and flow) from this test, it seems to address two
things. However, it is not really clear what happens in a Marshall test and how either
under-filled or over-filled mixtures will influence the results. This is an important
difference between the two methods, because although Hveem’s set-ups seem rather
archaic nowadays, their concepts are based on a sound approach towards mix
characterisation. Looking back, it is clear that the Marshall method, which provides
some numbers with no clear relation to mix characteristics or behaviour, could replace
the more fundamental Hveem procedure only because of the need for quick construction
during the war and the ensuing reconstruction period.

The link to road constructions is, in both cases, established trough intervals for the test
parameters (stabilometer value, cohesiometer value, stability etc.). These intervals are
based on experience. As a result, their applicability is limited to similar situations:
similar construction, climate, traffic conditions etc.. Even then, it remains unclear what
kind of pavement response can be expected on the basis of the mixture design results.
Eventually, this led to the development of pavement design procedures, which had to
provide a link between the mixture characteristics and the performance in a road.

1.1.3.2 Pavement design


To move from mixture design to pavement design requires knowledge about the loads
and construction geometry as well as material characteristics. The most well known
road design method is the Shell Pavement Design Manual (Shell 1978). The first
version of this design method was published in 1963. The method consists of a series of
charts, which were based on research and accumulated knowledge, such as the well-
known van der Poel nomograph, which gives the asphalt stiffness as a function of the
loading frequency, temperature and asphalt characteristics (Penetration Index and Ring
and Ball temperature). The design manual can be used as a recipe, using only limited
input parameters but it also allows the engineer to take initiative and incorporate his
own data. The method combines stresses and strains from linear-elastic multi-layer

Introduction 7
analyses (LEMLA) with performance criteria. These latter are expressed as relations
between the number of load repetitions and the initial strains or deformations that are
observed in laboratory tests. The misfit between the number of load repetitions observed
in tests and that found for actual constructions is accounted for by empirical correction
factors, which are explained by phenomena such as healing and traffic wander that do
occur in reality but are not part of the laboratory tests.

1.1.3.3 Limitations of the classic approaches


The three main drawbacks of the currently used rheological material representation are,
firstly, that these material models are one dimensional. As such, they do not account for
the confinement sensitivity of the material although it is well known that the carrying
capacity of the material increases with increasing confinement. Secondly, these model
do not distinguish between tension and compression behaviour. Yet this is important
because the damage caused by tension and compression loading differs. Thirdly, the
models do not include material deterioration and as a result they cannot be used to
investigate the damage mechanisms and development.

There are also two important limitations in using design approaches based on MLA. The
first one is the fixed geometry, which means that MLA can only be used to study
horizontally infinite layers. As a result, discontinuities such as joints or cracks cannot be
modelled. Also, reflective cracking or pavements with joints cannot be studied and
simulated with these techniques. For the same reason the analyses of laboratory tests,
another important aspect of research, is not possible with MLA. The second, more
serious, limitation is that the material characteristics within one layer cannot vary. In a
pavement the temperature and especially the strain rate and state of stress vary from
point to point, in horizontal as well as vertical direction. Since the stiffness, Poisson’s
ratio and strength are functions hereof, they also vary. These variations can lead to
damage at positions other than those where, according to analyses with uniform
characteristics, the highest stresses occur.

Because of these limitations, the results from the existing approaches and pavement
performance is usually based on empirical relations, which seriously limits the
applicability. Any changes in the materials, environmental, structural or loading
conditions such as higher vehicle speeds, increased traffic numbers and loads, new
materials and mixtures, require new empirical relations. In combination with the above
mentioned observations about areas which can not be covered by the available
techniques this led to a demand for new, additional methods to fill the gaps.

1.2 OBJECTIVES OF THE ACRe PROJECT


Form the previous section it is clear that there is a demand for improved design
procedures for both pavement structures and asphalt mixtures. In order to achieve this, it
is necessary to obtain more insight in the damage mechanisms that occur through a
better understanding of asphalt concrete response.

This requires analysis methods that can deal with geometrical discontinuities, variations
in material characteristics and damage development. A way to address those topics is
the use of the Finite Element Method (FEM), a numerical method that is used
successfully in many related engineering fields. In this method the effect of geometry
and material influences is dealt with separately. This alleviates the need for empirical
relations between laboratory results and actual pavement response. However, to exploit

8 Chapter 1
the possibilities of FEM to the fullest, a reliable material model capable of describing
the response to an arbitrary state of stress is needed. To arrive at such a model for
asphalt concrete, the response of the material to a number of states of stress must be
determined at various temperature and strain rate combinations.

In the Asphalt Concrete Response (ACRe) project such a material model as well as the
test set-ups and procedures needed to characterise the material are developed.

The model should include the confinement, temperature and strain rate sensitivity of the
material and distinguish between the damage that occurs due to tension and
compression loading, respectively. Furthermore, it must be able to describe the response
observed in the laboratory test used to determine the model parameters as well as tests
that were not used as input.

In order to obtain information on the material behaviour, it is necessary to subject the


material to tests that result in an a-priori known simple and preferably uniform state of
stress. By doing this for several states of stress while measuring the response, the
necessary input for the material model is found. The tests traditionally performed in
road engineering research are not suited for such a material characterisation procedure
because they lead to complicated internal states of stress. The observed response is then
related to that particular state of stress and can not be generalised to other situations.
Therefore, set-ups that allow material characterisation along the lines described above
had to be developed within this project.

1.2.1 THE MATERIAL MODEL


The material model used in this project is a so-called plasticity model. Such a model
describes the material response on a continuum level, which means that it considers a
volume of the material that is large enough to neglect the material heterogeneity. The
model consists of three components, first a limiting surface that separates the states of
stress that causes changes in response from those that do not. The second model
component is a set of relations that control the expansion and contraction of this surface.
The third component consists of a number of relations between stresses and strains for
both the states of stress inside a volume bounded by the limiting surface and those on
that surface. A simple kind of material model is the elastic-ideal plastic model (Figure
1.5). In this case there are only two components of material behaviour, linear-elasticity
until the strength is reached and ideal plasticity once the stress equals the strength. The
latter leads to increasing deformations without additional loading. In this case, the
model consists of two constitutive relations and a single, static limit surface.

Introduction 9
σ σ2 σ2
fy
σ1 σ1

σ3
ε

Constitutive relations: Inside the ellipse: Inside the cylinder:


σ<fy: 1) σ=Eε relation 1) relation 1)
σ=fy: 2) σ=E(ε−εp) On the ellipse: On the cylinder:
ε& p = λ& ∂f / ∂σ relation 2) relation 2)

Figure 1.5: A material model is 1: a limit surface that separates states of stress that
cause different kinds of response and 2: stress-strain relations for these responses
If the material behaviour is more complex, as is the case for asphalt concrete, the size
and shape of the limit surface will also change. In that case, the relations that control the
surface are also part of the model (Figure 1.6). The model that was developed in the
ACRe project is based on the flow surface developed by Desai (Desai et al., 1986). This
surface was selected because it is a single surface model (all the borderline states of
stress are described by one single, continuous expression), which is very flexible in
shape, size and position. Furthermore, the parameters that control the surface are related
to physically relevant phenomena or characteristics.

Since the damage that results from tension loading differs considerably from that caused
by compression, these phenomena are described separately. Compression damage
weakens the material in all directions, causing it to be less resistant to subsequent tensile
loading. This is described via isotropic softening in de the Desai-based model. Tension
damage, on the other hand, causes a plane of weakness perpendicular to the loading
direction. If this plane is later on loaded in compression, its compressive strength is not
affected. This is an important difference because in pavement structures the material is
alternately loaded in tension and compression and their different effects influence the
development of damage. Tension damage is described via the notion of fixed-cracking
by a cohesive-crack type description on the basis of the Hoffman model (Scarpas et al.
1998).

The complete ACRe model is discussed in detail in Chapter 7.

10 Chapter 1
σ2

fy 3
2
σ1
1
4
4
1
2
3
ε
Figure 1.6: A more complex material model with changing limit surfaces, this
requires additional relations to control the shape and size of the surface

1.2.2 TEST PROGRAMME


A test programme was developed to provide the parameters that control the response
surface. These parameters are different for different materials and have to be determined
from experiments. Each parameter is related to some component of pavement response
(e.g. strength or the onset of inelasticity). In order to obtain insight in the material
behaviour, it is necessary to separate the overall behaviour into geometry and material
components. Minimising the influence of geometry in a test, in order to arrive at a
simple and uniform internal stress distribution is the classical way to achieve this. If one
measures the material response to a series of uniform, a-priori known states of stress
one can generalise this to obtain a relation between stress and strain for any 3D state of
stress. Such a relation is a material or constitutive model and, as discussed above, it
describes the response of the material to any combination of states of stress. Using such
a model it is possible to analyse road constructions as well as laboratory tests such as an
Indirect Tension or Marshall test. The construction or test geometry can be modelled by
means of the Finite Element Method (FEM), which combines the geometry and loading
with the material model to analyse what happens in the combined system.

Obviously, the quality of the model depends largely on the reliability of the laboratory
tests, which in turn depends on the uniformity of the states of stress applied in the tests.
This requires rather elaborate test procedures and sophisticated set-ups and cannot be
achieved with standard road engineering tests. Therefore, the ACRe test programme
aimed to develop the set-ups as well as the test and data-acquisition procedures to
facilitate model parameter determination. Based on the input that was required, it was
decided to develop a uniaxial compression and tension set-up and a multi-axial testing
device. Because AC response is confinement sensitive, the material is stronger when
confined, the multi-axial device should preferably be able to apply tension-compression
rather than purely compressive states of stress. The four-point shear test set-up
developed in a previous PhD project in the Road and Railway Research Laboratory
(Bondt, de 1999), was selected as the basis for such a multi-axial set-up. The different
test set-ups, the instrumentation and data-acquisition used as well as the test results are
discussed in Chapters 4 through 6.

Introduction 11
1.2.3 THE INTERACTION BETWEEN MODEL AND TEST PROGRAMME
The model was developed iteratively during the project. The successive stages of this
development have been reported throughout the project (Scarpas et. al, 1997, 1998 and
1998(b), Erkens et al. 1998, 2000 and 2002). Each stage of model development was
implemented in the Finite Element Program CAPA-3D (Scarpas 1992) and used for
simulations. These simulations allowed assessment of the numerical stability of the
successive stages of the model as well as the interaction of the various components and
its agreement with the test results.

1.3 ORGANISATION OF THIS THESIS


Since the ACRe project consists of an experimental and a numerical/theoretical part,
these subjects can also be distinguished in this thesis. After this introductory chapter,
Chapter 2 provides a more detailed description of material models and the reasons to
develop a 3D, non-linear material model for asphalt concrete. In Chapter 3 the material
that was used in the test programme is described. Some road engineering tests
performed on the material are presented and the results are used to provide a frame of
reference for the material. Chapters 4 through 6 deal with the different tests that were
developed and performed, the uniaxial compression test is presented in Chapter 4, the
uniaxial tension test is discussed in Chapter 5 and Chapter 6 is dedicated to the
multiaxial tests. In these chapters the development of the set-ups and test procedures as
well as a series of tests performed on each set-up are discussed. In Chapter 7 the
material model development is discussed and the complete model is presented. The
determination of the model parameters using the results from the tests that were
discussed in the preceding chapters is also elaborately discussed. In Chapter 8 the ACRe
model is used for a series of simulations and the results are compared to those of the
laboratory experiments. Also, analyses of the Indirect Tension Test (ITT) are presented
in this Chapter. These simulations illustrate the complex stress state that exists in this
test and the necessity to describe both tension and compression damage in order to
accurately capture the complete response. Finally, in Chapter 9 the conclusions and
recommendations are presented.

1.4 REFERENCES

1.4.1 LITERATURE
Asphalt Institute, the (1965), Asphalt as a Material, the Asphalt Institute Information
Series no. 93, Revised Edition, College Park, Maryland
Bondt, A.H. de, (1999), "Anti-Reflective Cracking Design of (Reinforced) Asphaltic
Overlays", Ponsen & Looijen
Brzesowsky, R. and Nijsen, D., (1988), "65 years of Asphalt in the Netherlands (in
Dutch)", drukkerij Verweij bv., Mijdrecht
Desai, C.S., Somasundaram, S. and Frantziskonis, G., (1986), " a Hierarchical
approach for Constitutive Modelling of Geologic Materials", Int. Journal of
Numerical and Analytical Methods in Geomechanics, Vol. 10, No. 3pp. 225- 257
Erkens, S.M.J.G., (1998), The Uniaxial Compression Test - Asphalt Concrete Response
(ACRe), Delft University of Technology, Report number 7-98-117-2
Erkens, S. M.J.G., Liu, X., Scarpas, A. and Molenaar, A.A.A., (2000), 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

12 Chapter 1
Erkens, S. M.J.G., Liu, X., Scarpas, A. and Kasbergen, C., (2002), Issues in the
Consitutive Modeling of Asphalt Concrete”, Proceedings of the 3rd International
Symposium on 3D Finite Element in Pavement Engineering, Amsterdam, the
Netherlands, April 2-5 2002
Hindley, G., (1971), "a History of Roads", the Chausser Press Ltd., Bungay, Suffolk,
ISBN 432-06-7361
Hveem, F. and Davis, H., (1950), "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
Lee, A., (1953), "Latest Developments in the Asphalt Industry", text of a speech held at
the 25th aniversary of the Vereniging voor Bitumineuze Werken (VBW), (in
Dutch)
Scarpas, A., (1992), " CAPA -3D Finite Element System - Users manual, parts I, II and
III.", Delft University of Technology
Scarpas, A., Blaauwendraad, J., Al-Khoury, R.I.N. and Gurp, C.van , (1997),
Experimental calibration of a viscoplastic-fracturing computational model, Paper
presented at the Computational Methods and Experimental Measurements
(CMEM), Rhodos, Greece, May 1997
Scarpas, A. and Blaauwendraad, J., (1998), “Experimental Calibration of a Constitutive
Model for Asphaltic Concrete”, Proceedings of the Euro-C conference on
Computational Modelling of Concrete Structures, Badgastein, Austria, March 31st
April 3rd 1998
Scarpas, A. and Blaauwendraad, J., (1998(b)), Experimental Calibration of a
Constitutive Model for Asphaltic Concrete, Paper presented at the Euro-C
conference on the Computational Modelling of Concrete Structures, Badgastein,
Oostenrijk, 31 march- 3 april 1998
Shell (SPDM), (1978), "Shell Pavement Design Manual", Shell International Petroleum
Company Ltd., London

1.4.2 INFO FROM THE WORLD WIDE WEB


ArtsEdNet, accessed in November 1999, http://www.ArtsEdNet.getty.edu/, a site with
background information on Roman Roads
Auburn University Alabama, accessed in October 1999, http://www.auburn.edu/, site on
Roman Road construction, inc. pictures
Teaching Ideas, accessed in October 1999, http://www.teachingideas.co.uk/,
background information on Roman road constructions
Via Domitia, accessed in November 1999, http://www.viadomitia.org/uk/index.html,
site for the promotion of the Roman roads in the Meditertanean Region and
Europe, contains tourist information as well as many pictures

Introduction 13
14 Chapter 1
2. MODELLING OF PAVEMENT MATERIALS & STRUCTURES
Model: (2) simplified description of a system etc., to assist calculations and
predictions
The pocket Oxford dictionary, 1992
Equation Section (Next)

2.1 WHAT IS A GOOD MODEL?


A model is, by definition, a simplification used to explain, calculate or predict
something, for example a physical phenomenon. In the model the real situation is
simplified by leaving out those aspects that are not important for whatever is being
investigated. As a result, what is a “good” model depends on what it is used for and how
accurate the prediction or calculation should be. In engineering, models are used in the
design and analysis in order to arrive at safe and economic structures. The behaviour of
a structure depends on its geometry, the applied loads and the material(s) used. To
arrive at a good design method both the structural aspects and the material behaviour
must be modelled with sufficient accuracy.

Although road-engineering structures appear not to be very complicated since there are
no high-tech joints or complicated structural elements involved, they are difficult to
analyse. They are long, layered structures that are continuously supported in a way that
depends on the subgrade and, thus, may vary considerably over the length of the
structure. The position of the loads varies both in longitudinal and lateral direction
(wandering) and the stresses in a point of the pavement depend on the position of the
load with respect to that point. Finally, the materials used in a pavement are highly
confinement sensitive, many of them are strain rate dependent and the behaviour of
asphalt concrete is also a function of temperature. As a result, developing reliable
models for both the structure and the materials is complicated. The overwhelming
number of variables makes it hard to perform in-depth pavement analyses without
considerable computer power. Consequently, design and analysis procedures suffer
from significant simplifications.

2.2 PAVEMENT DESIGN METHODS


Until the 1920’s, a fixed asphalt concrete thickness was used for the wearing course,
regardless of the conditions. When the experience increased, methods were developed
to arrive at more suitable and economical structure for a given set of conditions. These
methods were usually the result of practical experience which was related to a given
climate or type of structure. Originally, these design procedures did not involve any
strength tests. An example is the thickness design system based on the soil classification
system. In this approach, the required subbase and pavement thickness was based solely
on the quality of the subgrade, expressed in terms of its yielding and plasticity indices.
The first type of design method using an indication of strength was the California
Bearing Ratio, introduced by the California Highway department. In this approach the
subgrade resistance to penetration is determined and compared to that of a standardised
crushed rock. The required pavement thickness is determined on the basis of the ratio of
the subgrade resistance to that of the crushed rock, using an experience-based estimate
of the overall quality of the pavement structure.

Modelling of pavement materials & structures 15


As mentioned before, pavements can not be schematised easily, beause their continuous,
layered nature requires (semi-) 3D analyses. For this reason structural modelling of
pavements became possible only with the development of the Multi-Layer theory by
Burmister (1945, 1945a and 1945b). His method was based on the well-known work by
Boussinesq (Timoshenko et al. 1951) who developed a solution for a linear-elastic half
space loaded by a point load at its boundary plane. Since the material of the half-space
is considered to be linear-elastic, the superposition principle can be used to extend this
solution to other types of loading, such as that of a distributed load (Figure 2.1).

2a
q

σzz

τzr θ
τrz
σrr

r z
σθθ
Figure 2.1: Distributed load on a semi-infinite half-space
In reality, many subgrades are layered systems and for this reason Burmister adapted the
Boussinesq model to describe a system of two (Burmister 1945 and 1945a) and later on
three layers (1945b). In this approach, the bottom layer of the system is again a linear-
elastic half space as used by Boussinesq, and the layer(s) above are infinite in width, but
finite in thickness. Thus, a better representation of the layered nature of many subgrades
is obtained. The layers are considered to be linear-elastic and the properties in a layer
are constant. However, even with those restrictions the calculations were complicated,
which initially limited the application to a series of influence charts.

The increased availability of computers resulted in the automation of the Multi-Layer


theory in Multi Layer Analyses (MLA) programs. This meant a considerable increase in
the possibilities for structural analyses, providing more flexibility of the load patterns
and pavement structures that could be analysed. For the first time, a theoretically sound
analysis of pavement structures was possible. As input for these programs, the elasticity
parameters of the materials had to be provided. Although linear-elasticity is the simplest
of material representations, determining its parameters for a material that exhibits elastic
behaviour in combination with viscous, plastic and fracturing response is extremely
complicated. Especially since, even if permanent deformations and fracture are not
considered, the viscous aspects cause the stiffness of the material to be strain rate
dependent. In combination with the temperature dependent nature of asphalt concrete
characteristics, this must be incorporated somehow in the “elasticity“ parameters used in

16 Chapter 2
MLA programs. The use of these programs led to the so-called mechanistic-empirical
approaches, where the MLA provided a mechanistic basis for analysis and empirical
relations are used to link the output of the analysis, laboratory tests and pavement
response.

Although the MLA offers a useful structural model for the analysis of pavement
systems, the material representation used is not very realistic. Boussinesq and
Burmister, whose work forms the foundation of the MLA, assumed that the material
was linear-elastic, in which case the superposition principle is valid. This means that the
response should be confinement insensitive, which is definitely not the case for granular
materials or asphalt concrete. For this reason, some researchers have incorporated non-
linear-elastic (state of stress dependent) and viscous characteristics in multi-layer
analysis programs. Although these changes lead to a more reliable material
representation the impact is limited. Since the MLA uses the same material
characteristics for a whole layer, horizontal variations in strain rate, temperature and
state of stress can not be taken into account. Basically, the material representation
remains one-dimensional and there is no strength or material degradation criterion
involved. As a result, current road design methods are still mechanistic-empirical in
nature, using separate criteria for, for example, fatigue cracking and rutting. This is not
in accordance with the situation in a pavement where usually several types of distress
occur and might interact.

2.3 THE ACRE PROJECT


The rapid increase in computer power in recent years enabled the incorporation of more
complex material representations in MLA packages and it also enabled the use of other
numerical methods. One of these methods, the Finite Element Method (FEM) is
successfully used in other engineering fields such as soil, rock, concrete and steel
engineering. In this method the structural and material influences are separated, using
standard mechanics principles. A structure is subdivided in small parts (the elements)
and the strain distribution is determined on the basis of the displacements. The stresses
that belong to these strains, which have to be in equilibrium with the external loads, are
then determined from the material model. Depending on the aims of the analyses this
material model can be more or less complicated.

Since it is usually difficult to extend a simple model, it was decided to develop in the
ACRe project a material model that included all the important influence factors. Since
asphalt concrete is the most complex material used in pavements, the model was
developed for this material. It can be used to describe other materials by turning off
those aspects (e.g. temperature dependence) that are not needed for that particular
material.

2.3.1 MATERIAL PARAMETER DETERMINATION


A material model describes the way in which a given material responds to an applied
load. Generally, such models are used to predict or analyse the response to complicated
states of stress on the basis of experiments in which simple and reproducible states of
stress exist. A simple state of stress in the experiment is a pre-requisite to link the
observed response to the applied state of stress. If the latter is not simple and uniform, it
is not possible to establish a one to one relation between stresses and strains, because in
that case the response will be the result of some complicated combination of stress
states.

Modelling of pavement materials & structures 17


For many confinement sensitive materials (concrete, soils) the triaxial test is used to
determine the material characteristics. If the model parameters for granular materials
have to be determined, this is the ideal test. However, asphalt concrete is particularly
sensitive to tension-compression states of stress and these are difficult to apply using
triaxial cells. For that reason the cohesive nature of the material is exploited and
uniaxial tension and compression tests are used as the basic tests in the ACRe model
parameter determination procedure. Most of the characteristics needed can be found
from these tests, additional information that requires a multi-axial state of stress is
obtained from four-point shear tests (Chapter 6. ). None of these tests is standard in road
engineering, while the standard road engineering tests lead to complicated internal states
of stress and are, thus, not suited for model parameter determination. For this reason the
set-ups and instrumentation needed to characterise asphalt concrete were developed
within the ACRe project. This work, the specific requirements set by the material, the
set-ups and test results obtained using them are discussed in Chapters 4, 5 and 6.

2.3.2 THE ACRE MATERIAL MODEL


The model developed in this project is a classic plasticity model. In such a model a
surface in the 3D-stress space, which is called a yield or flow surface, separates states of
stress that cause different responses. In its simplest form the yield surface separates
elastic from inelastic response and it is combined with constitutive relations that
describe the strains as a function of the stresses for both these types of response. The
most well known example is that of uniaxial ideal plasticity (Figure 2.2). For stresses
smaller than the yield stress the material behaves elastically, according to Hooke’s law,
and for a stress equal to the yield stress the material starts to yield. In other words, it
exhibits increasing deformations at the same stress level, which means that the relation
between stresses and strains has changed. The state of stress at which the transition
between the two types of response occurs is determined by the flow surface. This
surface is the collection of all the stress combinations that will cause this transition from
one type of response to the other. In the uniaxial case, this transition is a point, in a two-
dimensional situation it is a line and in the three-dimensional case it is a surface. The
flow surface is expressed as a function of the state of stress (f(σ)), if this function is
negative (f(σ)<0) the state of stress is within the yield surface and the response is
elastic. If the function is zero for the applied stress (f(σ)=0), that state of stress is
positioned on the yield surface and the response is inelastic. States of stress outside the
yield surface cannot exist.
σ σ2 σ2
fy
σ1 σ1

σ3
ε

Constitutive relations: Inside the ellipse: Inside the cylinder:


σ<fy: 1) σ=Eε relation 1) relation 1)
σ=fy: 2) σ=E(ε−ε p) On the ellipse: On the cylinder:
ε& p = λ& ∂f / ∂σ relation 2) relation 2)

Figure 2.2: In 1D states of stress for which transition from elastic to plastic occurs
form a point, in 2D a curve and in 3D a surface

18 Chapter 2
If the material behaviour is more complicated, there is more than one such surface.
Initially, there is the elastic limit surface that grows, a phenomenon known as
hardening, and after peak it again decreases, which is called softening (Figure 2.3). In
this case, the surfaces can no longer be regarded as yield surfaces, so the more general
term response surface or response envelope will be used.

σ2

fy 3
2
σ1
1
4
4
1
2
3
ε
Figure 2.3: More complicated material response can be described by a series of
successive response envelopes …………………………..

2.3.3 MODEL CAPABILITIES


The use of a 3D response surface allows the incorporation of stress dependency, the
transition of one type of response to another now depends on the full 3D stress state.
The model will describe the material response to any state of stress at any temperature
and strain rate. Because the state of stress for which this damage occurs is known, it is
also possible to predict the type of damage that will occur at a given position. For
example, in a volumetric tension region cracking will occur while volumetric
compression results in rutting. These phenomena can occur within one structure, at
different positions (Figure 2.4).

Modelling of pavement materials & structures 19


Volumetric

compression

tension

deviatoric

Figure 2.4: Different types of damage within the same structure can be described
The aim of the project is to arrive at such qualitative descriptions of structures on the
basis of realistic structural and material modelling, rather than quantitative predictions.
Due to the large variations that occur (in mixture composition, loading etc.) in a
pavement it would not be realistic to expect the exact prediction of rut depth. It is
possible, however, to predict the weaknesses of a given structure, the dominant failure
types and the way in which damage will develop.

In combination with the FEM any geometry can be modelled, whether it is a pavement
or a laboratory specimen (Blaauwendraad et al. 1986, Huebner et al. 1995, Erkens 1998
and 2000). The variations in stresses and strain rates will automatically be taken into
account for every point in the FEM model and this will influence the response that is
observed. As a result, different mixtures used in the same structure might give rise to
different failure patterns. The model and the physical meaning of the model parameters
are discussed in Chapter 7. In Chapter 8 some applications of the model are shown, to
illustrate the capabilities.

2.4 REFERENCES
Blaauwendraad, J., Borst, R. de and Custers, G., (1986), Finite Element method: non-
linearity module, PATO section Civil Engineering and GeoSiences (in Dutch),
Burmister, D., (1945), the general Theory of Stresses and Displacements in Layered
Systems I, Journal of applied Physics,
Burmister, D., (1945a), the general Theory of Stresses and Displacements in Layered
Soil Systems II, Journal of applied Physics,
Burmister, D., (1945b), the general Theory of Stresses and Displacements in Layered
Soil Systems III, Journal of applied Physics,
Erkens, S., Scarpas, A. and Liu, X., (2000), 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
Erkens, S., (1998), Some Notes on the Finite Element Method, Delft University of
Technology, Report number 7-98-117-2

20 Chapter 2
Huebner, K., Thornton, E.A. and Byrom, T.G., (1995), The finite Element Method for
Engineers, John Wiley & Sons Inc., USA
Timoshenko, S.P. and Goodier, J.N.(1951), “Theory of Elasticity”, McGraw-Hill, New
York

Modelling of pavement materials & structures 21


22 Chapter 2
3. THE ACRE MIXTURE
Equation Section (Next)
“This is one instance out of many that could be produced to show the absolute
necessity of correctly knowing the exact conditions under which any tests are made
before we can equitably compare results obtained from different quarters.”
-David Kirkaldy (1820-1897)-

3.1 SELECTION OF THE MIXTURE


At the beginning of the project, the physical meaning of the model parameters and the
methods to determine them were known for granular materials, for which the flow
surface and constitutive relations (Desai et al. 1986) were originally developed. Since
the model had to be adapted and extended to describe asphalt concrete, the expressions
for the model parameters also changed. The aim of this project was to establish suitable
model parameter determination procedures for asphalt mixtures. This implied the
development of the necessary tests and instrumentation procedures as well as the data
analysis methods. During the project, the model predictions were compared to the test
results whenever possible and if necessary, the model was again modified. On the other
hand, the model set the demands for what had to be measured in each test. As such,
model and test program development were highly interactive, which led to a time
consuming process. For this reason, it was decided that the tests would be performed on
a single mixture. Once the procedures are established and the set-ups available, they can
be used to characterise other mixtures as well.

The test results are used to develop a model that can predict the material response for
any stress situation and under all conditions. Since asphalt concrete is a temperature and
loading rate dependent material, this meant that the model had to be valid for many
temperatures and loading rates. Because interpolation of test results is usually more
reliable than extrapolation, the test conditions were chosen to cover a wide range.
Preliminary results at low temperatures and high loading rates (for specific test
conditions, see Chapter 4) showed that this may lead to very high compressive
strengths. To limit the necessary forces to a range that could be handled by the available
equipment the minimum specimen dimension (Dmin) was kept relatively small (50 mm
diameter). To prevent unnecessary variations in the test results, the maximum aggregate
size (Dg,max) must be small compared to the smallest specimen dimension (Dmin). This
requirement is usually set to something between Dmin ≥ 3Dg,max and Dmin ≥ 5Dg,max.
(Head 1982, Mier 1997 and Transportation Research Board 1997) based on the
maximum aggregate size that still gives reasonably reliable test results. Since this test
program would be used to develop a prototype material model, it was decided to
minimise the possible disturbances. This led to the requirement of a relatively
homogeneous and isotropic mixture with a desired maximum aggregate size of
approximately 1/10 of the minimum specimen dimension (=1/10 x 50= 5 mm). From a
previous PhD. project it was known that a rather greasy “sand asphalt” mixture fulfilled
these requirements (Jacobs 1995). Although similar materials were used for the ACRe
mixture, the actual composition is somewhat different because of different mass
densities (especially for the filler) and compaction methods.

The ACRe mixture 23


3.2 MIXTURE COMPOSITION
The mixture under consideration consisted of crushed rock, filler and bitumen 45/60.
The mixture composition is shown in Table 3.1.

Component Density Mass percentages Volume percentages


[kg/m3] [%] [%]
Crushed rock 2675 ± 15 77.1 ± 0.05 66.3 ± 0.4
Filler 2780 or 2770 ± 10 14.3 ± 0.05 11.8 ± 0.05
Bitumen 1025 ± 5 8.6 ± 0.05 19.3 ± 0.1
Air - - 2.6
Table 3.1: Density of the mixture components and their mass and volume percentages

3.2.1 CRUSHED ROCK


The coarser aggregates in the ACRe mixture were crushed rock particles that were
called “crushed sand” in the previous project. This is a rather inconvenient name since
this “sand” has a maximum grain size of 5 mm, while the largest grain size in sand is
usually set to 2 mm. Therefore, it will be called crushed rock throughout this thesis.
However, it is the same material referred to as “sand” or “crushed sand” in earlier
publications. A more detailed overview of the composition of this crushed rock is
shown in Figure 3.1 and Table 3.1.

percentage passing [%]


100

90
crushed rock
80
aggregate
70

60

50

40

30

20

10

0
0.01 0.1 1 sieve size [mm] 10

Figure 3.1: Sieve curves of the crushed rock and aggregate (crushed rock + filler)
used in the mixture

24 Chapter 3
Dry sieving Crushed rock 0/5 Aggregate (rock + filler)
Sieve size [mm] Percentage passing Percentage passing
[% m/m] [% m/m]
5.6 100 100
4 97.5 96.8
2.8 94.9 94.4
2 90.6 90.8
1 60.4 65.3
0.5 34.5 43.4
0.355 26.4 36.6
0.25 18.1 29.6
0.18 11.3 25.2
0.125 5.9 20.6
0.063 1.8 14.6
Table 3.2: Sieve data of the crushed rock and aggregate used in the mixture
For the crushed rock the density was determined through test 60.1: Density of non-
porous rock or rock-like materials (C.R.O.W. 1990) using the average of three samples.
This resulted in ρs = 2675 ± 15 kg/m3.

3.2.2 BITUMEN
The data provided on the bitumen was ρb = (1025 ± 2) kg/m3. Although the binder was
heated in small portions, increased ageing might have occurred due to prolonged or
repetitive heating from binder used in specimens that were produced at the end of the
day. Therefore, the uncertainty was enlarged to 5 kg/m3 in order to take the possibility
of ageing into account.

Additional information on the binder was obtained from standard penetration and ring &
ball tests (tests 32 and 47, C.R.O.W. 1990). The bitumen used throughout the project
consisted of two batches, each purchased at a different time. The information on the binder
characteristics in each batch are shown in Table 3.3.

Penetration (batch I) [x10-1 mm] Penetration (batch II) [x10-1 mm]


1e 2e 3e average 1e 2e 3e average
47 46 48 47 47 48 47 47.3
TR&B (batch I) [oC] TR&B (batch II) [oC]
thermostat ring TStart Result thermostat ring TStart Result
45 9 8.6 51/51 45 9 7.2 52/52.1
Table 3.3: Bitumen characteristics per batch
Based on these results the Penetration Index (PI) was determined using Equation 3.1.

20TR& B + 500 log( pen) − 1952


PI = (3.1)
TR& B − 50 log( pen) + 120

This yielded a PI of -1.10 for the first batch and PI = -0.83 for the second batch of binder.

The ACRe mixture 25


3.2.3 FILLER
Similar to the binder, the filler used throughout the project was also taken from
different, in this case three, deliveries (or batches). Based on the filler data for each
batch (Table 3.4), the uncertainty in filler density was determined, based on the
variation between the three batches. This resulted in ∆ρf = 10 kg/m3. The actual
densities of the different batches of filler were 2780, 2770 and 2770 kg/m3, respectively.

WI ANALYSES SI Value of batch no. Warranty


FILLERANALYSES I II III Min. Max.
231 GRAIN SIZE DISTRIBUTION ALPINE
+ 63 µm %(m/m) 19 15 14 5 25
+ 90 µm %(m/m) 11 9 9 0 15
+ 2 mm %(m/m) 0 0 0 0
232 BITUMEN NUMBER ml/100g 47 44 46 42 48
235 MASS LOSS 150 °C %(m/m) 0.34 0.6 0.1 0.0 1.5
ANALYSES AFTER DRYING AT 110 °C
222 DRYING AT 110 °C %(m/m) X X X
236 DENSITY kg/m3 2780 2770 2770 2675 2875
237 VOIDS %(v/v) 38 38 37 36 42
238 SOLVABILITY H2O %(m/m) 2.0 2.2 0.8 0 10
239 SOLVABILITY HCL %(m/m) 63.3 65.5 67.7 55 75
Table 3.4: Filler data as provided by the supplier (Steengroeve Laboratory
Winterswijk)

3.2.4 COMPARING THE MIXTURE COMPOSITION TO THAT OF STANDARD MIXTURES


The mixture composition of the ACRe mixture compared to that for SA and DAC 0/8 is
shown in Table 3.5.

Passing through SA ACRe-mixture DAC 0/8


8 mm - - 98 – 100
5.6 mm - - 70 - 90
4 mm - 97.9 -
2.8 mm - 95.7 -
2 mm 74 – 100 92.1 42 – 48
1 mm - 66.6 -
0.5 mm - 44.7 -
355 µm - 37.9 -
25 µm - 30.9 -
18 µm - 25.2 -
12.5 µm - 20.6 -
63 µm 4–8 14.6 7–9
Mass% bitumen (+) 3.5 – 5 9.4 6.6 - 7
Table 3.5: Composition of the ACRe mixture, a sand asphalt and a DAC 0/8
From Table 3.5 it can be seen that, with respect to its composition, the ACRe mixture is
a hybrid between Sand Asphalt and Dense Asphalt Concrete. The aggregate sieve curve
seems to correspond to something that could either be a very coarse SA or a downscaled
DAC mixture. However, the bitumen content is relatively high, which is in better
agreement with DAC than with SA.

26 Chapter 3
3.3 ROAD ENGINEERING TYPECASTING OF THE ACRe MIXTURE
In order to provide a road engineering characterisation of the ACRe mixture a series of
typical road engineering tests was performed on the material. Where possible, the
results are compared to those obtained or required for sand asphalt and dense asphalt
concrete, the two materials that are most similar to the ACRe mixture in with respect to
composition.

It must be kept in mind that, due to the complicated stress patterns that exist in the road
engineering tests, the results are not a true material characterisation. However, because
of the experience accumulated over the years, the tests results give road engineers an
indication of the kind of material they are dealing with. Since the tests that form the
basis of this project are not well known in the road engineering community and, thus,
the results do not fit the existing reference frames, these road engineering tests were
performed to provide an alternative typecasting of the material. The tests selected were
the Marshall test, the monotonic and cyclic Indirect Tension Tests (ITT) and the Four-
point Bending Fatigue test (FPBF). In the next sections these tests are discussed in some
detail.

3.3.1 THE MARSHALL TEST


Bruce Marshall, an Engineer with the Mississippi State Highway Department, originally
developed the Marshall Method for the design of asphalt concrete mixtures for
highways. The method was incorporated in an evaluation study of asphalt concrete
design methods by the Waterways Experiment Station (WES). The aim of the study was
to develop a mixture design method that would allow the development of mixtures able
to sustain the increasing loads to which runways were subjected. The Marshall method
was originally used to determine the optimum binder content for a mixture, but
nowadays it is also used for design and quality control. This is based on the vast amount
of experience with the Marshall test. Observations of the response in practice of many
mixtures have led to the formulation of acceptance intervals for the Marshall stability,
flow and quotient values for various mixtures (White 1985). In this respect, the results
from the Marshall test have become important in classifying materials.

The results from the Marshall tests on the ACRe mixture will be compared to the
intervals specified for SA and DAC 0/8. The tests were performed according to the
specifications as described in the Dutch pavement Design Guide (C.R.O.W. 1990).
Because different temperatures are specified for SA and DAC, respectively, the ACRe
mixture was tested at both temperatures. A standard Marshall test head was placed in an
ELE compression frame (Figure 3.2). The head was squeezed at a rate of 0.85 mm/s
(Marshall speed) and during the test the applied force was measured by a load cell,
while the deformation was measured by an externally mounted displacement transducer.

The ACRe mixture 27


Figure 3.2: Set-up for the Marshall tests on the ACRe mixture
The average results are shown in Table 3.6, along with the specifications for SA and
DAC 0/8.

TC = Traffic Class Sand Asphalt (SA) ACRe mixture Dense Asphalt


TC 2: < 500 ESAL100kN Concrete (DAC) 0/8
TC 3: <4000 ESAL100kN TC 2 TC 3 TC 2 TC 3
o
Test temperature [ C] 40 40 40 60 60 60
Marshall stability [N] 4000 5000 15620 7680 6500 7000
Marshall flow [mm] 1.5–3.5 1.5–3.5 5.6 5.3 2.0–5.0 2.0–4.0
Marshall quotient [N/mm] 2000 2500 2790 1530 2000 2500
Table 3.6: Comparing the Marshall characteristics of the ACRe mixture (Arif 1999
and El-Odaisy 1999) to those of SA and DAC 0/8 (R.O.W. 1990)
From Table 3.6 it can be seen that the stability of the ACRe mixture is high. Compared
to the specifications for SA it is more than three times as high as the specified maximum
value. Compared to DAC 0/8 it still exceeds the value required for Traffic Class (TC) 3.
The flow exhibited by the ACRe mixture is also relatively high, which can be explained
by the high binder content. The quotient, which is the stability divided by the flow, is
high compared to the specifications for SA, but compared to those for DAC 0/8 it is
low. Again it appears that this material should be classified as something between SA
and DAC 0/8, but from the above shown results it can be concluded that its Marshall
characteristics resemble a DAC 0/8 much more than a SA.

3.3.2 THE MONOTONIC AND CYCLIC INDIRECT TENSION TEST


The indirect tension test (itt), or Brazilian test, is a well-known test. The basic idea is to
submit a slice of the material under investigation to a compressive force along its
vertical axis. This results in a tensile stress along the horizontal axis, which causes
splitting failure (Figure 3.3).

28 Chapter 3
compression
tension compression
σ
yy, vert
σ
xx vert ,

σ
yy
, horz
σ ,
xx horz

tension
x
y
Figure 3.3: Indirect tension test, set-up (left) and stresses along the vertical and
horizontal cross section
Usually, the indirect tension test (itt) is used because it is easier to perform than a direct
tension test. First of all, the specimens are easy to produce, since they can be cut from a
core. Furthermore, no elaborate alignment and gluing procedures are needed. On the
other hand, the internal stress distribution is complex and the values obtained from the
test are not “real” tension values. If the horizontal and vertical strains at the centre of the
specimen are measured, a reliable Poisson’s ratio can be determined. The stiffness
values, on the other hand, are always approximations because there exists no direct
relation between the applied force and the stress at the centre, due to the complicated
internal stress distribution. A relation for the stiffness can be developed if plane stress in
the specimen and linear-elastic material behaviour are assumed. On the basis of such a
plane stress, elastic analysis also relations between the strains at the centre of the
specimen and the deformations over the diameter can be developed (Hadley et al. 1970).
Such relations have become very popular, since usually the deformations over the
vertical and horizontal axes are measured, rather than the strains at centre of the
specimen. It should, however, be kept in mind that these relations are valid only when
the assumptions on which they are based are met. This means that the thickness of the
specimen should be small compared to its diameter, in order to achieve plane stress.
Analyses ran during the Strategic Highway Research Program in the United States
(Lytton et al. 1993) indicated that for a specimen with a diameter of 4” (approximately
100 mm) the assumption of plane stress was warranted if the thickness was no more
than 1” (25.4 mm). Thus, the specimen thickness in the indirect tension test should be at
least four times as small as the diameter. As stated, the second assumption used in
determining the results of the indirect tensile test is that of linear-elasticity until brittle
fracture occurs. Whether this assumption is valid should also be considered when
interpreting the results of this test, especially at higher temperatures. These results
usually consist of the indirect tensile strength, the stiffness and the Poisson’s ratio.

In road engineering besides the monotonic (or destructive) ITT, also a cyclic indirect
tension test is used. The cyclic test is used to determine the stiffness modulus of the
strain rate dependent material. In this test, the material is subjected to a sinusoidal load,
with an amplitude of less than 10% of the estimated tensile strength to prevent damage
to the specimen. Either a continuous signal or one with rest periods in each cycle is

The ACRe mixture 29


used. At a number of temperatures, usually 3 to 5, a series of different frequencies is
applied. The stiffness values that are obtained in this way are used to construct a
mixture stiffness mastercurve, which expresses the stiffness as a function of temperature
and frequency. This approach is based on the equivalence of temperature and frequency
for linear visco-elastic materials.

f=1/T
S=F(ν+0.27)/(uH,resilient*t)
F
Total pulse width (TT): 2000 ms
t
Load pulse width Rest period D
(Tp=1/2 T)
T

time
Phase lag
uH,resilient

u
Figure 3.4: Signal used in the cyclic indirect tension test
The test conditions and results from both the monotonic and the cyclic indirect tension
tests on the ACRe mixture are shown in Table 3.7. In the monotonic test, a continuously
increasing vertical deformation is applied and the horizontal deformation as well as the
applied force is measured. In the cyclic test used at the Road and Railway Laboratory, a
half-sine force signal is applied. The total pulse width (defined as the duration of the
segment that is continuously repeated) is 2 seconds. That segment exists of the half sine
pulse and a rest period, the load pulse width (Tp) of the half sine pulse is ½ T, where T
is the duration of the full sine signal. In the test the load pulse width (Tp) can be
specified. This is considered to correspond to the frequency of a continuous signal
where f=1/T (=1/(2Tp), as indicated in Figure 3.4.

The cyclic test results are average values of 14 specimens. The values are rounded off
and the variation in the dynamic modulus was approximately 10% for all test conditions
(Medani 1999). The monotonic test results are based on only a single test per condition.

30 Chapter 3
type Deformation Frequency Temperature Stiffness Strength
rate [mm/s] [Hz] [oC] [N/mm2] [N/mm2]
- 8 20 4700 -
- 4 20 3200 -
- 2 20 2400 -
- 1 20 1700 -
C - 0.5 20 1400 -
Y - 8 15 6200 -
C - 4 15 4900 -
L - 2 15 3800 -
I - 1 15 3000 -
C - 0.5 15 2500 -
- 8 10 7400 -
- 4 10 6600 -
- 2 10 5600 -
- 1 10 4800 -
- 0.5 10 4200 -
M 0.85 - 15 250 1.4
O 0.1 - 15 55 0.7
N 0.85 - 5 960 2.7
O 0.1 - 5 590 1.6
Table 3.7: Test conditions and results for the indirect tension tests (Medani 1999)
The difference between the dynamic and monotonic stiffnesses is striking. This
difference can partially be explained by the fact that the static value reported is a secant
stiffness using the origin and the peak of the stress-strain curve rather than a tangent to
the first part of the curve (Figure 3.5). The tangent stiffness would be in better
agreement with the unloading modulus determined in the cyclic tests, which is a
resilient modulus. But apart from this difference it is a well-known phenomenon that the
stiffness observed in monotonic tests is lower than that obtained from cyclic tests, even
if for the former the tangent values are used. In the next section this phenomenon will
be discussed in more detail.
σ
tangent secant

ε
Figure 3.5: A secant stiffness (Medani 1999) is smaller than the corresponding
initial tangent stiffness

The ACRe mixture 31


3.3.2.1 Comparing the results from the cyclic test to those of other mixtures
Since the indirect tension test is not incorporated in any design method, it is not easy to
compare the results obtained for the ACRe mixture to those of other mixtures.
Groenendijk (1998) found that the results of this type of cyclic indirect tension test
agree rather well with those from direct tension test with a full sinusoidal load. Since
such results are available for DAC 0/8 as well as an ACRe-like mixture from Jacobs’
thesis (1995), these results are shown in Figure 3.6, along with the results from Table
3.7. In the graph also values from Verhoeven et al. (1993) and Molenaar (1983) are
shown. It appears that the ACRe mixture shows the kind of behaviour that would be
expected from a DAC-like mixture. Cyclic data from actual SA mixtures was not found,
but both the trend of the stiffness as a function of the frequency and the values
themselves are in good agreement with those from the other researchers. The ACRe
stiffness values at 10oC are higher than the other values, which is natural since it is the
lowest temperature incorporated in the overview. The ACRe stiffness values at 15 oC
fall between those found by the others for that temperature and the stiffnesses at 20oC
are in between those reported in literature at 15oC and at 25oC.
15'C, DAC, Verhoeven et al. 1993 15'C, DAC 0/8, Jacobs, 1995
15'C, SA, Jacobs, 1995 15'C, DAC, Molenaar, 1983
25'C, DAC, Verhoeven et al. 1993 25'C, DAC 0/8, Jacobs, 1995
25'C, SA, Jacobs, 1995 10'C, ACRe
15'C, ACRe 20'C, ACRe
10000
2
Smix [N/mm ]

f [Hz]
1000
0.1 1 10

Figure 3.6: Comparison of the cyclic itt ACRe stiffnesses with values from literature

3.3.2.2 Comparing the monotonic test results to those of other mixtures


The tensile strength values are compared to those obtained for various mixtures with
different types of tests. These results are obtained from several publications, direct
tension results were found in Jacobs (1995), indirect tension data on the ACRe mixture
was published by Medani (1999) and three point bending results on sand asphalt and
dense asphalt concrete were taken from Molenaar (1983). These strength values are
plotted as a function of the strain rate, for a number of temperatures. The strain rate is
used instead of the deformation rate because the material is strain rate sensitive. In the
deformation rate the specimen size and geometry have an influence, by using the strain
rate those influences are excluded. As such, using the strain rate enables the comparison
between these tests with very different geometries. Since the deformation rate is used as
a control parameter in the tests, the strain rate has to be computed and for the different
tests the relation between the specified deformation rate and the strain rate varies. So

32 Chapter 3
before comparing the results from the various tests, the relation between strain rate and
deformation rate is determined for each of them.

Direct tension test


The easiest relation exists for the direct tension test the direction of loading is the
direction in which tensile failure occurred. If the specimen cross section is constant over
the specimen height the strain rate can be found by dividing the deformation rate by the
specimen height.

u&
ε& = (3.2)
h
ε&, u& : strain rate and deformation rate, respectively
h: specimen height

Indirect tension test


In the indirect tension test the loading direction is perpendicular to the direction in
which failure occurs. The internal stress distribution in this test is extremely
complicated (Figure 3.3). The most reliable information about the strain rate in the
horizontal direction can be obtained from strain measurements at the centre of the
specimen (where the crack initiates). Since that information is not available, an
approximation is used. The relations for the stresses in the indirect tension test,
assuming linear-elastic material response and a state of plane stress, Equations 3.3 to 3.6
(Hadley et al. 1970), can be used to obtain relations for the strains and, by integrating
them over the specimen diameter, the deformations.
  r2   r2  
 1 − 2  sin(2α )  1 + 2  
2P   R   R  
σθ y =  − arctan  tan(α )   (3.3)
π at  2r 2
r 4   r 2 
 1 − cos(2α ) + 4   1− 
  R2 R    R2  
  
  r2   r2  
  1 − 2  sin(2α )  1 + 2  
2P   R   R  
σ ry = −  + arctan   tan(α )   (3.4)
π at  2r 2
r 4   r 2 
 1 − cos(2α ) + 4   1 − 
 2 
  R2 R   R  
  r2   r2  
 1 − 2  sin(2α )  1 − 2  
2P   R   R  
σθ x = −  + arctan   tan(α )   (3.5)
π at  2r 2
r 4   r 2 
 1 + cos(2α ) + 4   1 + 
 2 
  R2 R   R  
  r2   r2  
  1 − 2  sin(2α )  1 − 2  
2P   R   R  
σ rx =  − arctan  tan(α )   (3.6)
π at  2r 2
r 4   r 2 
 1 + cos(2α ) + 4   1 + 
 2 
  R2 R   R  

The ACRe mixture 33


Where:
P= the applied load
t = the specimen thickness
a = the width of the loading strip
α = half the top angle of the triangle between loadingstrip and specimen centre
R= specimen radius
r = distance to the centre of the specimen

Using that:
1 1
E E
(
ε x = (σ rx −νσ θ x ) and ε y = σ ry −νσ θ y ) (3.7)

and

R R R
1 1
E −R E −R
(
u x = ∫ ε x dr = ∫ (σ rx −νσ θ x ) dr ; u y = ∫ σ ry − νσ θ y dr ) (3.8)
−R
where:
ε x , ε y = strain in the horizontal and vertical direction, respectively
E = Youngs modulus
ν = Poissons ratio
u x , u y = total specimen deformations along the horizontal and vertical axis, respectively

a relation between the vertically applied deformation (rate) and the corresponding
horizontal strain (rate) at the centre of the specimen can be established. To do this, the
relation for uy was numerically integrated. The total axial deformation that was thus
found, relates to the horizontal strain at the centre as:

ε x,r =0 = 0.0036 u y ,tot (3.9)

for a specimen with D=100 mm, t=25 mm, ν=0.33 and a=12.7 mm. The same relation
exists between the time derivatives of both quantities:

ε&x,r =0 = 0.0036 u& y ,tot (3.10)

Three point bending fatigue test


For the three point bending test, the (bending) strain rate in the outermost fibre is
determined from:

h Fl h
M
σ= 2 = 4 2 = Flh ⇒ ε = σ = Flh (3.11)
I I 8I E 8EI

34 Chapter 3
σ , ε : stress and strain in the outer fibre at the centre of the beam
M: bending moment at the centre of the beam
h: beam height
l: beam span
I: moment of inertia
E: Young's modulus

Using the relation between load and deflection in a beam loaded by a point load at the
centre of its span:

Fl 3
w= (3.12)
48EI
w : deflection at the centre of the beam span
F : applied force
l: beam span
E: Young's modulus
I: moment of inertia

In combination with Equation 3.11 yields an expression for the strain as a function of
the deflection:
6h
ε= w (3.13)
l2
And the relation between the strain and deflection rate becomes:
6h
ε& = w& (3.14)
l2
Direct tension indirect tension three point bending
u u w
x u& time
time time
z x
h D h x
y z
l

D=100 mm
t=25 mm
a=12.7 mm
ν=0.33
u& 6h
ε& = ε&x = 0.0036 u& y ε& = w&
h l2
Figure 3.7: Relation between deformation and strain (rate) for the three tests
Equations 3.2, 3.10 and 3.14 and the principles used to derive them are illustrated in
Figure 3.7. On the basis of these relations the tensile strength values for various tests are
plotted as a function of the strain rates in Figure 3.8. Although the conversion to strain
rates makes the test conditions more comparable, the tests themselves still result in quite
different stress distributions. In the uniaxial tension and three point bending test, a

The ACRe mixture 35


uniaxial tension stress exists at the failure location, in the indirect tension test it is a two
dimensional tensile stress. Because of the confinement sensitive nature of Asphalt
Concrete it is to be expected that the latter test will, for the same temperature and strain
rate result in a somewhat lower load-carrying capacity. Furthermore, the uniaxial tensile
stress in a the tensile test is present along the whole cross-section while in the bending
test it occurs only in the outermost fibre. As a result, the change of imperfections in the
peak stress region in larger in case of a uniaxial tension test than in a bending test,
causing a higher load-carrying capacity in the latter test. For this reason it is not
expected to find perfect matching of the various test results, instead the comparison is
intended to provide a rough indication of the characteristics of the ACRe mixture when
compared to DAC and SA.
10
ft [N/mm2]

15'C, ACRe, dtt, Arif 25'C, DAC, dtt, Jacobs


15'C, DAC, dtt, Jacobs 25'C, SA, dtt, Jacobs
15'C, SA, dtt, Jacobs 25'C, 3pb*, DAB
15'C, ACRe, itt, Medani 25'C, 3pb*, SA
15'C, 3pb*, DAB, Molenaar 30'C, ACRe, dtt, Arif
15'C, 3pb*, SA, Molenaar
0.1
0.0001 0.001 e [mm/s] 0.01

Figure 3.8: Tensile strength values from different types of tests


As was mentioned above when looking at different tests, there is a lot of variation to be
expected. Therefore it is difficult to arrive at definite conclusions. It can, however, be
seen that the strength values for the ACRe mix do not differ considerably from the other
data. Most of the values are in the same range, generally increasing with the strain rate
and decreasing with temperature. It must be noted here that the results from the three
point bending test (3pb) are multiplied by a stress concentration factor because they
were performed on notched beams. In Figure 3.9 the relation for the stress concentration
factor for such a geometry (Tada 1985 and Herzberg 1996) and the values for the
various conditions are shown. The values calculated are related to the peak load found
in the experiments, but it remains unclear to which a/W this responds. It is however
well-known that after the peak the crack growth rate increases rapidly, indicating an
a/W of at least 0.5 and probably higher (K-a/W plot, Figure 3.9) at the peak. As such,
the stress concentration factor for a/W=0.5 is used. In his direct tension tests Jacobs
(1995) also used notched specimens. The stress concentration factor relations for this
test (Tada 1985), however, are usually based on freely rotating specimen ends and in
these tests the specimens are placed between fixed plates. From concrete mechanics

36 Chapter 3
(Mier, van 1994) it is known that the bending moment that develops in tension tests
between fixed plates results in higher tensile loads. Probably, this is because the stress
concentration effect is less pronounced in this geometry, but it is not clear to what
extend this is the case. Because the tensile strength values found by Jacobs agree well
with values generally found for asphalt concrete, it is assumed that for this geometry the
effect of the stress concentrations can be neglected.
K = Yσ π a P
σ = nominal stress at crack location a h
Y= dimensionless geometry factor
S
b
( )
1/2
3PS PS 3 a  a  3 a / h
K3pb = Y π a = 3/ 2 × Y π = f  h = ×
  2(1+ 2a / h)(1− a / h)
3/2
2
2Bh Bh 2 h
PS a
K3pb = 3/ 2 × f  
Bh  h
 (
1.99 −( a / h)(1− a / h) 2.15 −3.93a / h + 2.7( a / h)2 
 )
3/2
KI [N/mm ]
12
11
10
9
8
7
6
5
4
K_DAB_15 K_SA_25
3
2
K_DAB_25 K_SA_15
1
0
a/h
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Figure 3.9: Stress intensity factor for notched 3PB specimens

3.3.3 THE FOUR-POINT BENDING FATIGUE TEST


The four-point bending fatigue test (FPBFT) is the Dutch “standard” test for fatigue
testing. Despite this, the test conditions, set-ups and specimen dimensions vary widely.
An overview of the conditions used in several projects is presented in Table 3.8.

The ACRe mixture 37


temperature frequency lxhxb N #cycles / ε [µε] Source
[oC] [Hz] [mm]
5, 15 and 25 30 450x50x50 N≈ 5x104, 2x105 and Gurp, van et al. 1986
1x106
0 and 20 30 450x50xb N≈ 1x105, 5x105 and Eerden, van der, 1984
5x106
0 and 20 40 / 30* 100, 180 and 250 µε Gurp, van et al. 1985
(20oC)
100, 140 and 180 µε
(0oC)
0 and 20 30 450x50x50 N≈ 5x104, 2x105 and Gurp, van et al.
1x106 1986(b)
15 10 450x50x50 N≈ 5x103 and 1x106 Alphen, van et al.
1985
0 and 20 30 Wattimena (1991)
Ven, van de (1991)
*intial intention 40 Hz, later on 30Hz because of set-up limitations

Table 3.8: Overview of FPBFT conditions used in several projects


In this case, the tests were performed at 20oC and 10 Hz, using a full sine signal without
rest periods. Both the temperature and the signal were selected because they are often
used, the frequency is the selcted because it is the maximum frequency that can be
applied using the pneumatic four-point bending fatigue set-up currently used at the
Road & Railway Research Laboratory. The strain levels that were used are shown in
Table 3.9, along with the results. The strain levels were selected such that the number of
load repetitions until failure (defined as the load cycle in which the stiffness was
reduced to 50% of the original value) would lie between 104 and 106.

strain [m/m] N [#cycles] f [Hz] T [oC] S0 [N/mm 2] code


6.50E-04 25900 10 20 3556 3-6
5.50E-04 32400 10 20 3493 4-6
5.50E-04 54760 10 20 3268 3-5
4.50E-04 79920 10 20 3543 4-3
4.50E-04 135430 10 20 3582 3-4
3.50E-04 320890 10 20 3216 4-5
3.50E-04 284810 10 20 3530 3-3
2.50E-04 822780 10 20 3543 4-4
Table 3.9: Overview of the four-point bending test conditions and results

38 Chapter 3
Figure 3.10: Close-up of the pneumatic four-point bending set-up used at the Road &
Railway Research Laboratory
The four-point bending set-up (Figure 3.10) and the tests carried out on the ACRe
mixture are discussed in more detail in Erkens and Kalf (2000). The test results are
shown in Figure 3.11 it can be seen that the ACRe material exhibits a rather good
resistance against fatigue.

Gurp, van (1986), s15, 15'C&30Hz Eerden,van der (1984), DAB, 20'C&30Hz
ACRe Alphen, van (1985), 15'C&10Hz
Alphen, van (1985), 15'C&10Hz shifted to 20'C SPDM (1978), F1, Smix(ACRe)
SPDM (1978), F2, Smix(ACRe) RWS (2000), Smix(ACRe)
Gurp, van (1986), s15, 25'C&30Hz Gurp, van (1986), RW352, 15'C&30Hz
Gurp, van (1986), S3, 25'C&30Hz Gurp,van (1986), RW352, 25'C&30Hz
Gurp,van (1986), S3, 15'C&30Hz
1.E+07

1.E+06

1.E+05

N [#cycles]
1.E+04

1.E+03
1.00E-04 Tensile Strain 1.00E-03

Figure 3.11: Four-point bending fatigue results for different mixtures and conditions
The slope of the ACRe fatigue line is similar to those of the other data sets, but its
position indicates that the material can sustain a larger strain at the same number of load
repetitions compared to other mixtures. The data plotted in this graph are taken from
literature (Eerden 1984, Alphen and Molenaar 1985, Gurp 1986 and Jenkins 2000) and
the corresponding test conditions are mentioned in the legend.

The ACRe mixture 39


Additional information on the meaning of the fatigue results was obtained from
standards, using both the Shell Pavement Design Manual (SPDM 1978) and the Dutch
design guide (Rijkswaterstaat, Dienst Weg- en Waterbouwkunde (DWW) 2000). Using
the average initial stiffness found from the fatigue tests (Table 3.9), the corresponding
F1 and F2 characteristics were determined using charts M3 and M4 from the SPDM. In a
similar way the general fatigue chart for asphalt concrete from DWW was used. This
line is also shown Figure 3.11 and the numerical values for this line as well as the
SPDM lines are presented in Table 3.10.

The data obtained by van Alphen are, when the stiffness is corrected to account for the
difference in test temperature, rather similar. The mixture used in these (1985) tests was
a dense asphalt concrete (DAC) 0/16. The mixture used by van der Eerden was also a
DAC, but details about the composition are not known. The data from van Gurp were
obtained from tests on mixtures used for the top layers of pavements. It can therefore be
concluded that the ACRe mix shows fatigue characteristics similar to those kinds of
mixtures, with a better overall fatigue resistance. Comparison of the ACRe response
with the design guides also indicates that the material exhibits a good fatigue resistance.

N Corresponding strain
SPDM F1 SPDM F2 DWW 2000
1x104 5.4x10-4 4.4x10-4 3.2x10-4
1x105 3.6x10-4 2.6x10-4 2.1x10-4
6 -4 -4
1x10 2.4x10 1.7x10 1.3x10-4
1x107 1.6x10-4 1x10-4 8.7x10-5
8 -4 -5
1x10 1.2x10 6.4x10 5.6x10-5
Table 3.10: Fatigue data from design charts for the average ACRe mixture stiffness

3.3.4 COMPARING MONOTONIC AND CYCLIC STIFFNESS VALUES


From the results of the itt presented in Table 3.7 it could be seen that there was a large
difference between the stiffness values obtained from the monotonic and the cyclic
tests. In this case that was mainly due to the definition of the stiffness for the monotonic
tests (Section 3.3.2), but differences between cyclic and monotonic stiffnesses are
commonly observed. In this Section a comparison of monotonic and cyclic stiffness
values obtained from various tests is presented. Although it is difficult to compare the
results from various road engineering tests, because of the differences in internal stress
distributions, control parameters, temperature, frequencies, loading signals and
specimen geometries, a method to obtain some indication of the relative stiffnesses was
established by van der Poel (1954). Starting from the stress-strain relation for a purely
viscous material (Equation 3.15) he showed that the stiffness in a static creep test and a
force controlled cyclic sine load could be compared if for the one it was plotted against
the loading duration and for the other against the reciprocal angular frequency.

∂ε 1
= σ (3.15)
∂t 3η

For a static creep test the stress is constant in time, for a cyclic test it is a sine (in his
case):
creep test : σ (t ) = σ 0 (3.16)

40 Chapter 3
cyclic test : σ (t ) = σ sin(ω t ) (3.17)

Substituting Equations 3.16 and 3.17 in 3.15 and integrating over time yields:
σ0 σ
creep test : ε (t ) = ∫ dt = 0 t (3.18)
3η 3η
σ sin(ω t ) σ
cyclic test : ε (t ) = ∫ dt = − cos(ω t ) ⇔
3η 3ηω
σ π σ π
ε (t ) = − sin( − ω t ) = sin(ω t − ) (3.19)
3ηω 2 3ηω 2

Defining the stiffness as the ration of stress over strain at any value of time, t, yields:
σ σ 3η
creep test : S (t ) = 0 = 0 = (3.20)
ε (t ) σ 0 t t

σ sin(ω t ) σ sin(ω t ) sin(ω t )
cyclic test : S (t ) = = = 3ηω
ε (t ) σ π π
sin(ω t − ) sin(ω t − ) (3.21)
3ηω 2 2

The sinusoidal part expresses the delay (phase lag) between load and deformation that is
typical for a viscous material, for a purely viscous material this phase lag is π/2. The
stiffness in a cyclic test on such a material is usually determined by dividing the
amplitudes of the stress and strain signals, even though the maxima do not occur on the
same time. As a result, the time dependent stiffness in Equation 3.21 reduces to a
constant value, which depends only on the angular frequency:
σ σ
cyclic test : S (ω ) = = = 3ηω (3.22)
ε σ
3ηω
From these results, van der Poel concluded that both stiffness values could be compared
if he plotted them against load duration and the reciprocal angular frequency,
respectively. Extending this approach to monotonic loading yields:
∂ε $
monotonic disp. control : ε (t ) = ε$t ⇒ =ε
∂t
σ
ε$ = ⇔ σ = 3ηε$
3η (3.23)
σ 3ηεˆ 3η
S (t ) = = =
ε εˆ t t

Now t is the time over which the constant strain rate was imposed on the material. Since
asphalt concrete is not purely viscous, the actual stiffness exhibited by the material will
differ from the above derived relations. However, it seems a reasonable approximation
that the time-effects are due to the viscous material component. The apparent mixture
stiffness is then at least a combination of elastic and viscous components:
1
Smix = (3.24)
1 1
+
S E SV

The ACRe mixture 41


Where SE is the elastic stiffness and SV the viscous stiffness. For t=0 the viscous
stiffness goes to infinity, which means the mix stiffness is equal to the elastic stiffness.
If the viscous stiffness is also considered to take into account the temperature effects,
the elastic stiffness can be a “true” Young’s modulus, independent of strain rate and
temperature, the equivalent of the glass modulus for binders. The stiffness obtained
from tests is an apparent stiffness, which combines all the influences. In Figure 3.13
these apparent stiffness values obtained from various tests on the ACRe mixture are
plotted in one graph. Before discussing the results, the different tests and the ways in
which stiffnesses were determined from the results will be presented.

The graph contains results from cyclic direct tension tests, monotonic direct tension
tests, cyclic indirect tension tests and monotonic compression tests. The results from the
monotonic indirect tension tests (Table 3.7) are not incorporated, because they are
secant values (Figure 3.5) and cannot be compared with the other results.

Most of the cyclic tension tests, which were carried out to verify the strain gauges used
in the uniaxial tension tests (Chapter 5) were block signals. These signals can be
decomposed into three stages: loading, constant force (or displacement) level and
unloading. The loading phase was not taken into account, but the unloading phase was
considered elastic and the constant load phase was treated as a static creep test. This
resulted in a single “elastic” stiffness value and a time dependent creep stiffness. For
cyclic tests with a sinusoidal load the approach suggested by van der Poel is used, which
yields a single modulus that corresponds to the angular frequency that is used. For the
monotonic compression tests, the linear part of the response is used to determine an
“elastic” stiffness for that temperature and strain rate combination. From the tension
tests the unloading branch measured by the strain gauges is used to establish a stiffness
value. This value is not directly related to the test strain rate, because the unloading is
not controlled by the test strain rate. There is, however a time variable involved since
this unloading does not occur instantaneously. The time it takes can be determined from
the number of samples in (that part of) the unloading branch in combination with the
scan rate used for data acquisition. These procedures, used to determine the stiffness
from different tests, are illustrated in Figure 3.12.

Cyclic test Monotonic compression Monotonic tension


F S(0)=∆F/A l/∆u t = #samples/scan rate
S(t)=F/A l/u(t) σ
F σ
∆F

time S
t= u/(du/dt) S
ε
u(t) ∆u ε
u
Figure 3.12: Graphic representation of the stiffnesses obtained from different tests
The actual stiffness values obtained from the various tests are plotted in Figure 3.13.
From this graph it can be seen that the variation is considerable, but there is an obvious
trend in the data, showing the effect of temperature as well as the fact that the
differences between stiffness values from different test can largely be explained by their

42 Chapter 3
different loading times. The apparent stiffnesses obtained from the various tests are
grouped quite nicely per temperature. There is some difference between the values
obtained from compression test on the one hand and those from the tension tests on the
other, where tension yields higher stiffnesses than compression. In Chapter 5 it will be
shown that this effect can be explained by the influence of the boundary effects due to
the end caps in the tension test. These result in a three-dimensional tension state of
stress, which causes an over-estimation of the stiffness. The remaining variation in
stiffness values is mainly due to the fact that it is a derived quantity: on the basis of
deformation and force measurements strains and stresses are computed and these are
used to determine the stiffness. The reliability of the sensors, the accuracy of the
determination of the geometry, the scan rate of the data-acquisition equipment and the
assumptions used in computing the stiffness all influence the end result.

block force. contr. 1Hz 0'C, Erkens monotonic compression, 30'C, disp. Cont.
sine force contr. 1Hz 0'C, Erkens monotonic tension 15'C, Erkens
block, force contr. 15'C, Erkens monotonic tension 30'C, Erkens
block, force contr. 30'C, Erkens itt 10'C Medani
monotonic comp., 0'C, disp. Cont. itt 15'C Medani
monotonic comp., 15'C, disp. Cont. itt 20'C Medani
10000
2
Smix [N/mm ]

t or 1/ω [s]
1000
0 0.5 1 1.5 2 2.5 3 3.5

Figure 3.13: Stiffnesses obtained from different tests compared

3.4 CONCLUDING REMARKS


The mixture used in the ACRe project is not a standard pavement mixture. In order to
get an indication of the type of mixture with respect to road engineering properties both
its composition and its response to some road engineering tests are investigated and
compared to well-know mixtures.

From the mixture composition, it appears that the ACRe mixture is somewhere between
a SA and a DAC 0/8. The results from the Marshall test imply the same, with a closer
similarity between the ACRe mixture and the specifications for DAC 0/8. The results
from the monotonic and cyclic indirect tension tests are, due to the overall variation in
test results, difficult to use for such comparisons, but it appears that the stiffness
characteristics as well as the strength values found are rather similar to those for Dense
Asphalt Concrete mixtures. Finally, the fatigue behaviour was studied by means of four-
point bending tests. The results from these tests indicate that the material response is

The ACRe mixture 43


similar to that of DAC mixtures with respect to the strain level sensitivity (slope) while
the overall fatigue resistance is better. This is probably the result of the relatively high
binder content of the ACRe mixture compared to standard DAC mixtures.

Although the material does not exactly fit the specifications for dense asphalt concrete
mixtures, its composition and response to the tests indicated that it is a DAC-like, but
down-scaled and bitumen-rich mixture.

3.5 REFERENCES:
Alphen, C.A.M. and Molenaar, A.A.A., (1985), Vermoeiingsonderzoek middels het
beproeven van een asfaltmengsel met de vierpuntsbuig- en splijtproef, Delft
University of Technology, Report number 7-85-113-8
Arif, S.H., (1999), Simple tests for the Evaluation of Asphalt Mix Characteristics and
their Effect on the Pavement Behaviour, MSc. thesis Delft University of
Technology
C.R.O.W., (1990), RAW Standard Conditions of Contract for Works of Civil
Engineering Construction, VanderLoeff/Drukker B.V., Enschede,
ISBN 90-6628-102-2
Desai, C.S., Somasundaram, S. and Frantziskonis, G., (1986), a Hierarchical approach
for Constitutive Modelling of Geologic Materials, International Journal of
Numerical and Analytical Methods in Geomechanics, Vol. 10, Issue 3, pg. 225-
257
Eerden, M.A.C. van der, (1984), Vermoeiingsonderzoek aan een Dicht Asfalt Beton ten
behoeve van het Project RW15, Delft University of Technology, Report number
12-84-404-2
El-Odaisy, M.H., (1999), the Effect of Sealoflex® on the Mechanical Behavior of an
Asphalt Mixture, MSc. thesis International Institute for Infrastructural, Hydraulical
and Environmental Engineering in Delft, TR 082
Erkens, S.M.J.G. and Kalf, J.A.M., (2000), Four-point Bending Fatigue Tests - Asphalt
Concrete Response (ACRe), Delft University of Technology, Report number 7-00-
117-6
Groenendijk, J., (1998), Accelerated Testing and Surface Cracking of Asphaltic
Concrete Pavements, ISBN 90-804590-1-1
Gurp, C.A.P.M. van, Bientjes, J.W. and Elzenaar, J., (1985), Vermoeiingsonderzoek op
Grind Asfalt Beton, Delft University of Technology, Report number 7-85-408-2
Gurp, C.A.P.M. van, (1986), Vermoeiingsonderzoek Deklaagmengsels, Delft University
of Technology, Report number 7-86-401-3
Gurp, C.A.P.M. van, (1986b), Vermoeiingsonderzoek Zuid-Nederlandse Asfalt
Centrale, Delft University of Technology, Report number 7-86-406-1
Hadley, W., Hudson, W. and Kennedy, T.W., (1970), A Method of Estimating Tensile
Properties of Materials Tested in Indirect Tension, Center for Highway Research,
The University of Texas at Austin, 98-7
Head, K., (1982), Manual of Soil Laboratory Testing, Vol. 2: Permeability, Shear
Strength and Compressibility Tests, Pentech Press Limited, Estover Road,
Plymouth, Devon, ISBN 0-7273-1305-3
Herzberg, R.W., (1996), Deformation and Fracture Mechanics of Engineering
Materials, 4th edition, John Wiley & Sons Inc., New York, ISBN 0-471-01214-9
Jacobs, M.M.J., (1995), Crack Growth in Asphaltic Mixes, Delft University Press, ISBN
90-9007965-3

44 Chapter 3
Jenkins, K.J., (2000), Mix Design Considerations for Cold and Half-Warm Bituminous
Mixes with Emphasis on Foamed Bitumen, University of Stellenbosch, South
Africa
Lytton, R.L., Roque, R., Hiltunen, D. and Stoffels, S.M., (1993), Asphalt Concrete
Pavement Distress Prediction: Laboratory Testing, Analysis, Calibration and
Validation, Strategic Highway Research Program, National Research Council,
SHRP A005
Medani, T., (1999), a Simplified Procedure for Estimation of the Fatigue and Crack
Growth Characteristics of Asphalt Mixes, MSc. thesis International Institute for
Infrastructural, Hydraulical and Environmental Engineering in Delft, TRE 081
Mier, J.G.M. van, Vervuurt, A. and Schlangen, E., (1994), Boundary and Size Effects in
Uniaxial Tensile Tests: A Numerical and Experimental Study, in the proceeding of
Fracture and Damage in Quasibrittle Structures, Eds. Bazant et al, ISBN 0-419-
19280-8
Mier, J.G.M. van, (1997), Fracture Processes of Concrete, CRC Press, Boca Raton
New York London Tokyo, ISBN 0-8493-9123-7
Molenaar, A.A.A., (1983), Structural Performance and Design of Flexible Road
Constructions and Asphalt Concrete Overlays, PhD. thesis Delft University of
Technology
Poel, C.van der, (1954), Representation of Rheological Properties of Bitumens over a
Wide Range of Temperatures and Loading Rates, Paper presented at the 2nd
International Congress on Rheology, Oxford 26-31 July 1953, London, United
Kingdom
Rijkswaterstaat, Dienst Weg- en Waterbouwkunde, (2000), Handleiding Wegenbouw -
Ontwerp Verhardingen
Shell International Petroleum Company, (1978), Shell Pavement Design Manual, Shell
International Petroleum Company Ltd., London
Tada, H., Paris, P.C. and Irwin, G.R., (1985), The Stress Analysis of Cracks Handbook, ,
3rd Edition, ASME International, New York, ISBN 1-86058-303-0
Transportation Research Board, (1997), Design and Evaluation of Large-Stone Asphalt
Mixes, National Cooperative Highway Research Program (NCHRP), Texas
Transportation Institute & Texas A&M University System, Report number 386,
National Academy Press, Washington D.C.
Ven, M.F.C. van de, (1991), Vermoeiingsonderzoek G.A.B. van Meva II (Lintrack),
report 91489, Netherlands Pavement Consultants (NPC)
Verhoeven, J.H.M. and Gurp, C.A.P.M. van, (1993), Stijfheidsonderzoek op
asfaltkernen van plattelandswegen, Delft University of Technology, Report
number 7-93-406-12
Wattimena, J.S., (1991), Vermoeiingsonderzoek G.A.B. mei 1991 (Lintrack), report
91482, Netherlands Pavement Consultants (NPC)
White, T.D., (1985), Marshall Procedures for Design and Quality Control of Asphalt
Mixtures, Paper presented at the Association of Asphalt Pavement Technologists,
Feb. 11-13 1985, San Antonio, Texas, USA Vol. 54 pg. 265-284

The ACRe mixture 45


46 Chapter 3
4. THE UNIAXIAL COMPRESSION TEST
“…the friction on the contact surfaces must have a great effect on the stress
distributions, so that the results are not those of a true compression test.”
-Otto Mohr (1835-1918)-

4.1 INTRODUCTION
The tests performed in this project are used to determine material properties. The idea
behind material testing is simple: subject the material to a state of stress and measure
the response. By doing this for several states of stress, the relation between stress and
strain can be generalised into a model that can predict the response to any state of stress.
To obtain an accurate model, it is necessary that the state of stress in the specimen is the
intended state of stress and, preferably, uniform and that the response is measured
accurately. Since the stress in the specimen cannot be measured directly, it is important
to rule out any possible disturbance in the intended state of stress and to establish
indirectly that that state of stress is actually realised. As stated in the previous chapters,
standard road engineering tests are structural tests rather than material tests and
therefore not suited for material characterisation. In many cases triaxial cells are used
for material testing. That kind of equipment was available in the Road & Railway
Research Laboratory, but the cells were meant for either tests on soils or for cyclic tests
on asphalt concrete. As a result, they could only be used for a limited range of stress
levels. In combination with the information from practice that damage in pavement
structures is predominately caused by tension-compression states of stress, this led to
the decision to develop tests other than (expensive) high capacity triaxial equipment.
One of the set-ups that was developed as a result of this choice was a uniaxial
compression set-up. Such a test is basically a triaxial test with zero confinement, thus
the experience obtained with those tests in road engineering as well as other fields of
mechanics of materials (soil, concrete and rock mechanics) could be used as guide lines
in developing the set-up. However, the differences between these materials and asphalt
concrete, mainly the temperature and strain rate sensitivity of the latter, required new
solutions for known problems. In the first part of this chapter the requirements for the
set-up and the ways in which these requirements were met are discussed. In the last part
the test conditions and results are presented.

4.2 DEVELOPING THE TEST SET-UP


The response of a specimen during a test is determined by the combination of its
material characteristics and the state of stress to which it is subjected. If the imposed
state of stress is known, information on the material characteristics can be obtained from
the observed response. If, however, the actual state of stress differs from the intended
state of stress, the derived material characteristics are incorrect.

When performing an experiment, influences like alignment and boundary conditions are
not always considered in detail. Usually, the state of stress in a compression test is
considered to be the applied load divided by the specimen cross-section. However, the
true state of stress can be completely different. Examples of disturbing influences on the

The uniaxial compression test 47


state of stress are: skewed specimens, alignment problems, boundary friction and
temperature gradients. Some examples of disturbing influences and their effects are
shown in Figure 4.1.

u [mm]
externally applied

on specimen

A B t [s]
C D

T1

T1 > T2

T2

Figure 4.1: Examples of the effect of (A) alignment, (B) deformation of the frame,
(C) temperature effects and (D) boundary friction
In order to obtain reliable test results, these phenomena have to be prevented. In the
next sections the problems encountered during the set-up development are discussed.

4.2.1 FRAME DEFORMATIONS AND MOVEMENTS


Alignment and stability problems cause additional, unwanted, forces on the specimen.
Alignment problems can originate both from the set-up and from the specimen, while
stability problems are solely related to the set-up. Ensuring that the set-up is aligned
correctly and is sufficiently stable and rigid to prevent extensive movement or
deformation during a test can prevent these problems. In case of the set-up developed
for this project, an existing set-up had to be adapted. To prevent stability problems the
frame was extended to a space frame, providing a closed frame (Figure 4.2A) for the
load transfer. Alignment was ensured by carefully aligning all parts of the set-up in the
unloaded stage. To ensure that everything remained aligned during a test, guidance bars
between the loading plates were used. These bars were anchored in the bottom plate at
120o intervals and passed through linear bearings in the top plate (Figure 4.2B).

48 Chapter 4
ydraulic Actuator

Guidance Loadcell
bars

Specimen

Temperature
cabinet

Frame

Concrete block
(A) (B)
Figure 4.2: (A) A space frame was used to ensure a stable set-up (B) guidance bars
between the loading plates ensured alignment during the test
Another frame-related problem is the deformation in the loading frame during a test.
Since the loads are transferred to the specimen via the frame, there will always be frame
deformations. This can lead to two problems, the first is related to the deformations
measured by the displacement transducer of the actuator. This is the combination of the
frame deformation and the actual specimen deformation. As long as the frame
deformation is considerably smaller than the specimen deformation, this is no problem.
However, it must be certain that this is the case. The second problem is related to the
energy stored in the deformed frame. As soon as the specimen is degrading, the load is
diminishing, which will lead to rapid unloading of the frame. This will interfere with the
applied signal unless the frame deformations are very small. As a result, it is not the
strength but the stiffness of a loading frame that determines its suitability for a particular
application. In the compression set-up that was developed this potential problem was
handled by using a stiff frame (four columns, all inter-connected) in combination with
the use of the average signal of three displacement transducers, placed at 120o intervals
round the specimen and between the loading plates, as the control signal. This way, the
applied and measured signals are related directly to the specimen deformation,
excluding any frame deformations that may occur.

4.2.2 SET-UP CAPACITY


It is generally considered good experimental practice to ensure that the capacity of all
components of a set-up is sufficient for the series of tests performed on it. The strength
and stiffness of asphalt concrete, however, vary considerably with temperature and
strain rate. Yet it is not practical and with respect to variability in the results not
desirable to use different set-ups. Therefore, the frame was designed on the basis of the
extreme conditions. These conditions, a low temperature in combination with a high
strain rate, lead to brittle behaviour of asphalt concrete. To allow the intended high
strain rates to be applied from the beginning of the test a high-response hydraulic valve

The uniaxial compression test 49


was used in the set-up. The set-up could be used for the more moderate test conditions
as well, without compromising the accuracy. However, the strength of the material also
varies considerably, values between 2 N/mm2 and 55 N/mm2 were measured during the
tests described in this chapter. For the specimen dimensions used, this meant forces
between 4 and 110 kN, completely different ranges that would require different
actuators. Since using several actuators for one set-up in a single series of tests would be
quite time consuming and expensive, an actuator capable to cover the whole range was
used. To prevent loss of accuracy for tests performed at low load levels, a load cell and
test controller were used that could be set to and calibrated for several load levels (20,
50, 100 and 150 kN).

4.2.3 INFLUENCE OF BOUNDARY FRICTION

4.2.3.1 Effect of friction on the test results


Friction is a boundary condition that can have a significant disturbing effect on the state
of stress in a specimen, since it restrains the deformations. In a compression test,
friction depends on the contact between specimen and loading plates. During the test,
the specimen deforms in radial as well as in axial direction. At the contact surface
between specimen and the loading plates, the radial deformation is restrained due to the
fact that the plates and the specimen have a different ν/E ratio. The resulting friction
acts as a confinement for the top and bottom of the specimen, causing the well-known
barrel-shape of specimens in compression, as shown in (Figure 4.3).

Figure 4.3: Friction between plates and specimen causes barrelling in compression
tests, an indication of non-uniaxial states of stress
The influence of the loading plates on the response has been recognised for a long time.
As early as 1882 Otto Mohr (1835-1918) criticised a series of compression tests
performed by Bauschinger because he used cubical specimens. Mohr states that the
friction on the contact surfaces must have a great effect on the stress distributions, so
that the results are not those of a true compression test (Timoshenko et al. 1951). This

50 Chapter 4
problem is encountered wherever compression tests are used and it has to be addressed
in a way that is suitable for the material under investigation (Mier 1997, Vonk et al.
1989 and Head 1982). An example of the influence of the type of plate on the response
of cement concrete is shown in Figure 4.4. In this graph “dry platen” means a specimen
tested between steel loading plates without any friction reduction system, the other lines
show the effect of various friction reduction systems. For an explanation of these
systems, the reader is referred to Erkens (1998), van Mier (1997) or Vonk et al. (1989).
From this figure it is obvious that the confinement caused by the friction leads to an
over-estimation of the compressive strength and that in order to accurately measure the
response to a uniaxial state of stress it is necessary to minimise the influence of the
contact stresses. For specimens with a height to diameter ratio of 2:1 the middle third
part of the specimen is usually considered to be in a state of uniaxial compression.
Specimens with h/D ≥ 2 are therefore reputed to result in reliable strength values. For
this reason the h/D ration for the specimens was chosen equal to 2.

Figure 4.4: Influence of boundary friction on the response in a uniaxial compression


test on concrete (Mier, van 1997) ………………………………………….
However, despite this h/D ratio, considerable bulging occurred in the tests (Figure 4.3,
right). Besides the doubt this generated about the reliability of the strength values, it
also led to instrumentation problems. The radial deformations were measured by two
separate systems, a string and a circumferential kit. The first system consisted of a string
wrapped around the specimen and passing over two potentiometers. This system has a
very large range, but it must overcome some friction at the onset of the tests, causing it
to miss the first part of the deformations. The second system is a chain of very smooth
rollers in combination with an extensometer, this system is very sensitive but has a
limited range. In order to combine the data from both systems to an overall radial
deformation signal the systems must measure the same behaviour. Because they cannot
physically be at the same height, this requires uniform deformation.

4.2.3.2 Developing an effective friction reduction system


For the reasons mentioned in the previous section, it was decided to try to obtain more
uniform specimen deformations by reducing the influence of the friction. To achieve
this it was tried to minimise the friction itself. The existing anti-friction measures in
asphalt concrete research were related to (static or repeated-load) creep tests. The most
common approach is the use of a glycerine talc mixture with graphite powder (Geffen
1983 and C.R.O.W. 1990).

The uniaxial compression test 51


In uniaxial compression tests, no friction reduction was observed using this mixture. As
a matter of fact, at higher temperatures the mixture even worked as a glue. This effect
can also be observed in dynamic triaxial tests (Poot et al. 1995). An important
difference between these tests and classical creep tests is that the latter take place in
water. Since glycerine rejects water, this may result in a thin film of water between the
plates and the glycerine treated surfaces. This film would then account for the friction
reduction. Since this approach is not effective for the uniaxial compression test, another
solution had to be found.

A series of friction reduction systems for creep tests is mentioned in Bolk (1980), where
the effect of the reduction was expressed as the ratio between the average radial strain at
the contact surfaces ((εtop+ εbottom)/2) and the radial strain at the centre of the specimen
(εcentre). This is shown in Table 4.1. Although this table shows clearly that the
deformations of the specimen increase considerably when friction is reduced, it does not
provide information on the impact of friction on the measured strength. Furthermore,
the data is related to creep tests, which are quite different from monotonic compression
tests.

Friction reduction system (εtop+ εbottom)/2 : εcentre


dry platen 0.39
Teflon 0.67
Grease + graphite powder 0.70
Plastic brushes 0.75
Two rubber membranes with grease in between 1.05
Table 4.1:The effect of friction reduction systems on the uniformity of the radial
deformations in asphalt concrete in a creep test according to Bolk
From Figure 4.4 the effect of several boundary friction reduction systems on the
response of concrete can be seen. In concrete research, very good results were obtained
with a Teflon foil sandwich system: two layers of Teflon foil (100 µm each) with 50 µm
bearing grease in between (Mier, van 1997). In this system the two Teflon layers slide
over each other with very little friction. During the test, the top layer follows the
deformation of the specimen. A similar approach was tried for asphalt concrete, but
since the ultimate radial strain in asphalt concrete is much larger than in concrete (30-
50% vs. 2-5%), the asphalt is still confined too much. This agrees with the observation
in Bolk (1980), where Teflon turned out to be less effective than glycerine with graphite
powder in reducing the friction in creep tests on asphalt concrete.

In soil mechanics testing, rubber is often used as an intermediate layer between


specimen and plates (Head, 1982). Rubber has a higher ratio of ν/E than asphalt
concrete. Therefore, this will lead to a situation in which the rubber layer wants to
deform more than the specimen. This will cause tensile stresses in the contact plane,
which would lead to tension in the top and bottom of the specimen. The situation would
be the reverse of that shown in Figure 4.3A. This leads to increased deformation at the
contact surfaces, which results in cracks that initiate at these surfaces. This is further
corroborated by the results from creep test on asphalt concrete with rubber interlayers
(Table 4.1), which resulted in strains at the top and bottom that were larger than the
strain at the centre. That this also occurs in monotonic compression tests is shown in
Figure 4.5. Besides this surface-induced failure, rubber inter-layers have another

52 Chapter 4
disadvantage: although rubber can be deformed easily, it is an elastic material. Thus, it
wants to return to its original shape, which leads to confinement in the end.

Figure 4.5: Rubber layers reverse the friction effect: the deformations at the contact
surfaces are larger than at the centre…………………………
In the above text, nearly all the friction reduction systems shown in Table 4.1 are
discussed. The exception is the (plastic) brush plate. These plates are not rigid, but
consist of rods that are clamped at the plate side and free at the specimen side. During a
test, these rods will bend to facilitate the radial deformations of the specimen. This
system was not tested for two reasons. First of all these plates must be rather high to
obtain the desired effect, which made it difficult to integrate them in the set-up.
Secondly a major side effect is that it cracking at the contact surfaces between the rods,
due to the increasing resistance to radial deformations throughout the test.

From the evaluation of a series of friction reduction systems described above, it became
clear that for friction reduction in tests on asphalt concrete an interlayer with a low yield
strength, a high ultimate strain and a high resistance to rupture was needed. A material
with the required characteristics is a foil of Luflexen© from Basf (Table 4.2). It was
used in combination with a thin layer of soap, both between plate and foil and between
foil and specimen.

Density 0.903 G/cm3 ISO 1183


Fuse index (190oC, 2.16kg) 1.4 G/10min ISO 1133
G 50 MPa ISO 6721-2
E 75 MPa ISO 527
Yield stress 7 MPa ISO 527
Rupture resistance 32 MPa ISO 527
Ultimate strain 650 % ISO 527
Table 4.2: Material characteristics of Luflexen©
The Luflexen© foil that was used is 50 µm thick. As a result of this layer, the specimen
deformation is more uniform. The soap used is a soft soap with glycerine and comes

The uniaxial compression test 53


from the soapfactory Vandeputte in Moeskroen (Belgium). Using this Luflexen©-soap
system the contact surfaces could deform freely until well after peak stress, allowing the
specimen to remain cylindrical in shape (Figure 4.6). Once the specimen was well into
its softening regime, the radial deformation at the top and the bottom stagnated. Since
the foil had not reached its ultimate strain at that time, it is thought that this deformation
arrest was due to the reduced strength of the specimen, which no longer sufficed to
stretch the foil.

Figure 4.6: The Luflexen© and soap system resulted in uniform specimen
deformations
The axial deformations of the Luflexen© layers are measured by the axial LVDT’s as
part of the specimen deformation. This means that the specimen deformation is
overestimated. Since the interlayer follows the radial deformation of the asphalt
surfaces, assuming a constant volume (which is equal to assuming that the foil remains
intact), this can be used to approximate its reductions in thickness during the test. This
axial deformation can be used to correct the deformation signal throughout the test. As
expected, because of the small total thickness of the foil, this correction, which is about
7 µm, does not significantly alter the deformation signal.

The effect of the soap is limited to the beginning of the test where, despite the pre-load
that is applied, the applied deformation between the two plates at the beginning of the
tests consists of a combination of specimen deformation and inter-layer deformation
(soap and other settling effects). As a result the specimen deformation is smaller than
the applied deformation and, consequently, the force and radial deformations are
smaller than would be expected on the basis of the applied deformation. These signals
can be corrected by assuming that the steepest part of the force signal corresponds to the
linear-elastic response of the specimen. By extrapolating this slope to the beginning of
the test, the force signal as it would have been measured without inter-layer deformation
is found. If the same part of the signal is used to determine the slope of the radial
deformation signal, this can also be corrected. This approach is graphically represented
in Figure 4.7.

54 Chapter 4
#scans
0 20 40 60 80 100 120
12.5 0
Extensometer gap [mm] regression
12.6 interval (1) -2

Force [kN]
1
12.7 new orig in -4

12.8 2 -6
regression
12.9 interval (2) -8

13 - 10
Figure 4.7: Correction of the onset of the force (1) and the extensometer signal (2) (at
15 oC & 0.1 mm/s)

4.2.3.3 Assessing the effect of friction reduction


Once an effective friction reduction system was found, the reliability of the strength
values found without friction reduction for h/D=2 was investigated. This is of interest
since, if these values are reliable, application of friction reduction for this ratio is
necessary only when the radial deformations are measured. Furthermore, in road
engineering often low h/D ratio’s (1-0.5) are used, because the layer thickness in actual
structures is limited ( 40 – 60 mm) and cores taken from these structures are limited to
that height while the maximum grain size often requires a 100 mm diameter. Naturally,
this aggregate size versus minimum dimension relations should in principle also be
fulfilled for the specimen height, but this is often not possible. The resulting low h/D
ratios are particularly sensitive to friction effects, which can have a considerable effect
on the observed response. For this reason, the evaluation of the effect of the friction
reduction system was extended to other h/D ratio’s as well, assessing the applicability to
practical tests. Most of the specimen sizes that were used are shown in Figure 4.8 (the
exception is h=37.5 mm), the heights vary, but the diameter is 50 mm for all of them.
One combination of temperature and strain rate was used, to limit the number of tests.
Since information on the effect of both h/D ratio and friction reduction on the response
of cement concrete was available (Mier, van 1997), it was decided to use this response
as an indication of the expected behaviour of asphalt concrete under “brittle” conditions
(low temperatures and high strain rates). To obtain information on the behaviour under
conditions for which the response of asphalt concrete differs considerably from that of
cement concrete (“plastic conditions”, i.e. higher temperatures and strain rates) an
apropriate condition from the general test programme (Section 4.3.4) was selected.

D=50 mm Standard,

h=25 h=50 h=75 h=125


Figure 4.8: Different h/D ratios tested to establish the influence on the test results

The uniaxial compression test 55


The difference between asphalt concrete and cement concrete is caused by the
viscousity of the binder used in asphalt concrete. This characteristic is most pronounced
at higher temperatures, so the highest temperature from the test programme, 30oC was
selected. The displacement rates used in the standard programme are 0.1, 1, 5 and 10
mm/s, yielding strain rates of 0.1, 1, 5 and 10 x10-2/s.

T • D [mm] h [mm] h/D v [mm/s] Friction # specimens


[oC] ε s-1  reduction

30 0.05 50 125 2 ½ 6.25 Y 3


30 0.05 50 100 2 5 Y 3
30 0.05 50 75 1 ½ 3.75 Y 3
30 0.05 50 50 1 2.5 Y 3
30 0.05 50 37.5 ¾ 1.9 Y 3
30 0.05 50 25 ½ 1.25 Y 3
30 0.05 50 125 2 ½ 6.25 N 3
30 0.05 50 100 2 5 N 5
30 0.05 50 75 1 ½ 3.75 N 3
30 0.05 50 50 1 2.5 N 3
30 0.05 50 37.5 ¾ 1.9 N 3
30 0.05 50 25 ½ 1.25 N 3
Total number of tests: 38
Table 4.3: Test conditions used to asses the influence of h/D ratio and friction
Since the response is controlled by the strain rate rather than the deformation rate, these
deformation rates have to be adapted for the different heights. It was decided to use a
strain rate of 0.05 s-1 (=5 mm/s for the h=100 mm specimens), because that would lead
to deformation rates within the extremes of the standard programme (0.1 and 10 mm/s),
thus minimising the chance of encountering unexpected problems. The test conditions
are shown in Table 4.3.

The average stress-strain curves for each condition are shown in Figure 4.9 (with
friction reduction) and Figure 4.10 (without friction reduction). From these curves, it
can be seen that if the friction is reduced, the different h/D ratios result in approximately
the same strength while without friction reduction the measured strength value increases
with decreasing h/D ratio’s.

56 Chapter 4
-12
σ [N/mm2]
-11
-10
-9
-8
-7
-6
-5
h/D=0.5
-4
-3
h/D=0.75
-2 h/D=1
h/D=2
-1 h/D=1.5
h/D=2.5
0
0 -0.05 -0.1 -0.15 -0.2 -0.25 ε [m/m] -0.3

Figure 4.9: The effect of the h/D ratio on the overall stress-strain relation for tests
with friction reduction…………………………………….
-12
σ [N/mm2]
-11
-10
h/D=0.5
-9
-8
-7
-6
-5
h/D=2 h/D=0.75
-4
-3
h/D=1

-2
h/D=1.5
-1 h/D=2.5
0
0 -0.05 -0.1 -0.15 -0.2 -0.25 ε [m/m] -0.3

Figure 4.10: The effect of the h/D ratio on the overall stress-strain relation for tests
without friction reduction………………………………..
Unlike the increase in the observed strength the effect on the softening branch is,
although size related, not caused by friction. From the graphs, it seems that the slope of
the descending branches is less steep for smaller h/D ratios. This phenomenon has also
been observed in concrete and rock mechanics and there it is explained by the fact that
after the peak localisation of damage occurs, which means that the post-peak
deformations take place in a small part of the specimen. As a result, it is doubtful
whether these deformations can still be expressed as overall strains. In other words,

The uniaxial compression test 57


once larger cracks appear in the specimen, the assumption of a continuous material is
violated and as a result the definition for the overall strain is no longer valid. In concrete
mechanics, plots of the normalised stress versus the post-peak deformations are used to
show that the different softening branches are the result of damage localisation (Mier,
van 1997). In these plots the descending branches are more or less the same, indicating
that post-peak deformation leads to a similar reduction in strength for different
specimen sizes. However, the total reduction found for the different sizes still differs. A
similar pattern is seen if the post-peak results from Figure 4.9 and Figure 4.10 are
converted to normalised stress versus post-peak deformation diagrams (Figure 4.11).
Plotted in this way, the descending branch is initially more or less the same for all
specimen sizes and independent of the friction during the test. The level to which the
stress decreases depends on the specimen size, which is again similar to the response
observed for concrete. In a compression test the strength never really reduces to zero
because eventually the remains of the specimen are compressed. The shorter the
specimen, the sooner after reaching the peak load the remains, among which the larger
aggregates (in this case Dmax=5 mm) will be loaded. As a result, the minimum load level
for the shorter specimens is not as low as for the higher ones.
normalised
stress [σ/ft]
1
h/D=2.5 (Y)
0.9 h/D=2 (Y)
h/D=0.5 h/D=1.5 (Y)
0.8 h/D=1 (Y)
h/D=0.75 (Y)
0.7
h/D=0.5 (Y)
0.6 h/D=2.5
h/D=0.75 h/D=2
0.5 h/D=1.5
h/D=1
0.4 h/D=1 h/D=0.75
0.3 h/D=0.5
h/D=1.5 h/D=2.5
0.2 h/D=2
0.1
post-peak deformation [mm]
0
0 -5 -10 -15 -20 -25
Figure 4.11: Normalised stress versus post-peak deformation curves (Y=friction
reduction) …………………………………….
In Figure 4.12 the peak stress (=strength) for both the tests with and without friction
reduction is plotted as a function of the h/D ratio. From this graph it can be seen that the
strength increases with decreasing h/D ratio’s if no friction reduction is utilised. For
ratios smaller than one, the strength values rapidly increase, but even for h/D=1 and 1½
the obtained strengths are barely within the 95% reliability interval found for the tests in
the standard test programme with friction reduction and h/D=2 (this reliability interval
is used to give an indication of the significance of the deviations). If, on the other hand,
an effective friction reduction system is used, the strength values remain reliable, even
for very small h/D ratio’s. The stress-strain relations and the strain at peak strength
change with the h/D ratio for both situations (with and without friction reduction), but
again the effect is less when the friction is reduced.

58 Chapter 4
Because the tests were performed for a single combination of temperature and strain
rate, it is not certain if the effect of the h/D ratio varies with these parameters. However,
the test temperature was relatively high (30oC) so the materials viscous properties will
have an influence on the response. Since the behaviour and strength of asphalt concrete
at lower temperatures and higher rates is similar to that of concrete (Erkens et al. 1998
and Erkens et al. 2000), the test results were compared to those of concrete. To account
for the differences in strength the results were normalised with respect to the “true
strength”, defined as the strength found for h/D=2. The results are shown in Figure 4.13.
-15.0
2 NO REDUCTION
f c [N/mm ]
FRICTION REDUCTION
-13.0 95% reliability interval
fc(h/D=2)
-11.0

-9.0

-7.0

-5.0

-3.0

-1.0

0.00 0.50 1.00 1.50 2.00 2.50 h/D 3.00


Figure 4.12: The effect of the h/D ratio on the compressive strength for tests with and
without friction reduction(T=30oC and strain rate = 0.05 s-1)
3.0
fc,norm= asphalt concrete no reduction
fc/fc(h/D=2)
asphalt concrete friction reduction
2.5 cement concrete no reduction
cement concrete friction reduction

2.0

1.5

1.0

0.5

h/D
0.0
0.00 0.50 1.00 1.50 2.00 2.50 3.00
Figure 4.13: The normalised compressive strengths as a function of the h/D ratio for
asphalt concrete and cement concrete (Mier, van 1997)

The uniaxial compression test 59


Figure 4.13 illustrates that the effect of the h/D ratio is similar for concrete and asphalt
concrete, which leads to the expectation that this trend will be consistent over a wide
range of temperatures and strain rates for asphalt concrete. Furthermore, the effect
appears to be proportional to the “true” strength, leading to similar trends in the
normalised strengths. From these observations it seems that the friction reduction
system is not only useful for research purposes but definitely also for practical
applications, where it will allow reliable strength values to be determined on the basis of
specimens with low h/D ratio’s.

4.2.4 TEMPERATURE EFFECTS


It is well known that asphalt concrete is a temperature sensitive material and therefore
the set-up was equipped with a temperature-controlled cabinet. This cabinet, placed
within the frame, enclosed the loading plates, specimen and instrumentation. During the
preliminary phase it became clear that the specimens were very sensitive to temperature
gradients. Once the friction between specimen and plates was minimised (Section
4.2.3.2), local failure on one side of the specimens (Figure 4.14) was observed.

Figure 4.14: One-sided failure due to a temperature gradient in the set-up


In case of tests at temperatures above the laboratory temperature, failure occurred at the
top, in case of test temperatures below lab temperature failure occurred at the bottom of
the specimen. At that time, a single temperature sensor hanging in the cabinet was used
to control the temperature. On the basis of these observations, the loading plates were
also instrumented with temperature sensors. It turned out that a temperature gradient
developed in the cabinet and that a temperature difference of 0.8 to 1.5 oC between the
loading plates resulted in local failure at the warmer side of the specimen. Based on
these observations the airflow in the set-up was adjusted to ensure a uniform
temperature distribution. During the tests, the temperature of the loading plates as well
as the air temperature were monitored and kept within 0.5oC from each other. Before the
friction reduction system was introduced, this temperature effect was not observed.
Obviously, the friction, which is predominant under those conditions, suppresses it. This
means that in tests without friction reduction such temperature influences may occur
and influence the results without obvious effects on the specimen deformation pattern.

4.3 THE UNIAXIAL COMPRESSION TESTS

4.3.1 SPECIMENS
The specimen production started with the production of ∅150 mm cylinders of
approximately 85 mm height. These cylinders were compacted with the gyratory
compactor. This type of compactor uses a combination of compressive and shear forces,

60 Chapter 4
which results in a kneading compaction that is believed to be rather similar to the roller
compaction used for roads. The actual specimens where drilled from the centre of these
gyratory cores. The material used was the asphalt mixture that was discussed
elaborately in the previous chapter. An overview of the specimens and their
characteristics is shown in Table 4.4.

4.3.2 TEST SET-UP


The compression set-up consists of a 3D space frame that is build on an elastically
supported concrete block. On the lower part of the frame a pedestal is constructed to
which the bottom plate is connected. The bottom and top plate are kept parallel by
means of three massive Fortal (a strong aluminium alloy) bars of 16 mm diameter.
These bars are connected to the bottom plate and pass through linear bearings in the top
plate. The top plate itself is connected rigidly to the actuator. A schematic drawing of
the compression set-up is shown in Figure 4.2A.

An MTS 150 kN hydraulic actuator is mounted in the 3D frame. The force that is
applied with this actuator is measured with a 200 kN Lebow loadcell (model 3174-114)
that can be used in four different ranges (20, 50, 100, 150 kN. The actuator is controlled
by an MTS testcontroller (model 448.82) which allows the selection of the appropriate
force range for the test conditions. The test is performed in displacement-control and a
programmable function generator (MTS microprofiler model 418.91) generates the
control signal. Since some of the tests were performed at rather high deformation rates
and low temperatures, which resulted in high specimen stiffnesses, the system was fitted
with a 19 l/min high response valve (Moog, model D760-232A), which allowed the
actuator to apply the required forces at a high rates.

An insulated cabinet with dimensions 0.6x0.5x0.6 m. is placed within the frame. It is a


sandwich structure of wood and roofmate. The inside of the cabinet is covered with
aluminium/plastic insulation foil. To minimise temperature loss through the 3D-space
frame, a 1 cm. thick PVC plate is placed between the part of the frame inside and that
outside the temperature cabinet. Three PT 100 temperature sensors are used to control
the temperature. Two of them are connected to the bottom and top plate respectively
and the third measures the air temperature. The first two PT 100’s are connected to a
temperature control unit that maintains the temperature with an accuracy of ± 0.5 °C.
Either one of these sensors can be used to control this unit. The unit circulates air in the
temperature cabinet between two input channels and one output channel. One input
channel is close to the upper compression plate and the other is close to the lower
compression plate. With simple valves the ratio in airflow through the two input
channels can be varied. The unit is capable to maintain the temperature in a range of -5
°C till 35°C.

The uniaxial compression test 61


filename code T [oC] ε [%/s] D ± 0.05 [mm] h ± 0.1 [mm] ρ prfstk ± 3 [kg/m 3] V ± 0.5 [%]
com01_1 D66 30 0.1 49.7 100 2304 2.4
com01_2 D209 30 0.1 49.7 100.4 2304 2.4
com01_3 D344 30 0.1 49.9 100.5 2208 6.4
com01_4 D211 30 0.1 49.7 100.25 2303 2.4
com01_5 D345 30 0.1 49.9 100.45 2168 8.1
49.78 100.32 2257 4.4
com01_6 D206 15 0.1 49.7 100.1 2313 2.0
com01_7 D210 15 0.1 49.7 100.3 2281 3.3
com01_8 D221 15 0.1 49.8 100.1 2308 2.2
com01_9 D207 15 0.1 49.7 100.4 2296 2.7
com01_10 D215 15 0.1 49.7 100.2 2251 4.7
49.72 100.22 2290 3.0
com01_11 D63 0 0.1 49.7 99.95 2278 3.5
com01_12 D111 0 0.1 49.76 99.87 2303 2.4
com01_13 D120 0 0.1 49.7 99.6 2305 2.4
com01_14 D117 0 0.1 49.72 100.43 2300 2.5
com01_15 D122 0 0.1 49.67 99.37 2295 2.8
49.71 99.844 2296 2.7
com1_1 D202 30 1 49.7 99.8 2305 2.3
com1_2 D362 30 1 49.8 99.95 2314 1.9
com1_3 D67 30 1 49.75 99.95 2306 2.3
com1_4 D363 30 1 49.7 100.15 2301 2.5
com1_5 D62 30 1 49.75 100.25 2304 2.4
49.74 100.02 2306 2.3
com1_6 D219 15 1 49.7 100.1 2298 2.6
com1_7 D113 15 1 49.65 99.57 2297 2.7
com1_8 D203 15 1 49.7 99.9 2304 2.4
com1_9 D48 15 1 49.5 100.4 2311 2.1
com1_10 D204 15 1 49.7 99.8 2290 3.0
49.65 99.954 2300 2.5
com1_11 D133 0 1 49.72 100.05 2301 2.5
com1_12 D118 0 1 49.68 99.75 2310 2.1
com1_13 D131 0 1 49.7 100.07 2297 2.7
com1_14 D127 0 1 49.75 99.35 2308 2.2
com1_15 D201 0 1 49.7 99.5 2309 2.2
49.71 99.744 2305 2.4

Table 4.4a: Overview of the specimens

62 Chapter 4
o
filename code T [ C] ε [ %/ σ] D ± 0.05 [mm] h ± 0.1 [mm] ρ prfstk ± 3 [kg/m 3] V ± 0.5 [%]
com5_1 D364 30 5 49.8 100.2 2287 3.1
com5_2 D353 30 5 49.85 100.3 2293 2.8
com5_3 D346 30 5 49.8 100.5 2269 3.9
com5_4 D365 30 5 49.9 100.3 2280 3.4
com5_5 D342 30 5 49.8 100.5 2260 4.2
49.83 100.3 2278 3.5
com5_6 D368 15 5 49.8 100.1 2302 2.5
com5_7 D360 15 5 49.8 100.3 2272 3.7
com5_8 D359 15 5 49.8 100.2 2285 3.2
com5_9 D356 15 5 49.85 100.2 2278 3.5
com5_10 D354 15 5 49.8 100.3 2269 3.8
49.81 100.2 2281 3.3
com5_11 D366 0 5 49.8 100.0 2296 2.7
com5_12 D352 0 5 49.8 95.8 2278 3.5
com5_13 D349 0 5 49.9 100.5 2290 3.0
com5_14 D351 0 5 49.9 100.5 2296 2.7
com5_15 D369 0 5 49.8 100.0 2310 2.1
49.84 99.4 2294 2.8
com10_1 D50 30 10 49.55 99.8 2297 2.7
com10_2 D125 30 10 49.65 100.2 2315 2.0
com10_3 D140 30 10 49.75 100.5 2305 2.4
com10_4 D212 30 10 49.75 100.2 2304 2.4
com10_5 D138 30 10 49.7 100.2 2309 2.2
49.68 100.2 2306 2.3
com10_6 D137 15 10 49.73 99.8 2298 2.7
com10_7 D130 15 10 49.75 100.0 2311 2.1
com10_8 D65 15 10 49.75 100.2 2312 2.1
com10_9 D132 15 10 49.65 99.7 2310 2.1
com10_10 D135 15 10 49.75 100.3 2311 2.1
49.73 100.0 2308 2.2
com10_11 D116 0 10 49.85 100.0 2305 2.4
com10_12 D123 0 10 49.68 99.7 2301 2.5
com10_13 D61 0 10 49.75 99.2 2290 3.0
com10_14 D114 0 10 49.65 100.1 2301 2.5
com10_15 D218 0 10 49.7 100.3 2287 3.1
49.73 99.9 2297 2.7
Table 4.5b: Overview of the specimens (continued)

4.3.3 INSTRUMENTATION
Inside the temperature cabinet are the transducers, which measure the radial and axial
deformations of the specimen, as well as the applied force. A close up of the inside of
the compression set-up is shown in Figure 4.2B. The axial deformation of the specimen
is measured by six LVDT’s. Three of them have a range of ± 1 inch (= ± 25.4 mm) and
are able to cover the total axial deformation of 20 to 30 mm. These LVDT’s are used to
control the specimen deformation. This on-specimen control system ensures that the
applied deformation was the actual deformation of the specimen instead of a
combination of specimen and frame deformations. Since inductive transducers with a
large range are usually less accurate (their accuracy is a percentage of their total range),
three other LVDT’s are used to cover the first 1-2 mm of the axial deformation with
additional accuracy. All six LVDT’s are connected to the top plate and rest on the
bottom plate. During the test, the top plate is pushed down. The decrease in distance

The uniaxial compression test 63


between top and bottom plate that results from this movement is measured by the
LVDT’s.

The total radial deformation of a specimen that fails in uniaxial compression is rather
large. At the onset of the test, when the specimen is not yet damaged, the radial
deformations are small. In the linear-elastic stage of response, the radial strain is related
to the axial strain via Poisson’s constant. Once internal cracks start growing, the radial
deformations increase. This is due to dilation, which is volumetric expansion due to the
opening of internal cracks. The increase in specimen diameter at the end of a test can be
up to sixty percent. In order to measure the total change in diameter as well as the onset
of radial deformations, a combination of measurement systems is used here as well. The
first part of the radial deformation is accurately measured with an MTS circumferential
kit, which consist of a roller chain connected to a full bridge of strain gauges mounted
in an extensometer. The extensometer has a range of ± 3.75 mm, which in this case
corresponded to a change in specimen radius of approximately 1.5 mm. The total radial
deformation is measured by a Kevlar string, which is wrapped around the middle of the
specimen and two 10-turn potentiometers. The two signals are combined to get the
overall radial deformation signal.

The applied force is measured by means of the loadcell, which is positioned between
top plate and actuator.

All these devices are connected to a control panel that generates the supply voltage and
delivers the output voltage of the transducers. These output voltages are sent to an
input/output board, which consists of a BNC connection panel with 16 single ended or 8
differential connections. These channels are connected to the data acquisition board, a
DAS-1602 board from Keithley metrabyte, which is placed in a 486 PC (100 MHz). For
the compression test 10 single ended connections are used. The data acquisition board is
controlled through the DOS based data acquisition program ViewDac (Version 2.1).
The data acquisition is triggered by a programmable function generator, which also
generates the constant strain rate signal for the test, shortly before the test started.
During a test, the PC measures the output of the measurement systems in voltages.

4.3.4 TEST CONDITIONS


The tests were performed in the displacement control mode, at constant axial
deformation rates. Since asphalt concrete response is strain rate and temperature
dependent, these parameters were varied within the test program. The test conditions
can therefore be specified by the test temperature and strain rate. Three different
temperatures and four strain rates were used (T=0, 15 and 30 oC and v=0.1, 1, 5 and 10
%/s), which lead to a total of twelve different test conditions. In order to take the
expected variations in the test results into account, five tests were performed for each
condition, which led to a total of sixty tests. An overview of the conditions is shown in
Table 4.6.

64 Chapter 4
Temperature [oC] Loading rate [mm/s] Strain rate [s-1] # specimens
30 0.1 0.001 5
15 0.1 0.001 5
0 0.1 0.001 5
30 1 0.01 5
15 1 0.01 5
0 1 0.01 5
30 5 0.05 5
15 5 0.05 5
0 5 0.05 5
30 10 0.1 5
15 10 0.1 5
0 10 0.1 5
Total 60
Table 4.6: Overview of the test conditions for the unixial compression tests
Placing the specimen in the set-up, applying the friction reduction system and attaching
the instrumentation takes some time. During that time the temperature control unit is
open, causing the specimen to change in temperature. To ensure a uniform temperature
distribution in the specimen, the set-up was closed after finishing the placement
procedures and the temperature was allowed to stabilise for an hour. During this time
the top plate, which was connected to the actuator was lifted of the specimen to prevent
creep. To ensure that this would not lead to an unwanted additional pulse load due to the
actuator hitting the specimen at the beginning of a test the top plate was again
positioned on the specimen and a small pre-load was imposed before starting the test.

4.3.5 OVERVIEW OF THE TEST RESULTS


The overall axial and radial deformations and the forces were transformed into strains
and stresses, from which the compressive strength (fc), the axial strain at peak stress
(εax,max) and the radial strain at peak stress (εrad,max) could be determined. They were also
used to obtain the elasticity parameters, apparent stiffness (Smix) and Poisson’s ratio (ν).
Both Poisson’s ratio and Smix were determined using linear regression on the signals at
the beginning of the test, ν on the axial and radial strain and Smix on the axial strain and
the stress, respectively. The results of each test were presented in stress versus axial and
radial strain curves, such as shown in Figure 4.15.
σ [N/mm2] -25

com01_11
com01_12 -20
com01_13
com01_14
com01_15 -15

-10

-5
0.60 0.50 0.40 0.30 0.20 0.10 0.00 -0.10 -0.20 -0.30
0
ε radial [m/m] ε axial [m/m]
Figure 4.15: Example of test results, presented as stress versus axial and radial
strains (0oC and 0.1 mm/s)

The uniaxial compression test 65


Figure 4.16 to Figure 4.18 show the average signals for each test condition.

σ [N/mm2] -8

-7
10 mm/s
-6
5 mm/s
1 mm/s -5
0.1mm/s
-4

-3

-2

-1
0.6 0.5 0.4 0.3 0.2 0.1 0 -0.1 -0.2 -0.3
0
ε radial [m/m] ε axial [m/m]
Figure 4.16: Average test results at 30oC
σ [N/mm2] -25

10 mm/s
5 mm/s -20
1 mm/s
0.1mm/s -15

-10

-5
0.6 0.5 0.4 0.3 0.2 0.1 0 -0.1 -0.2 -0.3
0
ε radial [m/m] εaxial [m/m]
Figure 4.17: Average test results at 15oC
σ [N/mm2]-60

10 mm/s & 0'C -50


1mm/s & 0'C
0.1mm/s & 0'C -40

-30

-20

-10
0.60 0.50 0.40 0.30 0.20 0.10 0.00 -0.10 -0.20 -0.30
εradial [m/m] 0 εaxial [m/m]
o
Figure 4.18: Average test results at 0 C

66 Chapter 4
Besides the changes in strength and brittleness of response, there is another indication
that the behaviour changes with temperature and loading rate. From Figure 4.19 it can
be seen that the fracture surface differs. For high temperatures and low loading rates the
binder is obviously the weak link and the crack propagates through it, bypassing the
aggregates. This results in a very black fracture surface. For low temperatures and high
loading rates on the other hand, the binder is often at least as strong as the aggregate,
causing the cracks to go through the aggregates. This results in a grey fracture surface.

Figure 4.19: Difference in fracture surface at 30 oC (left) and 0 oC (right)


An overview of the test results is given in Table 4.7.

The uniaxial compression test 67


o
filename code T [ C] rate [%/s] fc [N/mm 2] εax
peak
[10-3m/m] ν E [N/mm 2]
com01_1 D66 30 0.1 -2.0 ± 0.3 -37.7 ± 0.03 0.41 ± 0.012 170 ± 5
com01_2 D209 30 0.1 -1.9 ± 0.3 -40.7 ± 0.03 0.43 ± 0.051 150 ± 25
com01_3 D344 30 0.1 -1.8 ± 0.3 -27.5 ± 0.03 0.37 ± 0.020 150 ± 10
com01_4 D211 30 0.1 -1.8 ± 0.3 -33.5 ± 0.03 0.34 ± 0.025 170 ± 35
com01_5 D345 30 0.1 -1.9 ± 0.3 -25.5 ± 0.02 0.33 ± 0.038 185 ± 10
-1.9 ± 0.1 -33 ± 7 0.37 ± 0.10 165 ± 20
com01_6 D206 15 0.1 -6.1 ± 0.3 -36.1 ± 0.03 0.25 ± 0.002 430 ± 5
com01_7 D210 15 0.1 -5.8 ± 0.3 -36.4 ± 0.03 0.25 ± 0.002 405 ± 5
com01_8 D221 15 0.1 -5.4 ± 0.3 -34.5 ± 0.03 0.29 ± 0.002 350 ± 10
com01_9 D207 15 0.1 -5.7 ± 0.3 -34.7 ± 0.03 0.32 ± 0.002 360 ± 5
com01_10 D215 15 0.1 -5.3 ± 0.3 -33.3 ± 0.03 0.25 ± 0.001 400 ± 5
-5.7 ± 0.5 -35 ± 2 0.27 ± 0.05 390 ± 45
com01_11 D63 0 0.1 -21.5 ± 0.3 -11.7 ± 0.02 0.25 ± 0.002 3650 ± 10
com01_12 D111 0 0.1 -22.2 ± 0.3 -12.8 ± 0.02 0.26 ± 0.002 4035 ± 15
com01_13 D120 0 0.1 -21.1 ± 0.3 -12.4 ± 0.02 0.19 ± 0.009 3640 ± 15
com01_14 D117 0 0.1 -21.9 ± 0.3 -11.5 ± 0.02 0.25 ± 0.008 4325 ± 20
com01_15 D122 0 0.1 -21.5 ± 0.3 -9.5 ± 0.02 0.27 ± 0.004 3980 ± 20
-21.6 ± 1.0 -12 ± 0.4 0.25 ± 0.05 3925 ± 360
com1_1 D202 30 1 -3.5 ± 0.3 -34.1 ± 0.03 0.30 ± 0.046 540 ± 20
com1_2 D362 30 1 -3.2 ± 0.3 -24.9 ± 0.03 0.33 ± 0.042 465 ± 35
com1_3 D67 30 1 -3.4 ± 0.3 -35.6 ± 0.03 0.47 ± 0.045 495 ± 80
com1_4 D363 30 1 -3.0 ± 0.3 -26.0 ± 0.02 0.39 ± 0.038 455 ± 60
com1_5 D62 30 1 -3.8 ± 0.3 -27.0 ± 0.02 0.31 ± 0.023 610 ± 60
-3.4 ± 0.5 -30 ± 6 0.36 ± 0.10 515 ± 80
com1_6 D219 15 1 -11.2 ± 0.3 -25.1 ± 0.02 0.25 ± 0.019 2335 ± 100
com1_7 D113 15 1 -11.3 ± 0.3 -26.3 ± 0.02 0.22 ± 0.005 1130 ± 10
com1_8 D203 15 1 -10.9 ± 0.3 -22.9 ± 0.02 0.20 ± 0.004 1030 ± 15
com1_9 D48 15 1 -12.7 ± 0.3 -24.6 ± 0.02 0.21 ± 0.005 1575 ± 10
com1_10 D204 15 1 -11.6 ± 0.3 -23.3 ± 0.02 0.15 ± 0.004 1020 ± 20
-11.5 ± 1.0 -24 ± 2 0.20 ± 0.05 1185 ± 330
com1_11 D133 0 1 -37.6 ± 0.3 -9.0 ± 0.02 0.25 ± 0.004 6980 ± 25
com1_12 D118 0 1 -37.6 ± 0.3 -7.9 ± 0.02 0.34 ± 0.003 8165 ± 40
com1_13 D131 0 1 -38.4 ± 0.3 -9.1 ± 0.02 0.29 ± 0.003 7560 ± 45
com1_14 D127 0 1 -37.9 ± 0.3 -9.1 ± 0.02 0.25 ± 0.005 7520 ± 35
com1_15 D201 0 1 -37.7 ± 0.3 -8.2 ± 0.02 0.28 ± 0.003 8215 ± 40
-37.8 ± 0.5 -8.7 ± 0.5 0.29 ± 0.05 7690 ± 640
Table 4.7a: Overview of the test results

68 Chapter 4
o
filename code T [ C] rate [%/s] fc [N/mm 2] εax
peak
[10-3m/m] ν E [N/mm 2]
com5_1 D364 30 5 -5.1 ± 0.3 -22.9 ± 0.02 0.36 ± 0.017 1140 ± 50
com5_2 D353 30 5 -5.3 ± 0.3 -31.2 ± 0.03 0.43 ± 0.024 1305 ± 50
com5_3 D346 30 5 -5.8 ± 0.3 -27.7 ± 0.03 0.36 ± 0.017 1280 ± 40
com5_4 D365 30 5 -5.2 ± 0.3 -26.8 ± 0.02 0.44 ± 0.036 1235 ± 40
com5_5 D342 30 5 -4.5 ± 0.3 -37.5 ± 0.03 0.34 ± 0.062 1260 ± 55
-5.2 ± 0.6 -29 ± 6 0.39 ± 0.10 1245 ± 80
com5_6 D368 15 5 -17.1 ± 0.3 -26.5 ± 0.02 0.31 ± 0.018 2135 ± 125
com5_7 D360 15 5 -16.3 ± 0.3 -24.0 ± 0.02 0.31 ± 0.002 1820 ± 25
com5_8 D359 15 5 -16.9 ± 0.3 -20.9 ± 0.02 0.32 ± 0.002 1980 ± 15
com5_9 D356 15 5 -16.7 ± 0.3 -18.4 ± 0.02 0.19 ± 0.003 1700 ± 15
com5_10 D354 15 5 -17.3 ± 0.3 -22.4 ± 0.02 0.22 ± 0.001 1815 ± 10
-16.8 ± 0.5 -22 ± 3 0.27 ± 0.10 1890 ± 215
com5_11 D366 0 5 -52.0 ± 0.3 -8.5 ± 0.02 0.40 ± 0.003 9755 ± 55
com5_12 D352 0 5 -50.8 ± 0.3 -8.9 ± 0.02 0.22 ± 0.004 9175 ± 70
com5_13 D349 0 5 -50.9 ± 0.3 -8.1 ± 0.02 0.23 ± 0.003 10240 ± 125
com5_14 D351 0 5 -44.0 ± 0.3 -8.2 ± 0.02 0.20 ± 0.003 8235 ± 80
com5_15 D369 0 5 -49.4 ± 0.3 -6.4 ± 0.02 0.36 ± 0.005 11095 ± 125
-49.4 ± 4.0 -8.0 ± 1.0 0.28 ± 0.15 9700 ± 1345
com10_1 D50 30 10 -8.6 ± 0.3 -27.3 ± 0.02 0.25 ± 0.010 930 ± 10
com10_2 D125 30 10 -6.8 ± 0.3 -29.2 ± 0.03 0.46 ± 0.006 1065 ± 15
com10_3 D140 30 10 -8.0 ± 0.3 -22.1 ± 0.02 0.43 ± 0.004 1425 ± 15
com10_4 D212 30 10 -6.5 ± 0.3 -28.2 ± 0.03 0.43 ± 0.007 1140 ± 10
com10_5 D138 30 10 -7.5 ± 0.3 -21.5 ± 0.02 0.48 ± 0.010 1310 ± 20
-7.5 ± 1.5 -26 ± 4 0.41 ± 0.15 1175 ± 245
com10_6 D137 15 10 -23.0 ± 0.3 -10.5 ± 0.02 0.47 ± 0.004 3545 ± 25
com10_7 D130 15 10 -21.6 ± 0.3 -12.0 ± 0.02 0.34 ± 0.005 3805 ± 15
com10_8 D65 15 10 -23.0 ± 0.3 -14.6 ± 0.02 0.47 ± 0.009 4655 ± 45
com10_9 D132 15 10 -22.0 ± 0.3 -14.8 ± 0.02 0.33 ± 0.004 3265 ± 15
com10_10 D135 15 10 -21.9 ± 0.3 -12.0 ± 0.02 0.33 ± 0.003 3900 ± 15
-22.3 ± 0.9 -13 ± 1.5 0.39 ± 0.10 3835 ± 650
com10_11 D116 0 10 -54.6 ± 0.3 -8.9 ± 0.02 0.35 ± 0.011 10165 ± 120
com10_12 D123 0 10 -57.0 ± 0.3 -9.0 ± 0.02 0.36 ± 0.009 11180 ± 180
com10_13 D61 0 10 -58.4 ± 0.3 -8.2 ± 0.02 0.25 ± 0.021 10735 ± 125
com10_14 D114 0 10 -57.1 ± 0.3 -8.2 ± 0.02 0.29 ± 0.014 10930 ± 105
com10_15 D218 0 10 -56.6 ± 0.3 -8.6 ± 0.02 0.42 ± 0.021 11175 ± 105
-56.7 ± 2.0 -8.6 ± 0.5 0.33 ± 0.10 10835 ± 525

Table 4.7b: Overview test results (continued)

4.3.6 TEST RESULTS AS FUNCTION OF TEMPERATURE AND LOADING RATE


Eventually, the model (described in Chapter 7) will be used to analyse laboratory tests
and pavement structures. For these applications, it is necessary to have general
expressions for the material characteristics and the model parameters. In this way, the
required input is limited to the type of mixture, or maybe the mixture composition, the
temperature profile and the loading rate. From this input the finite element program
CAPA-3D will determine the temperature and loading rate at each integration point in
the structure under analysis and from that information, using the appropriate values for
the material characteristics and model parameters, the overall response can be
determined. This means that the material characteristics as well as the model parameters
must be expressed as functions of loading rate and temperature and, eventually, also of
mixture composition. Since the model parameters cannot be determined on the basis of
compression tests only, in this chapter only the development of general expressions for
the material characteristics will be described. The model parameter determination and

The uniaxial compression test 69


the development of general expressions for these parameters will be presented in
Chapter 7.

4.3.6.1 Compressive strength


The aim was to develop a relation between the compressive strength (fc [N/mm2]),
temperature (T [oC]) and strain rate ( ε& [1/s]), based on the results. It was considered
important that this relation would be a physically relevant relation rather than a fit
function. Unfortunately, linear regression analyses is almost solely useful for data
fitting, while nearly all physical relations (e.g. growth laws, reaction processes) are non-
linear (i.e: one or more of the regression parameters are not linear with respect to the
dependent variable). Obviously, the relation between fc, ε& and T is also non-linear.
However, non-linear regression requires the choice of a relation prior to the regression
analysis, which is then used to determine the constants in that relation. Since it was not
clear what type of relation was needed, a step-by-step approach was used. At first the
basic shape that the relation would have was determined on the basis of the information
that is available. Basically, the relation must be able to describe the physical trends,
which were considered to be:
• ε& =0: this corresponds to not loading the specimen, leading to fc=Cmin
• ε& →∝: the strength will not increase indefinitely, but reach a limit value which
might vary with temperature: fc=Cmax. Based on the fact that bitumen reaches a glass
modulus for all temperatures it was assumed that this limit value is temperature
independent.
• T→-∝: for extremely low temperatures, asphalt will exhibit glass-like, linear-elastic
behaviour until sudden, brittle fracture occurs. This ultimate compressive strength
will be independent of the loading rate: fc=Cmax,
• T →∝: for extremely high temperatures (approximately 160oC) the bitumen will
become a fluid, which has no uniaxial compressive strength: fc=Cmin
This resulted in the expectation that an S-shaped (or transition) relation would be the
most realistic description. There are many transition relations, based on the analysis of
the test results as a function of the strain rate and taking into account the considerations
mentioned above. The following relation proved to be the most appropriate:
 
 1 
fc = a  d 
(4.1)
( )
1 + ε& e( b+ c (T ) ) 
 
Where ε& is the strain rate in s-1 and a, b, c and d are regression constants. Three of them
(a, b and d) are more or less the same for the three test temperatures, the fourth
parameter varies with temperature and will therefore be used to incorporate the
temperature influence. To incorporate this influence the relation between strength and
temperature has to be formulated.

To arrive at a relation between strength and temperature, the viscous component of


asphalt concrete behaviour is investigated. The viscosity of a material is expressed by η,
which gives the relation between the shear stress and the gradient of the shear rate
(Equation 4.2).
τ
η= (4.2)
∂γ
∂t

70 Chapter 4
With: η = coefficient of viscosity
τ = shear stress
γ = shear strain
t = time
The relation between stress and strain for the linear-elastic materials is the well-known
Hooke’s law (Equation 4.3).

σ = Eε ; E = Young's modulus (4.3)

For ideal, linearly viscous behaviour the relation between the coefficient of viscosity
and the shear modulus under the assumption of a constant shear strain rate is given in
Equation 4.4.

η SG = the time and temperature


SG = ; (4.4)
t dependent shear modulus

The relation between the shear modulus and the time and temperature dependent
stiffness modulus (SE) is shown in Equation 4.5.

2η (1 + ν )
S E = 2 SG (1 + ν ) = ; ν = Poisson's ratio (4.5)
t

If the strain rate and temperature effects are be supposed to occur solely due to the
viscous component, which is clearly a simplification but will help to get an indication of
the relation between strength and temperature, the stress in a viscous material can be
written as:

2η (1 + ν ) ε
σ = S Eε = ε = 2(1 + ν ) η (4.6)
t t

From Equation 4.6 it can be seen that the stress in a viscous material is proportional to
the strain rate and the viscosity. This influence of the viscosity represents the
temperature effect, but what is the relation between temperature and viscosity?
Heukelom found a linear relation between the logarithm of the viscosity and the
temperature for S-type bitumens (Lubbers 1985). This means that there also exists a
linear relation between the viscosity and the exponent of 1/T. It seems reasonable to
expect a relation between the strength on the one hand and the strain rate and exp(1/T)
on the other.

Based on these considerations the temperature effect was included in Equation 4.1 along
the lines shown in Equation 4.6. Since the temperature dependent regression constant c
was already part of the exponent, the following non-linear equation was found:

The uniaxial compression test 71


 
 
  ε& =strain rate in m/(m s) (=1/s)
 1 
fc = a 1 − d ; T= temperature in Kelvin (4.7)
   c
 b+   a,b,c,d= regression constants
 1 + ε& *e  
T

   
   

The term between the square brackets controls the transition between the limit strength
and zero, thereby representing the influence of strain rate and temperature on the
strength. The regression constants a through d might dependent on mix and/or bitumen
characteristics, this will become clear once more mixes are tested. In Equation 4.8 the
regression constants found when fitting Equation 4.7 to the test results using the
statistics package SPSS 9.0 are shown.

 
 
 
 1 
f c = −108  1 − 0.32 
(4.8)
  
 −86.3+
24260  
 
 1 + ε& *e  T 

   
   

The strength values predicted by this equation are compared to the test results in Table
4.8 (numerically) and Figure 4.20 (graphically).

T [oC] v [mm/s] dε/dt [s-1] fc,average [N/mm2] fc,predicted [N/mm2] (Eq. 4.8)
30 0.1 0.001 -1.9 -1.6
30 1 0.01 -3.5 -3.3
30 5 0.05 -5.2 -5.4
30 10 0.1 -7.5 -6.6
15 0.1 0.001 -5.7 -5.8
15 1 0.01 -11.5 -11.4
15 5 0.05 -17.0 -17.9
15 10 0.1 -22.5 -21.4
0 0.1 0.001 -21.5 -21.6
0 1 0.01 -38.0 -37.0
0 5 0.05 -49.5 -50.3
0 10 0.1 -56.5 -56.3
Sum of Squares: 3767 4.1
Table 4.8: Comparison between predicted and measured compressive strengths

72 Chapter 4
-1
strain rate [s ]
0 0.05 0.1 0.15 0.2 0.25 0.3
0
o
303K / 30 C
-10

-20 o
288K / 15 C
fc [N/mm ]

-30
2

-40

-50
o
273K / 0 C
-60

-70

Figure 4.20: Predicted compressive strength from Equation 4.8 compared with the
test data……………………………………………………

| fc [N/mm 2] |
100
273 K / 0oC

288 / 15oC

303 K/ 30oC
10

0.1
-1
0.0001 0.001 0.01 0.1 strain rate [ s ] 1
In
Figure 4.21: Predicted compressive strength from Equation 4.8 compared with the
test data, plotted on log-log scale ………………..
road engineering it is common practice to try to develop relations for a (material)
property as a function of temperature and frequency, the so-called master curves.
Equation 4.8, which expresses the uniaxial compressive strength as a function of strain
rate and temperature, fits within that tradition. Because master curves are usually plotted
on double logarithmic scales, the information shown in Figure 4.20 is repeated in on
log-log scale. To enable this double-logarithmic plot, the compressive strength is plotted
in absolute values.
The trend in the predicted compressive strength with respect to temperature for a larger
range is shown Figure 4.22. From this figure and Figure 4.20 it can be seen that the
relation for fc fulfils the boundary conditions mentioned earlier in this section.

The uniaxial compression test 73


[oC]
-40 -20 0 20 40 60 80 100
1200
233 253 273 293 313 333 353 373 393
-20 T [K]
ε
-40
.

-60

-80

-100

fc [N/mm2]
-120

Figure 4.22: Trend in predicted compressive strength (Equation 4.8) as a function of


temperature, for several strain rates [s-1]…………………..
The effect of a variation in the regression constants on the predicted strength values is
discussed in detail in Erkens et al. 2000.

4.3.6.2 Stiffness
Similar to the compressive strength, also the initial material characteristics: stiffness and
Poisson’s ratio must be expressed as general functions of temperature and strain rate for
application in numerical analyses. The observed trend in the modulus values was quite
similar to that in the compressive strength. Since it is known that asphalt exhibits a
maximum stiffness (“glass modulus”, Poel, v.d. 1954) for low temperatures,
independent of the strain rate, the same general relation as developed in the previous
section for the compressive strength was used to describe the stiffness data. The relation
itself is shown in Equation 4.9 and its fit to the test results is presented in Figure 4.23.

 
 
 
 1  ε& =strain rate in m/(m s) (=1/s)
log( Smix ) = 4.5  1 − 0.16 
; (4.9)
  27200   T= temperature in Kelvin
  −83.5 +  
 1 + ε& *e
T 

   
   

74 Chapter 4
2
S m ix [N /mm ]
1.E+05
T= 0 'C
T=15 'C
T=30 'C
eq. 4.9

1.E+04

1.E+03

ε [% /s]
1.E+02
0 2 4 6 8 10 12 14
Figure 4.23: Predicted stiffness (Equation 4.9) compared to the test results

4.3.6.3 Poisson’s ratio


Poisson’s ratio must also be expressed as a function of temperature and strain rate.
However, the trend in this particular parameter differs from that observed in the
compressive strength and stiffness. From the test results shown in Figure 4.24 it can be
seen that although the Poisson’s ratio increases with increasing temperature, there is no
clearly distinguishable trend between the values at a given temperature and the strain
rate. This is due to the rather large variation in the Poisson’s ratio, which is the result of
the small quantities on which this parameter is based (axial and radial strain at the
beginning of a test). Because of this, small variations in the original data have a
relatively large impact on the derived quantity.

Despite the variation, it appears that the Poisson’s ratio does increase with increasing
temperature while there is no clear inluence of the strain rate. For this reason, the
following general expression was derived:

ν = 0.0027 T − 0.49 (4.10)

with T the temperature in Kelvin

The predictions for the Poisson’s ratio on for each temperature are also plotted in Figure
4.24.

The uniaxial compression test 75


0.50
n
0.45
0.40
0.35 0.34
0.30
0.29

0.25 0.25

0.20
0.15 0'C, trek 15'C, trek 30'C, trek
0.10 0 'C, druk 15 'C, druk 30 'C, druk
0: 0.25 15: 0.29 30: 0.34
0.05
strain rate centre [% / s]
0.00
0.0 2.0 4.0 6.0 8.0 10.0

Figure 4.24: Poisson’s ratio from compression & tension data and general expression

4.4 CONCLUSIONS AND RECOMMENDATIONS

4.4.1 CONCLUSIONS
In developing the compression test a lot of attention was paid to the friction in the
contact surfaces and the temperature control in the test. From the results of this test, it
was clear that these two aspects, when neglected, have a large impact on the test results,
thus influencing any conclusion based on the results.
Friction in the contact surfaces between specimen and loading plates is the cause
of the well-known bulging or “barrelling” of specimens tested in compression. This
barrelling is the visual result of the confinement of the top and bottom of the specimen
due to the friction forces. Since the material is highly confinement sensitive, the friction
can also lead to an over-estimation of the strength. It is generally accepted that for
specimens with a height equal to or larger than twice their diameter this is not the case.
This rule of the thumb proved valid for asphalt concrete as well, but despite of this the
specimens still exhibited considerable barrelling. The system used to measure radial
deformation required uniform deformations and for this reason a friction reduction
system, consisting of a highly plastic layer of foil in combination with soft soap, was
developed. Besides its application in research, this system is also very useful for quality
control and assessment tests, because the specimens used for these tests usually do not
fulfil the requirement that the height should be twice the diameter. Tests showed that for
specimens with equal height and diameter this leads to an over-estimation of the
strength by 20%, increasing to 100% for even smaller heights. Using the friction
reduction system yields reliable strength values even for specimens that were only half
the diameter in height.
With respect to the test temperature, it was seen that a temperature difference of
only 1.5oC resulted in local failure at the warmer plate. Since the response (strength,
stiffness) itself also turned out to be very sensitive to temperature, it is absolutely
necessary that asphalt concrete is not tested without rigidly controlled temperature
conditions.
On the basis of the test results, which showed little variation, it could be
concluded that the compressive strength tended to a maximum value for increasing

76 Chapter 4
strain rates and decreasing temperatures. This plateau value is independent of
temperature and strain rate, similar to the glass modulus of bitumen. On the basis of the
test results and physical consideration concerning the influence of temperature and
strain rate on the material an expression for the compressive strength as a function of
temperature and strain rate was developed.
From the test results it also became apparent that the response varied from quite
plastic to extremely brittle, depending on the temperature and strain rate. This resulted
in a large variation in strength and stiffness values, which complicates the testing since
it requires a set-up that is accurate and reliable at both low and high load levels.
The final conclusion obtained from the compression tests is that aluminium is
too soft a material to be used for the loading plates if the friction is effectively reduced.
The large radial deformations in combination with the aggregates quickly damage the
plates to such an extend as to make it useless. When using aluminium for other set-up
components it must be taken into account that it is three times as flexible as steel and
the design should be adapted accordingly.

4.4.2 RECOMMENDATIONS
On the basis of the tests that were developed and performed in this project it is highly
recommended to perform compression tests only with an effective friction reduction
system when the deformation need to be recorded accurately. In case only the uniaxial
compressive strength has to be determined, friction reduction does not appear to be
necessary for height to diameter ratio’s of two and more. For lower ratio’s however,
friction reduction is again necessary.
Furthermore it is recommended to perform tests on asphalt concrete only under
temperature controlled conditions. The material is extremely temperature sensitive and
testing without temperature control induces additional variation in the response, making
it difficult to relate the response to a particular temperature.
To accommodate the variety in response, the test controller should allow the load
cell to be used and calibrated in fractions of its total range. This will allow testing of a
wide range of conditions at the same set-up.
Instead of aluminium high quality steel should be used for the loading plates to
prevent scratching. An alternative solution is the use of thin steel plates that can be
discarded when they get scratched. If aluminium is used for other set-up components, it
must be realised that it is three times more flexible than steel and the dimensions should
be adapted accordingly.

4.5 REFERENCES
Bolk, H., (1980), The Creep test, SCW (in Dutch)
C.R.O.W., (1990), RAW Standard Conditions of Contract for Works of Civil
Engineering Construction, VanderLoeff Drukker B.V., Enschede
Erkens, S.M.J.G., (1998), The Uniaxial Compression Test - Asphalt Concrete Response
(ACRe), Delft University of Technology, Report number 7-98-117-2
Erkens, S.M.J.G. and Poot, M.R., (2000), Additional Compression Tests - Asphalt
Concrete Response (ACRe), Delft University of Technology, Report number 7-00-
117-5
Geffen, M.C.G.J. van, (1983), Laboratory Testing of Asphaltic Concrete,
Head, K., (1982), Manual of Soil Laboratory Testing, Vol. 2: Permeability, Shear
Strength and Compressibility Tests, Pentech Press Limited, Estover Road,
Plymouth, Devon

The uniaxial compression test 77


Lubbers, H.E., (1985), “Bitumen in de Weg- en Waterbouw”, Nederlands Adviesbureau
voor Bitumentoepassingen (Nabit)
Mier, J.G.M. van, (1997), Fracture Processes of Concrete, CRC Press, Boca Raton
New York London Tokyo, ISBN 0-8493-9123-7
Poel, C. van der, (1954), “A general system describing the visco-elastic properties of
bitumens and its relation to routine test data”, Shell Bitumen Reprint No. 9,
reprinted from the Journal of Applied Chemistry, Vol. 4, part 5
Poot, M.R. , Kalf, J.A.M. and Houben, L.M.J., (1995), Dynamisch triaxiaalonderzoek
op asfalt - Onderzoek 2e fase op mengsels met standaard bitumen, , Delft
University of Technology, Report number 7-95-115-2
Timoshenko, S.P. and Goodier, J.N, (1951), Theory of Elasticity, McGraw-Hill
Kogakusha, Ltd.
Vonk, R.A., Rutten, H.S., Mier, J.G.M. van and Fijneman, H.L., (1989), Influence of
Boundary Conditions on softening of concrete loaded in compression, Paper
presented at the Fracture of Concrete and Rock - Recent Developments,,
London/New York

78 Chapter 4
5. THE UNIAXIAL TENSION TEST
“Make everything as simple as possible, but not simpler”
- Albert Einstein (1879-1955)-
45Equation Section (Next)

5.1 INTRODUCTION
As was mentioned in the previous chapter, the tests performed in this project are used to
obtain material properties. For this reason, the tests must subject the material to a known
and preferably simple state of stress while the response is measured. The previous
chapter dealt with subjecting the material to uniaxial compression, while this chapter
describes how the material is tested in uniaxial tension. Once again it is important that
the set-up and test-procedures ensure that the intended state of stress is realised. In a
tension test, failure occurs locally and after the peak load a combination of responses
occurs within the specimen. From that point on, the crack is opening while the
undamaged specimen halves unload. One of the objectives in developing the test set-up
and procedures for the tension test was to find a way to separate these different kinds of
response. The requirements for the set-up and test procedures and the ways in which
these requirements were met are discussed in the first part of this chapter. In the second
part the test conditions and results are presented.

5.2 DEVELOPING THE UNIAXIAL TENSION TEST


In the previous chapter it was shown that, when developing or performing a test, it is
important to consider the set-up and procedures in detail to ensure that the intended test
conditions are realised. Seemingly small details can have a major effect on the test
results. In this section the problems and challenges encountered while developing the
tension set-up and their solutions are discussed in detail.

5.2.1 ALIGNMENT
To perform a true uniaxial tension test the alignment of the set-up and the specimen is
of great importance. Two hinges, one at either side of the specimen ensure that the force
is applied along the specimen axis. Ideally, this would be sufficient to ensure alignment
but in reality the system must be flexible enough to accommodate imperfect specimens.
Usually, the specimens are connected to the hinges via end caps, which are glued onto
the specimen ends. These caps are screwed onto the hinges making it impossible to
accommodate deviations in the specimens prior to fracture. Once a crack appears, the
separating parts of the specimen can rotate with respect to each other, but before that
both caps have to be placed exactly in line. If a cap is slightly misaligned (or the
specimen skewed), the specimen will not fit in the set-up. This is due to the fact that at
both hinges the screw thread requires the cap to be placed perpendicular to the working
line of the set-up (Figure 5.1). If the caps are not completely parallel, this is not
possible.

The uniaxial tension test 79


traverse

loadcell

hinge

cap

specimen

cap

hinge

actuator

Figure 5.1: Two hinges give rotational freedom once a crack occurs, but they do not
compensate slight misalignments……………………………
To compensate for misalignment one of the hinges must be able to move horizontally.
In the set-up developed for the ACRe project this is realised by means of a third hinge
(Figure 5.2). Naturally, the specimen preparation procedure aims at preventing
misalignment in the specimens (Chapter 2), but it is impossible to achieve perfect
alignment. The three-hinge construction is used to accommodate any remaining (small)
deviations. If the deviations were large, the specimen would not be loaded in uniaxial
tension, but in bending.

2 hinges and horizontal


displacement

(a) (b) (c)


Figure 5.2: Alignment is important in an uniaxial tension test

80 Chapter 5
5.2.2 END EFFECTS
In a tension test the way to transfer the load onto the specimen is not as straightforward
as it is in a compression test. Somehow, the specimen must be attached to the set-up
using a system that can transfer tensile forces. The most common approach in road
engineering is to glue end caps onto the specimen. These end caps can then be
connected to the set-up. In this case, the caps are fitted with screw thread that allows
them to be screwed onto the set-up. However, the use of any connection system has an
effect on the stress distribution in the specimen. Whether the specimen is clamped or
glued, the state of stress near the specimen ends is disturbed. As a result it deviates from
the intended state of uniform tensile stresses. In case of gluing, the radial deformations
due to Poisson’s ratio are completely suppressed. These suppressed deformations lead to
stresses in the specimen in much the same way (but reversed) as discussed for friction in
compression tests (Chapter 4). This is one of the phenomena involved in cracking near
the end caps, which is often observed in tension testing. However, unlike for
compression tests, failure in a tension test is localised. Where 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. For this reason, the end effects do not influence the
observed fracturing response if the crack occurs far enough from the end caps. For
overall observations it must be noted that near the end caps a three-dimensional state of
stress exists. The effects this has will be discussed in more detail in Section 5.4.3.3.

5.2.3 SPECIMEN SHAPE

5.2.3.1 Localised failure in tension


As stated above, failure in a tension test is a local phenomenon. As a result, measuring
the overall response gives a combination of responses. Since the objective in this project
is to measure the response due to a known state of stress, it is important to separate the
different stress (and thus response) situations. In a tension test, the following responses
occur from the peak onwards:

• in the failure zone the material is cracking


• at both sides of the failure zone the material is unloading

By measuring both the overall response and that of the unloading, not-cracked parts it is
possible to reconstruct the individual responses (Figure 5.3). However, in order to do
that it must be known beforehand where the crack will occur (actually: where it will not
occur). This is achieved by a special specimen shape, which is shown in Figure 5.3.

The uniaxial tension test 81


F [kN]

l a =l/3 ∆l
a

l b =l/3 ∆l
b
w w
l c =l/3 ∆l
c

∆l

Figure 5.3: When the failure location is known, the different types of response in a
tension test can be separated.

5.2.3.2 A special specimen shape


Because tension failure is a local phenomenon, it is important for specimen
instrumentation purposes to know where the crack will occur. Otherwise data may be
lost due to a crack which disables the instrumentation. The asphalt mixture that is used
in the ACRe project (Chapter 3) is very homogeneous and as a result, fracture in
cylindrical specimens can occur anywhere, but it will most likely happen near the end
caps where the load is transferred onto the specimen. In any case, there is no guarantee
that fracture will occur under controlled circumstances, in case of failure in the
specimen the position is unknown and in case of failure at the caps the stress
distribution is unknown. If failure occurs near the caps, it is unclear if, and if so to what
extend, the stresses at the failure zone were influenced by the proximity of the end caps.

These problems can be solved by using a specimen geometry that results in a preference
for failure at a certain position. This can be achieved by a specimen with a diameter that
is maximal at the caps and reduces gradually towards the centre of the specimen. This
type of geometry had been used before in asphalt concrete research, but in those cases
the shape was achieved by cutting originally cylindrical specimens. That approach could
not be used in this case, since cutting might induce weak spots in the area of decreased
diameter, causing an underestimation of the tensile strength. To prevent this, a
completely new method was developed, based on the gyratory compactor. Using an
additional splitting mould inside the standard gyratory mould with 100 mm diameter,
the specimens were compacted in their parabolic, axi-symmetric shape (Figure 5.4). The
shape had to be such that it would provide a large enough difference between the
specimen diameter at the caps and the center to initiate failure at the center while this
difference had to be small enough to allow sufficient compaction of the specimen. The
various shapes that have been considered and the motivation for the eventual choice are
discussed Erkens et al. (2001).

82 Chapter 5
Figure 5.4: Split and gyratory moulds (right) and half the split mould (left)

5.2.3.3 Specimen production


Originally, the split mould was rigidly connected to the gyratory mould by means of
screws (Figure 5.4, left), but this caused a problem at the beginning of the compaction
process. When the mould is placed in the gyrator, the bottom plate moves upward (20-
25 mm), usually this results in the mixture moving up as well. But due to the shape of
the split mould, in this case it requires quite some pre-compaction of the asphalt in the
bottom part of the mould. As a result it takes a few minutes to get the mould in place.
Until the mix is poured in the mould, it is continuously kept at the right temperature for
compaction, but once it is inside the mould, the temperature is no longer maintained.
During the attempts to place the mould, the mix temperature drops, which causes a less
effective compaction. To solve this problem, the inner mould is left loose inside the
gyratory mould. That means that the mould can be placed easily in the gyrator,
preventing unnecessary loss of heat. Despite the energy that is wasted in moving the
inner mould inside the gyratory mould, the compaction achieved in the centre of the
specimens using this approach is considerably better than that found with the fixed inner
mould (Figure 5.5).

Figure 5.5: Difference in achieved compaction (expressed as percentage air voids, va)
for standard gyratory cores and parabolic ones with the mould fixed and loose
After the specimen is cooled down, the split mould is pressed out of the gyratory mould
by means of a press. The split mould remains in place until the specimen is cooled down
completely. After this, it must be detached from the specimen. One half of the split

The uniaxial tension test 83


mould could easily be loosened by using the other to get the necessary resistance.
Removing the remaining half caused problems because of the sticky characteristics of
bitumen. Many coatings were used to try to solve this problem: teflon spray, wax,
temperature resistant tape, a mixture of glycerine and talc but it still required excessive
force to loosen the mould, with the inherent risk of damaging the specimen. Eventually
impreglon, a factory-applied surface treatment, seemed to work. However, this coating
got damaged quickly and was eventually rubbed off by the sand particles in the mix
during compaction. The problem was finally solved by the development of a device that
allowed the two halves of the mould to be loosened slowly by using the excess material
of the specimen to deliver the necessary resistance. In combination with the use of a
mixture of glycerine and talc on the split mould surface this provided a successful
system to remove the split mould. The frame system is shown in Figure 5.6, where the
frames and the specimen are shown vertically. In reality, one side of the holding frame
is connected to a rigid plate, which results in a horizontal position of the specimen.

Holding frame

Excess
material
Split
mould
specimen
en

Excess
material

Push-off
frame

Figure 5.6: The split mould loosening system uses the excess material for support
The specimen and the mould are placed inside a holding frame that is connected to the
specimen by two screws at the top and bottom. Since there is a layer of 20 to 30 mm of
excess material on both sides, this does not damage the eventual specimen. Two smaller
push-off frames are connected to the two halves of the split mould. By means of
especially long screws that pass through the screw thread in the push-off frames and rest
in a cavity in the holding frame these frames are slowly pushed back from the holding
frame, thus loosening the mould. In Figure 5.7 an already cut specimen (without excess
material) is used to show the frame and mould more clearly.

84 Chapter 5
Figure 5.7: A device was developed which allowed loosening the split mould
Later on, an additional split mould that splits into four instead of two parts was
developed. For this mould a ring was made to take off the parts of the split mould by
using the other parts to provide the necessary resistance.

5.2.4 TEMPERATURE EFFECTS


The temperature sensitivity of asphalt concrete necessitates the use of temperature-
controlled cabinets for any test. Although the local failure in a tension test will not show
the presence of temperature gradients as clearly as was the case for the compression
tests (Section 4.2.4), it would still have an effect. For this reason the tension set-up was
constructed inside a professional temperature cabinet. This well insulated cabinet
maintained the temperature with an accuracy of 0.1oC between –40 and 100oC. The
airflow inside is such, that the temperature is uniform throughout the cabinet.

Figure 5.8: The door opening of the temperature cabinet was covered with plastic foil

The uniaxial tension test 85


5.2.5 SET-UP
The tension set-up that was developed consists of a rigid frame that is placed in a
temperature-controlled cabinet. The upper hinge is placed directly on the traverse of this
frame. The load cell is positioned underneath this hinge and immediately below the load
cell is the second hinge. The actuator is connected to the bottom part of the frame,
outside the temperature-controlled cabinet. The third hinge is placed on the actuator.
The specimens are connected between the two lower hinges. The fact that the load cell
is also placed between two hinges ensures that the measured force is the actual force
applied on the specimen (Figure 5.9).

LOAD CELL

HINGES

Figure 5.9: The TU Delft tension set-up uses 3 hinges to provide the required degrees
of freedom …………………………………………………….

5.3 THE UNIAXIAL TENSION TESTS

5.3.1 THE SPECIMENS


The specimens are compacted in their parabolic shape in a gyratory compactor. To
obtain the special shape, a split mould is used inside the gyratory mould (Section
5.2.3.3). Once the specimen is removed from the split mould, the excess material (all
the material outside the parabolic shape) is cut-off. To protect the specimen (and
especially the vulnerable mid section) from getting damaged in the process, the
specimen is placed in a cutting mould. The cutting mould is a Perspex mould with the
same shape as the split mould (Figure 5.10). The cut mould is bolted to the sawing table
during sawing and provides continuous support to the specimen. The non-parabolic
parts of the specimen are supported at the bottom, but not covered by the top of the
mould. These parts are cut off, leaving a specimen of 90 mm high. This specimen has
80 mm diameter at top and bottom, 50 mm at the centre and a parabolic shape in
between.

86 Chapter 5
Figure 5.10: A perspex cutting mould is used to protect the specimen during cutting
After cutting, the specimen is glued to the end caps that are used to transfer the load to
the specimen in the tension set-up. Despite the use of the cutting mould, the cutting
surfaces can still be somewhat non-parallel due to movements of the cutting blade and
some freedom in positioning it. To prevent gluing the caps non-parallel, since that
would lead to bending stresses in the specimen, a gluing mould is used. In this mould
the centre of the specimen, rather than one of the cutting surfaces is used as a reference.
The specimen is clamped at its centre, using a PVC ring of 40 mm high. The ring has
the same shape as the split mould and the cutting mould and the caps are placed parallel
to the centre of the specimen. The glue that is used is Araldite SW404 (an epoxy resin)
with HY2404 hardener. This is a two-component filling glue and it is smeared on both
the specimen surface and the cap to ensure proper contact with both. To ensure that the
glue can correct any deviation of the cutting surface, more glue than necessary is used.
This allows the glue to fill any gap there might exist between cutting surface and cap.
Excess glue spills out between cap and specimen and to prevent a layer of glue on the
topside of the specimen and the bottom cap, both are covered with tape. The glue spills
onto the tape and is later on removed, together with the tape (Figure 5.11).

Figure 5.11: Gluing mould ensures that caps are parallel to centre of the specimen

The uniaxial tension test 87


When the central clamping ring of the gluing mould is attached to the specimen, its top
is used to mark the specimen with horizontal lines at 120o intervals. These markers are
later on used to position the strain gauges (Section 5.3.3.2). This way, all strain gauges
are placed at the same height with respect to the centre of the specimen.

5.3.2 TENSION SET-UP IN DETAIL


The tension test set-up consists of a rigid loading frame in which an MTS 50 kN
hydraulic actuator is mounted. The actuator is connected rigidly to the bottom plate. The
frame itself is placed on a support frame that places the bottom plate at 1 m above the
floor. The bottom plate is a sandwich construction of a 600x330x100 mm steel plate
bolted onto a 600x600x50 mm steel plate. The top plate of this sandwich construction
provides the connection to the vertical bars and stiffens the bottom plate to the extend
necessary to support a 50 kN actuator. The vertical bars are 100 mm in diameter, made
of stainless steel. They are connected to the bottom plate via 8 M10 bolts each. The
actuator piston passes through the bottom plate. The first of the three hinges is screwed
directly on this piston. The other side of this hinge is screwed into the thread of one of
the specimen end caps. Between the other cap and the load cell, placed above the
specimen, is a second hinge and the third hinge is positioned between the load cell and
the traverse. This steel traverse has dimensions 530x150x200 mm and is clamped onto
the vertical bars. This clamping mechanism allows the traverse to be placed at any
height along the vertical bars. In the current set-up the traverse is placed close to the top
of the vertical bars.

During the test the load, the total specimen deformation and the axial and radial strain
are measured. The measurement systems are connected to a PC-based data acquisition
system, which produces a single ASCII output file for each test. These files are the basis
for the data-analysis procedures discussed in Chapter 4.

The force applied with the 50 kN MTS actuator is measured with a Lebow 50 kN load
cell (model 3157). The actuator is controlled by an MTS test controller (model 407)
which allows the selection of the appropriate force range for the test conditions. The test
controller is the central part of the closed-loop servo-hydraulic system used for this set-
up. The test is displacement-controlled and a programmable function generator (MTS
microprofiler model 418.91) generates its rate. Since some of the tests were performed
at rather high deformation rates and low temperatures, which resulted in high
stiffnesses, the system was fitted with a 3.8 l/min high response valve (MTS, model
252.41), which allowed the actuator to apply the required forces at a high speed.

The specimen is put in a temperature-controlled Weis-Enet cabinet with inside


dimensions 1.0x1.0x1.0 m. The unit can control the temperature between –40 and 100
o
C with an accuracy of ± 0.1 °C.

The axial deformation of the specimen was measured by three LVDT’s. The specimen
deformation was also controlled via these LVDT’s. This on-specimen control system
ensured that the applied deformation was the actual deformation of the specimen. The
axial and radial strains are measured at a fixed height (Section 5.3.3.2) using cross
shaped strain gauges. These gauges are connected to a series of Peekel amplifiers. The
on-specimen instrumentation (displacement transducers and strain gauges) is discussed
in more detail in Section 5.3.3.

88 Chapter 5
All measuring devices are connected to an input/output board, which consists of a BNC
connection panel with 16 single ended or 8 differential connections. These channels are
connected to the data acquisition board, a PCMCIA-12AI card from Keithley metrabyte,
which is placed in a Pentium II portable PC. For the tension test 10 single ended
connections are used (force, 3 LVDT’s, 3 axial and 3 radial strain gauges). The data
acquisition board was controlled through the Windows based data acquisition program
TestPoint. Through this program the number of scans per channel and the scan rate are
defined. The latter was varied with the test conditions, while the number of samples was
kept constant at 10.000. For the more brittle conditions, the maximum scan rate of the
measurement card was used, officially this is 100 kHz. For continuous sampling of a
single channel this can be achieved, but when sampling multiple channels the real scan
rate is slower. To prevent data-acquisition errors a scan rate of 8000 Hz (per channel,
yielding 80000 Hz in total) was used under these conditions. The data acquisition was
triggered by the microprofiler shortly before the test started. During a test, the output of
the measurement systems was measured in voltages and saved as columns in an ASCII
file. These files were then further analysed by means of Microsoft Excel.

5.3.3 INSTRUMENTATION
A side effect of the special specimen shape, is that it is difficult to instrument.
Limitations in the height of the gyratory mould meant that there was no room to use a
cylindrical area at the centre of the specimen. Only at the centre the diameter is the
minimum value of 50 mm, to simplify instrumentation it would have been practical to
have some height at the centre of the specimen over which the diameter would have
been 50 mm. On the other hand, creating a smooth transition between a curved and a
flat surface is notoriously difficult and a non-smooth transition leads to stress
concentrations. For these reasons, it was decided to settle for specimens without a
cylindrical area and accept the difficulties involved in instrumenting it. Later on, work
published by van Vliet (1994) showed that if a cylindrical center part is used stress
concentrations occur indeed in the rounded corners (in numerical modeling as well as
experiments), initiating cracking at those locations. Since the ACRe specimens did not
have this cylindrical part, this problem did not arise.

(a) (b)

Figure 5.12: ACRe shape (a) and a specimen a cylindrical part at the centre (b)

The uniaxial tension test 89


5.3.3.1 Displacement transducers
For the overall measurements, connecting displacement transducers between the caps
instead of on the specimen can solve the problems caused by the specimen shape.
Depending on the test conditions, the overall deformations vary between 0.03 mm and 3
mm, accordingly either ±5 mm or ±1 mm LVDT’s (Solartron AG 5 and AG 1) were
used. The influence of the glue and caps on the measurements is considered negligible
due to the thin layer and high stiffness, respectively. Both types of caps are considerably
stiffer than the specimen, even under the extreme conditions: Ecap=210000 MPa (for
steel) or 71000 MPa (for aluminum). The glue has a stiffness comparable to that of
asphalt concrete under extreme conditions (Eglue ≈ 9000 MPa and Especimen ≤ 10000
MPa) but is used in a thin layer. The connection between the caps is realised by a set of
rings. In the upper ring there are three holes at 120o angles, in which the LVDT’s are
clamped. The bottom plate is a massive ring on which the LVDT’s are placed (Figure
5.13).

5.3.3.2 Strain gauges


As mentioned before, the parabolic shape of the specimen is necessary to fix the
location of fracture without introducing stress concentrations, but it leaves no straight
parts that can easily be instrumented. Nevertheless information on the unloading of the
non-damaged parts of the specimen had to be obtained to allow separation of the
unloading and crack-opening that occur simultaneously in the specimen after the peak.
For this, strain gauges were used. The strain gauges used here are 90˚ 2-element cross
gauges (type PFC-10-11) from Tokyo Sokki Kenkyujo. These are foil-etched gauges
with a polyester backing. The gauge length is 10 mm and k=2.13. The gauges are glued
on the specimen with a cyanoacrylate single-component adhesive and coated with
PU120 from Hottinger Baldwin Messtechnik. The gauge signal is amplified by a Peekel
CA 110 universal precision amplifier. The CA-110 offers a total balancing range of ±
5000 µV/V.

Figure 5.13: Specimen with LVDT’s connected on the caps and strain gauges

90 Chapter 5
axial radial

tangential

Axial: along the cylinder axis


radial: along the specimen circumference
at a given height (in horizontal direction)
tangential: along the specimen circumference
in vertical direction

Figure 5.14: Definition of axial, radial and tangential with respect to the strain
gauges …………………………………………………
Three 2-element cross strain gauges are glued directly onto the specimen surface at the
same height at 120˚ intervals. The signal of the radial gauges could effectively be
considered to give the radial strains at a given height. The tangential gauges, however,
necessarily cover a range of heights in which the change in diameter was not negligible.
As a result, these gauges measured a combination of radial and axial deformations
(Figure 5.15). To obtain the “true” axial strains, these had to be extracted from the
tangential gauge signal.

Figure 5.15: The tangential strain gauge measures a combination of radial and axial
strain ………………………………………………………..

The uniaxial tension test 91


Before determining the way to calculate the axial strains from the tangential signal, the
shape of the signal over the gauge length had to be known. Since a straight line can be
used to approximate the shape over a limited height, it was assumed that this would also
be the case for the strains. If the specimen is considered to behave linearly elastic, the
same Stiffness and Poisson’s ratio can be used for the whole specimen and in that case
there is a direct relation between the specimen cross section, stress and strain (Section
5.4.2.2). Using the radial strain, the axial strain can be determined from the tangential
strain gauge signal. The relation between the axial and the radial component in the
tangential signal depends on the angle with the vertical axis under which the gauge is
glued. Since this angle varies with the height, the gauges should be glued at a fixed
height on the specimen.

The gauges themselves have a length of 10 mm. To ensure that the three 2-element
cross strain gauges are always glued on the same distance from the centre (defined as
mid-height) for all specimens, the clamping ring (Section 5.3.1) is used to mark these
positions. This mark is 20 mm from the centre of the specimen. The radial gauges are
aligned along this mark, so that the position is fixed. Since the geometry at this position
is known, the axial deformation can now be determined from the tangential gauge
signal. This method, presented in Erkens et al. (2001), yields the following expression
for the axial strain:
 2  1  
2

( 2h )  2
( ) 2 1
2
2 2 2 *
−2h + − 4h h − Lgauge +  R1 − R2 + ε R,C RC ( − )  
  R1 R2  

(ε axial )1,2 = 2
(5.1)
2h
Where:
h = part of the gauge that is aligned in axial direction, prior to the test
L*gauge = total gauge length (tangential) during the test (including elongation)
R1,2 = specimen radius at top and bottom of tangential gauge prior to the test
Rc = specimen radius at the position of the radial gauge prior to the test
εR,C = strain measured by the radial gauge
εaxial = that part of the tangential gauge signal that corresponds to the axial
specimen elongation

Unlike displacement transducers, strain gauges do not have to be calibrated since the
parameters that determine their output are provided directly by the manufacturer.
Calibration of strain gauges is therefore a matter of checking whether you measure what
you intended to. In this case, it meant checking whether the bitumen skin on the outside
of the specimen or the stiffness of strain gauge and glue influenced the results.
Comparing the strain gauge signal to that of a clip-on gauge outfitted with a bridge of
Wheatstone showed that the strain gauges did indeed measure the response accurately.
This verification process is presented in detail in the tension report (Erkens et al. 2001).

5.4 TEST PROCEDURES, CONDITIONS AND RESULTS

5.4.1 TEST PROCEDURES


The specimens were kept in a storeroom at the test temperature for at least 24 hours
prior to testing. This storeroom was placed next to the set-up and to ensure that the
temperature in the storeroom and the set-up was the same, an electronic reference
thermometer was used to check the temperature in both.

92 Chapter 5
Before testing the specimen was put in the temperature controlled set-up in a box of
sand, the sand provided continuous support which prevented damage to the specimen
(due to its shape the specimen is extremely sensitive to bending under its own weight).
In this box the strain gauges were connected to the precision amplifiers by soldering
wires to the soldering terminals on the specimen. After this the specimen was placed
between the hinges and instrumented with 3 LVDT's for on-specimen control. The
plastic curtain at the front of the temperature cabinet (Section 5.2.4) ensured that during
specimen instrumentation the temperature was kept more or less constant. Despite this,
the specimen was left in the set-up for about half an hour to account for slight
temperature variations before starting the test.

After instrumentation the LVDT's, the load cell and the strain gauges could be set in
range and the appropriate settings of the programmable function-generator and data-
acquisition system could be set (i.e. strain rate and scan rate). To be able to measure the
unloading of the specimen after cracking the actuator must not move back upwards after
specimen failure. This was accomplished by setting the 407 MTS servo controller to
valve clamping at +50 % of its full stroke. This setting was activated when a pre-set
error-level (the difference between the control signal and the actual on-specimen
deformation) was detected.

It was attempted to apply a small pre-load just before the test to ensure that there would
be no play in the set-up components (particularly the hinges). For 0 and 15oC this could
readily be achieved, but for 30oC the pre-load relaxed immediately. This causes a load
introduction effect at the beginning of the response curves at this temperature.

5.4.2 TEST CONDITIONS

5.4.2.1 Overview of the test conditions


The tests were performed in the displacement-control mode, applying a continuous axial
deformation rate. Since asphalt concrete response is temperature and strain rate
sensitive, these parameters are varied within the test programme. The test conditions can
therefore be specified by a combination of temperature and strain rate, or temperature
and deformation rate. The latter parameter, the applied deformation rate of the actuator,
is used to control the test but it is the strain rate in the material that determines the
response. For the specimen shape used in these tests the relation between the two is not
straightforward, therefore it will be discussed in more detail in Section 5.4.2.2.

The results from the tension tests will be combined with those from the compression
(Erkens et al. 1998 and Erkens et al. 2000) to determine model parameters. Therefore,
the results from both tests under similar conditions (temperature and strain rate) must be
known. This can be achieved by performing the tests at the same conditions or by
generalising the results by expressing them as a function of strain rate and temperature.
This must be done anyway, to allow implementation of the test results in the model.
Initially the same test conditions were chosen for the tension tests as for the
compression tests. During the preliminary tests, however, it became apparent that the
response in tension is so different from that in compression that simply maintaining the
same conditions would not result in the best characterisation of the material response
under tension. This is particularly the case for higher strain rates, in the compression
tests the plateau value of the strength was not reached even with the highest strain rates.

The uniaxial tension test 93


The existence of a plateau could only be deduced, on the basis of theoretical as well as
experimental observations. In tension it turned out that, especially at lower
temperatures, most of the rates used in compression sufficed to actually reach the
plateau strength and, thus, they did not provide any information on the changes in the
strength as a function of (temperature and) strain rate.
For this reason and because the strain rates for which the plateau was reached varied
with temperature, the test conditions were adapted and the strain rates were selected for
each temperature separately. As a result, the temperatures are the same as for the
compression tests, but the strain rates are different. These rates were selected for each
temperature in such a way as to get as much information as possible on the transition to
the plateau as well as on the plateau itself. The resulting conditions are shown in Table
5.1, together with the originally intended (compression) conditions.

94 Chapter 5
Tension Compression
• •
v [mm/s] ε [%/s] T [oC] # samples v [mm/s] ε [%/s] T [oC] # samples
0.000664 0.001 0 4 0.1 0.1 0 5
0.00664 0.01 0 5 1 1 0 5
0.0664 0.1 0 5 5 5 0 5
0.664 1 0 4 10 10 0 5
0.0664 0.1 15 5 0.1 0.1 15 5
0.33 0.5 15 4 1 1 15 5
0.664 1 15 4 5 5 15 5
0.0664 0.1 30 4 10 10 15 5
0.664 1 30 4 0.1 0.1 30 5
1.99 3 30 4 1 1 30 5
3.32 5 30 5 5 5 30 5
6.64 10 30 3 10 10 30 5
Table 5.1: Overview of the test conditions

5.4.2.2 Relation between strain rate and deformation rate


As stated above, the relation between the overall deformation rate and the strain rate at
the position where cracking occurs is not straightforward, due to the specimen shape.
This shape causes the stresses to vary over the specimen height. If a single stiffness is
assumed for the whole specimen, the strains can be found on the basis of the stress
distribution (geometry). This yields a relation between the overall deformation (rate)
and the strain (rate) at the centre of the specimen, or that at the position of the strain
gauges.
 F  F 1
u = ∫ ε dh = ∫   dh = ∫ dh (5.2)
 EA  E A
Where:
u= deformation
F= applied force
E= Youngs’ modulus
A= specimen cross section (at height h)

0.0035 0.25
ε [mm/mm] u [mm]
0.003
0.2
0.0025
0.002 0.15

0.0015 0.1
0.001
0.05
0.0005
h [mm]
0 0
0 10 20 30 40 50 60 70 80 90

Figure 5.16: Strain and overall deformation over the specimen

The uniaxial tension test 95


The overall deformation is found by integrating over the specimen height and the
deformations over the strain gauges by integrating over the strain gauge height. The
ratio between the two for any combination of F and E is:
h1
1
u gauge
∫ A dh
h0
= hspec
(5.3)
utotal 1
∫ A
dh
0
Where:
h0=height of the bottom of the strain gauge (with respect to the specimen)
h1=height at the top of the strain gauge (with respect to the specimen)
hspec=total specimen height

The strain measured by the strain gauges is the deformation that occurs over the gauges
divided by the gauge length. So if the gauge signal is multiplied by the gauge length and
divided by the above-mentioned ratio, it yields the overall specimen deformation. The
gauge length is 10 mm which is placed at an angle with the cylinder axis and the
“effective height” (along that axis) is 9.59 mm. As a result, the relation between the
total specimen deformation and the strain gauge data is:
*
ε gauge hgauge ε gauge × 9.59
utotal = h = = ε gauge × 83.22 (5.4)
1
1 0.1152
∫ A dh
h 0
hspec
1
∫ A
dh
0
Using the same approach, a relation between the strain at the centre of the specimen
(where the crack will occur) and the overall deformation can be established:
1 1
ε centre A A
= h centre ⇔ ε centre = utotal h centre = 0.0151× utotal (5.5)
utotal spec
1
spec
1
∫ A dh ∫ A dh
0 0
Or:
hspec
1
∫ A
dh
utotal = ε centre 0
= 66.411× ε centre (5.6)
1
Acentre
These same relations hold between the strain rate at the centre and the overall
deformation rate. Since the crack will occur at the centre, the strain rate at that position
is the one to which the strength will be related. The overall deformation rate should
therefore be determined on the basis of the intended strain rate at the centre. For the
strain rates mentioned in Table 5.1 this means that they must be multiplied by 0.66411
to obtain the deformation rate (strain rates were given in %/s). This yields the
deformation rates shown (Table 5.1).

96 Chapter 5
It must, however, be noted that this is an approximation of the real situation. In reality,
the relation between strain at the centre and overall deformation is complicated because
the response does not remain linear-elastic. The regions with higher stresses will first
start to exhibit inelastic response. The corresponding inelastic deformations will change
the relation between overall and centre strain, since now the strains at the centre will be
larger. Another aspect that complicates the relation between overall strain and strain at
the centre is the fact that the apparent stiffness of asphalt concrete is a strain rate
dependent quantity. As such, it will vary over the specimen height, just like the strains.
The higher strains at the centre lead to a higher stiffness, which in turn leads to smaller
strains. This is a continuous process that basically counter-acts the previously
mentioned one. Where plastic deformations will lead to larger strains at the centre, the
stiffness effect will tend to decrease those strains. Finally, the above presented relation
is based on the assumption of a uniaxial state of stress throughout the specimen. In that
case the strains can be computed directly from the applied load, the specimen geometry
and the stiffness. In the uniaxial tension test, however, the specimen is glued to the end-
caps, which results in a three-dimensional state of tension stresses. The radial tensile
stresses will, via Poisson’s ratio, lead to a smaller axial strain than expected on the basis
of a uniaxial state of stress. This effect will diminish over the specimen height, causing
the difference between the strains near the end caps and the specimen centre to be larger
than expected on the basis of the analytical solution. Also, the overall deformations will
be smaller than predicted on the basis of that analysis.

To get an impression of the effects of these phenomena, a test was carried out. In this
test a specimen was instrumented with gauges at different heights. Some results are
shown in Figure 5.17. The test was performed at T=15 °C, using a cyclic force signal
with an amplitude of 1.5 kN and a frequency of 1 Hz. The second strain gauge was
placed at the height used for the gauges in the actual experiments, the first one was
placed 10 mm closer to the centre of the specimen while the third one was placed 10
mm closer to the cap.

strain ( mm/m)
strain gauges at different heights

350 rad1 rad2 rad3 ax1 ax2 ax3


1

250

150

2 3
50
time [s]

-50 0 1 2 3 4 5 6 7 8 9 10

-150

Figure 5.17: Axial and radial strain signals for gauges at the original position (2),
closer to the centre (1) and closer to the cap (3)
The steeper slope of the average signal (indicated by the markers and straight lines) for
the gauges closer to the centre is due to increasing plastic deformations near the centre.
The difference in amplitude of the dynamic response is the result of the differences in

The uniaxial tension test 97


diameter between the three positions. When comparing the actual amplitude ratio
(A1:A2:A3=1.32:1.35:1) with that predicted on the basis of the linear-elastic stress
distribution (A1:A2:A3=1.5:1.3:1) it can be seen that they match rather well. Only the
amplitude of the signal closest to the specimen centre (A1) appears to be smaller than
expected when compared to the other two. This migth be due to the above mentioned
strain rate variation over the specimen. The observed effect, where the elastic strains on
level 1 and 2 appear to be equal, might be an indication of this effect, where the
difference in modulus compensates that in diameter as well as the influence of the end
effects. From the results presented in Figure 5.17 it appears that the deviations from the
assumed strain distribution are not significant. For this reason, it was decided to use the
relation between strains and overall deformation that was determined on the basis of
uniaxial stresses and linear-elasticity.

It must be mentioned that, if only the tensile strength needs to be determined, using
cylindrical specimens will prevent the uncertainty with respect to the strain rate at the
location of the crack, provided that the problem of fracture at the end caps can be
solved. In that case, the location of the crack will not be known beforehand, but the
strain rate distribution is straightforward. If, however, the unloading should also be
measured as was the case here, a pre-defined crack location is needed. In the Road and
Railway Research Laboratory satisfactory results were obtained using cylindrical
specimens of coarser mixtures (Ven, van de et al. 2002). For the relatively
homogeneous ACRe mixture, however, fracture at the end caps remains a problem.

5.4.3 TEST RESULTS


During each test the axial and radial deformations and the forces were recorded. The
accuracy of these signals was known from the calibration data of the displacement
transducers and loadcell, respectively. Using the specimen geometry, these signals were
transformed into strains and stresses, from which the apparent tensile strength (ft), the
axial strain at peak stress (εax,max) and the radial strain at peak stress (εrad,max) could be
determined.

5.4.3.1 Control parameter and measuring length


For stable experiments in which the post-peak response can be measured, the energy
required for crack growth must remain larger than the energy released by the unloading
of both the uncracked parts of the specimen and the loading frame. If this is not the
case, the crack will grow explosively, resulting in a vertical unloading branch or even a
snap-back (Figure 5.18). The latter is a phenomenon known from rock and concrete
testing. To minimise the chance that this happens, the loading frame should be stiff
compared to the specimen, to limit the energy stored in the frame. Another way to
minimise the chance of unstable crack growth is the selection of a proper control
variable. If the applied force is used to control the test, which is the classical way, the
test can not be continued after the peak. The reason is that the control variable should be
a continuously increasing quantity. After the peak, the specimen may still have
remaining strength but it is less than the peak value. Trying to increase the load past the
peak value leads to sudden failure. Using the axial deformation allows the registration
of the post-peak response and provides information on the remaining strength in this
post-peak region, which is an important factor in the design of structures since it
addresses the safety margins (“warning capacity”). If however the experiment becomes
unstable, the relation between force and deformation is no longer unique (one
deformation level corresponds to several force levels, depending on the stage of the

98 Chapter 5
test). In that case, in a displacement controlled test, a vertical drop of the force signal is
measured.

stress

Snap back

Registered in displacement control

Registered in force control


deformation
Figure 5.18: If too much elastic energy is released a snap back can occur
In the case of a uniaxial tension test, the length over which the specimen deformation is
measured is also important. By using on-specimen control, the effect of the unloading of
the set-up on the control signal is excluded from the signal. Because a crack is very
local and all the material on either side is unloading, the released energy that effects the
control variable increases with increasing registration length (Figure 5.19). If the
location of the crack is known, it is possible to use only a small part on either side of
that location to control the test. Another option is the use of many LVDT’s and a
sophisticated system to ensure that the one that measures the largest deformation is used
as a control signal (Mier, van 1997). This ensures a continuously increasing control
signal (Vliet, van 2000). In case of the highly strain rate sensitive nature of asphalt
concrete the latter solution might cause problems because during the post-peak response
local failure will disturb the strain rate that is imposed. Switching between LVDT’s at
various positions where the strain rates may vary, will further complicate the already
complex situation. Using the overall deformation at least ensures a consistent definition
of the overall strain rate.

With respect to the measuring length, for concrete values of 35 mm (Mier, van 1997)
and 70 mm (Vliet, van 2000) have been reported. In the latter case, it also involved
direct tension specimens with a gradually decreasing cross section. The specimens used
in this project are 90 mm in height and because, unlike the specimens used in concrete
research, they have a circular cross section, it is difficult to attach instrumentation to it
(there are no straight parts). For these reasons it was decided to use the full specimen
length as measuring length for the overall response.

The uniaxial tension test 99


Stress σ [MPa]
4
Measuring length lmeas [mm] lmeas
0 125 250 500
3 50
50
2 500 mm
ft=3.2 MPa E=30 GPa

Deformation w [µm]
0
0 20 40 60 80 100 120 140 160
Figure 5.19: Influence measuring length (Hordijk, 1991)

5.4.3.2 The different response components


Beside the overall response, in the post-peak regime the unloading of the uncracked
parts was measured also by means of strain gauges. The total deformation that is
measured during the tests consists of several components, namely instantaneous elastic,
delayed elastic (viscous) and permanent (plastic) deformation. These different
components of response are in reality not as easy to separate as in theory. This can be
seen from Figure 5.20 where the force and axial strain data are shown as function of the
number of scans (which is equivalent to a time base) for two different test conditions.
The graph at the top shows an example of sudden failure and the one at the bottom
illustrates ductile failure. From these graphs it can be seen that even in the case of
sudden failure the load diminishes gradually at first. In case of ductile failure this
gradual unloading continues to the end of the test. The fact that unloading, at least
initially, occurs gradually rather than instantaneously makes the distinction between
instantaneous and delayed elastic response quite difficult. Similarly, the recovery of
delayed elastic deformation continues well after the force has returned to zero, which
complicates the definition of the “remaining deformation”.

Because of these complications it was necessary to establish practical definitions for the
different components of response, to ensure that these were determined in the same way
for all test conditions, despite the different strain and scan rates that were used. For this
reason, the steepest part of the unloading branch is defined as the one corresponding to
“instantaneous” unloading. Furthermore, the point at which the force is zero is used to
define the end of the test and the remaining deformation at this point is considered the
permanent deformation. These definitions are illustrated in Figure 5.21, where a force-
deformation diagram for ductile failure is shown. If sudden failure occurred, the gradual
descend of the load-displacement diagram is suddenly disturbed, the deformation
increases and the load decreases rapidly (Figure 5.22).

100 Chapter 5
4000 10
micro strain sudden failure force [kN]
3500 9
8
3000
7
2500 strain
6
2000 5

1500 4

force 3
1000
2
500 1
scan #
0 0
1 501 1001 1501 2001 2501 3001 3501 4001 4501

9000 micro strain 5


ductile failure force [kN]
8000 4.5

7000 4
strain
3.5
6000
3
5000
2.5
4000
2
3000
force 1.5
2000 1
1000 0.5
scan #
0 0
1 501 1001 1501 2001 2501 3001 3501 4001

Figure 5.20: Signals for sudden (above) and ductile (underneath) failure
5
F [kN] 15_0.1_T200
4.5

3.5

2.5

2
instantaneous
permanent delayed
1.5

1 crack width (w)

0.5
utot [mm]
0
0 0.5 1 1.5 2 2.5 3

Figure 5.21: Different components of deformation in a test with ductile failure

The uniaxial tension test 101


12
F [kN]

10

6
permanent instantaneous

4
crack width (w)

2
utot [mm]
0
0 0.5 1 1.5 2 2.5

Figure 5.22: Different components of deformation in a test with sudden failure

5.4.3.3 Fracture energy


The area under the load-crack width diagram is usually considered as the fracture
energy. This is a parameter used in models for tensile fracture and the registration of the
unloading response was done to determine this parameter. If the deformation of the
unloading specimen halves is subtracted from the overall deformation, what remains is
the crack width. In Figure 5.21 this is the distance between the overall and the
unloading branches. The area between these branches is the fracture energy (Figure
5.23). The crack width at which the stress is again zero is defined as the maximum crack
width (wmax).

2.5 s [N/mm2]
15_0.1_T200

1.5

Gf
1
Gf

0.5

w [mm]
0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0

Figure 5.23: The fracture energy is the area under the stress-crack width diagram

102 Chapter 5
5.4.3.4 Test results
Graphic representations of the average test results per condition are shown, grouped per
temperature, in Figure 5.24, Figure 5.25 and Figure 5.26. If the response was stable,
allowing the post-peak response to be captured, the results could be averaged directly, if
brittle failure occurred the procedure had to be slightly adapted. This was necessary to
prevent a sudden change in the average values due to the disappearance of one of the
signals after sudden failure (in that case the stress level of that signal drops to zero). The
procedure used to arrive at the average signals is discussed in detail in Erkens et al.
(2001).

In the graphs, on the right side the stress (at the specimen centre) is plotted against the
axial deformation. From the peak, there are two responses plotted, the overall
deformation and the unloading path. This unloading of the specimen is calculated from
the strain gauge measurements on the basis of the elastic strain distribution in the
specimen, as described in Section 5.4.2.2. On the left side the stress is plotted versus the
radial deformations at the centre, also computed from the strain gauge registration using
the elastic strain distribution. The kinks that can be seen in some of the curves are the
effect of averaging, they are not present in the individual curves.

5
σ [N/mm2]

4.5 10%/s
4 5 %/s
3 %/s
3.5 1 %/s
0.1%/s
3
ε
2.5
2
1.5
1
0.5
0
-0.05 urad [mm] 0.45 0.95 1.45 uax [mm] 1.95

Figure 5.24a: Average response curves at 30oC and several strain rates

The uniaxial tension test 103


2.5
σ [N/mm2]

5 %/s
2 3 %/s
1 %/s
0.1%/s
1.5
ε

0.5

0
-0.05 urad [mm] 0.45 0.95 1.45 uax [mm] 1.95

Figure 5.24b: Average response curves at 30oC, leaving out the fastest rate
In these first two graphs the results at 30oC are shown. The first shows all five strain
rates used at this temperature, the second only the four slowest. It was attempted to
apply a strain rate of 10%/s to establish whether the plateau strength at this temperature
was the same as for the other temperatures. This appears to be the case, but it proved
impossible to impose a constant strain rate of 10%/s, in reality the strain rate varied
during the test and the actual rates at failure were much higher. Because of this and
because this strength value and those for the other strain rates differ considerably, a plot
showing only the slower rates is also included.

Because it was not possible to apply a pre-load at 30oC, the signals exhibit a peak at the
beginning. In the graphs shown here, the slope from the unloading branch was used to
correct for this effect of play in the set-up. The sudden change in slope between the
ascending branch and the remainder of the signal that can be seen for the slower rates is
a result of this effect.

6
σ [N/mm2] 0.1%/s
. 1%/s
5 ε 0.5%/s

0
-0.15 0 0.15 0.3 0.45 0.6 0.75 0.9 1.05 1.2 1.35 1.5
urad [mm] uax [mm]

Figure 5.25: Average response curves at 15oC and several strain rates (the kinks are
due to averaging, at those points one specimen exhibited brittle failure)

104 Chapter 5
6
σ [N/mm2]
0.001%/s
5 0.01%/s
0.1%/s
4 1%/s
.
ε
3

1
uax [mm]
urad [mm]
0
-0.1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Figure 5.26: Average response curves at 0oC & several strain rates (due to brittle
failure only a few points after the peak were captured for tests at higher strain rates)
In all graphs the basic aspects are conform the expectations, the strength and apparent
stiffness (slope of the ascending branch) increase with increasing strain rate. The
maximum strength value found is the same for the three temperatures, approximately 5
N/mm2, but it is reached for completely different strain rates. From the graphs it can
also be seen that some conditions did not result in stable tests, meaning that the post-
peak and unloading response were not captured for those conditions. This is the case for
the highest rates at 30 and 15oC and for all but the slowest strain rate at 0oC. In some
cases this was because failure occurred at a much higher rate than the test itself and the
data-acquisition rate was too slow to capture sufficient data points to reconstruct the
unloading part. In other cases, the response appeared to be brittle, yielding a vertical
drop in the overall signal, which was captured accurately.

Table 5.2 also contains information about the composition of the specimens. For these
specimens the degree of compaction and void percentage varies over the height of this
specimen. Since failure occurs locally, it is not correct to use the average mix
composition to indicate the composition of the failure zone. For this reason, slices of
material (approximately 20 mm thick) at each side of the failure zone were cut-off and
analysed after the test. The strength values (ft) shown are determined using the diameter
at the height of the crack. This information could be obtained from these slices, by using
the differences in weight to determine the distance of the crack to the actual centre of
the specimen. In most cases, the difference in diameter was negligible and it was never
large, since the cracks all occurred very close to the actual specimen centre. On basis of
these slices also the void percentage in the failure area (V%) was determined. The
specimen density shown was determined in the usual way by weighting the complete
specimens dry, wet and under water. Finally, the fracture energy (Gf) and maximum
crack width (wmax) that were determined from the unloading branches and overall
response in the way that was discussed earlier (Sections 5.4.3.2 and 5.4.3.3) are shown
for each test.

The uniaxial tension test 105


filename code T dε/dt ∆dε/dt ρprfstk ∆ρprfstk V% ∆V% ft ∆fc Gf ∆Gf wmax ∆wmax
o 2
[ C] [%/s] [kg/m3] [N/mm ] [N/mm] [mm]
0_0.001_T75 T75 0 0.0010 2126 ± 2.34 11.8 ± 0.4 2.7 ± 0.1 2.2E-01 ± 0.87 8.7 ± 4.15
0_0.001_T179 T179 0 0.0006 2240 ± 1.17 7.4 ± 0.4 2.5 ± 0.1 2.2E-01 ± 0.87 0.4 ± 4.15
0_0.001_T102 T102 0 0.0010 2163 ± 1.17 8.2 ± 0.4 2.8 ± 0.1 1.7E+00 ± 0.87 4.0 ± 4.15
average 0 0.0009 ± 5.E-04 2177 ± 145 9.1 ± 5.8 2.7 ± 0.3 7.2E-01 ± 2.16 4.4 ± 10.29
0_0.01_T119 T119 0 0.010 2258 ± 1.19 6.6 ± 0.4 6.3 ± 0.2 3.5E-01 ± 3.72 0.4 ± 0.77
0_0.01_T123 T123 0 0.010 2222 ± 1.18 8.4 ± 0.4 5.2 ± 0.1 4.5E+00 ± 3.72 0.9 ± 0.77
0_0.01_T211 T211 0 0.010 2271 ± 2.31 4.3 ± 0.4 5.3 ± 0.1 8.1E+00 ± 3.72 2.0 ± 0.77
0_0.01_T177 T177 0 0.010 2202 ± 1.16 9.6 ± 0.4 4.2 ± 0.1 *** ± 3.72 *** ± 0.77
0_0.01_T167 T167 0 0.010 2209 ± 2.29 8.8 ± 0.4 4.3 ± 0.1 3.4E-01 ± 3.72 0.2 ± 0.77
average 0 0.010 ± 4.E-05 2232 ± 40 7.5 ± 2.6 5.0 ± 1.1 3.3E+00 ± 4.63 0.9 ± 0.96
30_0.1_T162 T162 30 0.10 2237 ± 1.17 9.2 ± 0.4 0.3 ± 0.1 2.1E-01 ± 0.15 2.2 ± 0.93
30_0.1_T163 T163 30 0.10 2187 ± 1.16 11.1 ± 0.4 0.3 ± 0.1 1.3E-01 ± 0.15 3.0 ± 0.93
30_0.1_T171 T171 30 0.10 2195 ± 1.16 12.6 ± 0.4 0.3 ± 0.1 2.4E-01 ± 0.15 2.4 ± 0.93
30_0.1_T62 T62 30 0.10 2246 ± 1.18 9.3 ± 0.4 0.5 ± 0.1 4.8E-01 ± 0.15 4.2 ± 0.93
average 30 0.10 ± 1.E-04 2216 ± 50 10.6 ± 2.6 0.3 ± 0.2 2.6E-01 ± 0.24 2.9 ± 1.48
15_0.1_T101 T101 15 0.10 2243 ± 1.18 8.2 ± 0.4 2.3 ± 0.1 7.1E-01 ± 0.19 0.7 ± 0.37
15_0.1_T200 T200 15 0.10 2239 ± 1.18 6.9 ± 0.4 2.4 ± 0.1 1.2E+00 ± 0.19 1.5 ± 0.37
15_0.1_T205 T205 15 0.10 2216 ± 1.17 8.1 ± 0.4 2.0 ± 0.1 9.0E-01 ± 0.19 0.8 ± 0.37
15_0.1_T166 T166 15 0.10 2245 ± 1.94 9.0 ± 0.4 1.7 ± 0.1 8.9E-01 ± 0.19 1.3 ± 0.37
15_0.1_T105 T105 15 0.10 2255 ± 1.19 7.6 ± 0.4 2.6 ± 0.1 1.1E+00 ± 0.19 1.4 ± 0.37
average 15 0.10 ± 1.E-04 2240 ± 15 8.0 ± 1.0 2.2 ± 0.5 9.6E-01 ± 0.23 1.1 ± 0.46
0_0.1_T108 T108 0 0.10 2213 ± 2.32 8.1 ± 0.4 4.9 ± 0.1 *** ± *** *** ± ***
0_0.1_T121 T121 0 0.10 2214 ± 1.18 8.5 ± 0.4 4.2 ± 0.1 2.0E-02 ± 0.02 0.01 ± 0.01
0_0.1_T125 T125 0 0.10 2190 ± 1.18 8.5 ± 0.4 4.5 ± 0.1 1.9E-02 ± 0.02 0.02 ± 0.01
0_0.1_T148 T148 0 0.10 2281 ± 1.18 1.5 ± 0.4 7.1 ± 0.1 6.2E-02 ± 0.02 0.03 ± 0.01
0_0.1_T115 T115 0 0.10 2248 ± 1.19 7.1 ± 0.4 5.1 ± 1.6 4.5E-02 ± 0.02 0.01 ± 0.01
average 0 0.10 ± 4.E-04 2229 ± 45 6.7 ± 3.7 5.2 ± 1.4 3.7E-02 ± 0.03 0.0 ± 0.01
15_0.5_T201 T201 15 0.50 2276 ± 1.18 4.6 ± 0.4 4.6 ± 0.1 8.1E-01 ± 0.52 0.3 ± 1.19
15_0.5_T208 T208 15 0.51 2195 ± 1.17 10.3 ± 0.4 3.7 ± 0.1 8.1E-01 ± 0.52 1.4 ± 1.19
15_0.5_T203 T203 15 0.50 2258 ± 1.18 6.4 ± 0.4 4.5 ± 0.1 1.7E+00 ± 0.52 2.7 ± 1.19
15_0.5_T122 T122 15 0.50 2238 ± 1.19 7.8 ± 0.4 4.5 ± 1.3 4.6E-01 ± 0.52 0.1 ± 1.19
average 15 0.50 ± 5.E-03 2242 ± 60 7.3 ± 3.8 4.3 ± 0.6 9.4E-01 ± 0.83 1.1 ± 1.90
30_1_T70 T70 30 1 2230 ± 2.32 7.0 ± 0.4 1.1 ± 0.0 1.1E+00 ± 0.16 2.6 ± 0.47
30_1_T128 T128 30 1 2247 ± 1.19 7.8 ± 0.4 1.2 ± 0.1 9.7E-01 ± 0.16 3.6 ± 0.47
30_1_T173 T173 30 1 2219 ± 1.16 8.7 ± 0.4 0.7 ± 0.1 8.2E-01 ± 0.16 2.7 ± 0.47
30_1_T69 T69 30 1 2233 ± 2.37 7.7 ± 0.4 0.7 ± 0.1 1.2E+00 ± 0.16 2.7 ± 0.47
average 30 1 ± 3.E-03 2232 ± 20 7.8 ± 1.1 0.9 ± 0.4 1.0E+00 ± 0.26 2.9 ± 0.74
15_1_T207 T207 15 1 2209 ± 1.15 6.1 ± 0.4 5.0 ± 0.0 3.1E-01 ± 0.17 0.07 ± 0.03
15_1_T209 T209 15 1 2245 ± 1.18 5.2 ± 0.4 5.2 ± 0.1 5.2E-02 ± 0.17 0.02 ± 0.03
15_1_T206 T206 15 1 2217 ± 1.18 8.6 ± 0.4 4.9 ± 0.1 2.6E-01 ± 0.17 0.05 ± 0.03
15_1_T155 T155 15 1 2253 ± 2.00 5.7 ± 0.4 5.6 ± 0.1 4.7E-01 ± 0.17 0.10 ± 0.03
average 15 1 ± 2.E-01 2231 ± 35 6.4 ± 2.4 5.2 ± 0.4 2.7E-01 ± 0.27 0.06 ± 0.05
0_1_T64 T64 0 1 2229 ± 1.18 7.4 ± 0.4 4.5 ± 0.0 1.7E-02 ± 0.01 0.01 ± 0.00
0_1_T124 T124 0 1 2272 ± 2.38 3.7 ± 0.4 6.5 ± 0.0 3.1E-03 ± 0.01 0.00 ± 0.00
0_1_T126 T126 0 1 2243 ± 2.35 7.2 ± 0.4 5.4 ± 0.1 2.5E-02 ± 0.01 0.01 ± 0.00
0_1_T80 T80 0 1 2251 ± 1.19 6.1 ± 0.4 5.8 ± 0.1 2.8E-02 ± 0.01 0.01 ± 0.00
average 0 1 ± 2.E-01 2249 ± 30 6.1 ± 2.7 5.5 ± 1.4 1.8E-02 ± 0.02 0.01 ± 0.01
30_3_T67 T67 30 3 2225 ± 1.74 7.8 ± 0.4 1.9 ± 0.0 8.7E-01 ± 0.16 2.5 ± 0.18
30_3_T158 T158 30 3 2262 ± 1.17 6.3 ± 0.4 1.8 ± 0.1 *** ± 0.16 *** ± 0.18
30_3_T113 T113 30 3 2219 ± 1.19 8.2 ± 0.4 2.2 ± 0.1 1.1E+00 ± 0.16 2.2 ± 0.18
30_3_T160 T160 30 3 2228 ± 1.95 9.3 ± 0.4 1.3 ± 0.1 8.6E-01 ± 0.16 2.3 ± 0.18
average 30 3 ± 3.E-02 2234 ± 35 7.9 ± 2.0 1.8 ± 0.6 9.6E-01 ± 0.25 2.3 ± 0.28
30_5_T127 T127 30 5 2248 ± 1.18 6.1 ± 0.4 2.2 ± 0.0 1.6E+00 ± 0.69 3.1 ± 0.85
30_5_T175 T175 30 5 2191 ± 1.15 10.9 ± 0.4 1.5 ± 0.1 1.0E+00 ± 0.69 2.6 ± 0.85
30_5_T210 T210 30 5 2245 ± 1.18 6.4 ± 0.4 2.1 ± 0.1 2.3E+00 ± 0.69 2.4 ± 0.85
30_5_T149 T149 30 5 2246 ± 1.18 5.8 ± 0.4 1.8 ± 0.1 2.8E+00 ± 0.69 4.4 ± 0.85
30_5_T74 T74 30 5 2252 ± 1.19 7.4 ± 0.4 2.6 ± 0.5 1.5E+00 ± 0.69 2.6 ± 0.85
average 30 5 ± 3.E-02 2236 ± 35 7.3 ± 2.6 2.0 ± 0.5 1.8E+00 ± 0.86 3.0 ± 1.05
30_10_T169 T169 30 52 2223 ± 1.93 ** ± ** 4.6 ± 0.1 9.6E-02 ± 0.03 0.04 ± 0.01
30_10_T170 T170 30 41 2178 ± 1.16 11.2 ± 0.4 4.1 ± 0.1 4.4E-02 ± 0.03 0.02 ± 0.01
30_10_T76 T76 30 46 2245 ± 2.35 7.0 ± 0.4 5.5 ± 0.1 1.0E-01 ± 0.03 0.04 ± 0.01
average 30 46 ± 13 2215 ± 85 9.1 ± 7.4 4.7 ± 1.8 8.1E-02 ± 0.08 0.04 ± 0.03

Table 5.2: Overview test conditions and results for the individual specimens

106 Chapter 5
As expected, the results from the tension tests show more variation than those from the
compression tests. This can easily be understood by the kind of failure that occurs in
both tests. In compression failure is due to overall material degradation, which results in
a kind of average response from the specimen. In tension failure is localised and thus
more sensitive to local weaknesses in the specimen. Combined with the fact that the
variation in specimen composition was larger in the tension than in the compression
specimens (due to the shape and production procedure, Section 5.2.3), this explains the
larger variation.

5.4.4 GENERAL EXPRESSIONS FOR THE TEST RESULTS


Eventually, the model will be used to analyse laboratory tests and road constructions.
For these applications, it is necessary to have general expressions for the material
characteristics and model parameters. In a finite element program the temperature and
loading rate at each integration point in the structure under analysis can be determined
and with that information the appropriate model parameters and material characteristics
must be determined. For this reason general expressions for the test results as a function
of temperature and strain rate will be developed. In this chapter relations for the
material characteristics as they were obtained from the tension tests are discussed. The
determination of the model parameters and general relations describing them is
presented in Chapter 7.

5.4.4.1 Tensile strength


Basically, the relation must be able to describe the same physical trends as described for
the compressive strength in the previous chapter:

1. ε =0: this corresponds to not loading the specimen, because the strength
decreases with decreasing strain rate this is defined as: ft=Cmin

2. ε →∝: the strength will not increase indefinitely, but reach a limit: ft=Cmax.
Based on the “glass modulus” of bitumens, this limit is expected to be
temperature independent.
3. T→-∝: for extremely low temperatures, asphalt will exhibit glass-like, linear-
elastic behaviour until sudden, brittle fracture occurs. This ultimate tensile
strength will be independent of the loading rate: ft=Cmax.
4. T→∝: for extremely high temperatures (approximately 160oC) the bitumen will
become a fluid, ft=Cmin.

In Chapter 4 it was shown that these considerations can be met by a non-linear


transition relation (Equation 5.7):

   
   
   
 1   1 
ft = a 1 − d 
= 5.5 1 − 0.86 
(5.7)
   b+   
 c 
   −80.2+
 25050 

 
 1 +  ε
& e  T
  1 + ε
&  e  T 

       
     
Where ε& is the strain rate in s-1, T the temperature in Kelvin en a,b,c,d are regression
constants.

The uniaxial tension test 107


The relation was fitted to the tension data, shown in Figure 5.27. The non-linear
regression analysis was carried out using the statistical package SPSS 9.0. In these
analyses iterative estimation algorithms are used to find the most appropriate set of
regression constants. Due to the non-linear nature of the relations there are local minima
and maxima in the solution and as a result the solution found is not necessarily the best
one. The quality of the results can depend on the analysis method selected and on the
choice of the starting values for the regression constants. To assess if the solution was
really the optimal one, several sets of starting values were used and they all resulted in
the same solution. That the relation that was found described the test results quite well
can be seen from Figure 5.27 where the lines are the predicted strength values and the
markers the individual test results.

8.0 2
ft [N/mm ] Test data 0'C Eq. 5.7
7.0 Test data 15'C Eq. 5.7
Test data 30'C Eq. 5.7
6.0

5.0

4.0

3.0

2.0

1.0
strain rate centre [% / s]
0.0
0.0 1.0 2.0 3.0 4.0 5.0 6.0

Figure 5.27: Tension test results (markers) and Equation 5.7

1.E+01
2
ft [N/mm ]

1.E+00

1.E-01
Test data 0'C Eq. 5.7
Test data 15'C Eq. 5.7
Test data 30'C Eq. 5.7
strain rate centre [% / s]
1.E-02
1.0E-03 1.0E-02 1.0E-01 1.0E+00 1.0E+01 1.0E+02

Figure 5.28: Tension data and general expression plotted on double logarithmic scale
Because in road engineering general expressions for test results (mastercurves,
expressions for, for example, the stiffness as a function of temperature and strain rate or

108 Chapter 5
frequency) are usually plotted on double logarithmic scale, this was also done for the
tensile strength (Figure 5.28). In this plot also the high rates that were imposed when
trying to test the material at 30oC and 10%/s are shown. On the non-logarithmic plot
these values were omitted because incorporation changed the scale in such a way that
the trend in the data at lower rates could no longer be discerned.

5.4.4.2 Elasticity parameters


Unlike for the compression test, the tensile test data were not used to determine
elasticity parameters. The reason for this is that in the compression tests the end effects
could be minimised using the friction reduction system, while this is not possible in the
tension tests. The glue that connects the specimen to the end caps restrains the radial
deformation at the specimen ends. Although this does not effect the cracking process
that occurs near the specimen center, it does result in a three-dimensional state of
tension stresses near the caps. Due to the contraction that results from the horizontal
tensile stresses, the axial strains near the end caps are smaller than they would be in case
of a uniaxial stress-state. Weissenberg (Reiner 1960) illustrated this effect using two
specimens of cloth. Both cloths were stretched lengthwise and one of them was allowed
to contract freely at the ends by connecting it to parallel rollers. The other cloth
specimen was fixed rigidly at the ends. Once in a state of equilibrium, both cloths were
marked by rows of horizontal cuts in the top-half and vertical cuts in the bottom half.
The resulting deformations are shown in Figure 5.29, where the unrestrained cloth is
shown on the left and the restrained one on the right.

From Figure 5.29 it can be seen that in the unrestrained specimen, shown on the left, the
opening of the horizontal and vertical cuts is uniform throughout the specimen,
indicating a uniaxial state of stress. In the restrained specimen, however, the openings
vary. The horizontal cuts open wider near the specimen centre than near the end and the
vertical cuts are closed near the centre and open next to the specimen end. From the
smaller opening of the horizontal cuts close to the specimen end in the restrained case, it
can be concluded that the strains in this region will be under-estimated if the overall
data are used and, as a result, the stiffness will be over-estimated.

The uniaxial tension test 109


Figure 5.29: Visual representation of the end effect in a tension test
To get a qualitative impression of this effect an analytical analysis of the 2D case was
performed (Figure 5.30).

σ xx 1 ν 0 ε xx
E
σ yy = ν 1 0 ε yy
1 −ν 2
σ xy 1 − ν γ xy
0 0
2
y
fully restrained y
fully free

x x

E
εxx =0⇒σxx =νσyy σxx = (εxx +νεyy ) =0⇒εxx =−νεyy
1−ν2
1 1 E E
εyy = (−νσxx +σyy) = (−ν2σyy +σyy) ⇔ σyy = 2 (νεxx +εyy ) = 2 (−ν2εyy +εyy ) ⇔
E E 1−ν 1−ν
σyy σyy
E= E=
εyy (1−ν2) εyy
Figure 5.30: Analytical analysis of the influence of the end-effect in the 2D-case

110 Chapter 5
As can be seen, the stiffness is over-etimated in the restrained case and the amont of
over-estimation depends on the Poisson’s ratio. A series of three-dimensional numerical
analyses (one-element test) were performed to get an indication of the differences in
stiffness that would be obtained. The analyses consisted of a run without lateral
restrained, one where all nodes were restrained in the horizontal directions and finally
one where only the nodes in the bottom plane were restrained. The results are presented
in Figure 5.31. As can be seen, the influence of the end effect on the stiffness values is
considerable.

8-noded element, E=1000MPa, ν=0.3


z
y
x

Case: σzz σxx=σyy εzz εxx=εyy E

1) Fully free: 1 0 0.001 -0.0003 1000

2) Fully restrained 1 0.4286 0.000743 0 1346

3) Bottom restrained 1.13 -0.0665 0.00091 -3.08x10-4 1243 (top)


0.87 0.3665 0.00091 0 956 (bottom)
1098 (average)
Figure 5.31: Numerical representation of the end-effect in 3D
From the above it is clear that the compression stiffnesses, obtained from a test in which
the end-effects were minised, provides a more reliable indication for the stiffness than
the tension tests. For this reason, the compression values were used to arrive at a general
expression for the apparent mixture stiffness (Equation 5.8):

 
 
 
 1 
log( Smix ) = 4.5 1 − 0.16 
(5.8)
   −83.5+ 27200   
  T 
 1 + ε& e  
   

and Poisson’s ratio:

ν = 0.002733 T − 0.49 (5.9)

Where ε& is the strain rate in s-1, T the temperature in Kelvin en a,b,c,d are regression
constants.

The uniaxial tension test 111


5.5 CONCLUSIONS AND RECOMMENDATIONS

5.5.1 CONCLUSIONS
It is widely accepted that, in order to build a high quality tension test set-up the
alignment is very important. To accommodate this, the set-up developed in the ACRe
project was built according to precise specifications and equipped with high quality
hinges. A disadvantage of hinges is that they always lead to some play in the equipment,
which is difficult to compensate for in a tension test because of the small overall
deformations. A rigid test frame, to minimise other deformations in the set-up, and a
very fast hydraulic valve were used to compensate for this. Also, the deformation was
controlled and measured on the specimen, thus preventing influence of the play in the
hinges on the deformation signal. The alignment proved to be very good, causing
smooth fracture planes.
Another aspect of importance in the development of the tension test was the
specimen shape. In a tension test two different types of response occur simultaneously
after the peak load. On the one hand the crack is opening while on the other hand the
undamaged specimen parts are unloading. For modelling purposes, it is important to
differentiate between these types of response. This was achieved by measuring both the
overall response and the unloading of the undamaged parts, subtracting these two
responses results in the crack opening behaviour. The unloading was measured using
strain gauges, which were glued onto the specimen. To ensure that the crack would not
occur at the position of the strain gauges, and thus disable them, the specimen was given
a parabolic shape. This ensured that the crack would occur at or near the centre of the
specimen, leaving the remainder of the specimen for the strain gauges. Unlike notches,
the gradual change in specimen shape used in this project does not lead to the extreme
stress concentrations that obscure the actual tensile strength.
To achieve the required specimen shape, a special split mould was developed. This split
mould was placed inside a standard gyratory mould and enabled the compaction of the
specimen in the required shape. This procedure worked very well, although it would
have been even better if the temperature of the mixture could have been maintained
throughout compaction. Since this was not the case, the cooling of the mixture
interfered with the compaction, causing a relatively high variation in the specimen
composition at the centre. This method of specimen production resulted in specimen
with a thin bitumen rich layer (“skin”) at the outside. It was verified that this layer did
not interfere with the strain gauge measurements, ensuring accurate unloading
registration in the test.
Because the specimens are glued to the end caps, the deformation at the
specimen ends is restrained. This results in a three dimensional tensile state of stress
that is not suited to determine the material stiffness.

5.5.2 RECOMMENDATIONS
Despite the play that is introduced by hinges, it is recommended to use them in a tension
set-up since they minimise the influence of the set-up on the test results. By using a
rigid frame, a high response hydraulic valve and on-specimen instrumentation the effect
of the play can be adequately reduced.
The registration of unloading in the tension tests was done using strain gauges.
If these tests will be performed on open graded mixtures in the future, strain gauges
cannot be used. During the evaluation of different measurement systems, the available

112 Chapter 5
non-contacting systems, such as lasers proved to be too inaccurate. The rapid
developments in that field, however, will in the near future result in useful systems.
These can then be used to measure the unloading in open graded mixtures. Once these
systems are sufficiently accurate they also offer an alternative for the use of strain
gauges. The instrumentation of a specimen with strain gauges is extremely time
consuming and, thus, an easy-to-use non-contacting system is attractive even for the
dense asphalt mixtures.
Although the specimen shape does not lead to extreme stress concentrations, it
does have some effect on the stress state. If both cylinders and parabolic specimens
from the same mixture were to be tested, the effect of this influence could be assessed.
For the homogeneous ACRe mixture testing of cylinders was not successful, but for
coarser mixtures it has been done successfully.
To limit the variation in the composition of the fracture zone, a method to
maintain the temperature during gyratory compaction should be developed.
In order to arrive at a reliable stiffness for the material, tests with negligible end
effects, such as the compression test, should be used rather than a tension test.

5.6 REFERENCES
Erkens, S. and Poot, M., (2000), Additional Compression Tests - Asphalt Concrete
Response (ACRe), Delft University of Technology, 7-00-117-5
Erkens, S., (1998), The Uniaxial Compression Test - Asphalt Concrete Response
(ACRe), Delft University of Technology, 7-98-117-4
Hordijk, (1991), “Local Approach to Fatigue of Concrete”, PhD. Thesis Delft
University of Technology
Mier, J.G.M. van, (1997), “Fracture Processes of Concrete”, CRC Press Boca Raton,
New York, London, Tokyo, ISBN 0-8493-9123-7
Reiner, M. (1960), “Deformation, Strain and Flow, an elementary introduction to
rheology”, H.K. Lewis&Co. Ltd., London
Ven, M.F.C. van de, Erkens, S.M.J.G., Poot, M.R. and Rooijen, D. van, (2002),
“Embedded Rail in Asphalt (ERIA) Standaard en speciaal materiaalonderzoek
t.b.v. computersimulaties”, Delft University of Technology, 7-02-131-1
Vliet, M.R.A. van (2000), “Size Effect in Tensile Fracture of Concrete and Rock”, PhD.
Thesis Delft University of Technology, ISBN 90-407-1994-2, Delft University
Press, the Netherlands

The uniaxial tension test 113


114 Chapter 5
6. THE MULTIAXIAL TEST
“Errors using inadequate data are much less than those using no data at all”
-Charles Babbage (1791-1871)-

6.1 INTRODUCTION
The tests described in this dissertation are meant to provide the information necessary to
determine the model parameters used in the ACRe material model (Chapter 7). On the
basis of the uniaxial tension and compression tests described in the previous chapters,
four out of the five model parameters can be determined. The fifth parameter is related
to the influence of multiaxiality on the response. To determine this parameter the
influence of confinement on the material strength needs to be known. Since damage in
asphalt concrete is caused predominantly by tension-compression states of stress, the
effect of confinement on the response to these states of stress is of particular interest.
Shear can be decomposed into tension and compression, which makes a shear set-up an
interesting tool to use for this part of the investigation. As a result, a set-up developed
during a previous project on reinforcement in Asphalt Concrete (Bondt, de 1999) was
adapted to be used in this final part of the ACRe project. The adaptations are discussed
in the first part of this chapter and in the second part the tests and their results are
presented.

6.2 DEVELOPING THE TEST SET-UP


Shear tests have been used in many forms (Figure 6.1), such as the shear boxes used in
soil mechanics, the Leutner test used in road engineering and the constant height shear
test developed in the Strategic Highway Research Program. It is not surprising that there
exist so many different shear tests. Shear is a combination of tension and compression
and as such it is an interesting loading case for many materials. It is, however, difficult
to impose a pure shear load, in most shear tests the specimen is subjected to a
combination of shear and bending stresses (Bondt et al. 1993). As a result it is very
difficult to interpret the results.
Drill core
Parallel guidance system
D=150 mm
Moving part max τs=4/3 Ps/Fs

Fixing clamp P

spring

Fixed part τs = shear stress

Figure 6.1: There are many shear tests, like the Leutner (right) and shear box (left)

The multiaxial test 115


6.2.1 IDEAS BEHIND THE FOUR-POINT SHEAR TEST
In the four-point shear test it is attempted to circumvent this problem of combined shear
and bending stresses. Four-point shear tests have been used to study shear failure in
concrete, rock and metal. The idea is to load a beam in such a way that, according to
linear-elastic beam theory, a region of uniform high shear stresses is created while at the
centre of that region the bending moment is zero. This state of stress is achieved by an
appropriate test geometry. Although the actual construction of the set-ups can vary
considerably, the basic principle remains the same. In this section it is explained on the
basis of the geometry of the four-point shear set-up developed in the Road & Railway
Research Laboratory.

First of all, a hinge between the actuator and the loading block ensures that the forces
applied to the specimen are a function of their respective distance to the actuator (Figure
6.2).
• Factuator=F1+F2
• F2a2-F1a1=0
F2=F1a1/a2=2/13 F1
F1 (1+a1/a2)=Factuator
Factuator F1=Factuator/(1+a1/a2) = 13/15 Factuator
F2=Factuator a1/(a2(1+a1/a2)) = 2/15 Factuator
a2=195 mm
a1= 30 mm
Loading block

F2 F1 specimen

Bottom plate

Figure 6.2: Loading configuration in the four-point shear test


Secondly, the supports are placed in such a way that a zone of high shear stresses
develops between the middle supports while the bending moment at the centre of the
span between these supports is zero. As a result, in this cross section a combination of
high shear stresses exists without the disturbance of a bending moment.

116 Chapter 6
supports

Horizontal
actuator
specimen

supports
Horizontal
support
2/15 Factuator 2/15 Factuator

2/15 Factuator*(241.5-56.5)=24 2/3 Factuator


Figure 6.3: Shear and moment distribution in a four-point shear test

6.2.2 THE EXISTING SET-UP


The set-up in the Road & Railway Research Laboratory was initially developed to study
the effect of reinforcement in asphalt concrete. For these tests, large pre-cracked
specimens with and without reinforcement were tested in the four-point shear set-up
(Bondt 1999). The test was meant to determine the shear load that a specimen could
carry as well as the deformations over the crack at different levels of confinement. The
response to both monotonic and cyclic loading was investigated in this project. The
cyclic loading required a stable connection system for the specimens and the fact that
the specimen was pre-cracked meant that the centre zone (the high shear area) could not
be too small because otherwise the material near the support might fail. These
considerations resulted in the set-up shown in Figure 6.4. Supports were realised by
drilling holes trough the specimens and placing steel bars through them. This provides a
stable connection that can take both tension and compression. The hearth-to-hearth
distance between the middle supports was 65 mm. The specimen sizes tested were either
(lxhxw): 450x125x110 (non-reinforced) or 450x125x250 mm (reinforced).

The multiaxial test 117


Figure 6.4: The four-point shear set-up in its original configuration

6.2.3 CHANGES TO FACILITATE THE ACRe TESTS


For the tension and compression tests, the set-ups were developed in an iterative
approach. The initial design was made on the basis of the intended test and a series of
preliminary tests was used to determine whether the set-up functioned the way it was
intended to. If this was not the case, or if other problems were encountered, the set-up,
instrumentation or data-acquisition was modified to solve this. In case of the four-point
shear test this approach was not feasible because of time restraints. For this reason, only
those adaptations that could be thought out beforehand and carried out relatively easily
were done. Basically, this meant that the adaptations to the massive parts of the set-up,
such as loading block, actuators, frame and bottom plate were circumvented. These
changes were left as recommendations for later, using temporary solutions for the
current series of tests. In the following sections the changes are discussed, where
temporary solutions were used, this is mentioned along with the suggested final solution.
These final solutions are presented in more detail in the report about the four-point shear
tests (Erkens et al. 2002).

Also, because the tests were run on the set-up without an elaborate series of preliminary
tests and the corresponding data-analysis some problems were encountered afterwards,
during the data-analysis phase. These problems and their solutions are mentioned in
Section 6.3.3.3.

6.2.3.1 Temperature control


Like the tension and compression tests, the shear tests are used to determine material
parameters. For implementation in the Finite Element package CAPA-3D these
parameters will have to be expressed as a function of temperature and strain rate. In
order to do that, the tests must be performed for different temperatures and strain rates as
well. To facilitate the former, the shear set-up was equipped with a temperature control
cabinet similar to the one used for the compression test (Chapter 4). The cabinet consists
of an insulated, demountable cabinet of wood and roofmate (an insulation material) and
the same external temperature control unit as used for the compression tests was
attached to the cabinet to maintain the temperature during testing.

118 Chapter 6
6.2.3.2 Vertical tension using pendulum bars
Since the kind of tests for the ACRe project differed from those performed on the set-up
before, the set-up had to be adjusted accordingly. First of all, in this set up only
montonic tests would be performed. As a result, the specimen did not have to be
completely fixed in the set-up. Second, the specimens were not pre-cracked or notched
but would be loaded until failure in the set-up. This meant that, rather than deforming as
two halves the specimen would initially deform as a single entity. Pendulum bars were
used to provide the necessary freedom of movement and minimise the influence of the
rigidity of the set-up on the test. In order to use pendulum bars, the vertical load had to
be a tension rather than a compression load. Since both the actuator and the connection
to the loading block (a hinge) could facilitate tension loading, this did not pose a
problem.

Since the specimens would actually be broken in the set-up, their cross section was kept
small (50x70 mm) to limit the necessary load levels. This smaller size in combination
with additional space available in the frame of the set-up provided enough space for the
use of pendulum bars of 150 mm in length. These pendulum bars were connected to
either the bottom plate or the loading block at the top, via hinges. At each support, two
pendulum bars were placed, connected by a frame in which the specimen could be
positioned. Inside these frames small rolls were placed, these were the actual supports
for the specimen (Figure 6.5). The pendulum bars ensured freedom of movement for the
specimen throughout the test.

Pendulum bars

Frame

Rolling support (D=8 mm)

Figure 6.5: Inside the frames, small rolls provide the actual supports

6.2.3.3 Width of the shear zone


Because the ACRe specimens are not pre-cracked or notched the high shear area could
be smaller than was the case in the original set-up. The smaller this zone gets, the closer
it matches the central plane in which the bending moment is zero. Ideally in this test

The multiaxial test 119


failure should occur in that plane, so the smaller the shear zone, the better. The closest
distance between the centre supports was dictated by the width of the pendulum bars.
This resulted in a shear zone of 17 mm only. Initially, it seemed that the cracks were
growing from loading plate to loading plate. This is not illogical, since stress
concentrations occur at these places and they can easily be crack-initiators. Eventually,
during some slow tests it was possible to arrest the test before the specimens was broken
in two and from the crack pattern observed, it could be concluded that there definitely
developed a shear zone in the area between the middle supports. In the region over the
height of the specimen small shear cracks developed (Figure 6.6). These grow together
eventually to form one large crack.

Figure 6.6: Small shear cracks in the centre zone eventually grow together in a large
crack (complete specimen above and close-up underneath) ..
The placement of the supports was now dictated by the existing hole pattern in the
loading block and bottom plate. Changing these patterns would have quite time
consuming and therefore omitted. However, the required stress distribution in the
specimen could also be obtained with smaller distances between the outer supports and
because this would allow the use of smaller (less length) specimens it would reduce the

120 Chapter 6
costs of specimen production and result in specimens which are easier to handle. For this
reason, it is recommended to change the support positions for future applications. The
suggested set-up geometry is presented in the four-point shear test report (Erkens et al.
2001).

6.2.4 CONNECTION OF THE HORIZONTAL ACTUATOR


The existing connection of the horizontal actuator consisted of an angle steel at the base
of the actuator and an additional guidance clamp with roller bearings around the piston.

Hydraulic Actuator

Load cell
Hinge

Angle steel
Specimen Stretching rope
Hydraulic Actuator

Bottom plate
Figure 6.7: Existing connection of the horizontal actuator
The angle steel provided a good connection of the actuator to the bottom plate but,
because it is a hinge, it leaves much freedom of movement. The guidance clamp was
intended to provide additional stability. However, the position of this clamp at the piston
resulted in high lateral forces on the piston during testing. Besides the damage to the
piston itself, this can cause damage to the actuator. For this reason, the clamp was
removed. Eventually, it will be replaced by a steel frame that connects the actuator
housing to the bottom plate. Such a construction requires additional holes with thread in
the bottom plate and to save time, during this first series of tests a stretching rope was
used instead. In Section 6.3.2 the four-point shear test set-up that was the result of all
these modifications is shown and discussed in detail.

6.3 THE MULTIAXIAL TESTS

6.3.1 SPECIMENS

6.3.1.1 Specimen production


The specimens were beams of 500x70x50 mm (lxhxw). The specimen height was taken
higher than its width to agree as much as possible with the classical shear beam in which
it is assumed that the width of the beam is small compared to its height. Specimens of

The multiaxial test 121


this size and geometry are usually cut from asphalt concrete plates. The facilities to
prepare such plates are not available in the Road & Railway Research Laboratory. For
this reason, the plate preparation was done at KOAC WMD. They produced 8 slabs of
750x600x70 mm from which 9 specimens of 500x70x50 were cut at the Concrete
Laboratory of the Delft University of Technology. Although the specimens would fail
locally, as was the case in the tension tests, there was no reason to expect a large variety
in specimen composition in this case. The specimen composition was therefore
determined prior to testing. At that time, the dimensions of the specimens were also
determined. The width and height were determined at three positions over the specimen
length and the reported values are the average of these three measurements. The
specimens are coded with respect to the plates from which they were cut: S (for shear)
plate number – specimen number, e.g. S1-4. After this, the specimens were put in
wooden boxes filled with sand, which were stored in temperature controlled storage
rooms. An overview of the specimen geometry and composition is shown in Table 6.1a
and b.

Code h [mm] w [mm] l [mm] r [kg/m3] v%


S1-1 71.2 50.0 499 2233 4.7
S1-2 70.6 49.9 499 2257 3.7
S1-3 71.1 50.0 499 2258 3.6
S1-4 71.1 49.8 499 2258 3.6
S1-5 70.9 50.0 499 2258 3.6
S1-6 70.7 50.0 499 2256 3.7
S1-7 70.9 49.9 499 2257 3.7
S1-8 71.1 49.9 499 2259 3.6
S1-9 71.0 50.0 499 2256 3.7
S2-1 70.8 49.9 500 2250 4.0
S2-2 70.4 50.1 500 2258 3.6
S2-3 70.6 50.1 500 2260 3.6
S2-4 70.9 50.0 500 2257 3.7
S2-5 71.0 50.0 500 2257 3.7
S2-6 70.9 50.1 500 2257 3.7
S2-7 70.8 50.0 500 2256 3.7
S2-8 70.7 50.1 500 2259 3.6
S2-9 71.0 49.6 500 2264 3.4
S3-1 70.1 51.8 498 2260 3.6
S3-2 70.0 51.9 498 2257 3.7
S3-3 70.3 51.7 498 2259 3.6
S3-4 70.3 51.0 498 2260 3.6
S3-5 70.0 51.0 498 2271 3.1
S3-6 70.3 51.2 498 2244 4.2
S3-7 70.5 51.2 498 2257 3.7
S3-8 70.4 51.1 498 2259 3.6
S3-9 70.7 50.7 498 2261 3.5
S4-1 70.8 51.0 498 2258 3.6
S4-2 70.6 51.1 498 2262 3.5
S4-3 70.5 51.0 498 2259 3.6
S4-4 70.5 50.9 498 2258 3.6
S4-5 70.2 51.0 498 2256 3.7
S4-6 70.5 51.0 498 2255 3.7
S4-7 70.5 50.9 498 2257 3.7
S4-8 70.3 51.2 498 2258 3.6
S4-9 70.6 51.4 498 2260 3.5
Table 6.1a: Specimen geometry and composition (part 1)

122 Chapter 6
Code h [mm] w [mm] l [mm] r [kg/m3] v%
S5-1 70.9 50.4 500 2252 3.9
S5-2 70.8 50.5 500 2265 3.3
S5-3 70.7 50.6 500 2266 3.3
S5-4 70.7 50.2 500 2268 3.2
S5-5 70.8 50.3 500 2269 3.2
S5-6 70.7 50.2 500 2268 3.2
S5-7 70.6 50.4 500 2265 3.3
S5-8 70.4 50.5 500 2262 3.5
S5-9 70.7 50.4 500 2257 3.6
S6-1 70.1 51.3 501 2258 3.6
S6-2 69.9 51.1 501 2257 3.7
S6-3 69.9 51.0 501 2256 3.7
S6-4 69.8 50.9 501 2254 3.8
S6-5 69.9 50.9 501 2255 3.8
S6-6 69.9 50.7 501 2255 3.8
S6-7 69.9 50.9 501 2257 3.7
S6-8 70.2 50.2 501 2258 3.6
S6-9 70.3 50.6 501 2261 3.5
S7-1 70.6 51.0 499 2241 4.4
S7-2 70.8 50.9 499 2256 3.7
S7-3 70.7 50.9 499 2257 3.7
S7-4 70.8 51.0 499 2259 3.6
S7-5 70.7 50.8 499 2260 3.6
S7-6 70.9 50.8 499 2259 3.6
S7-7 70.9 51.0 499 2257 3.7
S7-8 70.7 50.8 499 2255 3.7
S7-9 70.7 50.9 499 2255 3.8
S8-1 70.7 50.5 500 2251 3.9
S8-2 70.7 50.5 500 2259 3.6
S8-3 70.6 50.5 500 2256 3.7
S8-4 70.5 50.3 500 2255 3.7
S8-5 70.4 50.2 500 2256 3.7
S8-6 70.4 50.1 500 2256 3.7
S8-7 70.5 50.4 500 2258 3.6
S8-8 70.8 50.3 500 2258 3.6
S8-9 70.9 50.3 500 2255 3.8
Table 6.1b: Specimen geometry and composition (continued)

6.3.1.2 Specimen preparation


At least 12 hours prior to testing, the specimens were taken from the storage room and
placed in a wooden box filled with sand in a temperature-controlled cabinet that stood
directly beside the set-up. This cabinet was set to the test temperature and the
temperature in the storage cabinet and that in the temperature controlled cabinet around
the set-up were checked with a reference temperature sensor to make sure that the
temperatures were the same.

Before the specimens were placed in the cabinet, the loading plates were glued onto
them. These loading plates were slightly bended aluminum plates that kept the
pendulum bars in place on the specimen and prevented the rolls from indenting the
specimen and causing local failure. A simple plastic strip marked with the positions of
the supports at top and bottom of the specimen was used to mark the appropriate
positions on the specimen. A fairly brittle two-component glue (X-60) was used to glue

The multiaxial test 123


the plates on the specimen. This glue was selected because it can relatively easily be
removed from the plates after the tests, so the plates can be re-used.

The specimens that would also be loaded in the horizontal direction were outfitted with
loading plates on their end surfaces as well. These plates were made from 5 mm thick
stainless steel instead of 2mm of aluminum, since they had to spread the load over a
much larger area. These loading plates were also glued to the specimen with X-60 and
recovered after the test. In Figure 6.8 both the loading plates for the vertical and
horizontal forces are shown. Underneath it, in Figure 6.9 a specimen with the loading
plates in place is shown.

Figure 6.8: The loading plates used to spread the vertical (curved aluminium plates)
and horizontal forces (flat steel plates) respectively

Figure 6.9: Specimen out-fitted with loading plates (picture taken after the test)

6.3.2 ACRe FOUR-POINT SHEAR TEST SET-UP IN DETAIL


The four-point shear set-up is built in a 3D space frame that is constructed on an
elastically supported concrete block. A series of steel profiles are mounted on this
block. The steel bottom plate (2400x400x30mm) for the set-up is connected to these
profiles. The horizontal actuator is connected to the bottom plate on the right hand side.
Around the other side of the bottom plate a portal of steel profiles is constructed, with a
cross connection at approximately two-thirds of its height. The vertical actuator is
connected to the top of this portal and the cross-connection is used to guide the actuator
and to connect the top of the insulated cabinet, which allows temperature controlled
testing. A schematic representation of the set-up is shown in Figure 6.10.

124 Chapter 6
Portal

Temperature Cabinet

insulated hoses, part of the


temperature controll system

Actuator Bottom plate


Support block

Loadc

Steel profiles Concrete block

Figure 6.10: Schematic drawing of the shear set-up


Inside the temperature cabinet a loading block, a steel HE 220 B profile of 500 mm in
length is connected to the vertical actuator via a hinge. This hinge ensures the
appropriate distribution of the forces, as a function of the respective distances of the
loading frames (pendulum bars plus frames) to the actuator (Section 6.2 and Figure 6.2).
The loading frames are connected to the bottom of the loading block. Two similar
loading frames are connected to the bottom plate. The pendulum bars rotate around an
axis at the connection with either the loading block or the bottom plate. They do not
rotate with respect to the frames, but the specimen can because it is placed on rolls in
the frame. In Figure 6.11 a picture of the inside of the set-up is shown.

Loading block

Pendulum bars

Pendulum bars
Horizontal frames
support

Figure 6.11: Specimen in the adapted four-point shear set-up

The multiaxial test 125


An MTS 111 kN (25000 pound) hydraulic actuator (MTS model 204.25, serial nr. 662)
is mounted in the vertical portal. The force that is applied with this actuator is measured
with a 100 kN LeBow loadcell (model 3116-106, serial number 3576) that can be used
in different ranges if an appropriate cartridge is used in the MTS test controller. On the
bottom plate a 25 kN MTS actuator (model 204.52, serial nr. 467) is mounted
horizontally. The force applied with this actuator is measured via an MTS 50 kN
loadcell (model 661.20 B-02, serial nr. 553). The test is performed with a displacement-
controlled vertical signal that is generated by a programmable function generator (MTS
microprofiler model 418.91). The vertical actuator is controlled through an MTS 258.20
Micro Console, using a 458.13 AC controller for the displacement transducer and a
458.11 DC controller for the force. The strain rates are similar to those used in tension
and compression, but since those rates were related to axial deformation over the whole
specimen while here the strain is related to the vertical deformation over the narrow
shear zone, the corresponding deformation rates are much smaller. As a result, no
special high-response valves were used in this system. The horizontal signal that is
applied in some of the tests, is a constant force. The horizontal actuator is controlled
through a 458.30 Station Control unit, the controllers used for the displacement and
force are the same types as for the vertical actuator. These controllers allow the
selection of the appropriate force and displacement ranges for the test conditions by
inserting a calibrated cartridge for that system. For the vertical displacement a ± 50mm
cartridge was used, the horizontal displacement was measured with an external
Solartron displacement transducer with a range of ± 10 mm, the force range for the
horizontal actuator was ± 10 kN and for the vertical force two cartridges were used: ±
50 kN for the tests at low temperatures and ± 20 kN for those at higher temperatures.

An insulated cabinet with dimensions 1000x700x400 (hxlxw) is placed around the


specimen. It is a sandwich structure of wood and roofmate. A PT 100 temperature sensor
that is connected to a temperature control unit that maintains the temperature with an
accuracy of ± 1 °C is used to control the temperature in the set-up. The unit circulates air
in the temperature cabinet between an input channel and an output channel. With simple
valves the ratio in airflow can be varied. The unit is capable to maintain the temperature
in a range of -5 °C till 35°C.

6.3.3 TEST PROCEDURES AND CONDITIONS

6.3.3.1 Test procedures


The specimen is only locally supported after it is placed in the set-up, so it can not be
left like that to regain its temperature or for elaborate instrumentation, because that
would lead to (plastic) deformations and damage. Since this test was only meant to
provide information about the changes in strength as function of the applied
confinement, instrumentation could be kept simple. Next to the force and displacement
information of the actuators only a single external displacement transducer was used to
provide information about the movements of the horizontal actuator. This transducer
was mounted in a frame and placed in contact with the load cell of the actuator. As a
result, no instrumentation had to be placed on the specimen. By placing the storage
cabinet next to the set-up, the specimen could be moved from the storage to the set-up
quickly, minimising the changes in temperature.

Two of the four supports are placed above the specimen and due to the gravity these
supports are active only if a load is applied on the specimen. Especially at higher

126 Chapter 6
temperatures, the lack of support at these places may lead to deformations of the
specimen. To prevent this, the frames at these positions were fitted with spring-plates at
the bottom part. These spring-plates, thin aluminium plates connected by two springs,
supported the specimen at those points prior to testing. Once the test is running, they
hardly influence the test because springs have an axial stiffness but they are very
flexible in the transverse direction. This system worked very well and after placing it
the specimen is resting on the bottom supports and the spring-plates.

Axial stiffness:
spring supports specimen

plates

springs

bolt

Transverse flexibility:
spring does not effect test

Spring plate
Figure 6.12: Plates supported by springs are used to support the specimen prior to
testing at those places where the supports are placed above the specimen
Once the specimen was placed in the set-up a small pre-load was applied and the safety
pins (used to immobilize the pendulum bars) were removed. If confinement was used, as
a final step the appropriate horizontal force was applied to the specimen and the test was
started.

6.3.3.2 Test conditions


As stated before, the conditions of the test were chosen in line with those for the tension
and compression tests. As a result, the same temperatures (0, 15 and 30 oC) were used
and strain rates that corresponded to those used in the tension test. In the previous tests,
the strain applied was a normal strain while in this test it is a shear strain. Using the
relation shown in Figure 6.13, the shear strain rate (dγxy/dt) is kept equal to the strain
rates used in the compression and tension tests (dεxx/dt). The strain rate in the shear tests
is then found by dividing the displacement rate of the vertical actuator by the width of
the shear zone (17 mm). This is the so-called “engineering shear strain” as opposed to
another definition of shear strain, where εxy=γxy/2. In this case the engineering definition
was used because of the analogy with the axial case, where σ=Eε and in shear it is
τ=Gγ, where E is the modulus of elasticity and, σ the normal stress, τ the shear stress
and G the shear modulus.

The multiaxial test 127


ε =[ ε xx ε yy ε zz γ xy γyz γxz ]

ε xx = ∆ x/x o ; εyy = ∆ y/y o γxy = ∆ y/x o + ∆ x/y o


x x ∆y

∆ y/2 ∆ y/2 ∆x

∆x x0 y

y0

x0
y0
y ∆ x*
y
y0
γ xy = ∆x* /y
0

Figure 6.13: Shear strain (rate) compared to strain in tension and compression
For each temperature, two strain rates were selected. The tests with horizontal force
were performed under the same conditions (temperature and strain rate), but with a
horizontal confinement equal to 15 and 40% of the maximum applied vertical force
during the test with zero horizontal confinement (Table 6.2). For each condition, three
repetitions were carried out.

T [oC] Strain rate [x10-2/s] Confinement [kN]


30 1 0
30 1 -0.3
30 1 -0.8
30 10 0
30 10 -0.6
30 10 -1.7
15 0.1 0
15 0.1 -0.7
15 0.1 -1.7
15 1 0
15 5 0
15 5 -1.8
15 5 -4.6
0 0.01 0
0 0.01 -1.3
0 0.01 -3.5
0 1 0
0 1 -3.3
0 1 -8.8
Table 6.2: Test conditions in the shear test

128 Chapter 6
6.3.3.3 Force or displacement control?
It must be noted that the strain rates mentioned in the previous section are related to the
displacement rate of the actuator. Because of the nature of the test the applied actuator
load is applied to the specimen via the two supports in a constant ratio (Section 6.2.1).
As a result, although the actuator is controlled via its displacement the specimen is
subjected to something half way between displacement and load control. The
distribution of the actuator load over the supports is known and constant, that of the
actuator displacement varies to ensure the proper load distribution. This means that the
movement of the individual supports cannot be controlled. As a result, this test cannot
really be performed at a given strain rate. The actuator rate can be specified, but the
strain rate in the shear zone must be determined afterwards.

In the current configuration this cannot be done, because only the actuator displacement
is measured. For future test series the instrumentation must be augmented to account for
this effect. The easiest way to do this is by instrumenting the loading block. Because
this is rigid and the distances between the supports and the hinge between block and
actuator are known, there is a direct relation between the displacement of the block, its
rotation and the displacement of the individual supports (Figure 6.14). Automatically
measuring the rotation of the block will therefore provide the additional information
needed to determine the shear strain rate.

To assess the strain rates applied in the current test, the relation between the movements
of the two supports was determined on the basis of the stress distribution in the elastic
stage of response. Due to pure bending the ratio between movement of the outer and
inner support would be approximately: uo:ui = 100:1, while shear yields a relation of
u’o:u’i = 1.6:1. The two can be combined by assuming a value for the Poisson’s ratio,
however it was felt that this would lead to an over-estimation of the difference because
of the large difference in bending. In reality, shear is predominant in this test but
bending is also present. For this reason a Finite Element analysis of the test was carried
out and this yielded a relation between the support movements of u”o:ui” = 1.9:1. This
means that initially the deformation rate in the shear zone is approximately 90% of the
actuator displacement rate. This value is used as the shear strain rate in the analysis of
the test results.

ui:uo ui uo
1:1.6 0.93u 1.48u
a1 a2 1:2 0.88u 1.76u
1:4 0.71u 2.86u
u
1:10 0.45u 4.54u

ui=u*(1-α*a2)
uo=u*(1-α*a1)
Figure 6.14: The relation between the support displacements depends on the rotation
of the loading block and the distances between supports and actuator

The multiaxial test 129


6.3.4 TEST RESULTS
During the tests, the applied vertical and horizontal load and the corresponding actuator
movements were measured. On the basis of the strength found without horizontal
confinement the confinement levels for that combination of strain rate and temperature
were determined, which were as stated previously, 15 and 40% of the maximum vertical
force applied in the unconfined shear test. The results are shown in Figure 6.15 through
Figure 6.20. In legend of the graphs, 0%, 15% and 40% refers to the applied horizontal
confinement. In the graphs the vertically applied force and, for the tests with horizontal
confinement, the horizontal deformation are plotted as a function of time. The results
are plotted as a function of time rather than shear strain because of the uncertainties in
the shear strain that were discussed in the previous section.
5
Fv [kN] 30oC & 10 %/s
4.5

4
Applied vertical force versus time
3.5

2.5

1.5

1
40%
15%
0.5
0% time [s]
0
0 2 4 6 8 10 12 14
o
Figure 6.15: Average response curves for three levels of confinement (30 C& 10%/s)
Fv [kN]
o
30 C & 1 %/s uh [mm]
2.5

2 0.1

1.5
0%
1 0.05
vertical load 40%

0.5
15% time [s]
0 0
0 10 20 30 40 50 60 70 80
-0.5
15%
-1 -0.05
horizontal actuator displacement
40%
-1.5

Figure 6.16: Average response curves for three levels of confinement (30oC & 1%/s)

130 Chapter 6
5 0.25
Fv [kN] o
15 C & 0.1 %/s uh [mm]

4 0.2

3 0.15
40%

2 0.1

vertical load 0% 15%


1 0.05

time [s]
0 0
0 100 200 300 400 500 600 700 800
-1 -0.05
15%
horizontal actuator displacement
-2 -0.1
40%
-3 -0.15

Figure 6.17: Average response curves for three levels of confinement at 15oC and
0.1%/s (actuator rate, actual strain rate approximately0.09%/s, Section 6.3.3.3)

o
15 C & 5 %/s
19 0.19
Fv [kN] 15% uh [mm]
17 0.17

15 0.15

13 0.13
vertical load
11 0.11

9 0.09

7 40% 0.07
40%
5 horizontal actuator 0.05
displacement
3 15% 0.03

1 0% 0.01
time [s]
-1 0 2 4 6 8 10 12 14 16 -0.01

Figure 6.18: Average response curves for three levels of confinement (15oC & 5%/s)

The multiaxial test 131


12 0.45
Fv [kN] o
0 C & 0.01 %/s uh [mm]

vertical load
10 0.35

8 0.25

6 0.15
15%
40%
4 horizontal actuator 0.05
displacement

15%
2 -0.05
0%
40%
time [s]
0 -0.15
0 1000 2000 3000 4000 5000 6000 7000 8000
Figure 6.19: Average response curves for three levels of confinement at 0oC and
0.01%/s (actuator rate, actual strain rate approximately 0.009%/s, Section 6.3.3.3)
35 0.07
Fv [kN] o
0 C & 1 %/s uh [mm]
30 0.06

25 0.05

vertical load
20 0.04

15 0.03

10 0.02

40%
5
0% 15% 0.01

time [s]
0 0
15%
0 10 20 30 40 50 60
-5 -0.01
horizontal actuator displacement
40%
-10 -0.02
o
Figure 6.20: Average response curves for three levels of confinement at 0 C and 1%/s
From the response curves it can be seen that asphalt concrete is indeed confinement
sensitive. As expected, the strength of the material increases with the applied
confinement. Whether this confinement sensitivity is temperature and strain rate
sensitive is investigated in the next section.
From Table 6.3 it can be seen that for a few conditions the third test was not completed.
In these cases a test went wrong (computer malfunction, a support (pendulum bar) that
did not stay in place) and because of a lack of specimens they could not be repeated.
This was not considered a problem because of the availability of two results and the
satisfactory repeatability.

132 Chapter 6
o 2 2
code T [ C] γ [%/s] Fh[kN] σ h [N/mm ] % support τav [N/mm ]
prfstk:S8-6 0 0.01 0.0 0.0 0 2.1
prfstk:S4-7 0 0.01 0.0 0.0 0 1.8
prfstk:S3-8 0 0.01 0.0 0.0 0 2.0
prfstk:S1-4 0 0.01 -1.3 -0.4 19 2.3
prfstk:S7-8 0 0.01 -1.3 -0.4 19 1.9
prfstk:S4-1 0 0.01 -2.1 -0.6 30 2.6
prfstk:S8-8 0 0.01 -3.4 -1.0 50 2.8
prfstk:S2-8 0 0.01 -3.4 -1.0 50 2.6
prfstk:S3-2 0 0.01 -3.5 -1.0 50 2.7
prfstk:S3-9 0 1 0.0 0.0 0 5.7
prfstk:S8-9 0 1 0.0 0.0 0 5.9
prfstk:S7-6 0 1 -3.1 -0.9 15 6.9
prfstk:S1-3 0 1 -3.3 -0.9 16 5.8
prfstk:S5-4 0 1 -3.3 -0.9 16 5.9
prfstk:S1-7 0 1 -8.8 -2.5 44 8.6
prfstk:S4-2 0 1 -8.8 -2.5 44 8.4
prfstk:S8-2 0 1 -8.8 -2.5 44 8.5
prfstk:S4-9 15 0.1 0.0 0.0 0 1.1
prfstk:S7-1 15 0.1 0.0 0.0 0 1.1
prfstk:S3-4 15 0.1 0.0 0.0 0 1.0
prfstk:S6-8 15 0.1 -0.7 -0.2 18 1.0
prfstk:S7-7 15 0.1 -0.7 -0.2 18 1.2
prfstk:S2-6 15 0.1 -0.7 -0.2 19 1.1
prfstk:S7-9 15 0.1 -1.7 -0.5 47 1.2
prfstk:S2-9 15 0.1 -1.7 -0.5 47 1.1
prfstk:S6-9 15 0.1 -1.7 -0.5 47 1.1
prfstk:Sx-5 15 5 0.0 0.0 0 3.0
prfstk:S8-4 15 5 0.0 0.0 0 2.9
prfstk:S1-6 15 5 -1.8 -0.5 17 3.6
prfstk:S8-3 15 5 -1.8 -0.5 17 2.9
prfstk:S7-5 15 5 -1.8 -0.5 17 3.5
prfstk:S8-1 15 5 -4.6 -1.3 44 3.3
prfstk:S6-4 15 5 -4.6 -1.3 44 3.4
prfstk:S5-6 15 5 -4.6 -1.3 44 3.5
prfstk:S4-4 30 1 0.0 0.0 0 0.5
prfstk:S7-2 30 1 0.0 0.0 0 0.6
prfstk:S3-7 30 1 0.0 0.0 0 0.5
prfstk:S1-2 30 1 -0.3 -0.1 16 0.7
prfstk:S6-6 30 1 -0.3 -0.1 16 0.6
prfstk:S5-7 30 1 -0.3 -0.1 16 0.8
prfstk:S2-4 30 1 -0.8 -0.2 40 0.7
prfstk:S8-5 30 1 -0.8 -0.2 41 0.7
prfstk:S1-8 30 1 -0.8 -0.2 41 0.7
prfstk:S7-3 30 10 0.0 0.0 0 1.1
prfstk:S4-3 30 10 0.0 0.0 0 1.0
prfstk:S3-5 30 10 0.0 0.0 0 1.1
prfstk:S2-3 30 10 -0.6 -0.2 17 1.2
prfstk:S6-7 30 10 -0.6 -0.2 17 1.2
prfstk:S5-8 30 10 -0.7 -0.2 17 1.4
prfstk:S1-9 30 10 -1.7 -0.5 44 1.4
prfstk:S6-3 30 10 -1.7 -0.5 44 1.3
prfstk:S2-2 30 10 -1.7 -0.5 44 1.3
Table 6.3: Individual test results (NB: γ refers to the actuator rate, Section 6.3.3.3 and
τav is the average shear stress (shear force divided by specimen cross section))

The multiaxial test 133


6.3.5 GENERAL RELATIONS FOR THE TEST RESULTS
A similar trend as observed for the tensile and compressive strength as functions of
temperature and strain rate is apparent in the shear strength, despite the different notions
(elongation versus shear) that underlie the tests. The general relation that was used to
describe all strength-related properties (Equation (6.1)) also enabled the expression of
the shear strength as a function of temperature and strain rate. In Figure 6.21 this
relation is plotted along with the test results. The shear stress used in this graph is the
average shear stress, which is the shear in the centre zone divided by the specimen cross
section. The shear strain rate is based on the engineering strain as shown in Section
6.3.3.2, Figure 6.13. The strain rate values are based on 90% of the applied actuator
deformation rate for the reasons mentioned in Section 6.3.3.3.

   
   
   
 1   1 
τ 0 = a 1 −  = 17.5 1 −  (6.1)
   •  b+ c   d      •  −93+ 26200   0.31  
 1 + γ e T      1 + γ e T 
 
      
      

      
     
Where: τ0=shear strength without confinement in N/mm2

γ =shear strain rate in s-1
T= temperature in Kelvin
a,b,c,d= regression constants

15
Eq.6.1 Eq.6.1 Eq.6.1
t0 [N/mm2] 0 15 30
12.5

10

7.5

2.5
γ [%/s]
0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15

Figure 6.21: General expression for the shear strength compared to the test results
In Figure 6.22 the shear strength values normalised with respect to the shear strength
without confinement are plotted against the confinement level. The confinement levels
plotted in this graph are determined, for every combination of temperature and strain
rate, by dividing the applied normal stress by the shear stress at failure if no confinement
is applied. These confinement levels are somewhat higher than the 15% and 40%
mentioned before because the shear in the failure zone is not equal to the applied

134 Chapter 6
actuator load (Section 6.2.1) and the levels mentioned earlier were based on the ratio of
the applied horizontal and vertical actuator forces. Since the data are used to develop a
relation between the shear strength and the confinement level, which is a material
characteristic, it must be represented in a test set-up independent way. Hence the
adapted definition of the confinement level, which now uses the stresses as they occur in
the specimen. The earlier definition is more convenient during testing since it is based
on quantities that are actually measured and used to control the test.

The increase in shear strength with confinement appears to be independent of


temperature and strain rate (Figure 6.22). For this reason a relation between shear
strength and confinement level is developed. This relation has to yield the unconfined
shear strength for zero confinement and it should go to a limit strength, since the shear
strength will not increase indefinitely with increasing confinement. These considerations
resulted in Equation 6.2, which is also plotted through the data points in Figure 6.22. As
can be seen it describes the trend rather well. This relation basically expresses the
confinement sensitivity of the ACRe mixture.

  σN
 − b* x100%   
   σN
 −0.005* x100%   

  τ     τ  
τ = τ 0  a − (a − 1) exp  0  =τ
 0  2.5 − 1.5exp
 0
 (6.2)
   
   
Where: τ=shear strength at that confinement level
τ0=shear strength at zero confinement
σN=applied confinement (normal stress)
a,b=regression constants

2
τ/τ0
1.8

1.6

1.4
0.01&0
1.2
1&0
1 0.1&15

0.8 5&15
1&30
0.6
10&30
0.4 Eq. 6.2
0.2
confinement [σN/t0x100]
0
0 10 20 30 40 50 60 70 80 90 100

Figure 6.22: Normalised shear stress as a function of the applied confinement

The multiaxial test 135


6.4 CONCLUSIONS AND RECOMMENDATIONS

6.4.1 CONCLUSIONS
The four-point shear test was developed to obtain additional information on the
confinement sensitivity of the material. It turned out that, for the mixture under
investigation, this confinement sensitivity was independent of temperature and strain
rate. For the ACRe mixture, the test did not really provide new information on the
confinement sensitivity, the results matched quite well with the predicted ones on the
basis of the uniaxial tests. On the basis of these results it could be shown that for this
mixture, the model parameter β=0.

6.4.2 RECOMMENDATIONS
Because of time restraints the adaptations on the existing four-point shear test were
limited to the components inside the frame, leaving the frame structure itself, the
loading block and the bottom plate intact. If these are adapted, the required stress
distribution can also be achieved using much smaller specimens. This would make
specimen handling much easier and, because of lower specimen production costs, it
would eventually pay for itself.
Although the test did not result in additional information for the ACRe mixture,
it is expected that it will do that for open mixtures. For this reason it is recommended to
perform this test as well as the uniaxial tests. If it turns out that the information on
confinement sensitivity obtained from this test does not provide enough data for the
determination of model parameter β, triaxial tests can be used to get additional
information.

6.5 REFERENCES
Bondt, A.H. and Scarpas, A., (1993), Shear Interface Test Set-Ups, Delft University of
Technology Report nr.7-93-203-12
Bondt, A.H., (1999), Anti-Reflective Cracking Design of (Reinforced) Asphaltic
Overlays, PhD. Thesis Delft University of Technology, Ponsen & Looijen, ISBN
90-6464-097-1
Erkens, S.M.J.G., Poot, M.R. and Moraal, J., (2002), MultiAxial Testing with the Four-
point Shear Test – Asphalt Concrete Response (ACRe), Delft University of
Technology Report nr. 7-02-117-7

136 Chapter 6
7. THE ACRE MATERIAL MODEL
“Mechanics is the paradise of mathematical science, because here we come to the
fruits of mathematics”
Leonardo da Vinci (1452-1519)

7.1 THE MODEL


A material model describes the response of a material to any state of stress. An example
of a simple material model is the elastic-ideal plastic model (Figure 7.1). In this case the
material behaves linearly elastic until its yield strength is reached and from there on it
behaves ideally plastic. The latter behaviour is characterised by increasing deformations
without additional loading (Figure 7.1, left). In this case, the material model consists of
two constitutive relations, one for the elastic and one for plastic response, and a single,
static limit surface. The limit surface separates elastic from inelastic response and is the
collection of all the stress combinations that will cause the transition from one type of
response to the other. The constitutive relations describe the strains as a function of the
stresses for both types of response.

σ σ2 σ2

fy
σ1 σ1

σ3
ε

Constitutive relations: Inside the ellipse: Inside the cylinder:


σ<fy: 1) σ=Eε relation 1) relation 1)
σ=fy: 2) σ=E(ε−εp) On the ellipse: On the cylinder:
ε& p = λ& ∂f / ∂σ relation 2) relation 2)

Figure 7.1: A material model consists of at least a limit and constitutive relations
If the material behaviour is more complex, as it is for asphalt concrete, the size and
shape of the limit surface will also change throughout the range of response. In that
case, the relations that control the surface are also part of the model (Figure 7.2). The
ACRe material model, which is based on the theory of rate dependent consistent
plasticity and provides a realistic, unified, phenomenological approach for materials that
exhibit strain rate dependent, inelastic deformations, works like this. The model is based
on the limit surface developed by Desai (1986) which is discussed in the next section.
The constitutive relations used in the material model are presented in Section 7.1.3 and
the procedures used to determine the model parameters on the basis of the tests
described in the previous chapters are shown in Section 7.2.

The ACRe material model 137


σ2
fy 3
2
σ1
1
4
4
1
2
3
ε
Figure 7.2: A more complex material model with changing limit surfaces

7.1.1 THE FLOW SURFACE


The material model developed in the ACRe project is based on the flow surface that was
proposed by Desai and co-workers. To provide a unified description of the behaviour of
soils they developed the Hierarchical Single Surface (HISS) models. This approach uses
a basic model for an isotropic material that exhibits isotropic hardening and associative
plasticity. When necessary, more complex behaviour (e.g. non-associative response,
kinematic hardening) can be super-imposed on the basic model. As part of the HISS
approach a single surface yield criterion was developed as an alternative for the use of
two intersecting surfaces. Starting from a complete polynomial in the invariants of the
stress matrix (Desai 1980 and 1984) eventually a truncated form for the single surface
yield function was proposed (Desai et al. 1986 and Frantziskonis et al. 1986):

  I − R n  I1 − R  
2
 −α  1
 +γ   
J   p   p  
f = 22 − 
a a
=0 (7.1)
pa (1 − β cos(3θ ) )
Where:
I1 = σ xx + σ yy + σ zz = the first stress invariant
σ xx + σ yy + σ zz
p= = the isotropic stress
3
J2 =
2
(
1 2 2 2 1
6
)
s1 + s2 + s3 = (σ 1 − σ 2 ) + (σ 1 − σ 3 ) + (σ 2 − σ 3 )  =
2 2 2

1
( ) ( )
σ xx − σ yy + σ yy − σ zz + (σ zz − σ xx )  + τ xy
2 2 2 2 2 2
+ τ yz + τ xz
6  
= the second deviatoric stress invariant
( )
J 3 = (σ xx − p ) σ yy − p (σ zz − p ) + 2τ xyτ yzτ xz − (σ xx − p )τ yz
2 2
− (σ yy − p )τ xz 2
− (σ zz − p )τ xy
= the third deviatoric stress invariant

138 Chapter 7
3 3 J3
cos(3θ ) = 3
2
( J2 )2
si = i - th principle deviator stress
pa = -1⋅105 Pa = -0.1 MPa= atmospheric pressure
(throughout this thesis mechanics notation, where tension is positive
and compression negative, is used)
α ,β ,γ ,n,R= model parameters, depending on material characteristics

Equation 7.1 represents the complete closed surface in the (I1, √J2, θ) space, Figure 7.3,
hence the use of the term single surface model.

Figure 7.3: Schematic of Desai response surface (I1, √J2, θ space, tension negative)
This particular flow surface was selected as the basis of the ACRe material model for
several reasons. It can be used to describe the response of all the materials typically
encountered in pavement profiles, which effectively reduces the number of models that
had to be incorporated in the Finite Element package CAPA-3D that was specifically
developed for the analysis of pavement structures (Scarpas et al. 1997). Furthermore,
the fact that the expression is that of a closed surface means that there is no need for cap
surfaces along the I1 axis. Finally, the model parameters ensure that the size, shape and
position of the limit surface are very flexible, allowing it to be used for a variety of
materials. Also, the parameters themselves are related to physical phenomena and can
be determined directly on the basis of laboratory experiments.

7.1.2 MODEL PARAMETERS


The model parameters α, β, γ, n and R each govern a specific aspect of the model. In
this section the role of each parameter will be discussed briefly. Throughout this section
the model is plotted in the I1-√J2 space, which allows the 3D model to be represented in
2D.

7.1.2.1 Influence of α
The model parameter α determines the size of the flow surface, the size increases with
decreasing α, so this parameter controls the hardening of the material. For elastic states

The ACRe material model 139


of stress it retains its original value, but as soon as non-linear states of stress occur
α decreases until α=0 at peak stress, when the hardening stops (Figure 7.4).

σ [N/mm2]
-6

-5

-4

-3

-2

-1

0
0 -0.05 -0.1 -0.15 -0.2
α=C 0≤α≤C α=0 ε [m/m]

Figure 7.4: The model parameter α controls the hardening response


The influence of α on the size of the flow surface is shown in Figure 7.5.

a0=0.01 a0=0.005 a0=0.001 a0=0 sqrt(J2)


120

100

80
α0
60

40

20
I1
0
-275 -245 -215 -185 -155 -125 -95 -65 -35 -5
Figure 7.5: Influence of α on the shape of the flow surface, plotted in I1-√J2 space

At peak stress, for α=0, the surface reduces to a straight line at the I1-√J2 surface. At this
point it can be compared to the well-known Mohr-Coulomb flow surface, although the
cross-section of the Mohr-Coulomb flow surface is six-facetted rather than smooth, or
to the smooth Drucker-Prager surface (Figure 7.6).

140 Chapter 7
σ
2 σ1=σ2=σ3
1) Mohr-Coulomb
2) Drucker-Prager
3) von Mises

σ1
2

3
1

σ3
Figure 7.6: For α=0 the Desai flow surface reduces to the Drucker-Prager surface,
which is similar to Mohr-Coulomb flow surface, but with smooth cross-section

7.1.2.2 Influence of γ
The model parameter γ determines the slope of the (ultimate) surface, which increases
with increasing γ (Figure 7.7). It is stress-state independent, but it can vary as a
function of, for example, temperature and loading rate.

g=0.1 g=0.2 g=0.4


160
sqrt(J2)
140

120
γ 100

80

60

40

20
I1
0
-250 -200 -150 -100 -50 0 50

Figure 7.7: Influence of γ on the shape of the flow surface, plotted in I1-√J2 space
After the peak load, when a=0, degradation of g can be used to obtain an overall
reduction in strength via isotropic softening. In this case g will be expressed as some
decreasing function of the non-linear strains.

The ACRe material model 141


7.1.2.3 Influence of n
The model parameter n determines the apex of the surface, it expresses the state of
stress after which the material starts to dilate. The apex is defined as that point of the
flow surface where the tangent is a horizontal line (∂f/∂I1=0, see Figure 7.8), indicating
a fully deviatoric state of stress. How this condition can be used to determine n, will be
discussed later on. Changes in the value of n do not only influence the shape, but also
the size of the surface. This is shown in Figure 7.8.
n=2 n=2.5 n=2.8 n=3 sqrt(J2)
90

80

70

60
n
50

tangent 40

30

20
apex
10

0
-185 -155 -125 -95 -65 -35 -5 I1

Figure 7.8 Influence of n on the shape of the flow surface, plotted in I1-√J2 space

7.1.2.4 Influence of β
The model parameter beta determines the shape of the model on the octahedral-plane.
For β=0 it is circular and with increasing β it becomes triangular. Since cos(3θ) is 1 for
uniaxial states of stress, the square root term in Equation 7.1 reduces to: √(1-β). The
effect of uniaxiality on θ is:

2  2 4 
cos(3θ ) = 1 ⇒ θ = 0 + k π ⇒ θ =  0, π or π  +2kπ , with k=1,2,3.. (7.2)
3  3 3 

These values of θ correspond to the position of the σ1, σ2 and σ3 axes on the octahedral-
plane, indicating that uniaxial test results are related to a state of stress where two
principal stresses are zero and the third is not. Whether or not the shape of the cross
section is triangular can be seen when results for different θ values at the same I1 are
available (Figure 7.9 and Figure 7.10). As long as that information is unavailable the
behaviour is assumed to be independent of θ, setting the cross section on the octahedral-
plane to a circle (β=0).

142 Chapter 7
25
sqrt(J2)

20

cos(3q)
15

10

I1
0
-175 -160 -145 -130 -115 -100 -85 -70 -55 -40 -25 -10 5
Figure 7.9: Influence of θ on the surface plotted in the I1-√J2 and π-plane (for β=0)

0
340350 2.5 10 20 β
330 30
320 2
40
310 50
1.5

300 60
1
290 70
280 0.5
80
270 0 90
260 100
250 110
240 120
230 130
220 140
210 150
200190 170160
180
Figure 7.10: Influence of β on the flow surface on the octahedral plane for a given I1

7.1.2.5 Influence of R
R is the three-dimensional tensile strength, which is an indication of cohesion. For
cohesionless materials R=0, which means that only states of stress for which I1≤0 are
possible within that material. For increasing R values, the flow surface moves in the
direction of the positive I1 axis (Figure 7.11).

The ACRe material model 143


R=0 R=6 R=12 sqrt(j2)
25

20
R

15

10

I1
0
-185 -170 -155 -140 -125 -110 -95 -80 -65 -50 -35 -20 -5 10

Figure 7.11: Influence of R on the shape of the flow surface in the I1-√J2 space (both
the R values in the legends and the axes are given in N/mm2)
Strain rate and temperature sensitivity can be incorporated into the classical Desai
surface by specifying the model parameters as functions of strain rate and temperature.
Since the parameters are determined from tests at different temperatures and strain rates
(see next section), the model parameters are automatically obtained for various
conditions.

7.1.3 CONSTITUTIVE FRAMEWORK


Within the ACRe project this flow surface is combined with constitutive relations that
describe elastic, visco-plastic and fracturing response. Thus, the whole range of
response observed in pavement materials can be described. From hereon the term
ACRe-model will be used to refer to this combination of flow surface and constitutive
relations. In this section, the constitutive relations are described.

The theory of rate dependent consistent plasticity is used as the constitutive framework
for the model (Scarpas et al. 1998). The plastic strain rate is defined as:
∂f
ε& p = λ& ⋅ (7.3)
∂σ
with λ& a constant of proportionality1, and f a response surface associated with a locus
of states of stress corresponding to a certain magnitude of inelastic response. Two main
phases of material response are distinguished by the formulation, Figure 7.12:
(a) “hardening”, spanning the range from linear-elastic to ultimate (peak) response and,
(b) “softening”, spanning from ultimate response to response annihilation.

The standard Kuhn-Tucker conditions are imposed:

1
For Perzyna type viscoplasticity λ& = Γ ⋅ Φ with Γ the material fluidity and Φ an overstress
function.

144 Chapter 7
λ& ≥ 0 , f ≤ 0 , λ& ⋅ f = 0

The evolution of plastic flow is determined by the consistency condition:


f& (σ , ε&, T , κ ) = 0 (7.4)
in which ε& is the strain rate, T is the temperature and κ is some measure of
hardening/softening.
σ

Response degradation
phase

hardening
phase

ε
Figure 7.12: Schematic of main phases of material model response

7.1.3.1 Hardening response


The flow surface described in Section 7.1.1 is used to simulate the hardening response.
As described in Section 7.1.2.1 the hardening response of the material is controlled by
the model parameter α . As α decreases, the size of the flow surface increases, Figure
7.5. By defining α as a decreasing function of plastic strain or plastic work, the
characteristics of the hardening response observed in the experiments can be
incorporated in the model.

7.1.3.2 Response degradation


For α = 0 the ultimate response of the material is attained. In the proposed model, for
deformation levels beyond those corresponding to α = 0 two independent mechanisms
are activated to control the subsequent response. The need for two independent
mechanisms arises because in compression and tension different response degradation
mechanisms are activated. Tensile damage does not weaken the compressive strength
after stress reversal, while compressive loading does weaken the tensile strength of the
material. As a result two types of response degradation can be observed, isotropic
softening due to compressive loads and another degradation mechanism acting on
cracking planes as a result of tensile loads (Erkens et al. 2000). In typical road
engineering applications the material is alternatively loaded in tension and compression,
so for an accurate representation of the damage development these different
mechanisms are of importance. The mechanisms incorporated in the model to describe
these different phenomena are discussed in the next sections.

7.1.3.2.1 Isotropic softening via γ degradation


An isotropic measure of response degradation has been implemented by specifying γ ,
after the peak load, as a decaying function of some monotonically increasing physical

The ACRe material model 145


quantity (e.g. equivalent post fracture plastic strain, post fracture plastic work etc.), the
deformation rate δ& and the temperature T (Figure 7.13).

√J2

I1
Figure 7.13: For α=0, γ reduction controls isotropic degradation of the Desai surface
Since isotropic softening is meant to simulate the overall material degradation that is
observed experimentally as a result of compressive loading γ softening is introduced
for compressive stress paths only. After response degradation is initiated the principal
values of the plastic strain vector are computed. An equivalent plastic strain measure is
constructed:
ε p,c = ∫ dε p ,c (7.5)
consisting only of the increments of compressive principal plastic strain components:
d ε p ,c = ( d ε I ⋅ d ε I )
12
: dε I < 0 (7.6)
Isotropic response degradation due to the development of compressive principal strains
is obtained by specifying:
(
γ = γ δ&, T , ε p ,c ) (7.7)
In the model, an additional softening as described in the next section complements
isotropic softening.

7.1.3.2.2 Cracking
The tensile softening response is described along the lines of the classical notion of
fixed cracking. For states of stress equal to the tensile strength, a plane of cracking is
introduced perpendicular to the principal tensile stress direction if the following
condition is satisfied:
ε n ≥ ftu E (7.8)
in which εn is the strain normal to the plane of potential cracking and ftu the tensile
strength at crack initiation.

146 Chapter 7
σ

τ τ
s t

Figure 7.14: Schematic of Hoffman surface on crack plane


On the crack plane, a Hoffman type criterion, Figure 7.14 (Scarpas et al. 1998), is
specified to control the subsequent softening response:

( )
σ 2 + q ⋅ τ s2 + τ t2 = ft 2 (δ&, T , κ ) (7.9)

in which σ is the normal stress on the crack plane, Figure 7.15, τs and τt are the shear
stress components, ft the actual tensile strength after crack initiation and κ some
measure of softening.

σ
τt τs

Figure 7.15: Stresses on crack plane


The two models (Desai and Hoffman) are used in series. First of all, the overall stress
state is transformed into principal stresses. If a principal tensile stress component equal
to or larger than the tensile stress that would lead to (additional) non-linear response is
present, that stress is entered into the Hoffman-part of the model. Once the principal
stress components are analysed in this way, the resulting state of stress is transformed
back to the global reference system. If the resulting state of stress is located within the

The ACRe material model 147


Hoffman regime, the analysis is finished. If on the otherhand it falls within the Desai
regime, the new state of stress is further reduced according to the Desai-part of the
model (Figure 7.16).

The numerical techniques utilised to describe the material response are not part of the
work presented in this dissertation. It has been presented in other publications (Scarpas
et al. 1998, Erkens et al. 2000 and 2002).

Apply Hoffman Finally in Hoffman region?


ready

Back to xyz Finally in Desai-region?


Principle stress space Submit new stresses to Desai:
INTERACTION!

Hoffman regime=
cracking

Desai regime=
Isotropic damage

Figure 7.16: The two model components are used in series

7.2 MODEL PARAMETER DETERMINATION USING ONLY UNIAXIAL TEST RESULTS


In this section the parameters of the model described in the first part of this chapter will
be determined using the results of the uniaxial compression and tension tests only. The
methodology was originally presented in Scarpas et al (1997). In the context of this
thesis the influence of the strainrate and temperature was introduced in the parameter
determination process. The test set-ups, procedures and results were described in detail
in Chapters 4 and 5 and the asphalt mixture used was presented in Chapter 3. Since the
tests were carried out at several combinations of temperature and strain rate, the model
parameters could be expressed as functions of strain rate and temperature. In this way,
strain rate and temperature sensitivity was incorporated into the model.

7.2.1 FLOW SURFACE FOR UNIAXIAL STATES OF STRESS


For uniaxial states of stress, the expression for the flow surface (Equation 7.1)
simplifies considerably. In that case:
σ1 = σ , σ2 = σ3 = 0 ⇒
I1 = σ1 + σ 2 + σ 3 = σ (7.10)

148 Chapter 7
1 
2 2 2
I   I   I 
J2 =  σ 1 − 1  +  σ 2 − 1  +  σ 3 − 1  =
2  3  3  3 
(7.11)
1  2σ   σ   σ 
2 2 2
1 2
  +−  +−  = σ
2  3   3   3   3
 I  I  I  2
J3 =  σ1 − 1   σ 2 − 1   σ 3 − 1  = σ 3 (7.12)
 3  3  3  27

2 3 2 3
3 3 J3 3 3 27
σ 27 27
σ
cos(3θ ) = 3
= 3
= = ±1 (7.13)
2 2 2 1
( )
J2 2  1 2 2 σ
3
 σ  27
3 

Substitution in Equation 7.1 yields:

1 2
σ   σ − R n σ − R  
2
1
3 =  −α + γ  ⋅ (1 ± β ) −
(7.14)
    2
pa2   pa   pa  
 

7.2.2 MODEL PARAMETER β


This parameter controls the shape of the model on the octahedral plane. The shape can
vary from circular to triangular (Section 7.1.2.4). To determine this parameter test
results obtained at the same I1 but different θ values are needed. Uniaxial tension and
compression tests are related to different θ ’s, but also result in different I1-values, so
they cannot be plotted on the same octahedral plane. For this reason it is assumed that β
is zero, which corresponds to a circular cross section of the flow surface. Using this
assumption, Equation 7.14 reduces to:

σ2   σ − R n σ − R 
2 
=  −α   +γ    (7.15)
3 pa2   pa   pa  
 

7.2.3 MODEL PARAMETERS R AND γ


It was already explained that the hardening parameter α is zero at peak stress (Section
7.1.2.1). As a result, the first term within brackets in Equation 7.15 is zero if it is
evaluated for the uniaxial tensile and/or compressive strength values:
1 2
fc = γ ( fc − R )
2
3
The only model parameters left in this expression are R, which is the three dimensional
tensile strength, and γ. R can be found from uniaxial tension and compression data as
the intercept with the I1 axis of a line through the tensile and compressive strength,
plotted in I1-√J2 space (Figure 7.17).

The ACRe material model 149


Uniaxial compressive strength (fc, √(1/3fc2))
6 √J2 [MPa]
1/3ft2∆x
5 R= ft +
∆y
4 or
1/3fc2∆x
3 R= + fc
∆y=|sqrt(1/3*fc2)-sqrt(1/3*ft2)|
∆y
2 Uniaxial tensile strength
∆x=|ft-fc| (ft, √(1/3ft2))
1

0
-11 -9 -7 -5 -3 -1 1 3 5
I1 [MPa]
R
Figure 7.17: R can be determined from uniaxial tension and compression test results
Using the results from the tension and compression tests, R can be found as:

 ( f − fc )   ( f + fc ) 
R = fc  t − 1 = f c  t − 1 (7.17)
 f c − ft   f c − ft 

And g is found from rewriting Equation 7.16 and substituting Equation 7.17:
1 2 1 2 1 2
fc fc fc
γ= 3 = 3 = 3 (7.18)
( f c − R )2   ( ft + f c )  
2
 ( ft + fc ) 
2

 f c − f c  − 1 
  − fc 
  f c − f t   f c − f t 

Since general expressions for ft and fc as functions of strain rate and temperature are
available (Chapters 5 and 4), both R and g can now be expressed in a similar way.
Basically, R and g are the intercept and slope of the ultimate surface, the line through fc
and ft on the I1-√J2 plane. For the conditions tested some of these lines cross, indicating
that the expressions for R and g as a function of temperature and strain rate are not
straightforward. From the graph it can be seen that for higher temperatures and slow
strain rates the trend in the data is as expected. For these conditions both fc and ft
increase with decreasing temperature and increasing strain rate, causing the response
envelope to shift upwards. If temperatures keep decreasing and strain rates increase, a
different phenomenon can be observed. Instead of shifting upwards, the response surface
appears to rotate around a point in the tensile region. This point is the plateau tensile
strength (Chapter 5) and the phenomenon results from the fact that this plateau tensile
strength is reached for combinations of temperature and strain rate for which the
compressive strength is still increasing (Chapter 4). Naturally, lines through these
increasing fc and constant ft values grow steeper, affecting both R (the intercept with the
horizontal axis) and γ (the square of the slope).

150 Chapter 7
0 & 10 0&5 0&1 0 & 0.1
0 & 0.01 0 & 0.001 15 & 10 15 & 5
15 & 1 15 & 0.5 15 & 0.1
35 30 & 10
30 & 5 30 & 3 30 & 1 30 & 0.1

30
2
SQRT(J2) [N/mm ]
25

20

15

10

5
2
I1 [N/mm ]
0
-60.0 -50.0 -40.0 -30.0 -20.0 -10.0 0.0 10.0 20.0

Figure 7.18: Ultimate slopes (lines through fc and ft ) for the test conditions
The expressions for R and γ as functions of temperature and strain rate, found by
substituting the general expressions for fc and ft in Equations 7.17 and 7.18, are plotted
in Figure 7.19 and Figure 7.20, along with the parameter values determined for the
individual test conditions.

25 30'C
R [N/mm2] 15'C
0'C
20 Eq. 7.17

15oC
15
0oC
10
30oC
5

strain rate [%/s]


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

Figure 7.19: Data points and general expressions for R

The ACRe material model 151


0.35
γ 30'C
0.30 15'C
0'C
0.25 Eq. 7.18
0oC
0.20

0.15
15oC
0.10
30oC
0.05
strain rate [%/s]
0.00
0 1 2 3 4 5 6 7 8 9 10 11 12

Figure 7.20: Data points and general expressions for γ

7.2.4 MODEL PARAMETER n


The model parameter n is related to the onset of dilation in the specimen (Section
7.1.2.3). Dilation is the increase in volume that results from the opening of internal
cracks. Once dilation starts, the volume increases. Therefore, the onset of dilation is
related to the point at which the volumetric strain changes from decreasing to
increasing. Like stresses, strains can be decomposed in a volumetric and a deviatoric
part. The volumetric strains are related to changes in volume and the deviatoric strains
are related to changes in shape. In this case, only the volumetric strains are of interest
and in particular the plastic volumetric strains, since dilation does not occur in the
elastic region. The beginning of dilation can therefore be defined as the minimum of the
plastic volumetric strain curve, or mathematically:

p
∆ε vol =0 (7.19)

With
p ∂f
∆ε vol = ∆Γ Φ δ ij (7.20)
∂σ ij
where:
∆Γ :plastic multiplier (has to be determined experimentally)
<Φ> :plastic flow function, the McCauley brackets indicate that plastic flow only
occurs for positive
values of the flow function (<Φ>=0 for Φ<0 and <Φ>=Φ for Φ>0)
∂f
:an indication of the direction of plastic straining
∂σ
δij :Kronecker delta, 0 for i≠j and 1 for i=j

Since the flow surface f is defined as a function of the stress invariants rather than the
stress vector, Equation 7.20 becomes:

152 Chapter 7
p
 ∂ f ∂ I1 ∂ f ∂ J 2 ∂ f ∂ J 3 
∆ε vol = ∆Γ Φ  + + δ
 ∂ I1 ∂σ ij ∂ J 2 ∂σ ij ∂ J 3 ∂σ ij  ij
(7.21)
 

The derivatives of J2 and J3 to the stress vector reduce to zero after multiplication with
δij. This leads to:

p ∂f
∆ε vol = 3∆Γ Φ =0 (7.22)
∂ I1

The derivative of the flow function to the first stress invariant is zero for:

 I1 − R    I − R n−2   I1 − R 
n−2

 =0 ∨  nα  1  − 2γ  = 0 ⇒   = (7.23)
 pa    pa    pa  nα
 

The left-hand solution is a trivial one, so the right-hand one will be used in the
remainder of this derivation. Rewriting the expression for the flow surface results in:

 I − R  2   I − R  n − 2 
 1 
 −α 
1
 + γ 
 pa    pa  
J2   
f = − =0 (7.24)
pa2 (1 − β cos(3θ ) )

Substitute Equation 7.23 in 7.24 to get:

2
n= (7.25)
 J 2 (1 − β cos(3θ ) ) 
1 − 
γ ( I1 − R )
2
 

Evaluating Equation 7.25 for uniaxial states of stress yields:

2 2 (7.26)
n= =
 σ2  2  f c ( ft − f c ) 
 1 − σ dil  fc − 
2 f c − ft 
 3γ (σ − R )  1− 
 f ( f − fc ) 
f c2  σ dil − c t 
 f c − ft 

The state of stress in this expression must be evaluated at the beginning of dilation
(σdil).

At the beginning of a compression test, the axial strain is larger than the radial strain,
which leads to a decrease in volume. The onset of dilation is the point at which the
volume starts to increase and the expression for n has to be evaluated at the stress that
corresponds to that transition. This point can be determined by means of the axial versus

The ACRe material model 153


volumetric strain plot and the axial strain versus stress plot. The first plot is used to
determine the axial strain at minimum volumetric strain and with this strain the
corresponding state of stress can be determined using the second plot (Figure 7.21).

σ [Ν/mm2]
-1.0
-0.8
-0.06
-0.04
-0.02
0
0 - 0.005 -0.01 -0.015
-0.001
-0.002 εaxial [m/m]

-0.003 εvol [m/


Figure 7.21: Determine the state of stress at the onset of dilation
This information is obtained from the compression tests discussed in Chapter 4 (Erkens
et al 1998 and 2000).

From Equation 7.26 it is clear that the only term in the expression for n that is not yet
known as a function of the temperature and strain rate is the stress at the onset of
dilation, σdil. For this reason, it was attempted to express σdil as a function of strain rate
and temperature, using the same general relation as for the tension and compression
strength. The result is shown in Figure 7.22, where Equation 7.27 is plotted through the
data points.

   
   
   
σ dil = a 1 −   
1 1
d  = −56 1 − 0.43 
(7.27)
   b+ c       −96+ 27813   
  T    T 
 1 + ε& e    1 + ε& e  
       
Where:
ε& =strain rate [m/m]
σ dil = stress at the onset of dilation [N/mm 2 ]
T = temperature [K]
a,b,c,d= regression constants

154 Chapter 7
0.0
0.00 2.00 4.00 6.00 8.00 10.00 12.00
-5.0 ε [%/s]
-10.0

-15.0

-20.0

-25.0 30'C&0.1%/s 15'C&0.1%/s 0'C&0.1%/s


30'C&1%/s 15'C&1%/s 0'C&1%/s
-30.0 30'C&5%/s 15'C&5%/s 0'C&5%/s
30'C&10%/s 15'C&10%/s 0'C&10%/s
-35.0 303K 288K 273K
-40.0

-45.0
2
s dil [N/mm ]
-50.0

Figure 7.22: Individual values of the stress at the onset of dilation, combined with the
general expression (Equation7.27 ) ………………………..
The expression for n (Equation 7.26), which exhibits a vertical asymptote for σ≈fc after
a long horizontal branch, leads to very high n-values if dilation starts at a stress close to
the uniaxial compressive strength. A plot of n as a function of the stress is shown in
Figure 7.23.

Figure 7.23: Plot of n as a function of the stress at the onset of dilation (Eq. 5.16), the
asymptote near fc explains the larger variation in n-values for states of stress near fc
This plot shows that for states of stress close to the compressive strength a small
variation in the stress at the beginning of dilation results in a large variation in n. On the
other hand, if the state of stress at the onset of dilation differs considerably from the
compressive strength, a variation in the state of stress causes hardly any variation in n.
This results in some very high n-values (Table 7.3, page 169). Because of the strong
inter-relation between α and n, these high n-values will be addressed in combination
with the corresponding α-values in the next section.

The ACRe material model 155


7.2.5 HARDENING RESPONSE PARAMETER DETERMINATION

7.2.5.1 Parameter α
The last parameter to be determined, α, controls the hardening. Since all other
parameters are known (β is assumed to be zero, as stated before) α can be computed for
each stress level in a compression or tension test using the expression for the flow
surface (Equation 7.28).

  I − R n  I1 − R  
2
 −α  1
 +γ   
J   p   p  
f = 22 − 
a a
=0⇒
pa (1 − β cos(3θ ) )
2
J  I −R
− 22 (1 − β cos(3θ ) ) + γ  1 
α=
pa  pa  (7.28)
n
 I1 − R 
 
 pa 
Based on Equation 7.28 values for α for all states of stress throughout the stress strain
curves can be found. However, based on the definition for α, only the values between
the onset of non-linearity and peak strength have to be determined. The value found for
the state of stress at the onset of non-linearity gives α0, while α=0 at the peak stress and
onwards until complete annihilation of strength. To find α0, Equation 7.28 must be
evaluated at the state of stress at which non-linearity is initiated (σplas,c). All the
parameters in the equation are known for any combination of temperature and strain
rate, with the exception of the state of stress.

0.0
0.00 2.00 4.00 6.00 8.00 10.00 ε [%/s] 12.00
-5.0

-10.0

-15.0

-20.0

-25.0

-30.0

-35.0

-40.0 30'C&0.1%/s 15'C&0.1%/s 0'C&0.1%/s


30'C&1%/s 15'C&1%/s 0'C&1%/s
-45.0 30'C&5%/s 15'C&5%/s 0'C&5%/s
s plas [N/mm2] 30'C&10%/s 15'C&10%/s 0'C&10%/s
-50.0 303K 288K 273K

Figure 7.24: Individual values of the stress at the onset of non-linearity in


compression, combined with the general expression (Equation 7.29)
Similarly as for σdil in the previous section, the same general relation as for the tension
and compression strengths was used to describe σplas,c. The result is shown in Figure

156 Chapter 7
7.24, where the individual data points are shown along with the relation that was found
(Equation 7.29).

   
   
   
 1   1 
σ plas ,c = a 1 − d 
= −42 1 − 0.48 
(7.29)
   b + c       −92+ 26541   
  T 
 1 +  ε& e
T 
 1 + ε& e

  
       
  
Where:
ε& =strain rate [m/m]
σ plas,c = stress at the onset of plasticity [N/mm 2 ]
T = temperature [K]
a,b,c,d= regression constants
Using the expression for σplas,c the α0-values for every combination of temperature and
strain rate can be found from Equation 7.28:

 2   
2
 − σ plas ,c + γ  σ plas ,c − R  
 3 pa2  pa  
   
α0 = n
(7.30)
 σ plas ,c − R 
 
 p 
 a 

The degradation of α from α0 at the onset of plasticity to zero at the peak can be
expressed as either a function of the equivalent plastic strain ξ:

ξ = ∑ ε Tp , ij ⋅ ε p , ij ; i , j = x, y , z (7.31)

or the plastic work Wp:

Wp = ∫ σ dε p (7.32)

In earlier publications (Scarpas et al. 1997 and 1998), based on the preliminary tests, Wp
was used since this parameter does not require information on the radial deformations.
Currently, α is controlled via ξ, which incorporates the strains in all three dimensions
and is therefore believed to be a more powerful indicator of the triaxial response. The
relation used to describe the degradation is:
α = α 0 e−κ cξ (7.33)
Where:

The ACRe material model 157


ξ = the equivalent plastic strain
α 0 = the initial α value (material parameter)
κ c = controls the rate of degradation (material parameter)

Both the parameters in this expression are heavily influenced by the n-values. Especially
α0 becomes very small for large n-values. However, the combination of these large n’s
and small α0‘s still yields appropriate response envelopes for the initiation of non-
linearity. This is illustrated in by curve “1” in Figure 7.25, which corresponds to the
initiation of non-linearity at 0oC and 5%/s (n=12.3 and α0=9.49x10-26). In the graph the
ultimate response envelope for this conditions is also shown (“curve “2”). To visualise
the effect of n and α0 a third curve (“3”) is plotted. This curve corresponds to the same
γ, β and R as curve “1”, but n is set to 4 and α0 is determined using Equation 7.28 in
combination with the same point of initiation of non-linearity (σplas) as curve “1”. As
can be seen, curves “1” and “3” are comparable in size, but the shape is quite different.
The steeper descend of the original envelope influences the ratio between I1 and √J2 at
that point. For that reason and because n is directly related to an aspect of the material
response (dilation, Section 7.2.4) the values are used as they were determined. It must be
observed that the low α0 values are a potential source of numerical problems. In the
analysis ran until now (Chapter 8), no problems of this kind were encountered and
therefore the numerical implementation follows the analytical formulations presented in
this section.
compressive n=12.3, a=9.49E-26 (0&5%/s)
strength n=4, corr. Alpha
alpha=0, n=12.3
10.0
VJ2 [N/mm2]
9.0
8.0
2
7.0
6.0
5.0
1
4.0
initiation of 3.0
non-linearity
2.0
3
1.0
I1 [N/mm2]
0.0
-20 -10 0 10
Figure 7.25: High n-values in combination with low α0 yield appropriate envelopes
The exponent in the expression for the hardening parameter (Equation 7.33), varies with
temperature and strain rate (Figure 7.26). It can be described by Equation 7.34:

κ c = (-700T+213500)ε& + 3x1019 e( −0.1337T ) (7.34)

158 Chapter 7
With: T the temperature in Kelvin
ε& the strain rate in s-1

10000 κc

1000

100
0
15
30
Eq. 7.33 ε [%/s]
10
0 2 4 6 8 10 12

Figure 7.26: κ-values from the test data compared to Equation 7.34 (the vaules next
to the markers in the legend give the temperature in oC) ...
To illustrate the effect of Equation 7.34, some normalised degradation curves from the
tests are compared to the curves found using the κ-values from the equation. This is
shown in Figure 7.27, the test results are plotted using markers and the predictions
found using Equation 7.34 are shown as continuous lines. As can be seen, the
degradation of the hardening parameter is described quite well.
1
α/α01 15'C&1%/s 15'C&0.1%/s
0.9 30'C&10%/s 30'C&5%/s
30'C&1%/s 30'C&0.1%/s
0.8 n=8 n=5
n=3.3 n=2.4
0.7 n 2

0.6

0.5

0.4

0.3

0.2
x
0.1

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045

Figure 7.27: Normalised degradation curves for the hardening parameter (α), test
results and Equation 7.34 ………………………………
In order to ensure a correct transition from hardening to softening, it must be clear when
the hardening phase ends. Since the increase in stress at the end of this phase occurs
rather slow it is more effective to use the strain as a criterion. Because of the three-
dimensional nature of the model, the equivalent plastic strain is used. In Equation 7.35
the general expression for the limit value of the plastic strain (ξplim, the equivalent

The ACRe material model 159


plastic strain at the peak load) is presented. Figure 7.28 shows this relation as well as the
data points.

0.06
ξ p lim [m/m] 0'C 15'C 30'C Eq. 7.34
0.05

0.04

0.03

0.02

0.01
.
ε [% /s]
0.00
0.0 2.0 4.0 6.0 8.0 10.0 12.0

Figure 7.28: Values for ξplim compared to Equation 7.35 for the test conditions
p
ξlim =-0.249+9.122 × 10-4 T-3.508 × 10-8 T 2 × ln(ε&) (7.35)
T = temperature in Kelvin
ε& = strain rate in s-1

7.2.5.2 Hardening in uniaxial tension


From the tensile tests presented in Chapter 5 it could be seen that for most test
conditions the material exhibited a distinct non-linear response prior to the peak. This
hardening response could not be described using the hardening parameter found for the
compression tests, because this is initiated at very high stress levels compared to the
uniaxial tensile strength. For this reason a separate tensile hardening law had to be
introduced. It was already mentioned that tensile failure is of a different nature than
compressive failure, the former occurs localised on a plane, while the latter is more
distributed. For this reason the Hoffman model was introduced to control tension
softening on a crack plane (Section 7.1.3.2). Initially, it was attempted to incorporate
tension hardening in the overall Desai-based model but this required different hardening
parameters as well as n-values for tension and compression (Erkens et al. 2002). Since n
is related to a specific part of material response, it was decided not to use this approach.
Instead, it was decided that for a realistic representation of states of stress with
prevailing tension, a tension hardening mechanism independent of the Desai surface
was needed. As mentioned before, (Section 7.1.3.2), the Hoffman response surface was
utilized for simulation of tensile response degradation on crack planes.

On a crack plane, the Hoffman surface is defined as:


( ) (
σ 2 + q ⋅ τ 2 + τ 2 = f 2 δ&, T , κ
s t t ) (7.36)
in which σ is the normal stress on the plane, τs and τt are the shear stress components, ft
the tensile strength and κ some measure of softening. This formulation was extended
with a hardening component in order to capture the tension hardening.

160 Chapter 7
On the basis of the tension tests described in Chapter 5, a hardening law was developed
for ft in Equation 7.36. Before response degradation, any state of stress for which I1≥0
is resolved into principal stress space. The positive stress components are reduced
according to Equation 7.36 (with q=0). The resulting state of stress is rotated back to the
original stress space and compliance with the Desai criterion is checked.
Since the Hoffman criterion describes the response on a plane perpendicular to the
principal tensile stress, the hardening relation can be determined directly from the part
of the stress-strain curve between the end of linearity and the peak. This led to the
expression:
  εp 
− b 
  ε p −ε p  
ft = σ plas ,t + a  1 − e  lim   (7.37)
 
 
 
where : a = ft − σ plas ,t
p
Smixε lim
b=
ft − σ plas ,t
 
 
 
 1 
σ plas ,t = ft − 3.8 1 − 0.85 
   −105.2+ 32090   
 1 + ε&e
 T 

   
  
Smix = the apparent stiffness (Equations 4.9 and 5.8)
p
ε lim = the plastic strain at the peak

As can be seen, all constants in this relation are determined on the basis of material
properties such as apparent stiffness and tensile strength. This reduces the potential to
closely describe the hardening curvature, but it enhances numerical stability by
providing continuity of both the stress and the slope of the stress curve at the onset and
end of the hardening phase.

Similarly to isotropic hardening, for tension hardening also the end of the hardening
phase needs to be known. In this case the uniaxal strain rather than the equivalent three
dimensional strain is used. On the basis of the plastic strain at the peak a general
expression was developed (Equation 7.38). In Figure 7.29 this relation is shown
together with the data points.

p
ε lim = -8.363 × 10-2 +2.872 × 10-4 T-1.277 × 10-8 T 2 ln(ε& ) ≥ 1× 10-5 (7.38)
With: T the temperature in Kelvin and ε& the strain rate in s-1

From the form of the expression it can be seen that it is a linear regression relation. It
does not have a specific physical meaning, it just describes the data obtained from the
tests. As a result, when extrapolating it negative values for εplim are predicted. Of course,
this is not physically realistic. For that reason the values obtained from Equation 7.38

The ACRe material model 161


are required to be at least 1x10-5, which appears a realistic minimum value on the basis
of the available data.

0.014
ε plim [m/m] 0'C 15'C 30'C Eq 7.37
0.012

0.01

0.008

0.006

0.004

0.002

0
0 2 4 6 8 ε [%/s] 10

Figure 7.29: Equation 7.38 compared to the data for εplim for the test conditions

7.2.6 SOFTENING PARAMETER DETERMINATION


During softening the surface shrinks, via degradation of γ as a function of the post-peak
strain. The response degradation via γ reduction affects the whole surface, expressing a
degradation in strength for all states of stress. This isotropic softening is a realistic
representation of the degradation due to compressive loads, which cause a distributed
damage that weakens the material in all directions.
Tensile stresses, on the other hand, lead to a localised form of damage perpendicular to
the direction of loading i.e. a plane-of-weakness. If this plane is later on loaded in
compression, the strength is not affected by the tensile damage. This kind of response
cannot be modelled by isotropic softening, yet it is important in road engineering where
the material is alternatively loaded in tension and compression. For this reason another
degradation mechanism, based on the Hoffman plasticity model acting on a plane
perpendicular to the principal tensile stress, is used to describe this form of response
degradation.
In this section the procedures and parameters used for both types of response
degradation are discussed.

7.2.6.1 Isotropic softening via γ degradation


Throughout the hardening phase γ is constant. As mentioned earlier, at peak stress α
reduces to zero. For α=0, the flow surface becomes an open surface. On the I1-√J2 plane
it is represented by a line through the uniaxial tension and compression strengths (in
case β=0 as assumed here). The isotropic softening response, after peak, is modelled by
a gradual decrease in γ. The area under the softening part of the response curve controls
this reduction (Figure 7.30). Currently, the following expression is utilised to control γ
during response degradation (Scarpas et al. 1998, Erkens et al. 2000, 2001 and 2002):
γ = ηγ f + (1 − η )γ r (7.39)

162 Chapter 7
Where:
η = (1 + κ γ ξ pf ) e
−κ γ ξ pf

γ f = γ 0 , the initial value, as found from Equation 7.18


γ r = Cγ 0 , the ultimate (minimum) value, in this case C =0
ξ pf = the equivalent post-failure strain
κ γ = material parameter

The material parameter in this expression can be obtained from the softening branches
of the uniaxial compression tests, since in these tests the material exhibits isotropic
softening. Since the model describes the three dimensional response, it should also
incorporate material degradation in three dimensions. For that reason the different strain
components are combined into an equivalent scalar quantity: the equivalent post-failure
strain (ξpf, the strain after the peak), in the same way as for the plastic strain in the
section about hardening. In Figure 7.30 the normalised γ-value (i.e. divided by γ0) is
plotted against this equivalent post-fracture strain for all test conditions. As can be seen,
the ultimate value for γ is approximately zero. Under this assumption, the above relation
was fitted to the softening branches (Equation 7.40). It turned out that for all conditions
except the most brittle (5 and 10%/s at 0oC, see Chapter 4) approximately the same
material parameter (κγ) was found. Since it seems questionable that using a temperature
and strain rate dependent κγ will lead to a large increase in reliability, it is taken as a
constant. The resulting expression is plotted through the normalised softening branches
for the various test conditions in Figure 7.30. The strain rate and temperature
dependence of the actual softening curves is accounted for via γ0.

γ = (1 + κ γ ξ pf ) e
−κ γ ξ pf
γ 0 = (1 + 15 ξ pf ) e
(
− 15 ξ pf )γ (7.40)
0

1
γnorm=γ/γ0 general expression
0.9
Test results
0.8
0.7
0.6
0.5
0.4
0.3
0.2
xpf [mm/mm]
0.1
0
0 0.1 0.2 0.3 0.4 0.5 0.6

Figure 7.30: A single relation can be used for the normalised softening branches

The ACRe material model 163


7.2.6.2 Tension softening
The tension softening model follows the lines of the classical notion of fixed cracking,
where for states of stress exceeding the magnitude of the fracture surface, a plane of
cracking is introduced perpendicular to the principal tensile stress direction. At all
subsequent steps, the state of stress is resolved along the fixed planes of cracking and
compliance is ensured with the Hoffman criterion (Equation 7.36).

Similarly to the approach described for the tension hardening (Section 7.2.5.2), a stress-
degradation relation has to be determined on the basis of the uniaxial tension tests. As
discussed in Chapter 5, in the tension tests not only the overall response but also the
unloading of the uncracked parts was measured. The combination of this information
allowed the construction of stress versus crack-width (w) diagrams. These diagrams
describe the tension softening behaviour, which again had to be generalised for
integration in the material model.
Unlike the response degradation in compression, that in tension varies considerably
between the various test conditions, from very brittle, sudden fracture to a gradual
unloading. This makes it more difficult to generalise the cracking response, but a
relation that could describe all the varieties in response was developed on the basis of
the following considerations:
• Transition from ft to zero between w=0 and w=wmax
• Steepness and shape of the transition flexible enough to describe brittle as well as
gradual strength reduction
This yielded the following relation for tension softening:
   
   
 1   1 
ft = ft  c 
= ft  3
(7.41)
  w     w  
 1+  a ⋅ w + b    1 +  0.1385 ⋅ w + 0.0028  
  max     max  

where: ft =tensile strength


w=crack − width
wmax =maximum crack − width
a,b,c = regression constants

The maximum crack-width is, like the tensile strength, expressed as a function of
temperature and strain rate:

d 3
wmax = = (7.42)
 g  21345 
f+   −69.3+ 
1 + eε& e T  1 + 3ε& e T 

where : ε& = strain rate in s −1


T = temperature in Kelvin
d , e, f , g = regression constants

Relation 7.42 and maximum crack-widths obtained from the test data are shown on a
linear scale in Figure 7.31 and in on a logarithmic scale Figure 7.32.

164 Chapter 7
5.0
wmax [mm] Test data 0'C
4.5 Test data 15'C
Test data 30'C
4.0 Eq. 7.37

3.5

3.0

2.5
.

2.0

1.5

1.0
strain rate centre [% / s]
0.5

0.0
0.0 1.0 2.0 3.0 4.0 5.0 6.0

Figure 7.31: Maximum crack-width from test data and Equation 7.42

1.0E+01
wmax [mm]

1.0E+00

1.0E-01
.
Test data 0'C
Test data 15'C
Test data 30'C
1.0E-02 Eq.7.37

strain rate centre [% / s]


1.0E-03
0.0 1.0 2.0 3.0 4.0 5.0 6.0

Figure 7.32: Maximum crack-width from test data and Eq. 7.42 on logarithmic scale
That Equation 7.41 describes the various types of response degradation reasonably is
illustrated in Figure 7.33, where three stress versus crack-width relations that vary from
brittle to ductile as found from the experiments are shown together with the
corresponding softening branch predicted by Equation 7.41.

The ACRe material model 165


1
σn=σ/ft
0.9

0.8
fit
0.7 30o&3%/s
data
0.6
fit
0.5 o
0 &0.0.01%/s
0.4 data

0.3
data
o
0.2 15 &0.5%/s
fit
0.1 w [mm]
0
0 0.5 1 1.5 2 2.5

Figure 7.33: Tension softening by Eq. 7.41 versus some actual softening branches

7.2.7 OVERVIEW OF THE MODEL RELATIONS AND PARAMETERS


In this section the model expressions discussed in the previous sections are listed
together. In the elastic phase of material response the model relations are (Chapter 5):

 
 
 1 
log( Smix ) = 4.5 1 − (7.43)
 0.16 
 1 + ε& exp  −83.5 + 27200   
   T   

ν = 0.0027 T − 0.49 (7.44)


2
Where Smix is the mixture stiffness in N/mm , ν the Poisson’s ratio, T the temperature in
Kelvin and ε& the strain rate in s-1

The linear-elastic response is bounded by the initial response envelopes. From the
initiation of non-linearity until the annihilation of strength the response is controlled by
the Desai-based model and the Hoffman based model, as discussed earlier in this
chapter. In Table 7.1a and b the relations for both these model components are listed. In
the first table the expressions for the envelope and the parameters that correspond to its
initial form (initiation of non-linearity) are listed and in the second table the relations
that govern hardening and softening are shown. In all relations T represents the
temperature in Kelvin, ε the strain in m/m, ε& the strain rate in m/(m*s) and all stress-
related terms are given in N/mm2.

166 Chapter 7
Desai-based model Hoffman-based model
  I − R n  I1 − R  
2

J
 −α  1
 p
 a 
 +γ 
p
 a  
  ( )
f = σ 2 + q ⋅ τ s2 + τ t2 − ft 2 = 0
f = 22 −  =0
pa (1 − β cos(3θ ) )
Initial envelope:
Initial envelope: q=0
β =0

f t ≤ σ plas ,t
f t = 5.5 ×
 
  σ plas ,t = ft −
1 − 1 
 0.86   
 1 +  ε& exp  − 80.2 + 25050     
  T   3.8 1 −
1

    0.85
   32090  
 1 +  ε& exp  − 105.2 + T   
f c = − 108 ×
  f t = 5.5 ×
 
1 − 1   
 0.32   
 1 +  ε& exp  − 86.3 + 24260    1 − 1 
  T    0.86 
    
 1 +  ε& exp  − 80.2 + 25050   
  T  
   
 ( f − fc ) 
R = fc  t − 1
 fc − ft 

1
fc2
γ = 3 =
0
( fc − R )2
1
fc2
3
2
  ( f − fc )  
 f c − f c  t − 1 
  fc − ft  

2
n=
 σ dil 2 
 1 − 2
 3γ (σ dil − R ) 

Table 7.1a: Model relations for each phase of response, initial envelope part I

The ACRe material model 167


Initial envelope:

σ dil = − 56 ×
 
 
1 − 1 
 0.43 
 1 +  ε& exp  − 96 + 27813   
   T  

 σ2  σ plas ,c − R  
2
− plas , c
+γ   
 3 pa2  p  
  a  
α0 = n
 σ plas ,c − R 
 
 pa 
 
σ plas ,c = −42 ×
 
 
1 − 1 
 0.48 
 1 +  ε& exp  −92 + 26541   
 T   
  
Table 7.2b: Model relations for each phase of response, initial envelope part II

168 Chapter 7
Desai-based model Hoffman-based model

Hardening: Hardening:
α = α0 e −κ c (ξ p )   εp
− b p

p 
  ε − ε  
Where: ft = σ plas , t + a  1 − e  lim ;
19 ( −0.1337T )  
κ c = (-700T+213500)ε& + 3 x10 e
 
ξ p =equivalent plastic strain= a = ft − σ plas ,t

(∑ ε ijp ε ijp ) , i,j=x,y,z b=


p
S m ix ε lim
with 0 ≤ ξ p ≤ ξlim
p f t − σ plas ,t
p p
p
ξlim =-0.249+9.122 × 10-4 T w ith : 0 ≤ ε ≤ ε lim
-3.508 ×10-8 T 2 × ln(ε&) p
ε lim = -2
-8.363 × 10 +2.872 × 10 T
-4

-1.277 × 10 -8 T 2 ln( ε& ) ≥ 1 × 10 -5

Softening: Softening:
γ = (1 + κ γ ξ pf ) e
−κ γ ξ pf
γ0  
 
( )γ  1 
( )
− 15ξ pf
= 1 + 15 ξ pf e 0
ft = ft  3
 1+  w  
with ξ pf the equivalent    
  0.1385wmax + 0.0028  
post-failure strain
3
wmax = ;

21345 
 −69.3+ 
1 + 3 ε& e T 

Table 7.1b: Model relations for each phase of response, hardening and softening
A qualitative overview of the model parameters mentioned in this overview is given in
the Table 7.3 and Table 7.4. In these tables the model parameter values for both models
are listed as functions of the test conditions.
p
T rate fc ft R γ 0 κ γ s dil n σ plas,c α0 κc ξ lim
2 2
[K] [%/s] [N/mm ] [N/mm2] [N/mm ] [N/mm2] n [N/mm2] [m/m]
273 0.1 -21.5 5.2 13.6 0.13 15 -19.7 29.4 -10.9 1.93E-67 -5500 1.5E-02
273 1 -38.0 5.5 13.0 0.19 15 -36.4 93.6 -23.9 2.94E-237 -17211 7.5E-03
273 5 -49.5 0.8 1.5 0.31 15 -44.2 278.3 -30.6 8.95E-111 -7000 6.9E-03
273 10 -56.5 3.0 6.2 0.27 15 -44.2 37.8 -18.6 4.25E-51 -7641 7.8E-03
288 0.1 -5.7 2.4 8.2 0.06 15 -3.8 5.0 -1.1 5.18E-08 -168 4.1E-02
288 1 -11.5 5.2 18.8 0.05 15 -9.2 8.1 -2.1 2.83E-16 -413 3.0E-02
288 5 -17.0 0.1 0.2 0.32 15 -14.2 369.8 -5.5 9.49E-26 -949 2.5E-02
288 10 -22.3 2.1 4.7 0.23 15 -17.4 21.6 -12.1 9.54E-23 -1259 1.7E-02
303 0.1 -1.9 0.3 0.8 0.16 15 -0.4 2.4 -0.1 5.62E-02 -111 4.5E-02
303 1 -3.5 0.9 2.5 0.12 15 -0.8 2.4 -0.6 2.69E-02 -102 4.1E-02
303 5 -5.0 2.0 6.8 0.06 15 -1.5 2.4 -1.1 8.10E-03 -119 3.9E-02
303 10 -7.5 1.6 4.1 0.14 15 -3.7 2.0 -2.7 8.72E-05 -156 3.3E-02
Table 7.3: Overview of the Desai-based model parameters for the test conditions

The ACRe material model 169


p
T ra te ft σ plas,t ε lim w max
2 2
[K] [%/s] [N/m m ] [N/mm ] [m/m] [m m ]
273 0.001 2.7 0.44 4.2E-03 4.4E+00
273 0.01 5.1 0.96 3.5E-03 8.7E-01
273 0.1 5.2 0.91 5.7E-04 1.6E-02
273 1 5.5 2.93 1.9E-04 7.0E-03
288 0.1 2.4 0.80 8.7E-03 1.1E+00
288 0.5 4.3 2.07 4.4E-03 1.1E+00
288 1 5.2 1.61 4.6E-03 6.1E-02
303 0.1 0.3 0.18 1.2E-02 2.9E+00
303 1 0.9 0.64 1.0E-02 2.9E+00
303 3 1.8 1.48 4.6E-03 2.3E+00
303 5 2.0 1.69 6.4E-03 3.0E+00
Table 7.4: Overview of the Hoffman-based model parameters for the test conditions

7.3 MODEL PARAMETER DETERMINATION FROM UNIAXIAL & MULTIAXIAL TESTS

7.3.1 RELATION BETWEEN β, γ AND R


If for each temperature and strain rate results from more than two stress paths are
available, the parameters that control the ultimate surface can be determined via
regression analysis with respect to the strength values. In that case the expression for the
flow surface (Equation 7.1) reduces to (using α=0 at peak stress) Equation 7.45. In this
expression I1, J2 and cos(3θ) are stress invariants and γ, β and R are model parameters.

γ
f = J2 − ( I1 − R )2 = 0 (7.45)
(1 − β cos(3θ ) )

Using this expression, where β no longer has to be zero, means that the flow surface can
no longer be represented on the I1 - √J2 plane, since now the cos(3θ) values have to be
taken into account. This also means that the results from the different tests are no longer
positioned on a straight line, since even the tension and compression results are related
to different cos(3θ) values (Table 7.5). As a result, the expressions developed in the
previous section under the assumption β=0 cannot be used.

Type of test cos(3θ)


Compression -1
Tension 1
Shear (no confinement) 0
Table 7.5: Different tests yield different cos(3θ) values
Instead, Equation 7.45 was rewritten and used for a least squares analyses:

γ
J2 =
4 1− β
( I1 − R ) (7.46)
cos(3θ )
Using this relation in the statistics package SPSS resulted in R values of zero.
Obviously, a material that exhibits uniaxial tensile strength can not have an R-value of
zero (see also Figure 7.11 and Figure 7.17). When restraints for the regression constants

170 Chapter 7
were specified, based on the information obtained from the uniaxial tests and the
physical meaning of the parameters (R≥0, γ≥0 and 0≤β<1) this resulted in β=0. In order
to gain understanding about the inter-relation between the parameters, Equation 7.46
was rewritten:

J2 1
J 2 − γ ( I1 − R ) (1 − β cos ( 3θ ) )
−1/ 4
=0 ⇒ R+ = I1 (7.47)
14442444 3 H γ
H
Rewriting the above as:

 
 J   1 
 2 1  γ =I (7.48)
{ H    {1
 ai   R  bi

If N test data are available, Equation 7.48 becomes


 a1 1  b1 
   1   
 a1 a2 .... aN   a2 1   =  a1 a2 .... aN   b2  ⇒
 1 1 .... 1   M M   γ   1 1 .... 1   M  (7.49)
   
   R   
 aN 1 bN 
N 2 N  N 
 ∑ ai ∑ i    ∑ ai bi 
a 1
 i =1 i =1   γ  =  i =1  (7.50)
N    N 
 ∑ ai N   R   ∑ bi 
 i =1   i =1 

Solving for R and γ:


 N  N 
 1   N − ∑ ai   ∑ ai bi 
 γ = 1 i =1   i =1  (7.51)
  ∆ N N  N 
 R   −∑ ai ∑ ai2   ∑ bi 
 i =1 i =1   i =1 
2
N  N 
where: ∆ = N ∑ ai2 −  ∑ ai  (7.52)
i =1  i =1 
and hence
1 1 N N N 
=  N ∑ ai bi − ∑ ai ∑ bi  (7.53)
γ ∆  i =1 i =1 i =1 
1 N N N  N
 −∑ ai ∑ ai bi + ∑ a i ∑ bi 
2
R= (7.54)
∆  i =1 i =1 i =1 i =1 
Both ai and bi contain H, which means that R and γ are expressed as functions of β.

Equations 7.53 and 7.54 are higher order expressions in β. As a result they have many
local extremes which makes the solution sensitive for the selected starting values. As

The ACRe material model 171


mentioned above, this had a large impact on the SPSS analyses. For that reason a least-
squares procedure using Newton-Raphson iteration and the partial derivatives of the
above mentioned parameters was programmed in Maple V, as an alternative for the
procedures provided by SPSS. In these analyses as well as in the SPSS analyses β,
which should fulfill the criterion: 0≤ β < 1 (Equation 7.45), tended to go to 1. This
resulted in parameters that differed considerably from those found under the assumption
that β=0. Although the overall fit appeared to be quite good the low R values, which
were not only smaller than to those found for β=0, but even smaller than the uniaxial
tensile strength, were a cause for concern. When the data and the model were plotted on
the σ1-σ3 plane, a useful representation since in all the tests σ2=0, it became clear how
different sets of parameters could lead to a good fit (Figure 7.34).

2 step linearised LSq


1 step least squares
data points

Figure 7.34: Different sets of model parameters fit the test results
From Figure 7.34 it is clear that the good fit for the least squares procedure is the result
of the two branches of the model expression fitting each a part of the data. Since only
the left-hand branch is physically relevant, this is not the intended solution. However, it
explains the difficulties encountered in the least squares analysis. The existence of two
branches is the effect of the square root in the model expression. In order to limit the
solution to the physically relevant branch a two-step least squares approach was used. In
the first step Equation 7.46 is written as:
γ
J2 =
4 1− β
( I1 − R ) = P1I1 + P2 (7.55)
cos(3θ )
From this step P1 and P2 are found, where:
P γ
R = − 2 and P1 = − (7.56)
P1 4 1 − β cos(3θ )

In the second step a linearised form of P1 is used in the least squares analyses. This
linearisation eliminates the effect from the square root:
γ 1 
P1 = − → −  cos(3θ ) P1  β − γ = 0 (7.57)
4 1 − β cos(3θ ) 4 
Using the P1 values found from step 1, this second step results in the values for β and γ.
This procedure led indeed to a fit of the data using only the physically relevant branch
of the flow surface (Figure 7.34). Again the β values tended to go to an extreme value,
but this time it was zero. For the data under consideration, this is understandable. From

172 Chapter 7
Figure 7.35 it can be seen that larger β values lead to a larger difference between
uniaxial and biaxial compressive strength, indicating that β is related to the confinement
sensitivity of the material. For β=0, this difference is slightly less than a factor two, for
larger β values it rapidly increases. For the material under investigation more than a
factor two difference between uniaxial and biaxial compressive strength seems
unrealistic. Furthermore, the influence of β on the shape of the response surface in the
bi-axial tension and tension-compression regions is very limited (Figure 7.35). As a
result, on the basis of the available information β appears to be zero for this material. To
establish whether β is truly zero or rather a fairly small value requires the results of
actual triaxial compression tests. For this reason it might be worthwhile to integrate
such a test in the test program, especially with respect to the characterisation of other,
more confinement sensitive (open-graded), mixtures.

Figure 7.35: Effect β on the shape of the surface


The γ and R values found using the approach described above are shown in Table 7.6
along with those found using the approach described in Section 7.2.3. As can be seen
there are some differences but these are small. The approach described in Section 7.2.3
results in general relations for R and γ because they are expressed as functions of fc and
ft. The two-step least squares approach results in discrete values for the model
parameters and because only results at a limited number of conditions are available and
the trend in R and γ is rather complex (Figure 7.19 and Figure 7.20), it will be difficult
to develop a general relation. For numerical implementation two approaches are
possible, a general relation can be programmed, but it is also feasible to enter several
discrete values for given conditions and let the program interpolate to determine the
values for other conditions. The latter approach could be used on the basis of the
available data.
direct calculation 2step least squares
T rate fc ft R g b R γ β ∆R ∆γ
o 2 2 2 2 2
[ C] [%/s] [N/mm ] [N/mm ] [N/mm ] [N/mm ] [N/mm ]
0 0.005 -9.4 4.4 16.7 0.04 0 19.2 0.03 0 -2.4 0.01
0 0.5 -31.8 5.5 13.2 0.17 0 13.4 0.17 0 -0.2 0.00
15 0.05 -4.7 1.8 5.9 0.06 0 5.2 0.07 0 0.7 -0.01
15 0.5 -9.4 4.3 15.9 0.05 0 18.6 0.04 0 -2.7 0.01
15 2.5 -14.8 5.1 15.8 0.08 0 16.3 0.07 0 -0.5 0.01
30 0.5 -2.6 0.4 1.1 0.17 0 1.3 0.15 0 -0.3 0.02
30 5 -5.4 2.1 7.1 0.06 0 6.2 0.08 0 0.9 -0.02
Table 7.6: R and γ from 2 step least squares procedure and those from Section 7.2.3

The ACRe material model 173


7.3.2 CONSEQUENCES FOR THE OTHER MODEL PARAMETERS
Since it turned out that for the ACRe asphalt mixture β truly is zero, the other material
parameters do not change. If β had not been zero, the two-step least squares method
would have yielded different R and γ values and in that case some of the other
parameters would also change. In this section the differences that might exist between
the model parameters determined under the assumption that β=0, as described in
Section 7.2 and those found without this assumption (Section 7.3) for a mixture where β
is not equal to zero are investigated. For the ACRe mixture this problem does not arise,
but it is considered worthwhile to list the points at which the procedure outlined in
Section 7.2 will have to be adapted for mixtures with β≠0.

The tension hardening and softening relations are not influenced by the β-value because
they depend solely on the results from the tension tests. Also, the isotropic softening is
not influenced because it is expressed as the degradation of a normalised gamma (γ/γ0)
on the basis of the descending branches in the compression tests. The parameter related
to the onset of dilation, n, would also remain the same because it is determined on the
basis of the volumetric versus axial strain curve, which is a direct result from the
compression tests and thus independent of the β-value.

The hardening parameters, however, are determined on the basis of the stress-strain
relation in combination with the flow surface expression and would be effected by non-
zero β values and the changes in R and γ. The hardening parameter for both tension and
compression is determined by evaluating Equation 7.58.

  I − R n  I1 − R  
2
 −α  1
 +γ   
J   p   p  
f = 22 − 
a a
=0⇒
pa (1 − β cos(3θ ) )
2 (7.58)
J  I −R
− 22 (1 − β cos(3θ ) ) + γ  1 
α=
pa  pa 
n
 I1 − R 
 
 pa 

Obviously, this relation will yield different results if β changes from zero to non-zero.
Since the values for R and γ will also differ between those two cases, it is probable that
the initial hardening parameter will be different for non-zero β’s. The degradation of the
hardening parameters towards the peak where they are zero, however, is determined on
the basis of the normalised hardening parameter and the equivalent plastic strain in the
compression test. So these relations are not effected by a change in the initial value.

From these considerations it can be concluded that only the procedure for the
determination of R and γ changes when the assumption β=0 is not used. In that case,
these two parameters and β have to be determined on the basis of a least squares
analyses on the available data set (Section 7.3.1). The procedures used in Section 7.2 to
determine the other model parameters all remain valid. The only other parameter that
changes is α0, but its detemination procedure is not affected.

174 Chapter 7
7.4 CONCLUSIONS AND RECOMMENDATIONS
On the basis of the combined tension and compression test results, general expressions
were developed for the elasticity parameters (stiffness and Poisson’s ratio) as functions
of temperature and strain rate. It turned out that the apparent stiffness values obtained
from the tension tests were considerably higher than those found from the compression
tests. This disagrees with the often-proposed idea that the asphalt concrete responds
stiffer in compression than in tension because in the first case the aggregate skeleton is
loaded and in the second the binder. It is argued in this thesis that this may be valid for
mixtures with a true aggregate skeleton, but not for dense mixtures like the ACRe
mixture. For those it is expected that the stiffnesses would be roughly the same. The
higher values found from tension can be explained by the influence of the end caps.
These cause a state of three dimensional tension, which reduces the axial strain and that
leads to an over-estimation of the stiffness. Without friction reduction, in the
compression test the same phenomenon would occur, but that is counter-acted in this
case by the friction reduction system. To exclude the three dimensional tensile effect on
the stiffness, this parameter should only be obtained from tension data measured at the
centre of a specimen. If this is not possible, it is better to determine it on the basis of
compression data from tests in which friction reduction is used to minimise the end
effect.
The model differentiates between the damage caused by tension and
compression loads, where the isotropic damage that results from compression is
incorporated in the Desai-based plasticity model and the loacalised tension damage is
described using a Hoffman-like model. The parameters that controll the response in
both of these models can be determined as afunctions of temperature and strain rate on
the basis of the ACRe test programme.

7.5 REFERENCES
Desai, C.S. and Omar Faruque, M., (1984), Constitutive Model for (Geological)
Materials, Journal of Engineering Mechanics, Vol. 110, No. 9, Pg. 1391-1408
Desai, C.S., Somasundaram, S. and Frantziskonis, G., (1986), a Hierarchical approach
for Constitutive Modelling of Geologic Materials, International Journal of
Numerical and Analytical Methods in Geomechanics, Vol. 10, No. 3, Pg. 225-257
Desai, C.S, (1980), A General Basis for Yield, Failure and Potential Functions in
Plasticity, International Journal for Numerical and Analytical Methods in
Geomechanics, Vol. 4, Pg. 361-375
Erkens, S.M.J.G. and Poot, M.R., (1998), The Uniaxial Compression Test - Asphalt
Concrete Response (ACRe), Delft University of Technology Report 7-98-117-4
Erkens, S M.J.G. and Poot, M.R., (2000), Additional Compression Tests - Asphalt
Concrete Response (ACRe), Delft University of Technology Report 7-00-117-5
Erkens, S. M.J.G., Liu, X., Scarpas, A. and Molenaar, A.A.A., (2000), 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
Frantziskonis, G., Desai, C.S. and Somasundaram, S., (1986), Constitutive Model for
Non-associative Behaviour, ASCE Journal of Engineering Mechanics, Vol. 112
Scarpas, A., Blaauwendraad, J., Al-Khoury, R.I.N. and Gurp, C. van , (1997),
Experimental calibration of a viscoplastic-fracturing computational model, Paper
presented at the Computational Methods and Experimental Measurements
(CMEM), Rhodos, Greece, May 1997

The ACRe material model 175


Scarpas, A., Gurp, C.A.P.M. van, Al-Khoury, R.I.N. and Erkens, S.M.J.G. , (1997b),
Finite Element Simulation of Damage Developement in Asphalt Concrete
Pavements, 8th International Conference on Asphalt Concrete Pavements, Seattle,
Washington, USA, 1997
Scarpas, A. and Blaauwendraad, J., (1998), Experimental Calibration of a Constitutive
Model for Asphaltic Concrete, Paper presented at the Euro-C conference on the
Computational Modelling of Concrete Structures, Badgastein, Oostenrijk, 31
march- 3 april 1998

176 Chapter 7
8. MODEL VERIFICATION AND APPLICATION
“In theory, there is no difference between theory and practice. But in practice there is."
-Jan L.A. van de Snepscheut-
Equation Section (Next)

8.1 INTRODUCTION
Before the model can be used to reliably simulate asphalt concrete behaviour it must be
verified. For this purpose some of the compression and tension tests described in
Chapters 4 and 5 are simulated using the model and model parameters presented in the
previous chapter. Based on these simulations, stress-strain curves are determined and
compared to the ones found from experiments. Also, the stress and damage patterns
obtained from the simulations are compared to the failure patterns observed in
laboratory tests. From these simulations it can be seen that the model provides a reliable
prediction of the response of asphalt concrete.

Once the model has been verified, it is used to simulate a standard road engineering test
in which both tension and compression damage occurs: the indirect tensile test. By
alternately activating the tension and compression damage components, the necessity of
modelling both types of response degradation is investigated. The results from the
simulation are again compared to test data. These simulations demonstrate the model
capabilities with respect to conditions and geometries that were not included in the
model parameter determination procedures and they provide information about the
damage processes that take place in the tests.

Although the capabilities of a realistic 3D material model like the ACRe model can only
be fully exploited using powerful numerical tools such as the Finite Element Method
(FEM), those are not available to anyone. Firstly, these methods require a certain level
of expertise in order to arrive at the correct model and to interpret the data and secondly
the analyses are rather time consuming. As such, non-linear Finite Element analysis is
still mainly a tool for those involved in developing fundamental knowledge and
understanding of Asphalt Concrete. However, the ACRe test program and model can
also be used in a more pragmatic, analytical, manner and although this will not provide
the detailed information and insight that can be obtained from FEM analyses, it provides
important information on material characteristics and construction safety. This
simplified approach will be presented in the last section.

8.2 DAMAGE DEFINITION


The simulations presented in this chapter are non-linear analyses, using the model
presented in the previous chapter. In non-linear analyses the material is degrading and
damage occurs. In order to understand the results, it is necessary to introduce the
damage definition that was used, because in plasticity and viscoplasticity related
literature various definitions of damage can be encountered. In the framework of this
investigation, damage has been defined as the magnitude of permanent (i.e.
irrecoverable) strain in the material. Mathematically this notion of total damage can be
expressed as:
ξ = ∫ d εˆ p (8.1)

Model verification and application 177


in which d εˆ p = deijp ⋅ deijp = de xx xx yy yy
p ⋅ de p + de p ⋅ de p + ... (8.2)
in which the term inside the square root is the sum of the squares of the plastic strain
increments. In the above expression all these plastic strain increments are used to arrive
at a skalar that expresses the total damage at a given location. This total damage can be
subdivided in a volumetric component, Figure 8.1(a) and a deviatoric one,Figure 8.1(b).
(a) Compressive
volumetric
deformation

Tensile
volumetric
deformation

(b)

Deviatoric
deformation

Figure 8.1: Schematic of types of damage


This subdivision of damage types can be useful to understand what kind of damage is
developed. Compressive volumetric damage is typically associated with deformations
leading to inelastic compaction of the material (rutting). Tensile volumetric damage is
typically associated with deformations leading to cracking. Deviatoric damage is the
result of tensile-compressive states of stress and can lead to Mode II associated
cracking.

8.3 MODEL VERIFICATION: SIMULATIONS OF THE ACRe TESTS


The simulations were performed using CAPA-3D, a Finite Element package developed
by Scarpas and co-workers at the Delft University of Technology. The model described
in the previous chapter is implemented in the package, which facilitates three-
dimensional, non-linear analyses using static, dynamic (monotonic) and cyclic loading.
Because the computation time in a Finite Element calculation is directly related to the
number of degrees of freedom (DOF) and, thus, the number of elements it is important
to reduce the number of elements where possible, without loosing accuracy. As a result,
in the simulation of the uniaxial tension and compression test, where the specimen and
the loads are symmetrical with respect to the cylinder axis and the specimen mid-plane,
only one-eight of the specimen is actually modelled. The elements used for all the
simulations discussed in this chapter are three-dimensional, twenty-nodes brick
elements.

8.3.1 SIMULATIONS OF THE UNIAXIAL COMPRESSION TEST


The model described in the previous chapter has been used to simulate a typical
compression test from the ACRe test programme. The mesh used for the simulation of
the compression tests is shown in Figure 8.2. The test is simulated by forcing the top
plane of the specimen to move downwards at a specified displacement rate. Since the
top surface is completely free to expand laterally, the simulation is that of an ideal,
frictionless compression test. The effect of friction will be addressed later in this

178 Chapter 8
section. Because only half the specimen height is modelled, using the mid-plane of the
specimen as a symmetry plane, the applied deformation rate is only half of that used in
the actual test. Due to the symmetry, the specimen half that was not modelled behaves
exactly the same, resulting in the required total deformation.

Figure 8.2: Finite Element mesh for the compression test undeformed (left) and
deformed (right)………… …………..
The predicted curves as well as the average curves found from the experiments are
shown in Figure 8.3. Two types of predicted response curves are shown. Curves
generated by the model by specifying as input parameters values which were actually
determined from the tests (these are labelled as “model using parameter values”) and,
values obtained via the general relations for the model parameters presented in the
previous Sections. Both sets of input parameters produce almost identical results.
From Figure 8.3 it can be concluded that the proposed model can simulate satisfactorily
both, the ultimate and the dilatant response of the asphaltic material. Nevertheless, some
differences in response do exist and attempts are currently under way to explore the
possibility of using more intricate response functions for post-peak response simulation.
The differences between predicted and measured response can be also partially
attributed to the formulation of the constitutive model. As mentioned in Chapter 7, the
Desai function is controlled via the artifact of “equivalent plastic strain”. This means
that for model fitting purposes, an average measure of plastic strain is developed on the
basis of the experimentally recorded inelastic radial and axial strains. This, in
combination with the plasticity formulation, implies that the model abolishes the
capability to actually control individual components of strain. Instead emphasis is
shifted on the adequate simulation of the multiaxial material response.

Model verification and application 179


-25
15C&0.1%/s
15C&1%/s
σ [N/mm2]
15C&10%/s -20
ACRe model

-15

-10

-5

0.6 0.5 0.4 0.3 0.2 0.1 0 -0.1 -0.2 -0.3


ε rad [m/m] 0 ε ax [m/m]

Figure 8.3: Test results compared to simulations using specific parameters values
and parameter expressions ………………………….
This conjecture is supported by the plot in Figure 8.4 in which the vertical stress is
plotted against the equivalent strain. It can be seen that the matching between
experimental and predicted response is almost ideal.

-25
σ [N/mm2]
-20

-15

-10

-5
ξpeq [m/m]
0 0.1 0.2 0.3 0.4 0.5 0.6
0

Figure 8.4. Compressive stress versus equivalent strain


Because of the predominant role that dilation plays in asphaltic materials, a constitutive
model capable of describing satisfactorily the multiaxial response is deemed more
important than one, which can provide perfect matching in a single loading direction.
This is particularly true in simulations of traffic loading in which stress rotation occurs.

To illustrate the effect of friction in the contact surfaces, a topic that was treated
elaborately in Chapter 4, a simulation where expansion of the top elements was
restrained was carried out. The effect on the stress distribution and specimen
deformation is shown in Figure 8.5.

180 Chapter 8
σ

Figure 8.5: Stress distribution in a compression test with friction, vertical (σyy,
Magnification Factor (MF)=1, left) and horizontal (σxx, MF=0, right)
From Figure 8.5 it can be seen that the specimen deformations are not as uniform as in
the unconstrained, frictionless simulation (Figure 8.2, right). The specimen is barrelling
and the distribution of vertical stresses is less uniform. In case of an unconstrained test,
the horizontal stresses are, apart from some numerical fluctuations, zero. In case of
friction, it can be seen that horizontal compression causes confinement of the material
near the loading plate. This confinement reduces towards the specimen centre.

8.3.2 SIMULATION OF THE UNIAXIAL TENSION TESTS


For the tension tests again 1/8 specimen was modelled. In this case, the contact surface
is restrained in the horizontal direction to incorporate the effect of the cap that is glued
onto the specimen in the simulation. The mesh is shown in Figure 8.6, along with the
stress distribution in the specimen. It can be seen that the specimen shape leads to high
stresses in the centre region, where failure occurred.

Figure 8.6: Mesh (left) for the tension test and elastic stress distribution (right)

Model verification and application 181


The capability of the model to simulate the tensile response of asphalt concrete is shown
in Figure 8.7. As can be seen, both the pre-peak and the softening response of the
material are simulated satisfactorily.
6 o
15 C
σ [N/mm ]
2

1%/s, tests
5 0.5%/s, tests
0.1%/s, tests
1%/s, model
4 0.5%/s, model
0.1%/s, model

1
uax [mm]

0
urad [mm]
-0.2 0.0 0.2 0.3 0.5 0.6 0.8 0.9 1.1 1.2 1.4 1.5

Figure 8.7: Tension test simulation

8.4 MODEL APPLICATION: SIMULATION OF THE INDIRECT TENSION TEST (ITT)


From the simulations in the previous section it is clear that the ACRe material model
can succesfully simulate the tests used in the model parameter determination. In this
section it is taken one step further by using the model to simulate a test that was not
used in and not suited for the model parameter determination: the indirect tensile test
(Erkens et al 2002). Results from this tests on the ACRe mixture are available (Chapter
3) and can be used to verify the simulations. On the other hand, this tests results in
complicated stress patterns and is difficult to interpret. The Finite Element simulations
show what happens inside the specimens during the test, so they can be used to gain
understanding of the phenomena that occur.

In the ITT a slice of asphalt (D= 100 or 150 mm and t=20 to 25 mm) is subjected to a
load or deformation along its diameter. The load is applied at the top, on a small loading
plate and the specimen is supported at the bottom (Figure 8.8a). The applied
deformation and load as well as the deformation along the diameter perpendicular to the
load are measured. The specimens were subjected to a monotonically increasing
deformation rate of 0.85 mm/sec over a length of 14 mm at the top at a temperature of
15oC.

The mesh used for the simulations, again 1/8 of the actual specimen is shown in Figure
8.8b. The loading strip used to apply the deformation is also modelled and the contact
between specimen and strip is modelled by an interface. The steel block consists of
linear-elastic elements with a Youngs modulus of 220x103 N/mm2. The normal stiffness
of the interface is equal to the stiffness of the steel and the in-plane stiffnesses of the
interface (both directions) are set to 0.1 N/mm2 to allow the material to slide from

182 Chapter 8
underneath the loading strip when it expands. The asphalt itself is modelled as a non-
linear, temperature and strain rate dependent material, using the model described in the
previous chapter. Since the stresses and strains in the ITT vary from point to point, the
strain rate throughout the specimen is also non-uniform. As a result, the response will
vary from point to point because of the variation in stress state and strain rate.

Figure 8.8: Specimen in the indirect tension set-up (left) and mesh of the ITT (right)

In order to demonstrate some salient features of material response simulation, analyses


results will be presented in the following obtained by selectively (de)-activating
particular features of the model. It is hoped that on the basis of these results, it will
become possible to demonstrate the highly interdependent nature of fundamental
material response phenomena like cracking and yielding and also, to stress the
importance and the difficulty of incorporating them in a unified constitutive material
model.

8.4.1 RESPONSE SIMULATION VIA THE HOFFMAN CRITERION


Scarpas et al (1997) proposed the utilisation of smeared cracking for the simulation of
the tensile response of asphaltic materials. The model is also known as the “cohesive
crack” model and was originally proposed by Hillerborg (1979) for simulation of the
fracturing response of plain concrete. In CAPA-3D, the Hoffman criterion (Chapter 7) is
the three-dimensional computational implementation of the “cohesive crack” model.

As a first step in the evaluation of the contribution of the various fundamental response
phenomena in overall asphaltic concrete response under complex loading conditions,
finite element analyses were performed of the ITT specimen by de-activating the Desai-
based part of the model and specifying Hoffman cracking as the only inelastic material
phenomenon. The successive stages of cracking development at the centre of the ITT
specimen as they are predicted by the Hoffman criterion are shown in Figure 8.9. The

Model verification and application 183


cracking shown is the post-peak in-elastic deformation in the direction perpendicular to
the crack plane.

(a) (b)

(c) (d)
Figure 8.9: Successive stages of cracking development at specimen centre-line
Figure 8.10a shows the distribution of horizontal stresses in the specimen at the point of
crack initiation ( Figure 8.9a) before cracks occur in the centerline of the specimen. As
expected, high tensile stresses occur at the specimen centre. Figure 8.10b shows the
distribution of horizontal stresses in the specimen corresponding to the cracking
distribution of Figure 8.9b. Because of cracking, the tensile stresses in the region near

184 Chapter 8
the centre of the specimen start decreasing. Figure 8.10c and d show the distribution of
horizontal stresses in the specimen corresponding to the cracking distributions of
Figure 8.9c and d, respectively. From these figures it can be seen that additional
reduction of the tensile stresses along the centre line of the specimen occurs.

compression
0
temsion
(a) (b)

(c) (d)
Figure 8.10 Horizontal tensile stress release at the centre of the specimen due to
cracking ………………………………………………….

Model verification and application 185


From the above it is clear that using only the Hoffman criterion leads to the prediction
of crack initiation almost exclusively along the specimen centre line. These cracks
eventually grow towards the area directly beside the loading strip. In reality, the high
shear stresses that exist in this region will initiate damage early on in the test.
The inability of the Hoffman criterion to measure damage other than tension cracking is
considered a major disadvantage and raises questions about recent attempts in literature
to investigate asphalt concrete response by the exclusive use of “cohesive cracking”
models. Such attempts can only be successful in the investigation of rather simplistic
states of stress not representative of those typically encountered in pavement
engineering practise.

8.4.2 RESPONSE SIMULATION VIA THE DESAI CRITERION


Finite element analyses of the ITT specimen were also performed by de-activating the
Hoffman criterion and specifying the Desai surface criterion as the only source of
inelastic response. The total damage development in the body of the specimen at load
annihilation is shown in Figure 8.11.

As can be seen frm these pictures the Desai-based criterion predicts the shear-
compression damage next to the support that is typically encountered in ITT tests and
that was also observed in the TU Delft tests.
The constitutive model was also capable of predicting the large dilation of the
compression region of the specimen, Figure 8.11e and f, a feature that was also
observed in the TU Delft tests. As a result of this dilation and the associated contraction
of those regions of the specimen that are subjected to tension, warping of the specimen
faces occurs. Warping is accompanied by the generation of a self-equilibrated field of
stresses along the thickness direction of the specimen. The presence of this stress field
further complicates interpretation of the specimen response.

Very little damage localization criterion along the centre line of the specimen is
predicted by means of the Desai model. This contradicts experimental evidence and
raises questions about the utilisation of isotropic failure criteria for the simulation of
discrete fracture associated with cracking.

186 Chapter 8
(a) (b)

(c) (d)

εpeq

(e) (f)
Figure 8.11. Total damage distribution predicted via the Desai response surface
(displacement magnification factor is 1) …………….

Model verification and application 187


8.4.3 RESPONSE SIMULATION VIA THE ACRE CONSTITUTIVE MODEL
Finite element analyses of the ITT specimen performed by means of the proposed
ACRe constitutive model, combining the Hoffman-based cohesive-crack approach with
the Desai-based model, are presented in this Section.

In similarity to the predictions of the Desai criterion, a dominant field of shear stresses
is predicted in the area adjacent to the loading strip, Figure 8.12.

negative
0
positive

Figure 8.12. Shear-stress distribution (no displacement magnification factor)


The presence of a dominant shear field in combination with the vertical and horizontal
stresses results in two different modes of damage in the regions of the specimen near the
loading strip.
As shown in Figure 8.13 (left), an overall deformation pattern develops similar to that
predicted by the Desai criterion and typical of highly sheared regions. At the same time,
the presence of the field of shear stresses activates the Hoffman criterion, which
indicates the development of cracking in the same as above region of the specimen,
followed by cracking along the centre line of the specimen in a later stage of the test
(Figure 8.13, right).

188 Chapter 8
w
εpeq

………………

Figure 8.13. Prediction of total damage (left) and crack opening (right) via the
ACRe model …………………………………………

Model verification and application 189


8.4.4 COMPARISON OF THE SIMULATION RESULTS WITH LAB TESTS
The combination of cracking and shear-compression damage in the loading regions of
ITT specimens is well documented in literature and was also typically observed in the
TU Delft tests. Also the warping predicted by the Desai-based model component due to
the compressive stresses near the loading area was observed in the tests. Typical
examples of the damage seen in the specimens are shown in Figure 8.14.

Figure 8.14: Damage patterns observed in ITT laboratory tests


When the damage patterns shown in Figure 8.14 are compared to those predicted
numerically, it can be seen that it is indeed important to describe both the localised
damage that occurs due to tensile loads and the distributed compression damage in order
to fully capture the response of asphalt concrete. It is shown that the ACRe model can
correctly predict the response in a test in which different damage types interact,
although that test was not used as input for the model. Because the analyses presented in
this section are dynamic analyses, reflection at the mesh boundaries occurs. In the near
future it will be varified whether or not this has an effect on the observed response by
simulating the tests presented here using larger meshes.

By simulation of more tests and test conditions, additional verification of the model is
achieved. By studying the damage development in simulations that agree well with the
observed response, more insight in damage mechanisms will be obtained. Naturally,
also the damage development due to repetitive loading on pavement structures can be
simulated and analysed with this model. In an earlier stage of the project it was already
shown that the model accurately predicted the different damage patterns that are
observed for stiff and flexible pavement structures, respectively (Erkens et al. 2000).

8.5 A SIMPLIFIED MODEL APPLICATION


The tests described in this thesis provide information on the monotonic material
response, from the virgin material to ultimate failure. This information is used as input
for an elaborate material model that is capable of describing all these aspects of asphalt
concrete response. To fully exploit the capabilities of such a model advanced numerical

190 Chapter 8
methods such as the Finite Element Method (FEM) are needed. The possibilities of such
an approach are illustrated in the previous sections. It allows the study of any geometry
and loading and provides insight in the stress patterns and damage initiation and
development. As such, it is an indispensable tool to gain insight and knowledge about
the performance of asphalt concrete. However, non-linear FEM analyses are time
consuming and require specialised knowledge, both to run the software and to interpret
the results. Since the required time and expertise is not always available, it is important
to note that the scheme developed in the ACRe project can also be applied in a
simplified manner. Naturally, this does not provide as much information, but it is
definitely a powerful extension of the available analysis methods for asphalt concrete.
In this section that simplified approach will be described and illustrated.

8.5.1 TESTING
As explained in the previous chapter, most of the model information can be obtained
from the tension and compression tests. If only failure is considered, neglecting the
hardening phase and not considering the softening response, the series of successive
response surfaces reduces to a single surface, corresponding to α=0. Under the
assumption that β is also zero, an assumption which was proven for the ACRe mixture
and which will be acceptable for any asphalt mixture in a simplified approach, only two
more parameters need to be determined, R and γ. These can be found from a straight
line through the tensile and compressive strength values when plotted on the I1-√J2
plane.

To obtain only those strength values, one can use a simpler test procedure than for the
determination of all the model parameters. It is, off course, still of utmost importance
that the set-ups are reliable and that the tests are performed accurately, but the
instrumentation can be much simpler. In case of the compression tests this means that
the friction reduction is still applied, but the radial measurements can be omitted since it
is only the strength that has to be measured.

From the results presented in Chapter 4 it can be concluded that for sufficiently high
specimens (h/D>2) the friction reduction could even be omitted. For the tension tests
the simplification means a return to a cylindrical specimen shape, if necessary with PVC
blocks to increase the glueing surface and ensure a gradual load introduction. Using a
device developed in an earlier stage of the ACRe project (Scarpas et al. 1997) the
tension tests can even be carried out on a compression set-up (Figure 8.15). Also, the
strain gauges are omitted.

Model verification and application 191


Figure 8.15: Specimen geometry in the simplified tension test

From these tests, the tensile (ft) and compressive strength (fc) at various temperatures
and strain rates are obtained. By plotting them on the I1-√J2 plane, R and γ can be
determined as the intercept with the I1 axis and the slope of the line through fc and ft.

8.5.2 APPLICATION
Once the ultimate surface is known, it can be used to approximate the strength of the
material under any stress state or to evaluate the safety of a stress state with respect to
failure. This is illustrated in this section for both monotonic and cyclic tests.

8.5.2.1 Evaluating monotonic test results


In Chapter 3 the indirect tension test (ITT) or three point bending test (TPBT) were
discussed as tests that are often used to obtain an indication of the tensile properties of a
mixture. The state of stress in both tests, based on the assumption of linear-elastic
behaviour, is shown in Figure 8.16. Such a representation is, of course, limited. The
stresses in these tests vary throughout the specimen and the specimen response is the
result of the resulting combination of states of stress. Because this multitude of stress
states can not be taken into consideration in order to determine the failure strength, the
stress at the critical point is selected. For the ITT this is the stress state at the centre of
the specimen, for the TPBT that at the bottom of the beam, at midspan.

When the failure stress is reached at this position, it does not mean that the specimen
has actually failed, since the remainder of the specimen has at that point not reached the
critical load so there is still carrying capacity left. Response degradation of the material
in the damage zone and stress redistribution throughout the specimen will ensure that
the actual failure load is higher than the one determined on the basis of reaching the
failure stress in the critical location. As a result, the predicted load will be on the safe
side, marking the initiation of damage rather than actual collapse.

192 Chapter 8
How this works out for the load carrying capacity depends on the test geometry. For the
ITT, for example, the tensile stress at which damage is initiated at the critical location
will be lower than the uniaxial tensile strength. This is due to the tension-compression
state of stress that exists in this location. But in a uniaxial tension test the critical stress
is reached in the complete cross-section, which means first of all that if a flaw is present
somewhere in that cross section, failure will be triggered. Secondly, there is no
opportunity for stress redistribution in the specimen. Because this does happen in the
ITT, and because the critical stress state occurs only in a limited part of the specimen
cross section, which limits the change that a flaw is present, the load carrying capacity
may still be higher.

In the three point bending test (TPBT) the outer fibre is loaded in uniaxial tension, but
again this is a very limited region. Because the chance of a flaw in that region is much
smaller than in the case of the uniaxial tension test, the stress at which damage is
initiated is higher for the bending test (Herzberg, 1996).

Tension compression
compression

σ yy, vert P
b
σ xx,vert t

σ yy,horz h
D
σ xx,horz
L
tension

P
-6P
x
2P πDt
y πDt
P/2 L/2
1/6bh2

Figure 8.16: State of stress at critical location of the ITT and TPBT
On the basis of these stress states at the critical locations, values for I1-√J2 were
determined, plotted and compared to the failure line for corresponding conditions. This
was done for ITT tests on the ACRe mixture (Chapter 3 and Medani et. al 1999) and
TPBT results for a Sand Asphalt (SA) and Dense Asphalt Concrete (DAC) obtained
from Molenaar (1983). The test conditions and results are shown in Table 8.1. The
strain rates were determined from the deformation rates using the procedures described
in Chapter 3. It must be mentioned that the SA and DAC mixture are not exactly the
same as the ACRe mixture. However, the results described in Chapter 3 indicate that
these mixtures are quite similar so it is believed that the model parameter relations for
the ACRe mixture should also provide a useful indication in these cases.

Model verification and application 193


mix T def.rate strain rate stress type R γ
[oC] [mm/s] [1/s] [N/mm2] [N/mm2]
ACRe 15 0.85 0.00306 1.4 ITT 4.6 0.08
ACRe 15 0.1 0.00036 0.66 ITT 14.7 0.04
SA 15 0.85 0.001259 0.8 TPBT 10.7 0.05
DAC 15 0.85 0.001259 0.8 TPBT 10.7 0.05
Table 8.1: Test conditions, results and corresponding model parameters
Those states of stress that correspond to the failure conditions shown in Table 8.1 are
plotted in Figure 8.17 by larger markers. In the graph also the ultimate lines are shown
and it can be seen that for the ITT conditions the plotted points are indeed quite close to
the failure lines that correspond to the test conditions. For the TPBT, plotted as open
squares this is not the case. However, these tests were performed as fracture toughness
tests, using a notch in the critical area. In Chapter 3 the stress concentration relation for
the three point bending geometry was presented and that information is used to
compensate for the effect of the notch. The resulting stress states, which correspond to
the same load levels as the original values, are plotted as smaller filled squares. With
this correction, the failure stress is located outside the ultimate response area that was
determined on the basis of the uniaxial tests. This is due to the above described effect,
which causes the bending strength to be higher than the tensile strength.

ITT TPBT
sqrt(J2) 5 15'C&0.3%/s(ITT) TPBT-notch
2 15'C&0.036%/s(ITT) 15'C&0.126%/s (TPBT)
[N/mm ] 4.5

4
3.5
3
2.5
2
1.5 ITT
1
TPBT
0.5
ITT
0
2
-5 0 5 10 I1 [N/mm ] 15

Figure 8.17: Stress at critical location in the ITT and TPBT versus ultimate surfaces
By specifying a minimum distance between the failure strength from the ITT and the
corresponding flow surface a criterion for mixture evaluation is obtained. This distance
is an indication of the “safety against failure”. However, from the above examples it
will be clear that in order to evaluate the strength from tests with a non-uniform internal
state of stress still an empirical relation that links the test geometry to the response
found for that stress state is needed.
However, once a criterion for, for example, the ITT is known and a mixture has been
characterised using the tests discussed in the previous section, the ultimate line for that
mixture can be determined for any combination of temperature and strain rate. Thus, an
evaluation criterion for that mixture is available and for future evaluation only the

194 Chapter 8
simple test (ITT) on the mixture needs to be performed and checked against the
available ultimate parameters.

8.5.2.2 Evaluating cyclic test results


Because of the typical cyclic nature of the traffic loads, cyclic tests are often used in
road engineering to determine the fatigue behaviour of a mixture. As discussed in
Chapter 3, the four-point bending fatigue test (FPBFT) is the test that is used for this
purpose in the Netherlands. Using the same approach as in the previous section the state
of stress in the critical location is determined. For a four-point bending test this is:
M ± Pa
σ crit = = (8.3)
W 1 bh 2
6
Where: P= the applied load
a= distance between load and nearest support (118.5 mm in this case)
b= specimen width (50 mm)
h= specimen height (50 mm)
Because the test is performed using a full sine deformation signal, the outer fibres of the
beam are alternately loaded in tension and compression. Since the material is weaker in
tension than in compression, the former loading case will be the critical one. The sine
signal makes it more difficult to determine the strain rate, because this is not constant.
Under the applied deformation, the strain (and stress) in the outer fibres also varies in a
full sine and as a result the strain rate follows a cosine. Because it is the intention to
establish a simple analytical application, a simple method is used to determine a
representative strain rate (Figure 8.18). The sine is converted to a saw tooth signal with
the same maxima and frequency. The strain itself is, in an absolute sense,
underestimated except for the minima, maxima and cross-sections with the time-axis,
but the strain rate (the dotted horizontal lines) is over-estimated for part of the test and
under-estimated for the remainder. This is considered to provide an acceptable
approximation. The strain rate used to determine the ultimate surface can be computed
as:
ε& = 4εˆ f (8.4)
Where: εˆ =strain amplitude
f =strain signal frequency
As a result, each strain level corresponds to a different ultimate line. This is easily
understood by considering that, for lower strain levels one would expect a longer fatigue
life.

Model verification and application 195


cosine sine approximation rate approximation
1
0.8
0.6
0.4
0.2
0
-0.20.00 3.14 6.28
-0.4
-0.6
-0.8
-1

Figure 8.18: Conversion from a sine signal to a saw tooth signal


The fatigue test results obtained for the ACRe mixture, along with their corresponding
ultimate lines are shown in Figure 8.19. The test conditions and the model parameters
used to obtain this plot are presented in Table 8.2.

T [oC] ε [µε] 2
N [#cycles] rate [%/s] fc [N/mm ]
2
ft [N/mm2] R [N/mm ] γ
15 650 25900 2.6 -15.0 5.2 15.7 0.08
15 550 32400 2.2 -14.3 5.1 15.9 0.07
" " 54760 2.2 " " " "
15 450 79920 1.8 -13.5 5.0 16.0 0.07
" " 135430 1.8 " " " "
15 350 320890 1.4 -12.6 4.9 16.2 0.06
" " 284810 1.4 " " " "
15 250 822780 1 -11.4 4.8 16.3 0.06
Table 8.2: Test conditions and corresponding model parameters for the FPBFT

10
sqrt(J2) [N/mm 2]
9 FPBFT
ultimate surface 2.6%/s
8 ultimate surface 2.2%/s
ultimate surface 1.8%/s
7
ultimate surface1.4%/s
6 ultimate surface 1%/s
.
ε 5

2
I1 [N/mm 2]
1

0
-20 -15 -10 -5 0 5 10 15 20

Figure 8.19: Data from FPBFT compared to their ultimate response lines

196 Chapter 8
The distance between each data point and the corresponding ultimate response line is an
indication of the safety of that stress state. This distance is measured along the line
through all the data points (Figure 8.19). Under the assumption of linear-elastic
response until failure the maximum stress will increase along this path with increasing
load. Since this distance (indicated by the distance ∆ in the graph) is an indication of the
safety margin, it must be related to the number of load repetitions until failure. The
failure criterion used in the four-point bending fatigue test is a 50% reduction of the
stiffness while the ultimate response line indicates the maximum carrying capacity of
the material. Assuming that under all conditions 50% stiffness reduction corresponds to
a similar amount of damage, a relation between ∆ and the number of load repetitions
can be developed. In Figure 8.20 a plot of ∆ versus the number of load repetition until
failure is presented, along with a relation between the two. Ideally, such a relation
should go through ∆=0, N=1. A direct fit yields the following relation:

N = 1.0147e 2.9227 ∆ ; R 2 = 0.992 (8.5)

It can easily be seen that with a minor adaptation this relation fulfils the above-
mentioned condition. As a result the relation plotted in the graph is:

N = 1.0 e 2.9 ∆ (8.6)

Once general expressions for the model parameters for a mixture are available, this
information can be used over and over again. For fatigue results determined using
another geometry or failure criterion, a similar relation can be developed. This provides
a platform for the development of useful comparisons between different tests,
something that can not be done using the existing analysis methods.

N [#cycles]
1.E+07

1.E+06

1.E+05

1.E+04


1.E+03
0.00 1.00 2.00 3.00 4.00 5.00 6.00

Figure 8.20: Relation between ∆ and N, data point and Equation 8.6

8.6 CONCLUDING REMARKS


The model was used to simulate both some of the laboratory experiments performed for
this project and the Indirect Tensile Test (ITT). On the basis of the former it could be

Model verification and application 197


concluded that the model is capable to describe the main types of response observed in
asphalt concrete very well. The latter demonstrated that, using Finite Element and a
realistic material model, any test or structure can be simulated and analysed. The
simulated tests agreed well with actual laboratory responses and the simulation results
also provided insight in what actually happens in these tests.
By simulation of other tests and test conditions, additional verification of the
model is achieved. By studying the damage development in simulations that agree well
with the observed response, more insight in damage mechanisms will be obtained.
From the last part of the Chapter it can be seen that the approach that is
described in this thesis is not only a powerful research tool. Using a simplified version
in combination with elastic analysis, it provides a useful reference frame for mixture
evaluation. Once the model parameters for a mixture are known, the analyses presented
in the previous sections can be repeated for any test. This allows not only the
development of response criteria for simple tests, it also opens possibilities to compare
the results from various tests within an independent reference frame. In view of the
current trend towards performance-based specifications, these approaches may well turn
out to be of great importance.

8.7 REFERENCES
Erkens, S.M.J.G., Liu, X., Scarpas, A. and Molenaar, A.A.A., (2000), 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
Erkens, S.M.J.G., Liu, X., Scarpas, A. and Kasbergen, C., (2002), Issues in the
Constitutive Modeling of Asphalt Concrete, Paper presented at the 3rd
International Symposium on 3D Finite Element in Pavement Engineering,
Amsterdam, the Netherlands

Hillerborg, A., The Fictitious Crack Model and its Use in Numerical Analysis,
International Conference on Fracture Mechanics in Engineering Application,
Bangalore, 1979
Medani, T., (1999), A Simplified Procedure for Estimation of the Fatigue and Crack
Growth Characteristics of Asphalt Mixes, MSc. thesis International Institute for
Infrastructural, Hydraulical and Environmental Engineering in Delft, TRE 081
Scarpas, A., Gurp, C.A.P.M. van, Al-Khoury, R.I.N. and Erkens, S.M.J.G. , (1997),
Finite Element Simulation of Damage Developement in Asphalt Concrete
Pavements, 8th International Conference on Asphalt Concrete Pavements, Seattle,
Washington, USA, 1997

198 Chapter 8
9. CONCLUSIONS AND RECOMMENDATIONS
“The truth is the one thing that nobody will believe”, George B. Shaw (1856-1950)

9.1 INTRODUCTION
In this chapter the conclusions and recommendations that resulted from the research
described in this thesis are presented. Besides these general conclusions the various
stages in the project also resulted in conclusions and recommendations related to
specific topics. These sub conclusions and recommendations are presented at the end of
the chapters to which they relate, but for convenience sake they are listed at the end of
this chapter as well. A more elaborate description can be found in the respective
chapters.

9.2 GENERAL CONCLUSIONS AND RECOMMENDATIONS

9.2.1 GENERAL CONCLUSIONS


To arrive at new, generally applicable design procedures for pavements and asphalt
mixtures a better understanding of asphalt concrete response is necessary. The work
presented in this thesis clearly shows that to obtain that understanding one needs both
sophisticated tests and realistic material models. Sophisticated tests that result in
uniform states of stress are needed to characterise the mixtures and a material model,
fed with the information from the tests, is necessary to apply the knowledge on asphalt
concrete characteristics to pavement analysis. Also, the material models can relate the
response obtained from quality control tests to the actual material characteristics,
providing a reliable basis for performance based specifications.

The ACRe material model is a plasticity model, based on the model developed by Desai
for granular materials that was adapted to express the response of dense asphalt concrete
as a function of the three dimensional state of stress, the temperature and the strain rate.
Furthermore, it distinguishes between the distributed damage that results from
compressive loading and the localised damage-planes that occur due to tension. The
interaction between these two forms of damage in case of tension-compression states of
stress is also taken into account.

The model was used to simulate some of the laboratory tests that were used as input for
the model and it proved that the observed response was described accurately.
Simulations of the Indirect Tensile Test, which was not used as input for the model,
showed that an accurate prediction of the type and location of damage, as it is observed
in laboratory tests, is obtained only if both tensile and compressive damage and their
interaction is taken into account.

In order to provide the necessary input for the material model, tests that result in an a-
priory known, simple and preferably uniform state of stress were needed. For this
reason, a uniaxial compression and tension set-up for the testing of asphalt concrete

Conclusions and recommendations 199


were developed. Together, these tests provide nearly all the information that is needed
for the model.
The input for the tension response is taken from the tension tests and the input for the
basic three-dimensional model is determined on the basis of both the tension and
compression tests.
The difference between tension and compression response already results in a state of
stress dependent response, but to investigate additional confinement sensitivity of the
material a third test, the four-point shear test, was developed. In this test the effect of
increasing confinement on the shear strength is tested. For the asphalt mixture used in
this project, this test did not result in additional information. However, for open
mixtures, which are more sensitive to confinement, this could be different.

9.2.2 GENERAL RECOMMENDATIONS


Since the complicated interaction of geometry, climate, loading and material
characteristics that determines the pavement performance can never be captured in a
single test, future developments should aim at developing and applying tests for
material characterisation. That information should then be used as input for material
models that will allow the simulation of pavement performance. Once the tools are
available, studying the effect of changes in any of the influences on pavement
performance will be possible and much less time and money consuming than it would
be experimentally.
Since tests for material characterisation are generally not suited for quality control
purposes, separate tests should be employed for this aim. The material models can be
used to analyse these tests and relate their output to the actual material characteristics
that control pavement response.

Now that a realistic material model is available it should be used to analyse more tests
and test conditions to see if the predicted and observed response also match in those
case. By studying the damage development in these simulations, a better understanding
of the damage mechanisms in asphalt concrete will be obtained.

Having a test programme for material characterisation, it should now be applied to all
commonly used mixtures in order to allow the investigation of their sensitivity and
resistance to specific types of damage.
Furthermore, it is recommended to extend the test program described in this dissertation
with cyclic tests. The current tests are all monotonic, leading to failure in a single load
“repetition”, but pavement failure is usually due to a large number of small loads. In
order to run useful cyclic tests, the monotonic response must be known. This will allow
the application of load levels at a percentage of the monotonic strength. Using the
monotonic strength data in combination with plasticity theory already enables some
cyclic analyses, but to incorporate the cyclic response degradation into the material
model it is necessary to develop cyclic versions of the ACRe tests.

9.3 SUB CONCLUSIONS AND RECOMMENDATIONS

9.3.1 SUB CONCLUSIONS


Related to the uniaxial compression test (Chapter 4):
• Friction in the contact surfaces (barreling) and temperature distribution have a large
impact on the test results.

200 Chapter 9
• The uniaxial compressive strength tends to zero for high temperatures and strain
rates and to a maximum plateau strenght for high strain rates and low temperatures
• The response varies from brittle to plastic with changes in temperature and strain
rate, which sets high demands on the set-up.
• Aluminum turns out to be too soft a material to use for loading plates in
compression test with effective friction reduction.

Related to the uniaxial tension test (Chapter 5):


• Hinges in a tension set-up ensure proper alignment and ensure that the influence of
the set-up on the result in minimised, but they also result in play in the set-up.
• The parabolic specimen shape, used to fix the crack region in order to be able to
measure the unloading after the peak with strain gauges, works very well. The shape
does however result in additional variation in the composition of the failure zone
because of the more complicated compaction process in combination with the lack
of temperature control during gyratory compation.
• The bitumen skin on the specimen surface doesnot interfere with the strain gauge
measurements
• Because the specimens are glued to the end caps, the deformation at the specimen
ends is restrained. This results in a three-dimensional tensile state of stress that is
not suited to determine the material stiffness.

Related to the multiaxial test (Chapter 6):


• For the ACRe mixture the effect of confinement on the shear strength proved to be
independent of temperature and strain rate. Also, the effect of confinement for this
material did not differ from what could be predicted on the basis of the tension and
compression tests, which means that for the ACRe mixture β=0.

Related to the ACRe material model (Chapter 7):


• The stiffness values obtained from the tension tests proved to be considerably higher
than those found from compression. This could be traced to the end effect due to the
caps. These cause a 3D state of stress that disturbs the stiffness determination. Since
fracture occurs at a distance from the caps, this is not effected.
• The isotropic response degradation that results from compressive loads and the
localised damage due to tension are descirbed separately. Alle model parameters are
expressed as functions of temperature and strain rate on the basis of the test
programme described in this thesis.

Related to model verification and application (Chapter 8):


• The response observed in the uniaxial tension and compression tests is described
accurately by the model.
• The response in the indirect tension test (ITT) can also be described quite well.
Simulations of this test prove that both tension and compression damage need to be
modelled to describe the response observed in laboratory tests.

9.3.2 SUB RECOMMENDATIONS


Related to the uniaxial compression test (Chapter 4):
• Close attention should be payed to the end effects and other possibel disturbances
when designening tests and test set-ups. In compression testing, special attention

Conclusions and recommendations 201


must be focussed on minimising the friction in the contact planes. In general when
testing asphalt, the temperature should be maintained and monitored carefully.
• To facilitate testing a wide range of temperatures and strain rates on one set-up, it
should consist of a rigid framesuited for the most extreme conditions and have a test
controller that allows it to be used and calibrated in several ranges.
• Aluminum should not be used for the loading plates and if it is used in other parts of
the set-up because of its corrosion resistance, the fact that it is three times as flexible
as steel should be carefully considered in the design.

Related to the uniaxial tension test (Chapter 5):


• Despite the play they cause, it is recommended to use hinges in a tension set-up in
order to minimise the influence of the set-up on the observed response.
• If the temperature of the material in the mould could be maintained during gyratory
compaction, the variation in the composition of the failure zone could be reduced
considerably.
• Although strain gauges worked well in measuring the unloading of the ACRe
mixture, it will not work for porous mixtures. For such mixtures non-contacting
measurement systems of sufficient accuracy are needed. If these are available, they
would also be an interesting alternative for dens mixture because of the labour-
intensive task of accurately attaching the strain gauges.
• In order to arrive at a reliable stiffness for the material, tests with negligible end
effects, such as the compression test, should be used rather than a tension test.

Related to the multiaxial test (Chapter 6):


• It is expected that the effect of confinement will vary with the type of mixture, in
particular porous mixtures are expected to be much more confinement sensitive than
the dense ACRe mixture. It is therefore recommended to maintain the four-point
shear test as a component of the ACRe mixture characterisation programme.

Related to the ACRe material model (Chapter 7):


• Stiffness values should either be obtained from strain measurements at a distance of
the end caps or from compression tests with an effective friction reduction system.
• For any material the test conditions (temperatures and strain rates) and number of
repitions must be selected in such a way that the resulting data-grid can be used to
determine general expressions for the model parameters. This is necessary to use the
model for arbitrary conditions later on.

Related to model verification and application (Chapter 8):


• Since tension and compression damage differ it is logical to use tension and
compression tests as the basis for the characterisation programme. The four-point
shear test can be used to obtain additional information.
• By simulating more laboratory tests and pavement structures, the insight in the
occuring damage mechanisms will further increase, which will eventually result in
design procedures that directly address these damage mechanisms.

202 Chapter 9
CURRICULUM VITEA

April 6th 1971: Born in Nijmegen as Sandra Maria Johanna Grada Erkens

sept.1983-june 1989: VWO-ß, Stedelijke Scholengemeenschap Nijmegen


(SSGN)

sept. 1989-june 1994: Civil Engineering, Delft University of Technology.


Graduated at he Section of Structural Mechanics on the
numerical simulation of cracking in masonry due to
prevented shrinkage. Obtained an additional exam from
the Section of Road Engineering.

July 1994 – dec. 1996: Research Engineer at the Road&Railway Research


Laboratory of the Delft University of Technology, focused
on the study of fatigue in Gravel Asphalt Concrete for the
Dutch Strategic Highway Research Plan (SHRP-NL) and
the preparation of the PhD. project.

Since Jan. 1997 : PhD. project at the Road&Railway Research Laboratory


of the Delft University of Technology, carried out in co-
operation with the Section of Structural Mechanics, which
resulted in this thesis.

Since Jan. 2002: Part time project leader at the Materials Laboratory of the
Road and Hydraulics Division of the Dutch Ministry of
Transportation (60%) and part time assistant professor at
the Road & Railway Research Laboratory of the Delft
University of Technology (40%).
STELLINGEN

a witty saying proves nothing - Voltaire (1694-1778)-


Wit is educated insolence - Aristotle (284-322 B.C.)-

1. Het testen van een temperatuurgevoelig materiaal als asfalt zonder de temperatuur
nauwkeurig te regelen onder de aanname dat het effect op de resultaten
verwaarloosbaar is getuigt van grenzeloos optimisme.
Testing of temperature dependent materials as asphalt concrete without accurate
temperature controll under the assumption that the effect on the results is negligible,
is a sign of unlimited optimism. [Hfstk. 4]

2. De invloed van wrijving in de contactvlakken bij drukproeven is bij platte


proefstukken zeer groot. Resultaten verkregen uit onderzoek waarbij deze invloed is
veronachtzaamd zeggen dan ook meer over de kwaliteit van het onderzoek dan over
dat van het onderzochte materiaal.
The influence of friction in the contact surfaces during a compression test on flat
specimens is very large. Results that are obtained from research in which this effect
is negelected give therefore more information about the quality of the research than
on that of the material. [Hfstk. 4]

3. Als er geen uniaxiale spanningstoestand heerst, kan de stijfheid van een materiaal
niet bepaald worden door simpelweg de spanning en de rek in een richting op elkaar
te delen.
If the state of stress is not uniaxial, the stiffness of a material cannot be determined
by directly dividing the stress and strain in a given direction. [Hfstk. 5]

4. Daar het gedrag van asfalt sterk afhangt van de aangebrachte steundruk, schiet de
huidige generatie rheologische modellen per definitie te kort.
Since asphalt concrete response strongly depends on confinement, the current
generation of rheological models is not suited to describe it.

5. Het onderling vergelijken van asfaltmengsels op basis van een proef bij één
temperatuur of belastingsnelheid/frequentie om de relatieve geschiktheid voor
toepassing in een wegconstructie te bepalen levert alleen schijninformatie op.
Comparing asphalt mixtures on the basis of a test at a single temperature or strain
rate/frequency in order to determine their relative suitability for application in a
pavement provides only phoney information.

6. Proeven die tot een uniforme (“simpele”) interne spanningstoestand leiden zijn
lastig om uit te voeren en te instrumenteren, terwijl eenvoudig uitvoerbare
(“simplistische”) proeven leiden tot ingewikkelde spanningstoestanden waardoor ze
moeilijk zijn te interpreteren.
Tests that yield a uniform (“simple”) internal state of stress are difficult to perform
and instrument, while tests that are simple to perform (“simplistic”) result in
complicated stress patterns that are difficult to interpret.
7. De roep om eenvoudig uit te voeren proeven zonder voorafgaand tijdrovend
fundamenteel onderzoek gaat voorbij aan het feit dat juist voor de interpretatie van
deze proeven fundamenteel inzicht in het materiaalgedrag vereist is.
The call for simple tests without time consuming fundamental research ignores the
fact that the interpretation of these tests requires fundamental knowledge about the
material response

8. Gedragsgebonden specificaties voor de wegenbouw zijn alleen dan haalbaar als er


naast proeven ook geïnvesteerd wordt in de ontwikkeling van materiaalmodellen.
Performance based specifications for road engineering are feasible only when
besides the development of tests also the development of material models is boosted.

9. Het ontwikkelen van een materiaalmodel zonder het te verifiëren is vergelijkbaar


met het moeizaam maken van een prachtige fotocompositie zonder ooit het rolletje
te ontwikkelen.
Developing a material model without verifying it can be compared to laboriously
aranging a beautifull photo composition without ever developing the film.

10. Met een stelling neemt men duidelijk stelling in een op inzichten gebasseerde
dicussie.
With a proposition one takes a clear position in a discussion based on insights.

11. Als een onderzoeker er van overuigd raakt alle antwoorden te hebben, wordt het tijd
om van baan (of onderzoeksgebied) te veranderen.
When a researcher believes to know all the answers, it is time to find another job (or
research area).

12. Promoveren zonder frustraties vraagt een zodanig grenzeloze naïviteit dat het
kritisch onderzoek onmogelijk maakt.
Obtaining a PhD without frustrations requires such naivity that it disqualifies a
person from performing critical research.

13. Gezien de multidisciplinaire aard van civieltechnische vraagstukken is het essentieel


dat civielers goede bruggenbouwers zijn.
Since civil engineering problems are multi-disciplinary, it is essential that civil
engineers are qualified bridge builders.

14. Zowel in onderzoek als bij volleybal presteert een team het beste als het bestaat uit
mensen met verschillende kwaliteiten en met voldoende onderling vertrouwen om
elkaar de bal toe te spelen.
In research as well as volleybal a team performs best when it consists of people with
different qualities and enough confidence in one another to “play ball”.

15. Het leven is een optimaliseringsproces met slecht gestelde randvoorwaarden.


Life is an optimisation process with ill-posed boudary conditions.

View publication stats

You might also like