You are on page 1of 396

Performance related

characterisation of the mechanical


behaviour of asphalt mixtures

ii

iii

Road and Hydraulic Engineering Institute

Performance related
characterisation of the mechanical
behaviour of asphalt mixtures

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,
in het openbaar te verdedigen op 20 januari te 10:30 uur
door
Jacobus Michal Maria MOLENAAR
metaalkundig ingenieur
geboren te Edam

iv

Dit proefschrift is goedgekeurd door de promotor:


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

Samenstelling promotiecommissie:
Rector Magnificus,
Prof. dr. ir. A.A.A. Molenaar
Prof. dr. ir. Ch.F. Hendriks
Prof. dr. R.L. Lytton
Prof. dr. U. Isacsson
Dr. ir. J. Zuidema
Dr. P.C. Hopman
Ir. L.A. Bosch
Prof. dr. ir. F. Molenkamp

Technische Universiteit Delft, voorzitter


Technische Universiteit Delft, promotor
Technische Universiteit Delft
Texas A&M University, Verenigde Staten
Royal Institute of Technology, Zweden
Technische Universiteit Delft
Netherlands Pavement Consultants
Rijkswaterstaat, Dienst Weg- en
Waterbouwkunde
Technische Universiteit Delft, reservelid

Dit onderzoek is gedeeltelijk in dienstverband bij de Dienst Weg- en


Waterbouwkunde van Rijkswaterstaat uitgevoerd. De auteur is de DWW zeer
erkentelijk voor de geboden mogelijkheid.

ISBN 90 369 5556 4


DWW-2003-129
Published by:
Dienst Weg- en Waterbouwkunde, Rijkswaterstaat
Road and Hydraulic Engineering Institute
P.O. Box 5044
2600 GA Delft
The Netherlands

Jaap Molenaar, Zoetermeer, 2003


All rights reserved. No part of the material protected by this copyright may be
reproduced or utilised 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.
Printed in The Netherlands

To Miente,
Lieuwke, Niels,
Beppeke, and Willemijn

vi

Keywords:
Asphalt mixture, performance related characterisation, mechanical
property, creep, permanent deformation, crack-growth, fracture,
constitutive modelling.

vii

Foreword
At this place I want to thank Cos van Teylingen, former head of the
department Realisation and Maintenance Infrastructure I am working in,
for stimulating me to write this thesis, Peter Hoogweg, director of the
Road and Hydraulic Engineering Division, for giving his consent, and
Andr Molenaar, professor at the Road and Rail Road Laboratory of the
Delft University of Technology, for his willingness to be my promotor. I
want to thank also my former colleagues and project leaders in the
Scientific Asphalt Research Project (Technisch Wetenschappelijk Asfalt
Onderzoek, TWAO), Harry Verburg and Rutger Krans. I want to thank Jos
van der Heide, representative of the Dutch Asphalt Producers and
Contractors Association, VBW-Asfalt, for his sharpening views from the
other side in the project team. I want to thank all those who contributed
directly or indirectly to the realisation of this thesis, those who participated
in or contributed to the TWAO-project teams, Arthur van Dommelen
(RHEI), Bernard Eckmann (first Exxon, later Nynas), Louis Francken
(Belgian Road Research Centre), professor Andr Gastmans (first Exxon,
later Nynas), Piet Hopman (first TU-Delft, later NPC), Andr Houtepen
(KOAC), Maarten Jacobs (first TU-Delft, later KOAC), Jasper van der
Kooij (RHEI), Piet Kunst (NPC), Hans Nugteren (RHEI), Ad Pronk
(RHEI), Carl Robertus (Shell), Theo Terlouw (Shell), Fedde Tolman
(NPC), Kees Valkering (Shell), Martin van de Ven (NPC), Ann
Vanelstraete (Belgian Road Research Centre), and Gerrit Westera
(KOAC). I want to thank Jan Zuidema (TU-Delft) for his cooperation and
his graduate students, Carel Kleemans, Jan Boone, Johan Schulte, and
Marlies Arbouw who were willing to do their Master of Science thesis on
crack-growth of asphalt mixture. I thank Ad Pronk for his critical reading
and comments on the analysis in appendix 1. And I want to thank Tom
Dingjan for his help with the illustrations.
The TWAO-project started in 1990. It was started because there was a
wide spread desire to develop so-called functional, or performance
related requirements for asphalt pavements and asphalt mixtures. The
contractors eagerly developed new products. However, the application of
new products was, and is, hindered by the difficulty that the traditional
requirements are not applicable to newly developed products. It was felt
that a fundamental approach to the characterisation of asphalt mixtures
was necessary to help find an answer to this difficulty. The TWAO-project
was intended to be scientific, to improve the knowledge and expertise of
the RHEI, but also to show that the RHEI is a road authority which
cooperates with the road building industry to enhance innovation. The
ambitious goal of the project meant among other things that new tests and

viii

test protocols had to be developed from scratch. Research into the


mechanical behaviour of asphalt mixture is particularly difficult for a
number of reasons. The behaviour is time dependent and stress dependent.
The material is heterogeneous, and the test specimens are relatively small,
compared to the material heterogeneity a necessity from a practical
viewpoint, but a disaster from a scientific (interpretative) viewpoint
which means that the interpretation of the test results is troubled by size
effects of the specimens. Scientific research can grow mature if conditions
are favourable. Looking back, the whole project was a big adventure,
because the research infrastructure had to be developed, and, more
importantly, expectations among those from whom support of the project
was needed were diverse. In spite of that, management took the visionary
decision to support the project. Meanwhile, the 150 million dollar United
States Strategic Highway Research Project was finished by the end of
1993, leaving results that were received with mixed enthusiasm in both the
United States and in Europe. In Europe, Technical Committees of CEN
were struggling to achieve consensus about the testing conditions of
performance related tests for asphalt mixture. The whole atmosphere
around performance related specification of bituminous materials
sometimes breathed an air of disappointment.
The TWAO-project ended 1997. After I had written a 70 page summary
report about the activities and results, Cos asked me if I was interested to
write a dissertation about the subject. Although I had great doubts as to
whether that would be wise, I finally agreed to his proposition, expecting
that the effort already invested would be to my advantage. It turned out
that I was to learn that a scientific treatment of a subject demands a greater
depth of interpreting test results, and consequently a greater investment in
time than is considered acceptable in a customer-oriented organisation.
The cost of science is sometimes considered a burden rather than an
investment in the future. I hope this study will help to prove that the
opposite is true. I finished this study out of the conviction that it pays a
contribution to indispensable innovation in road building. I believe this
study shows it to be possible to evaluate or judge the cost-effectiveness
and the risk of failure of a newly developed paving material, based on a
characterisation based on constitutive models. Thus, I believe that the
acceptation of newly developed paving materials is feasible based on
laboratory testing and some accelerated pavement testing but can do
without a monitoring of its behaviour during the service-life. Then, this
study will ultimately prove to give value for money by its contribution to
the enhancement of innovation of paving materials.
I remember a discussion I had with Cos, saying how little requirements for
asphalt mixtures had changed over thirty years time, and that something
had to be done to draw attention to innovation of paving materials. After

ix

all, durable paving materials are an asset to our nations transport


infrastructure, economy, and prosperity.
Last but not least, I want to thank my wife and children for the countless
hours invested in this study instead of in our family life.

Zoetermeer,
December, 2002

Samenvatting
Het onderzoek is gedaan ter ondersteuning van innovaties op het gebied
van asfaltverhardingsontwerp en materiaalkeuze, om het van risico van
falen en de kosteneffectiviteit van nieuw ontwikkelde verhardingsmaterialen aantoonbaar te maken, en daarmee die materialen toepasbaar te
maken. Om het risico van falen en de kosteneffectiviteit te bepalen is
informatie nodig over de kwaliteit van de verharding. Om de kwaliteit van
de verharding te definiren is het nodig het gedrag van de toegepaste
materialen te kennen. Om het gedrag van de materialen te kennen is het
nodig de eigenschappen van de materialen te kennen, die voor het gedrag
in de weg relevant zijn. De volgende aspecten van het mechanisch gedrag
van asfaltmengsels werden onderzocht, omdat op basis van ervaring
bekend is dat het de bepalende fenomenen zijn: het viscolastisch en
viscoplastisch spanning-rekgedrag, en het scheurgroeigedrag. Daarbij
werden analytische en numerieke methoden toegepast.
Er zijn testmethoden ontwikkeld die voor gebruik in een praktische
context geschikt zijn om de stijfheidsmodulus, de weerstand tegen
permanente vervorming, en de weerstand tegen vermoeiing en scheurgroei
te bepalen.
Het onderzoek leidt tot de conclusie dat de volgende proeven geschikt zijn
voor de karakterisering van het mechanisch gedrag van asfaltmengsels in
een praktische context: een vierpunts buig frequency sweep test, ter
karakterisering van het lineair dynamisch viscolastisch spanningrekgedrag, een dynamische triaxiale kruipproef, ter karakterisering van het
niet-lineair dynamisch elasto-viscoplastisch spanning-rekgedrag, een
trekproef, ter karakterisering van de weerstand tegen scheurgroei, en een
breuktaaiheidsproef, ter karakterisering van de weerstand tegen breuk.
De genoemde eigenschappen zijn van belang voor de functionaliteit van
de verhardingsconstructie, die wordt uitgedrukt in draagvermogen,
oppervlakkenmerken en lange-termijngedrag.
Verder kan worden geconcludeerd dat het op basis van de genoemde
proeven mogelijk is, een systeem van gedragsgerelateerde specificaties te
ontwikkelen, op basis waarvan asfaltmengsels relatief op toepasbaarheid
kunnen worden getoetst, dat wil zeggen in vergelijking met standaardasfaltmengsels waarvan het gedrag bekend is.
Met analytische methoden kan sneller dan met empirische methoden
worden bepaald of verbeteringen van verhardingsmaterialen nut hebben,
en is het mogelijk sneller over de informatie te beschikken die nodig is om
het gedrag van een nieuw ontwikkeld, niet-gestandaardiseerd materiaal op
risico van falen en kosteneffectiviteit te beoordelen. Daarom zullen
analytische methoden de toepasbaarheid van innovatieve, niet-gestandaardiseerde verhardingsmaterialen versnellen.

xi

Summary
The investigation was undertaken to support innovations in the field of
asphalt pavement design and material selection, and to be able to evaluate
or judge the risk of failure and cost-effectiveness of newly developed
paving materials in order to justify their application. To be able to
determine the risk of failure and cost-effectiveness, information is needed
about the quality of the pavement. In order to define the quality of the
pavement, it is necessary to know the behaviour of the applied materials.
In order to know the behaviour of the materials it is necessary to know the
properties that are relevant for the behaviour of the material in the
pavement. The following aspects of the mechanical behaviour of asphalt
mixture were investigated, because it is known based on experience that
these are the important phenomena: the viscoelastic and viscoplastic stress
strain behaviour, and the crack-growth behaviour. Both analytical and
numerical approaches were followed.
Test methods were developed that are suitable for use in a practical
context for the determination of the stiffness modulus, the resistance to
permanent deformation, and the resistance to fatigue and crack-growth.
It is concluded that the following tests are suitable for the characterisation
of the mechanical behaviour of asphalt mixtures in a practical context: a
four point bending frequency sweep test, to characterise the linear
dynamic viscoelastic stress strain behaviour, a dynamic triaxial creep test,
to characterise the nonlinear dynamic elasto-viscoplastic stress strain
behaviour, a tensile test, to characterise the resistance to crack-growth, and
a fracture toughness test, to characterise the resistance to fracture.
Those properties are important to the functionality of the pavement
structure that is defined in terms of bearing capacity, surface
characteristics, and long-term behaviour.
It is concluded that it is possible, based on the tests mentioned, to develop
a set of performance related specifications, which will allow newly
developed asphalt mixtures to be tested for applicability relative to
standardised asphalt mixtures for which the behaviour is known.
Analytical methods will allow one to determine useful improvements to
paving materials faster than empirical methods, and to obtain the
information required to judge a newly developed and non-standardised
paving material for its risk of failure and cost-effectiveness. Therefore, the
use of analytical methods will facilitate the acceptation for application of
innovative, non-standardised, paving materials.

xii

About the author


The author was born in Edam, December 9, 1952. He attended the
Waterlant College in Amsterdam, studied chemistry and physics at the
University of Amsterdam, and received a diploma Kandidaatsexamen
Chemistry and Physics in 1973.
He then went to study metallurgy at Delft Technical University and
graduated as a Metallurgical Engineer, MSc, in January 1980. He served
the Royal Navy as a marine corrosion specialist, and returned to Delft
University as a research associate in September 1981. There he worked in
a research project aimed at refining recycled aluminium alloys by means
of rapid solidification, and specialised in the stirred solidification of semisolid Al-Cu alloy.
In September 1986, he joined the Road and Hydraulic Engineering
Institute and worked as a research project leader in the field of bituminous
construction materials. From 1989 until 1997 he was project leader of the
Scientific Asphalt Research Project aimed at developing performance
related test methods for asphalt mixtures. He started with the present study
in May 1998.

xiii

Contents
Samenvatting
Summary
About the author
List of symbols
List of units
List of abbreviations

x
xi
xii
xxi
xxiv
xxv

Chapter 1: Introduction
1

4
5

Motive for this study

1.1 Socio-economic developments


1.1.1 Control of development of transport infrastructure
1.1.2 The growth of road traffic
1.1.3 Sustained use of materials and energy
1.1.4 Design-Build-Maintain contracts
1.1.5 Road building, changing road authorities, and the
knowledge economy
1.2 Innovation and product quality The need for a rationalised quality
control methodology

1
2
4
5
7
7
9

Scope of this study

12

2.1 Introduction
2.2 This studys subject

12
14

Aim of this study

15

3.1 General objective


3.2 Practical goal
3.3 Research goal

15
15
15

Selected topics
Outcome of this study

16
16

5.1 General objective


5.2 Practical goal
5.3 Research goal

16
17
18

Importance of this study to the current practice

18

Chapter 2: The current pavement design methodology


1
2

Introduction
The current methodology of asphalt pavement design,
material selection, and asphalt mixture design

21

2.1 The pavement design method


2.1.1 Bearing capacity
2.1.2 Surface characteristics

21
22
23

21

xiv

2.1.3 Two functional requirements for pavements:


longitudinal evenness and skid resistance
2.1.4 Summary
2.2 Material selection
2.2.1 Asphalt concrete
2.2.2 Porous asphalt
2.2.3 Stone mastic asphalt
2.3 The asphalt mixture design method

24
24
24
25
28
31
31

Discussion and conclusions

33

Chapter 3: Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture
1
2
3

Aim
Methodology
Theory

37
37
38

3.1 Linear viscoelastic creep susceptibility of asphalt mixture


3.2 Viscoplastic creep susceptibility of asphalt mixture

38
42

Experimental details

42

4.1
4.2

43
43
43
44
45
45
48
49

4.3

5
6

Results of frequency sweep tests Complex modulus


Results of static creep tests in the absence of confinement
Dependence of the creep of asphalt concrete on time,
temperature, stress, and mixture composition
6.1
6.2
6.3

Test methods
Test set-up and testing conditions
4.2.1 Four point bending test
4.2.2 Static uniaxial compression creep test
4.2.3 Dynamic uniaxial compression creep test
4.2.4 Dynamic triaxial compression creep test
4.2.5 Modified friction reduction system
Materials

Dense graded asphalt concrete 0/16, unmodified


Dense graded asphalt concrete 0/16, polymer modified
Comparison of static creep of DAC 0/16, unmodified and
DAC 0/16, polymer modified

Results of dynamic creep tests in the absence of confinement


Dependence of the creep of asphalt concrete on time,
Temperature, stress, and mixture composition
7.1

52

58
59
61
63

65

Dense graded asphalt concrete 0/16, unmodified, using a


block-wave form of applied stress
66
7.1.1 Time dependence of creep properties constant rest-time
66
7.1.2 Time dependence of creep properties constant loading time 76
7.1.3 Temperature and stress dependence of creep properties
77

xv

7.2

7.3

7.4

7.5

7.6
7.7
7.8
7.9

Dense graded asphalt concrete 0/16, polymer modified, using a


block-wave form of applied stress
7.2.1 Time dependence of creep properties
7.2.2 Temperature and stress dependence of creep properties
Dense graded asphalt concrete 0/16, unmodified, using a
sinusoidal applied stress
7.3.1 Time dependence of creep properties
7.3.2 Temperature and stress dependence of creep properties
Dense graded asphalt concrete 0/16, polymer modified, using a
sinusoidal applied stress
7.4.1 Time dependence of creep properties
7.4.2 Temperature and stress dependence of creep properties
Block-wave form of applied stress versus sinusoidal applied stress
7.5.1 Time dependence of creep properties
7.5.2 Temperature and stress dependence of creep properties
Comparison of dynamic creep to static creep
The interrelatedness of J 10 and z of equation (21) and the
parallelism of the creep curves
Change of volume of the specimen
Barrelling

10

97
97
102
103
103
109
109
109
111
112
115
117
119

Results of dynamic creep tests in the presence of confinement


Dependence of creep properties on time, stress,
specimen height, and mixture composition
120
8.1 Creep susceptibility of dense graded asphalt concrete Transition
of creep according to eq. (15) to creep according to eq. (16)
8.2 Creep susceptibility of dense graded asphalt concrete and
porous asphalt Time and stress dependence
8.3 Change of volume of the specimen
8.4 Dependence on the specimen height

80
80
97

120
122
131
132

Results and interpretation of finite element computations of


the creep of a heterogeneous viscoelastic-viscoplastic creep
specimen
134
Modelling the time dependence of linear viscoelastic stress strain
behaviour
137
10.1
10.2
10.3
10.4

Linear viscoelastic stress strain behaviour in the Burgers model


Static creep
Dynamic creep using a block-wave form of applied stress
Dynamic viscoelastic stress strain behaviour, using a
sinusoidal wave-form of applied stress
10.5 Complex modulus and linear viscoelastic creep susceptibility
of asphalt mixture
10.6 Discussion
10.7 Limitations of the linear viscoelastic model

138
139
139
141
143
144
145

xvi

11
12

13
14

Discussion I Dependence of creep properties on the


specimen geometry (height)
Discussion II Relatedness of dynamic viscoelastic and
elasto-viscoplastic (creep) properties of asphalt mixture

149

12.1 Relatedness of complex modulus and creep properties


12.2 Creep of asphalt mixture in the absence of confinement
12.3 Creep of asphalt mixture in the presence of confinement

149
149
150

146

Discussion III Physical meaningfulness and predictive value of


viscoelastic and elasto-viscoplastic properties of asphalt mixture 151
Conclusions
153

Chapter 4: Characterisation of the resistance to crack-growth


and the resistance to fracture of asphalt mixture
1
2
3

Aim
Methodology
Theory

155
155
158

3.1 The stress intensity factor


3.2 ASTM minimum size requirements
3.3 Fatigue crack-growth and creep crack-growth
3.4 Indirect methods of characterising the resistance to crack-growth
3.5 Definitions of tensile strength

158
161
161
163
163

Experimental details

164

4.1
4.2

164

4.3

Test methods
Test set-up and testing conditions Influences of stress
condition, shape of the waveform of applied stress, and
specimen geometry
4.2.1 Crack-growth test using the centre-cracked tensile
(CCT) specimen
4.2.2 Crack-growth test using the four point bending
(4PB) specimen
4.2.3 Determination of the fracture toughness test using
the semi-circular bending (SCB) specimen
4.2.3.1 The ASTM-method for metals
4.2.3.2 A modified ASTM-method for asphalt mixture
4.2.4 Three tensile tests
4.2.4.1 Uniaxial tensile (UT) test
4.2.4.2 Semi-circular bending (SCB) test
4.2.4.3 Indirect tensile (IT) test
Materials Influence of the material composition
4.3.1 Sand asphalt test in the CCT-tests and in the 4PB-tests
4.3.2 Asphalt mixtures tested in the SCB fracture toughness
test and indirect tensile test
4.3.3 Asphalt mixtures used in the SCB tensile tests,

166
166
170
170
172
174
175
177
179
179
179
179
179

xvii

uniaxial tensile test and indirect tensile test


4.3.4 Asphalt concrete mixtures used in the indirect tensile test

Results and analysis of crack-growth tests


5.1
5.2

5.3
5.4
5.5
5.6
5.7
5.8

Results and analysis of fracture toughness tests


using the SCB-specimen
6.1
6.2
6.3
6.4

Obtaining a graph of da / dN versus K from raw test data


Crack-growth test using the CCT-specimen Influence of
frequency, applied stress, and specimen thickness
5.2.1 Constant load amplitude tests
5.2.2 Constant K test
5.2.3 Creep crack-growth tests
Crack-growth test using the 4PB-specimen Influence of
frequency and specimen thickness
Analysis of the parameters A and n of the Paris equation
Discussion
Prediction of dynamic crack-growth based on static
creep crack-growth
Discussion
Critical stress intensity factor

Dependence of the fracture toughness on the specimen size


Dependence of the fracture toughness on the displacement rate
Influence of the material composition
Analysis of the results
6.4.1 Dependence of the fracture toughness on the specimen size
6.4.2 Dependence of the apparent fracture toughness on the
specimen size at 15C
6.4.3 Validity of the fracture toughness
6.4.4 Dependence of the apparent fracture toughness on the
displacement rate

Results and analysis of tensile strength using the uniaxial


tensile (UT) specimen, the semi-circular bending (SCB)
specimen, and the indirect tensile (IT) specimen
7.1

7.2

7.3

7.4

Case 1: Fine porous asphalt Uniaxial tensile strength


and bending tensile strength Influence of temperature
and diplacement rate
Case 2: Porous asphalt Uniaxial tensile strength and bending
tensile strength Influence of mixture grading, bitumen
content, and type of bitumen
Case 3: Various asphalt concrete and porous asphalt
mixtures Uniaxial tensile strength, bending tensile
strength, and indirect tensile strength Influence of
temperature and displacement rate
Case 4: Asphalt concrete Indirect tensile strength
Influence of specimen size, loading strip, temperature,

180
180

181
181
186
187
193
195
195
195
208
212
214
216

219
219
219
221
221
221
228
228
231

233

233

234

236

xviii

and mixture composition


7.5 Analysis of the results
7.5.1 Uniaxial tensile strength and bending tensile strength
at low temperature (0C and lower)
7.5.2 Uniaxial tensile strength and bending tensile strength
at high temperature (15C and higher)
7.5.3 Indirect tensile strength at low temperature
(0C and lower)
7.5.4 Indirect tensile strength at high temperature
(15C and higher)
7.5.5 Summary
7.6 Relationship between fracture toughness and bending tensile
strength in the SCB-test

9
10

240
244
246
249
250
253
253
254

Finite element model of the SCB-specimen

256

8.1
8.2
8.3
8.4
8.5

256
257
257
258
261

Specimen geometry
Definition of damage
Material
Results of computations
Discussion and conclusions of computational results

General discussion
Conclusions

262
264

Chapter 5: Performance judgement of asphalt mixtures


1
2
3

Introduction
Composition-relatedness: An impediment to innovation in
road building
Property-related requirements for asphalt mixture The key
to enable innovation in road building

267

3.1
3.2

271

3.3

3.4

Definition of performance relatedness


Physical meaningfulness and predictive value of constitutive
model variables and parameters
Controlling pavement performance
3.3.1 Cost-effectiveness and risk of failure
3.3.2 Pavement performance
3.3.3 Performance related asphalt mixture properties
Summary

Asphalt mixture: The complicating factor in the development


of performance related requirements
4.1
4.2
4.3
4.4

Dependence of asphalt mixture properties on specimen size


and shape
The stress state of the specimen
Relative (qualitative) predictive value
Summary

268
271

272
273
273
281
281
283

284
284
285
288
288

xix

Interrelatedness of constitutive model variables and parameters 289


5.1
5.2
5.3
5.4
5.5
5.6

6
7

Complex modulus
Linear viscoelastic creep
Creep of asphalt mixture
Crack-growth in asphalt mixture
Fracture toughness and tensile strength
Summary

Discussion: Importance of the interrelatedness of constitutive


model variables and parameters
Conclusions

289
291
294
298
300
300

301
302

Chapter 6: General discussion


1
2
3
4

Introduction
General objective
Practical goal
Research goal

303
303
304
306

Chapter 7: General conclusions


1
2
3
4
5

Introduction
General objective
Practical goal
Research goal
Property-related requirements versus composition-related
requirements for asphalt mixture

309
309
310
313
313

Appendix 1: Influence of the shape of the wave-form of the


applied stress on the response strain of the Burgers model
1
2

Introduction
Analysis

315
316

2.1
2.2
2.3
2.4
2.5
2.6
2.7

317
320
320
321
322
323

Constant applied stress


Sinusoidal applied stress
Unidirectional sinusoidal applied stress (haversine)
Alternating block-wave of applied stress
Unidirectional block-wave of applied stress
Half sine waveform of applied stress
Influence of the shape of the waveform of the applied stress
on the creep strain rate

325

Detailed solutions of the Burgers equation

326

Case 1
Case 2
Case 3

327
329
332

xx

Conclusion

333

Appendix 2: Influence of the friction reduction system


1
2
3

Introduction
RHEI investigation into friction reduction
Discussion

Appendix 3: List of creep models

335
335
336
339

Appendix 4: The material model of the FEMMASSE finite


element model of a heterogeneous viscoelastic-viscoplastic
creep specimen
1
2
3

FEMMASSE finite element code


The aggregate grains
The bituminous matrix

347
347
348

Appendix 5: The Asphalt Concrete Response (ACRe) Model


1

The material model

351

1.1
1.2
1.3
1.4
1.5

351
351
353
354
355

Definition of damage
Plastic flow criterion
Simulation of the hardening process
Simulation of the degradation process
Simulation of crack-growth

Determination of material parameters

356

Appendix 6: Computer programme to compute the


da / dN K -relationship

359

Bibliography

361

xxi

List of symbols
a
A
A0
B
BTS
C
D
e
E
E1 , E 2
f
Fm
I
I1
ITS
J
J*
J
J
J0
J1
J 2
k
K
Kc
KI
K Ic
K IQ
m
n
N
Nf
P0
Pm
Qm
R
S
S*

crack-length
constant (coefficient of the Paris equation)
constant
specimen thickness
bending tensile strength
constant
specimen diameter
base of the natural logarithm
elasticity modulus
spring constants of rheological model
frequency
Marshall flow
constant
first stress invariant
indirect tensile strength
compliance
complex compliance
storage compliance
loss compliance
constant (instantaneous elastic compliance at t = 0
compliance at time t = 1 s
second deviatoric stress invariant
constant
constant
stress intensity factor
critical stress intensity factor
mode I stress intensity factor
critical mode I stress intensity factor, or fracture toughness
apparent fracture toughness (before test of validity)
slope of the complex compliance on log-log scale
constant (exponent of the Paris equation)
number of load repetitions
fatigue-life
pertinent force
Marshall stability
Marshall quotient
stress ratio
stiffness modulus
complex modulus

xxii

S
S
S
S+
S mix
t
tl
tr
T
TR
UTS
W
z
z
~
z
z

storage modulus
loss modulus
stiffness modulus in compression
stiffness modulus in tension
mixture stiffness
time
loading time
rest-time
temperature
reference temperature
uniaxial tensile strength
specimen width
slope of the creep compliance on semi-ln scale
slope of the creep compliance on semi-log scale
slope of the creep compliance on ln-ln scale or log-log scale
slope of the creep compliance versus ln t on ln-ln scale, or
the creep compliance versus log t on log-log scale

&

constant
constant
shear rate
Gamma-function, fracture energy
loss angle
deformation rate
difference between minimum and maximum stress intensity
factor, K = K max K min
strain
strain amplitude
strain rate
minimum permanent strain rate
permanent strain rate
strain amplitude in compression
strain amplitude in tension
dashpot viscosities of rheological model
reference time
time dependent analogue of z
time dependent analogue of z
time dependent analogue of ~
z
relaxation time of rheological model, i = i / E i
Poissons ratio, kinematic viscosity
3.14159..
damage

&

$
&
&0
&perm
$
$+
1 , 2

xxiii

$
0
1
3
R
ys

stress
stress amplitude
pertinent stress
axial stress
radial stress
rupture strength
yield strength
sum
shear stress
angular frequency

cos
exp
sec
tan
tanh

cosine
exponential
secans (1/cos)
tangent
tangent hyperbolicus

is proportional to

xxiv

List of units
%m/m
C
g
K
m
N
Pa
rad
s

mass percent
degrees Celsius
gram
Kelvin
meter
Newton
Pascal, 1 Pa = 1 N/m2
radian
second

Prefixes
G
k
m
M

giga, 1 giga = 109


kilo, 1 kilo = 103
milli, 1 milli = 10-3
mega, 1 mega = 106

xxv

List of abbreviations
App.
ASTM
BTS
Ch.
Chs.
CROW
CGAC
DAC
eq.
eqs.
fig.
figs.
GAC
ITS
ln
log
LVDT
OAC
PA
SMA
RHEI
SAL
SAL100
sec.
tc
UTS

appendix
American Society for Testing Materials
bending tensile strength
Chapter
Chapters
Bureau for Contract Standardisation and Research for Civil
Infrastructure
crushed gravel asphalt concrete
dense graded asphalt concrete
equation
equations
figure
figures
gravel asphalt concrete
indirect tensile strength
logarithmus naturalis, natural logarithm
logarithm
linearly variable displacement transducer
open graded asphalt concrete
porous asphalt
stone mastic asphalt
Road and Hydraulic Engineering Institute, Ministry of
Transport, Public Works, and Water Management
standard axle loads
equivalent 100 kN standard axle loads
section
traffic class
uniaxial tensile strength

xxvi

1
Introduction

1 Motive for this study


The commonly used road paving materials are asphalt and cement
concrete. In The Netherlands, asphalt pavements make up approximately
95% of the main road network. There is a need for a different approach to
control the quality of asphalt pavements and asphalt mixtures than the
traditional one. This need arises from a variety of socio-economic
developments.
1.1 Socio-economic developments
By virtue of the constitution, the State protects the prosperity and welfare
of its citizens. Constitutional rights are, for example, the protection of the
populations means of living, the nation-wide distribution of prosperity,
the enhancement of room for living and employment, and the right of
health-care and education. Important for the fulfilment of constitutional
rights is the competitiveness of the national economy. The greater the
productivity, the lower the unit cost of production and the greater the
competitiveness of the economy. The competitiveness of the economy is
enhanced by economic growth and innovation, which, in turn, are
necessary for the preservation of employment, and, by that, for the
preservation of the populations means of living and prosperity.

Introduction

Ch. 1

160

traveller km (billion km)

140
120
100
private car

80

public transport

60
40
20
0
1950

1960

1970

1980

1990

2000

year

Figure 1. Use of the private car and public transport between 1950 and 1995.
[Data from: National Bureau of Statistics].

1.1.1 Control of development of transport infrastructure


Important pillars of the economy are the different transport infrastructure
networks; the road network, the waterways network, the rail network, and
the air and seaports. The economy strongly depends on an efficient goods
transport sector. An indication of the quality of the national transport
infrastructure, in comparison to surrounding countries can be obtained
from the World Competitiveness Report of IMD/World Economic Forum.
In this yearly study, an international panel of entrepreneurs give a rating of
different sorts of infrastructure. The rating of the Dutch road infrastructure
has decreased during the last few years, and is clearly lower than that of
neighbouring countries. A similar trend can be observed in the rating of
the Dutch rail infrastructure. The ratings of the air and sea ports have been
better than those for the surrounding countries.
In 1996, it was recognised that the execution of the Second Traffic and
Transport Framework Programme 19881, FPTT II (MT 1988) needed
intensification. This condensed in the bill Beating Congestion2 (MT
1996a) and the action plan Balancing Transport3 (MT 1996b). The main
policies mentioned in the Balancing Transport plan are:
. the enhancement of the competitiveness of sustained transport, in
particular by rail, canal, and short sea,
. the reduction of the burden on the environment caused by road traffic,
. improved access to economic areas for road transport of goods.
1

Tweede Structuurschema Verkeer en Vervoer, SVV II.


Samen Werken aan Bereikbaarheid, SWAB.
3
Transport in Balans, TiB.
2

Sec. 1

Motive for this study

Economic growth has as an effect a growth of transport of persons and


goods. The development of the use of private cars as compared to public
transport since 1950 is shown in figure 1. The Second Traffic and
Transport Framework Programme forecasted an auto mobility increase by
72% from 1986 until 2010. In FPTT II and the National Environmental
Programme 19884, NEP (MH 1988), it was targeted to limit this increase
to 35%. However, this was realised as early as 1993. In view of the
enormous increase in congestion (70% increase of vehicle-loss hours since
1986), it is clear that the road infrastructure is incapable of accommodating the extrapolated auto mobility in the next fifty years.
Over the past decades, the developments in traffic and transport have been
largely autonomous, being the resultant of decisions by individuals in an
individualised society. The national authority saw its influence reduced to
control the utilisation of the different modalities. Extension of the road
infrastructure has not provided an adequate answer to the developing
congestion. Typically, the supplied road capacity suffices during most of
the daytime, and is insufficient during rush hours. The average utilisation
of the private car during rush hours has remained constant to just 1.2
persons, despite measures to stimulate drivers to use different modalities.
The average utilisation of trucks has decreased since the mid eighties until
approximately 50% at present (as a result of external factors). Thus, the
existent road infrastructure is not utilised optimally, whilst further
extension requires a lot of money, time and green space, and in fact creates
additional over-capacity.
The property of the different types of transport infrastructure is in the
hands of the national authority. The infrastructure networks are a national
collective good, in the sense that the construction and maintenance has
been financed collectively. The road user is not charged directly in relation
to his use of the road infrastructure. As a result, there is at present no
relationship between the users use of road infrastructure and his
compensation for this use. This yields abundance and temporary scantiness
of road capacity uncontrolled. With the National Traffic and Transport
Framework Programme 20015, NTTFP (MT 2001), the original aims of
FPTT II and NEP have been reconsidered. The growing need for mobility
is acknowledged. The aim is to make possible that further growth of road
traffic can be accommodated. To manage the growing demand, three main
goals have been formulated in NTTFP:
. to improve utilisation of infrastructure,
. to enhance capacity (by smart utilisation as well as physical extension),
. road pricing.
It is anticipated that these goals are within reach thanks to advances in
information and communication technology.
4
5

Nationaal Milieubeleidsplan, NMP.


Nationaal Verkeer- en Vervoersplan, NVVP.

Introduction

Ch. 1

domestic road transport of goods


(billion tonkm)

25

20

15

10

0
1975

1980

1985

1990

1995

2000

year

Figure 2. Domestic road transport of goods.


[Data from: National Bureau of Statistics].

intern. road transport of goods


(mln ton)

60
50
40
30
20
10
0
1984
D

1986

1988
B/L

1990

1992
F

1994

year

1996
I

1998

2000
SP

Figure 3. International road transport of goods. D = Germany, B/L =


Belgium and Luxembourg, F = France, I = Italy, SP = Spain.
[Data from: National Bureau of Statistics].

1.1.2 The growth of road traffic


The tremendous increase in the use of the private car over the past fifty
years has had mainly environmental implications, such as loss of green
space, noise emission, and air pollution. However, the damaging effect of
the private car to the road pavement is practically negligible, in
comparison to that of the truck. Therefore, if the growth of traffic is
considered in relation to design and maintenance of roads, then it suffices
to consider only truck traffic. Truck traffic also has been growing beyond

Sec. 1

Motive for this study

120

cumulative percentage

100

80

60
1968

40

1979
1986

20

1993
1999

0
0

50

100

150

200

axle load (kN)

Figure 4. Axle load spectrum from 1968 to 1999. The median axle load,
indicated by the arrows, increases steadily with time. [Data from RHEI6].

expectations both in number and weight. Figure 2 shows the domestic road
transport of goods from 1980 to 1998. Figure 3 shows the international
road transport of goods from 1986 to 1998. In figure 4, the axle load
spectrum shows a gradual increase of the median with time. Thus, it
follows that in particular the combination of number and weight of trucks
causes a significant increase of the traffic load of the road network. Figure
5 shows the increase of the percentage of trucks equipped with super
singles. The super single has a greater tyre pressure and a smaller
tyre/pavement contact area than the traditional dual wheel configuration.
Therefore, the damaging effect on the pavement is greater, in comparison
to the dual wheel. This effect may increase further, should the European
Community decide to allow a higher maximum axle load without limiting
simultaneously the tyre pressure. This trend is supported by economic and
environmental advantages, because bigger trucks are more efficient in fuel
consumption, and mean fewer trucks, fewer wasted tyres and lesser
congestion.
1.1.3 Sustained use of materials and energy
Sustained Development is an embedded policy, which compels economic
use of materials and energy to protect the environment. It means many
things. It means a continuous effort to re-use materials to save energy of

The curve for 1999 represents the cumulative result of axle load measurements
performed over the period between 1-1-1993 and 1-1-2000. Data from RHEI 2001.

Introduction

Ch. 1

percentage super singles

100

80

60

40

20

0
1980

1985

1990

1995

2000

2005

year

Figure 5. Top (a): Super single and dual wheel. Bottom (b): Increase in the use
of super singles between 1980 and 2000. [Photographs from RHEI; data from
RHEI 1996].

production and reduce emission of green house gasses. It means that new
materials are developed, as well as production and maintenance
techniques, to make possible a more efficient use of materials and energy.
In the road building industry, important contributions to this can be
realised by re-use of building materials in road bases, and re-use of old
asphalt. Hot recycling of asphalt is developed by stimulating increased
recycling percentages, and re-use of old asphalt in the original application
(e.g. porous asphalt in porous asphalt) if possible. Cold re-use of old
asphalt is developed in various ways, for example, by means of polymer
modified bitumen emulsions, and foam bitumen. Further contributions can
be realised by developing new, more durable paving materials.

Sec. 1

Motive for this study

A continuous effort is developed to control traffic noise emission, by the


development of silent running surfaces. A second generation of silent
running surface is made of two-layered porous asphalt, which consists of a
top-layer of fine-graded polymer modified porous asphalt, on top of a
lower layer of standard porous asphalt. Thus, up to 6-7 dB(A) noise
reduction can be achieved instead of the normal 3 dB(A).
1.1.4 Design-Build-Maintain contracts
A recent development in road building is the development of DesignBuild-Maintain contracts. In these contracts, the contractor is offered the
opportunity to submit a tender incorporating not only the building but also
the design and maintenance of a road or pavement for a long period, e.g. a
substantial portion of the pavements service-life, the entire service-life, or
even a longer period. This type of contract aims among other things to
stimulate innovation in road building. The incorporation of design, build
and maintenance in a single contract implies that the contractor takes
responsibility for the design including the material selection, and the
consequences of that for maintenance. This type of contract causes
responsibilities to shift from the client (the road authority) to the
contractor. This, in turn, demands from the contractor an in-depth analysis
of the project risks involved, not only from a contractual viewpoint, but
also from an insurance point of view.
1.1.5 Road building, changing authorities, and the knowledge economy7
In modern society, the national road authorities are no longer the sole
organisations where the knowledge to build roads is concentrated.
Relevant knowledge is more and more distributed in the economy.
National authorities, not just road authorities, have indicated they wish to
be more efficient. One way to achieve this is to make use of relevant
knowledge available in the private economy, rather than to develop
relevant knowledge inside the public organisation. It seems, authorities
prefer to lose intentionally capacity of developing new knowledge in
favour of developing a capacity of arranging knowledge available in the
private economy. Real innovating power lies in enterprises, where the
development costs of innovations can be regarded as an investment with a
chance to become profitable. It is important to permit room for innovative
developments, also in road building and maintenance, in order to keep up
7

A knowledge economy is an economy in which the production factors labour


and capital are strongly aimed at the development and application of new
technology. Romer (1986, 1990) has proposed a change to the neo-classical
model of economics by seeing technology (and the knowledge on which it is
based) as an intrinsic part rather than an exogeneous factor of the economic
system. In Romers theory, knowledge is the basic form of capital. Ecomic
growth is driven by the accumulation of knowledge.

Introduction

Ch. 1

with socio-economic development in general and to hold the cost of


building and maintenance of the road network on an acceptable level. To
achieve that, it is necessary to put available knowledge to work anywhere
in the economy in an efficient manner. How can that be achieved? It
requires more than one thing to change. One thing necessary is to remove
restrictive systems that inhibit the application of knowledge available in
the economy. Imagine businesses and road authorities come together to
communicate about the performance and the price of objects of public
infrastructure. The current system of contractual requirements and
technical specifications works satisfactorily as long as it is operated within
its framework of standardised technology. With any new development the
road authority asks if current requirements are applicable, and if not, to
develop new requirements. This question unfolds the current systems
restriction. The restriction lies in its empirical character. This causes the
limited applicability of the current requirements and specifications to
newly developed products, and the long time needed to evaluate the
performance of new products, and, because of that, also a long time to
develop new requirements. The time needed to develop new requirements
let alone the time needed to develop the knowledge to be able to develop a
more fundamental approach, causes the implementation of innovative
techniques and materials to stay at a low pace, until a system is developed
which permits development of more generally applicable requirements.
That the pace of innovation is low can be observed from the rate at which
new standardised materials and test methods have been introduced in
standard regulations and requirements (CROW 2000) over a period of say
thirty years. Thus, the present restrictive system of requirements and
specifications prevents economic parties from communicating over the
performance quality of innovative asphalt pavements and asphalt
mixtures.
To improve this situation, to make possible that available knowledge in
the economy can be put to work in an efficient manner, it is necessary to
modernise pavement design concepts and asphalt mixture design concepts.
It is necessary to replace current requirements by requirements that permit
economic parties to communicate over performance quality of new
products. It seems that requirements have to be more generally applicable,
not just to standardised technology but to new technology as well. The
question may be posed which form such requirements should take. When
businesses and road authorities communicate over performance quality,
they discuss many things including road comfort, traffic safety,
environment, and health and safety of road workers. If the performance
quality of the object of infrastructure itself is considered, the essence of
what both parties are concerned with is the cost-effectiveness and the risk
of failure. The cost-effectiveness and the risk of failure of standardised
technology, asphalt pavements, paving materials, production technology,

Sec. 1

Motive for this study

and paving technology, are known, based on experience, laid down in the
empirical requirements and technical specifications. With a new
development, these certainties fall away as soon as normal requirements
and technical specifications are not applicable. Suddenly, in order to be
able to judge its cost-effectiveness and risk of failure, there is the need to
predict the pavements behaviour. Technicians start building computer
models to predict the pavements response to loading. That is where a
significant change takes place in comparison to the current pavement
design methodology. The current system with its empirical test methods to
determine optimal mixture compositions is inadequate. Typical examples
of empirical methods are the Marshall test, to some extent the fatigue test,
the wheel tracking test, various ravelling tests; the empirical parameters
a Marshall stiffness, a fatigue-life as commonly reported, a permanent
deformation in the wheel-tracking test, a mass-loss by abrasion in a
ravelling test are not the sort of properties required to predict the
pavements behaviour. In a functional or performance related approach
concerned with the prediction of pavement behaviour, and the evaluation
of the cost-effectiveness and risk of failure, the material composition is
irrelevant; relevant are only the properties needed to predict or judge the
cost-effectiveness and the risk of failure.
As long as the road network is a national asset and a public interest, and
thus the responsibility of a road building authority, it is the responsibility
of that authority to impose (performance related) requirements. This
means that the answers to questions regarding the facilitation of
innovation cannot come from the responsible authority if that has lost its
knowledge to judge the cost-effectiveness and risk of failure of pavements
and the applied paving materials.
1.2 Innovation and product quality - The need for a rationalised quality
control methodology8
A challenge of the future is to unify economical and environmental goals
in improving road infrastructure utilisation and enhancing its capacity. It
means, for example:
1 a further growth of traffic has to be accommodated, while noise levels
and pollution must be reduced,
2 more durable pavements must be developed, so that maintenance
frequencies are reduced, to avoid congestion by maintenance and the
negative effects of congestion on road safety,
3 new materials, production techniques, or maintenance techniques have to
be developed, which save materials and energy.
8

The word methodology is used to indicate a body of methods. The quality


control methodology covers an asphalt production control method, an asphalt
pavement design method, a material selection method, an asphalt mixture design
method, and a mixture constituents selection method.

10

Introduction

Ch. 1

Asphalt production
control method

pavement design
method
material selection
method
mixture design
method

mixture constituents
selection method

Figure 6. Scheme, illustrating the total quality control methodology (light) and
the pavement design methodology (dark).

The realisation of these goals requires a technological innovation, which


has to be developed at a higher pace than in the past. Considering the
growth of road traffic over the past five decades, it is probable that roads
are now built for future traffic loads that are far beyond our present
experience. The developments in traffic and transport make it probable,
that we will have to use pavement structures and materials with which we
have presently no experience.
Apart from new products, new instruments for quality judgement of
products are needed. The present study is concerned with the instrument to
judge the quality of the asphalt mixtures. Figure 6 illustrates the asphalt
production quality control methodology and the pavement design
methodology. The asphalt production quality control methodology can be
considered to consist of five methods: a method for the selection of the
asphalt mixture constituents, an asphalt mixture design method to design
an optimised mixture composition, a material selection method to select
the asphalt mixtures to be applied in a pavement design, and an asphalt
production control method to control the production quality (which
includes the quality of the asphalt after paving, i.e. after compaction). Not
considered in the present study are the method for the selection of the
asphalt mixture constituents , and the asphalt production quality control

Sec. 1

Motive for this study

11

Intermezzo 1
A characteristic of an empirical method is that it relies on practical experience
rather than theories. This makes an empirical method descriptive rather than
explaining. An empirical law can describe a phenomenon without providing an
understanding, although the empirical law itself could be considered a sort of
understanding; yet, this differs from an understanding in terms of
fundamental principles, which have more general predictive value. An
empirical law is predictive merely in its own reference system. The following
example can illustrate this. An empirical law could be, for example, the moon
moving from the east to the west across the south in the northern hemisphere.
This law is predictive in the northern hemisphere, but not in the southern
hemisphere, where the moon moves from the east to the west across the north.
To design a similar law which is predictive for both hemispheres, a deeper
understanding of the system of the moon and the two hemispheres is required.

method. The remaining three methods, the pavement design method, the
material selection method, and the asphalt mixture design method (the
dark section of figure 6), is conveniently called the pavement design
methodology. Considering the methods indicated in figure 6, it can be
observed that currently all five methods are almost entirely or entirely
empirical. The pavement design method is the only method, which uses
more or less fundamental design criteria, but apart from that is mainly
empirical. The asphalt production control method, the material selection
method, the mixture design method, and the mixture constituents selection
method, are based on the compositions of the asphalt mixtures. The fact
that the methods are composition-based causes the methods to be
empirical. What this means is explained in Intermezzo 1.
An empirical methodology requires renewal of empirical reference data,
based on practical experience. To gain practical experience with a new
pavement design, or a new type of asphalt mixture, requires monitoring of
the nominal service-life, to gather reference data, and to verify the
performance (cost-effectiveness with respect to standard pavement
designs, respectively asphalt mixtures). This leads to a delay of innovation
that is no longer acceptable. Thus, to date, the current empirical design
methods have been of very little value for the development and
acceptation of new types of asphalt pavement and asphalt mixture.
The alternative of the empirical method that relies on practical experience
is the fundamental method, which relies on theory. However, one may
wonder what can be the added value of a fundamental method since it is
not possible to predict pavement performance quantitatively, i.e. to predict

12

Introduction

Ch. 1

the type and amount of a specific type of damage as function of the time
during the pavements service-life. The main reason is the complexity of
the road system and the unpredictability of a number of influence factors,
such as the traffic, the climate (during paving and service), the variability
of mixture constituents, the variability of production and paving, and the
spill of chemical agents (leaking motor oil, solvents). Furthermore, one
should realise that apart from the complexity of the road system and the
unpredictability of influence factors, the available theories are in fact
oversimplifications of the reality. This is caused by the following
assumptions:
. homogeneous and isotropic pavement material,
. linear elastic stress strain behaviour, instead of nonlinear viscoelastoplastic stress strain behaviour
. simplified dynamic loading by traffic,
. a one-dimensional uniform contact pressure distribution, instead of a
three-dimensional nonlinear distribution,
. a simplified temperature distribution in the pavement (mean annual
asphalt temperature).
A fundamental approach requires that:
1 materials are characterised by means of true material properties,
2 tests are available to determine those properties.
By definition, a true material property is a property which is independent
of the geometry (size and shape) of the specimen, and which is not
influenced by the measurement itself. In a popular way of saying, a true
material property is reproducible in different tests. The behaviour of an
asphalt mixture is really too complicated to be described in detail by any
available fundamental model. It is shown in this study that fundamental
models exist, in which the material is assumed to be homogeneous and
isotropic, and that these models can be used as approximate models when
applied to an asphalt mixture. Thus, a truly fundamental method to control
the quality of asphalt is not feasible to date. What is feasible, is a rational
approach, i.e. an approach based on reason instead of belief. It is not
meant by this, that the traditional empirical method is irrational. However,
it is based on empirical fact rather than reason. A new method can be more
adaptable to new developments, if reason gains importance as one of its
pillars relative to (historical) fact.

2 Scope of this study


2.1 Introduction
A pavements main functions are: bearing capacity, surface characteristics,
and long-term performance. Long-term performance can be defined as the
gradual loss with time of the pavements bearing capacity and surface

Sec. 2

Scope of this study

13

characteristics. The total pavement response9 to loading can be regarded as


the resultant of the actual response to loading, and the long-term
performance. In terms of response to loading, the long-term performance
can be defined as the gradual decline of the actual response as a function
of the time during the pavements service-life. The pavements main
functions are controlled by the pavements functional properties. These are
indicated in table 1.
Table 1. Functional properties of asphalt pavements.
Functional properties of an asphalt pavement
bearing capacity
Surface characteristics
Long-term performance
evenness
stiffness
stiffness
resistance to fatigue resistance to skidding resistance to:
(crack-growth from
slant
* permanent deformation
bottom of pavement noise emission
* fatigue
upwards)
hydraulic conductivity * surface cracking
light reflectivity
* ravelling
resistance to:
* disbonding (stripping)
* ravelling
* ageing
* rutting
* surface cracking
Bold-faced properties are called functional properties of asphalt mixtures.

The pavements functional properties are controlled by the functional


properties of the applied materials. In the case that the applied materials
are asphalt mixtures, the functional pavement properties are controlled by
means of the functional asphalt mixture properties. In the following, table
1 is explained in more detail.
The pavements bearing capacity is controlled by the pavements stiffness,
which, in turn, is controlled by the thickness of the asphalt layers and the
stiffness of the applied asphalt mixtures. It is important that the pavement
stays intact, and that no cracks form. Therefore, it is important that the
pavement has resistance to fatigue.
The pavements surface characteristics can be divided into two types:
1 the characteristics at the very pavement surface, which the pavement
derives from the properties of the asphalt mixtures constituents at the
pavement surface,
9

The word response is used to indicate the mechanical reaction to mechanical


loading. The total pavement behaviour might be more generally defined as the
resultant of the actual behaviour and the long-term behaviour, whilst the word
behaviour is used to indicate any type of reaction (mechanical, thermal, or
chemical), to the corresponding type of loading. However, the scope of this
study is limited to the mechanical behaviour of asphalt mixtures.

14

Introduction

Ch. 1

2 the characteristics determined by the mechanical properties of the


asphalt mixture lying in the top pavement layer.
For a full performance related approach to pavement design it is necessary
to consider the pavements long-term performance. Important for the
pavements long-term performance is the characterisation of the asphalt
mixtures physiochemistry, i.e. the adhesion and disbonding of mineral
aggregate and bitumen, and the ageing of bitumen10. The asphalt mixtures
long-term performance depends on the adhesive properties of the mixture
constituents, how these are influenced in the presence of an agent11, and
the ageing properties of the asphalt mixture and the mixture constituents12.
2.2 This studys subject
This study is concerned with the mechanical properties of the asphalt
mixtures, which are relevant to the actual pavement response. The asphalt
mixtures mechanical properties are bold-faced in table 1. A reason for the
selection of this subject can be given as follows:
1 mechanical test methods are needed to characterise both an asphalt
mixtures actual response and long-term performance,
2 assuming that the asphalt mixtures resistance to fatigue, resistance to
fatigue crack-growth, and resistance to ravelling (i.e. the mechanical
aspect of ravelling) are a part of the characterisation of the mixtures
actual response, then the characterisation of this actual response requires
mechanical testing protocols,
3 assuming that the asphalt mixtures resistance to disbonding and
stripping, and its resistance to ageing, are a part of the characterisation of
the mixtures long-term performance, then the characterisation of this
performance requires testing protocols to test disbonding and stripping
properties, and ageing properties, in addition to mechanical testing
protocols.
Thus, the characterisation of the actual response can be considered critical
to the development of a performance related asphalt mixture design
method. That is, if mechanical tests to characterise a mixtures actual
response are lacking, then a performance related mixture design method is
not feasible, and a performance related pavement design methodology is
10

These aspects have been, and are still investigated: cf. Elphingstone (1997),
Groenendijk (1998), Voskuilen et al. (1996), Kuppens (1997), Mes (2003).
11
Agents can be additives to improve adhesive properties, or can be agents
which affect the adhesion in a negative sense: water, ice, de-icers, oil-spill,
chemical solvents, clay in mineral aggregate, and other contaminations.
12
Ageing is thought to have a physical component (e.g. time-hardening, loss of
volatiles from the binder, bitumen), and a chemical component (e.g. oxidation
by the air, interaction with ultra-violet radiation in day-light).

Sec. 3

Aim of this study

15

also not feasible. Testing protocols for characterisation of adhesion,


disbonding, and ageing can always be embedded later into a performance
related quality control methodology.

3 Aim of this study


3.1 General objective
The general objective of this study is to make a characterisation of the
mechanical behaviour of an asphalt mixture possible in as much as that is
relevant to the pavements main functions, bearing capacity, surface
characteristics, and long-term performance.
The ultimate aim is to facilitate the acceptance of new and nonstandardised paving materials that are needed to enhance the durability of
our heavily trafficked main road network.
The general objective is limited to the actual response to loading, i.e. the
mechanical behaviour, of an asphalt mixture.
3.2 Practical goal
The practical goal of this study can be formulated as: to make possible a
characterisation of an asphalt mixtures mechanical behaviour relevant to
the pavements main functions, bearing capacity, surface characteristics,
and long-term performance, allowing the use of tests that are suitable for
the practical purposes of material selection in pavement design, asphalt
mixture design (type testing), and production quality control.
3.3 Research goal
Mechanical testing of an asphalt mixture is particularly difficult, because
of the combination of the following features:
. the material heterogeneity,
. the different mechanical properties of the constituent phases,
. the time dependence and stress dependence of the mechanical behaviour,
. the use of a specimen which is relatively small in comparison to the
material heterogeneity (i.e. the maximum grain size).
A relatively small specimen is convenient from a practical viewpoint, but
complicates the testing for the reason that the interaction of the material
heterogeneity and size and shape effects of the specimen may influence the
measured property. Therefore,
The research goal of this study can be formulated as: to develop a method
for the validation of simple tests for practical purposes.
The ultimate aim is to limit the empirical character of the methods used for
pavement design and asphalt mixture design (type testing), and to find
methods for validation of methods other than practical experience.

16

Introduction

Ch. 1

4 Selected topics
The following topics are discussed: In chapter 2, the current pavement
design method, including the material selection method, and the asphalt
mixture design method are discussed. Chapter 3 contains the experimental evidence and an analysis of the viscoelastic properties and creep or
viscoplastic properties of asphalt mixture. Chapter 4 contains the
experimental evidence and an analysis of the crack-growth and fracture
properties of asphalt mixture. Chapter 5 discusses elements of a method
for the evaluation or judgement of the performance of an asphalt mixture
in a pavement in relation to cost-effectiveness and risk of failure. This
thesis ends with a general discussion in chapter 6, and the general
conclusions in chapter 7.
The topics in chapter 3 and chapter 4 were selected because permanent
deformation (rutting in the pavement) and crack-growth are important
elements of the method discussed in chapter 5, which is needed to quantify
the bold-faced properties in table 1. These properties can be quantified by
quantifying the following aspects of the mechanical behaviour:
. linear viscoelastic stress strain behaviour,
. time dependence of the stress strain behaviour,
. stress dependence of the stress strain behaviour,
. fatigue and crack-growth behaviour.

5 Outcome of this study


5.1 General objective
In general terms, the outcome of this study is that it is possible to
characterise the mechanical behaviour of asphalt mixture by means of
constitutive models. Constitutive model parameters define mechanical
properties of an asphalt mixture. Constitutive models serve to attribute a
physical meaning to an asphalt mixtures mechanical properties. It is
important to be able to attribute a physical meaning to constitutive model
parameters. This is important for the predictive value of the model and its
parameters. A predictive value is indispensable for a quantification of the
cost-effectiveness and the risk of failure.
However, mechanical properties of asphalt mixture, determined in the
laboratory, lack quantitative or absolute predictive value for the behaviour
of asphalt mixture in the pavement. Two reasons can be given for this:
1 The dependence of, for example, a creep property on the specimen
geometry is an intrinsic property of a granular bituminous material that
exhibits nonlinear stress strain behaviour. This means, that this
dependence cannot be avoided or eliminated.
2 The stress strain behaviour depends on the shape of the waveform of the
applied stress. The shape of the waveform of the stress induced in the

Sec. 5

Outcome of this study

17

pavement by a passing wheel-load differs from the shapes of waveform


used in the laboratory.
Nonetheless, the predictive value of the constitutive models and the model
parameters is not entirely lost. A qualitative or relative predictive value
can be retained by comparison of the behaviour of newly developed (types
of) material to that of standardised (types of) material, and by proper
standardisation of test methods.
5.2 Practical goal
Referring to the practical goal in 3.2, the outcome of this study is that the
following three tests can be described in the form of the following
propositions:
1 The mechanical aspects of the pavements bearing capacity, surface
characteristics and long-term performance can be controlled by the
following pavement properties:
. stiffness,
. resistance to permanent deformation,
. resistance to crack-growth,
. resistance to ravelling,
. resistance to fatigue.
2 The mechanical aspects of the pavements bearing capacity, surface
characteristics and long-term performance can be controlled by means of
fundamental or phenomenological constitutive models, which describe the
mechanical behaviour of the applied asphalt mixtures. Variables and
parameters of constitutive models represent mechanical asphalt mixture
properties. For the practical purposes of asphalt mixture design and
production quality control, the following mechanical asphalt mixture
properties can be specified:
. complex modulus (in vector notation, this includes the loss angle),
. creep susceptibility,
. tensile strength or fracture toughness, and fracture energy13.
3 The following three test methods are proposed:
. frequency sweep test,
. creep test,
. tensile test.
The number of tests (the number of measured quantities) can be limited
because of the interrelationships between the asphalt properties. For more
details, the reader is referred to Ch. 5, 5.

13

The fracture energy is considered important but was not investigated. For
more information, the reader is referred to Irwin (1977) and Jacobs (1995).

18

Introduction

Ch. 1

5.3 Research goal


Referring to the research goal in 3.3, the outcome of this study is that it
was found possible to validate the use of a simple test for the practical
purposes of asphalt mixture design or production quality control, based on
finite element analysis and investigations using advanced tests. The
combination of finite element analysis and advanced tests enables one to
verify the validity of a constitutive model to describe a given aspect of the
mechanical behaviour of an asphalt mixture.

6 Importance of this study to the current practice


The importance of this study to the current practice is that it is a
significant contribution to the development of performance related
specifications for asphalt mixtures. It is concluded that such specifications
can enhance the pace of innovation in the road building industry. As such,
this study can be seen in the light of the historical developments in the
field of pavement design and material selection. Although the quality of
asphalt mixtures in the Netherlands is specified in terms of compositionrelated requirements, there have been different events in the post-war
history that called for specifications based on properties.
The first event was the publication of the Shell Pavement Design Method,
developed by Shell from the 1950s up into the early 1970s, which
required mechanical asphalt mixture properties as input data for a
pavement design, such as stiffness and the resistance to fatigue, and
permanent deformation. Tests developed by Shell are, amongst other, a
fatigue bending test, a static creep test, and a wheel-tracking test.
Furthermore, Shell developed very useful practical procedures to estimate
mechanical mixture characteristics from the volumetric mixture
composition, and characteristics of the bitumen.
The second event was the development of the re-use of old asphalt in the
Netherlands in the period 1975-1985. It was realised that to be able to
justify application of partially recycled asphalt mixtures in all layers of the
pavement in terms of cost-effectiveness and risk of failure, it was
necessary to know the mechanical behaviour of the mixtures, and that this
behaviour cannot be known on the basis of composition-related
requirements.
For that reason fatigue, stiffness, and permanent deformation testing
played a vital role in the acceptance of recycled asphalt mixtures. It was
this understanding that paved the way for the Scientific Asphalt Research
project (Technisch Wetenschappelijk Asfalt Onderzoek) that started in
1989, which, although ended in 1997, finds its final report in the present
in-depth study.
The aim of the Scientific Asphalt Research Project was to develop
practical test methods for the determination of performance related

Sec. 6

Importance of this study to the current practice

19

mechanical asphalt mixture properties. The project served also as a basis


from which contributions were made to the European standards
harmonisation process. By January 2007 it will be possible for the first
time in the Netherlands to specify the quality of asphalt mixtures in terms
of property-related requirements.
To date, one observes a fourth period of growing interest in propertyrelated requirements. In fact, it is quite clear that property related
requirements are urgently needed. The reason for that is that traffic has
grown both in volume and weight well beyond experience. In order to
maintain the serviceability of the main road network as a vital asset of our
economy, it is crucial to get a longer life from our pavements. Moreover,
heavy demands are made on skid resistance and noise reduction.
Altogether this asks for more durable pavement structures and paving
materials. It is strongly believed that such structures and materials can no
longer be developed solely on the basis of practical experience.
Fundamental research is needed in order to be able to successfully develop
materials with improved performance, and to develop specifications to
ensure their successful production.
The importance of the present study is both fundamental and practical. A
lesson learnt from the developments of the re-use of old asphalt, porous
asphalt, and presently, polymer modified asphalt, and two-layered porous
asphalt is that it is absolutely necessary to know the behaviour of paving
materials in terms of mechanical properties relevant to the behaviour of
the material in the pavement. Otherwise it is not possible to justify
application of a newly developed material in terms of cost-effectiveness
and risk of failure.
The present study is a contribution to the fundamental knowledge of the
behaviour of asphalt mixtures. The results of it allow paving materials to
be judged using sound engineering principles. The results show that this is
not possible based on empirical research alone.
A practical aspect of this study is the understanding of the behaviour of the
material in mechanical tests in the laboratory, which should make
specification in terms of property-related requirements possible.

20

Introduction

Ch. 1

2
The current pavement design
methodology
1 Introduction
The purpose of this chapter is to describe the pavement design
methodology currently practised by the RHEI. In the remainder of this
study this method is used as a reference (because it is beyond the scope of
this study to discuss differences between different pavement design
methods). The methodology consists of a pavement design method (RHEI
1998), a material selection method (RHEI 1998), and a mixture design
method (CROW 2000). In 2, each of these methods as currently practised
is reviewed briefly. In 3, the weaknesses and limitations are discussed.
This chapter is concluded with suggestions for improvements.

2 The current methodology of asphalt pavement design,


material selection, and asphalt mixture design
2.1 The pavement design method
An example of a road structure is shown in figure 1. The asphalt layers of
a pavement are normally three layers: a base layer, a binder layer, and a
top layer. Each layer may consist of several sub-layers, normally of a
similar type of asphalt mixture. According to current practice, no binder
layer is required in a pavement consisting of a porous asphalt top layer on
a crushed gravel asphalt concrete base layer.

22

The current pavement design methodology

Ch. 2

Figure 1. Representative structure of road in the Netherlands.

2.1.1 Bearing capacity


The bearing capacity is controlled by the pavement stiffness, which, in
turn, is controlled by the thickness of the asphalt layers and the stiffness of
the materials. It is important that the pavement stays intact and that no
cracks form. A most serious damage results if cracks grow from the
bottom of the pavement upwards, as that may lead to an early
reconstruction of the entire pavement. Based on experience (Claessen et
al. 1977, Clard 1977), it has been generally agreed that, with respect to
the structural design of a pavement, two failure criteria are important:
1 the horizontal tensile strain in the lower fibre of the bituminous base,
2 the vertical compressive strain at the top of the sub-base, or sub-grade.
Pavement design aims to control these variables. The principle, illustrated
in figure 2, is the reduction of the stresses to acceptable levels by socalled load spreading. This load spreading is a lateral stress redistribution
that is enhanced by the use of a granular material with a dense granular
structure.

Sec. 2

The current methodology of asphalt pavement design,


material selection, and asphalt mixture design

23

Figure 2. Principal design criteria for road pavements.

For the design of an asphalt pavement, normally a linear elastic multilayer


programme is used (Claessen et al. 1977, Clard 1977, RHEI 1998). The
pavement material is assumed to be homogeneous, isotropic, and linearly
elastic. However, for the thickness design of an asphalt pavement, linear
viscoelastic complex moduli of the applied asphalt mixtures are used. The
selection criteria for the applied materials are:
1 stiffness
2 resistance to fatigue.
The resistance to fatigue is considered relevant only to the material
applied below the neutral plane of the pavement. The required
characteristics of the applied materials depend on the traffic class. Since
1995, four traffic classes are in effect, indicated respectively as traffic
class 2, 3, 4 and 5, see table 1.
2.1.2 Surface characteristics
In the RHEIs pavement design method, no design criteria have been
developed for the surface characteristics; i.e. the resistance to surface
cracking, the resistance to rutting, or the resistance to surface
deterioration, or ravelling. The function of the binder layer, if
incorporated, is to provide added resistance to permanent deformation in
the layer between circa 4 and 8 cm below the pavement surface, where the
shear stresses reach a maximum (Eisenmann and Hilmer 1987).

24

The current pavement design methodology

Ch. 2

Table 1. Traffic classes, according to CROW (2000).


Traffic class
SAL100 Explanation
2
< 500 lightly trafficked pavements
3
< 4000 normally trafficked pavements
4
> 4000 heavily trafficked pavements
5
> 500 slow and heavy traffic,
driving speed less than 15 km/hr
SAL100 = equivalent standard axle loads, 100 kN.

2.1.3 Two functional requirements for pavements: longitudinal evenness


and skid resistance
Two functional requirements with respect to the pavements surface
characteristics are in effect:
1 a requirement with respect to the longitudinal evenness of the pavement,
2 a requirement with respect to its skid resistance.
Typically, these properties are pavement surface characteristics, which
cannot be specified as mixture design properties. The pavements
longitudinal evenness is created during paving. The pavements skid
resistance is controlled by the materials surface texture, i.e. its macro
texture and micro texture. The macro texture is controlled by mineral
aggregate size and angularity. The micro texture is controlled by means of
the polished stone value of the aggregate.
2.1.4 Summary
The present pavement design method is focused on the pavements
bearing capacity. No design criteria are in effect to control the pavements
surface characteristics.
2.2 Material selection
Different types of mixture are used for different pavement layers. This is
indicated in table 2. A detailed overview of the types of mixture, and their
application, is given in table 3. To understand the differences between the
various mixtures, the following explanation is given.
An asphalt mixture can be considered to consist of two phases; a mineral
aggregate phase, and a binder phase. Depending on the type of mixture,
the mineral aggregate phase may be a stone skeleton, a sand skeleton, a
filler skeleton, or a transition between these types of skeleton. The binder
phase may be a mastic - i.e. a mixture of sand, filler and bitumen - or a
mortar - i.e. a mixture of filler and bitumen. Depending on the type of
mixture, the sand is a part of the mineral aggregate phase (e.g. the stonesand skeleton, or a part of the binder phase (e.g. the mastic, or the mortar).
The mineral aggregate phase may be underfilled, filled or overfilled, with

Sec. 2

The current methodology of asphalt pavement design,


material selection, and asphalt mixture design

25

Table 2. Types of asphalt mixture and application.


Type of mixture
aggregate structure
wearing course mixtures
porous asphalt
stone skeleton mixture, underfilled
stone mastic asphalt
stone skeleton mixture, filled
dense graded asphalt concrete
stone-sand skeleton mixture, filled
1
Guss-asphalt
sand-filler skeleton mixture, filled
binder course mixtures
open graded asphalt concrete
stone-sand skeleton mixture, slightly underfilled
base course mixtures
crushed gravel asphalt concrete
stone-sand skeleton mixture, filled
gravel asphalt concrete
stone-sand skeleton mixture, filled
1
for bridge-decks.

binder. An asphalt mixture is underfilled, if not sufficient binder is


available to fill the voids of the mineral aggregate skeleton. The pores in
the mixture form a network. An asphalt mixture is filled, if just enough
binder is available to fill the voids of the mineral aggregate skeleton. The
voids still there, are isolated. A mixture is overfilled if more sand and
mortar is available than can be stored in the voids of the stone skeleton.
An example of an underfilled type of asphalt mixture is porous asphalt.
This type of mixture contains approximately 20% voids. Examples of
filled mixture types are dense graded asphalt concrete and stone mastic
asphalt. Depending on the amounts of sand and mortar, the stones may
even be separated completely, to form a stone-sand skeleton, filled with
mortar. An example is Guss-asphalt. Representative grading curves of
different asphalt mixture types are shown in figure 3. The mixtures are
discussed in more detail in 2.2.12.2.3.
2.2.1 Asphalt concrete
Up to 1990, almost exclusively asphalt concrete mixtures were applied in
all pavement layers. Major adjustments of the standard compositions of
asphalt concrete mixtures were introduced in 1978, and in 1985, while
minor adjustments were introduced in 1987, 1990, and 1995. The whole
process can be regarded as a process of optimisation of mixture
compositions and mixture design properties in relation to the development
of traffic, taking account of the workability of the mixtures. The
adjustments of 1978 were necessary, as traffic had increased dramatically
during the sixties and seventies, while no major changes to the mixture
compositions and the mixture design procedure had occurred. This need
was enhanced by severe rutting on the main road network, which occurred
during the unusually hot summer of 1976. It is important to note that the

26

The current pavement design methodology

Ch. 2

Table 3. Types of asphalt mixture, and applications.


Layer thickness, mm
Mixture
Application
min opt max
porous asphalt
0/16
wearing course 50 mm
35 45 65
0/11
wearing course 40 mm
25 35 45
stone mastic asphalt
25 35 50
wearing course 35 mm, tc 5
0/11 type 1
wearing course 35 mm
0/11 type 2
25 35 50
wearing course 25 mm
0/8
20 25 30
wearing course 20 mm
0/6
15 20 25
dense graded asphalt concrete
35 45 65
wearing course 40 mm, tc 3, 4
0/16
wearing course 35 mm, tc 2,3
0/11
25 35 45
wearing course 25 mm, tc 2,3
0/8
20 25 30
open graded asphalt concrete
40 60 80
binder course > 50 mm
0/22
binder course < 50 mm
0/16 type 2
35 45 65
binder course/temporary wearing course
0/16 type 3
35 45 65
crushed gravel asphalt concrete
0/22
base course 50 mm
40 60 80
0/16
base course < 50 mm
35 45 60
gravel asphalt concrete
50 70 90
base course 50 mm, tc 2, 3
0/32
wearing course tc 2
0/16 type 1
50 55 65
wearing course/base course < 50 mm, tc 2, 3 35 45 60
0/16 type 2
0/16, etc., indicates the minimum respectively maximum grain size, in mm.
tc = traffic class, see table 1.

whole process of optimisation of mixture compositions and design


properties took place on an empirical basis. This was shown, for example,
by the adjustments implemented in 1985. Subsequently it was found that
these adjustments had had a negative influence on the workability of
mixtures, which led to corrections in 1987.
A simplified overview of the changes over the years is shown in table 4.
Table 4 shows that the percentage of stone was increased at the cost of the
percentage of mastic (sand, filler, and bitumen), and at the cost of the
percentage of bitumen. The asphalt concrete mixtures are designed
following the concrete principle, which means that the grain skeleton is
designed to be as dense as possible, with voids as low as possible. Ideally,
this is achieved by means of a continuous grading curve, which, on
logarithmic scales, transforms to a straight line. This is called the Fller
grading curve. In reality, the grading of asphalt concrete mixtures deviates
from the Fller grading curve. It can be seen in figure 4a that for DAC
0/11 and DAC 0/16 that the amount passing sieve 2 mm is somewhat
greater in comparison to the Fller grading. This is to enhance the
workability of these mixtures.

Sec. 2

The current methodology of asphalt pavement design,


material selection, and asphalt mixture design

27

% m/m passing through sieve

100

10

1
0.01

CGAC 0/16
OAC 0/16
DAC 0/16
PA 0/16
SMA 0/11

0.1

10

mesh (mm)

Figure 3. Examples of grading curves of standardised asphalt mixtures.


[Data from: CROW 2000].

Table 4. Simplified overview of desired asphalt mixture


compositions (% m/m) of gravel asphalt concrete, open graded
asphalt concrete, and dense graded asphalt concrete according
to the requirements published in 1967, 1972, and 1978.
1967
1972
1978
gravel asphalt concrete
57
55
50
stone
37
39
44
sand
6
6
6
filler
4.5
5
5.5
bitumen
open graded asphalt concrete
stone
65
65
67
sand
29
29
27
filler
6
6
6
bitumen
6
5.5
5.4
dense graded asphalt concrete
stone
57
57
57
sand
35
35
35
filler
8
8
8
bitumen
7
6.5
6.4
1
After 1972, % stone was measured on sieve 2 mm, and
% sand on sieve 63 m. Before 1972, % stone was
measured on sieve 2.4 mm, and % sand on sieve 75 m.

100

28

The current pavement design methodology

Ch. 2

For different reasons, dense asphalt mixtures are particularly suitable for
application in wearing courses. Dense mixtures are durable. The
extremely dense packing of mineral aggregates provides a high resistance
to deformation. The low voids content is beneficial, as it yields high
resistance against the disintegrating influences of penetrating water. Water
causes damage, inside an asphalt mixture, not only if turned to ice. It
influences the thermodynamic potential of adhesion between mineral
aggregate and bitumen, causing gradual loss of adhesion. In addition,
pressurisation by traffic of water in voids contributes to early fatiguecrack initiation. A low void content is beneficial for the resistance to
ageing, i.e. loss of volatiles and oxidation by the air. Finally, dense
wearing courses provide a protection to the lower asphalt layer from being
penetrated by water and against oxidation.
Open graded asphalt concrete is intended for application in binder layers.
It is relatively stone rich and open graded. The idea is to increase the
internal friction, which provides a greater resistance to permanent
deformation, in comparison to gravel asphalt concrete and to dense graded
asphalt concrete.
Gravel asphalt concrete is intended for application as bituminous base. Its
characteristics are the rounded, i.e. uncrushed, aggregate, and the
relatively low binder content of approximately 4.5%. As a result, the
internal friction is relatively low, and consequently the resistance to
permanent deformation. Although this is compensated by a harder
bitumen, 40/60, gravel asphalt concrete is not recommended for heavily
trafficked class 4 or 5 roads.
Crushed gravel asphalt concrete is a more recent development, intended
for application in the bituminous base. Its general characteristics are
similar to those of gravel asphalt concrete, except that crushed gravel is
used, which provides greater resistance to permanent deformation.
2.2.2 Porous asphalt
Porous asphalt was developed originally to reduce splash and spray.
Today, it is applied mainly to reduce noise emission of pavements. In
order to fulfil both functions, a design minimum of 20% air voids is
required. This is achieved by means of a so-called gapped grading, i.e. a
grading in which the sand and small stone fractions are reduced, cf. figure
5. The mixture structure can be characterised as an underfilled stone
skeleton mixture.
Porous asphalt is valued for its resistance to permanent deformation. It
owes this to the stone skeleton, provided there is sufficient lateral confinement. It was observed that porous asphalt helps to lower the surface
temperature of the pavement during the summer, by approximately 5C. It

The current methodology of asphalt pavement design,


material selection, and asphalt mixture design

% m/m passing through sieve

Sec. 2

29

100

10
DAC 0/8
DAC 0/11
DAC 0/16
1
0.01

0.1

10

100

% m/m passing through sieve

mesh (mm)

100

10

1
0.01

OAC /11
OAC 0/16 t2
OAC 0/16 t3
OAC 0/22

0.1

10

100

% m/m passing through sieve

mesh (mm)

100

CGAC 0/16
CGAC 0/22
GAC 0/16 t1
GAC 016 t2
GAC 0/32

10

1
0.01

0.1

10

100

mesh (mm)

Figure 4. Grading curves of standardised asphalt concrete mixtures. Top (a): dense
graded asphalt concrete, DAC; middle (b): open graded asphalt concrete, OAC;
extensions t2 and t3 indicate type 2 respectively type 3; bottom (c): crushed gravel
asphalt concrete, CGAC, and gravel asphalt concrete GAC; extensions t1 and t2
indicate type 1 respectively type 2. [Data from: CROW 2000].

The current pavement design methodology

% m/m passing through sieve

30

Ch. 2

100

10

PA 0/11
PA 0/16
1
0.01

0.1

10

100

mesh (mm)

% pavement area porous asphalt

Figure 5. Grading curves of standardised porous asphalt mixtures.


[Data from: CROW 2000].
70
60
48

50

51

54

57

43

40

35
30

30
21
20
9.5

13

24

16

10
0
1990 1991 1992 1993 1994 1995 1996 1997 1998 1999 2000 2001

year

Figure 6. Application of porous asphalt on the main road network


since 1991. Numbers above bars indicate km2 porous asphalt.
Total pavement area of main road network: approximately 80 km2.

was suggested this is caused by traffic induced convection in the air voids.
This is expected to have a favourable effect on the temperature
distribution in the whole pavement. A temperature reduction of 5C is
sufficient to cause a significant reduction of the rutting susceptibility of
the pavement.
The large-scale application follows the government policy to develop
noise reducing surface layers, which has been in effect since 1989. Figure
6 depicts the application since 1991. By 2010, the entire main road
network shall be covered with porous asphalt.

The current methodology of asphalt pavement design,


material selection, and asphalt mixture design

% m/m passing through sieve

Sec. 2

31

100

SMA 0/6
SMA 0/8 t1
SMA 0/8 t2
SMA 0/11 t1
SMA 0/11 t2

10

1
0.01

0.1

10

100

mesh (mm)

Figure 7. Grading curves of standardised stone mastic asphalt


mixtures. [Data from: CROW 2000].

2.2.3 Stone mastic asphalt


Stone mastic asphalt was developed in the sixties and seventies, in
response to the need for a more durable pavement material, as traffic loads
increased. It is applied mainly for its durability and resistance to
permanent deformation. In addition, it provides improved noise reducing
capacity and drainage capacity, in comparison to dense graded asphalt
concrete. The mixture structure can be characterised as a filled stone
skeleton mixture.
The experiences with stone mastic asphalt were studied and published by
the CROW-working group Stone mastic asphalt (CROW 1998). Stone
mastic asphalt was developed on an empirical basis in Germany. Positive
experiences led to standardisation in the German standard requirements,
ZTV bit-St. B84, in 1984. In the CROW-Standard 1990, SMA 0/11,
SMA 0/8 and SMA 0/6 were introduced as standard mixtures. In general,
but with SMA 0/11 in particular, the mineral aggregate grading may be
critical, especially in combination with low voids content. This may easily
cause the mixture to become overfilled. Therefore, SMA 0/11 type 2 was
developed, showing a greater gap grading, and corresponding higher
design voids content, cf. figure 7. Simultaneously, the requirements with
respect to shape and hardness of stone were sharpened. Presently,
Rijkswaterstaat
prudently
advises
its
application,
awaiting
recommendations from a running investigation into the volumetric
mixture design.
2.3 The asphalt mixture design method
Asphalt mixture design as officially standardised in The Netherlands, is
based on the Marshall method (CROW 2000). This is a method to

32

The current pavement design methodology

Ch. 2

Figure 8. Marshall test. Top (a): Test set-up. A cylindrical test specimen with a
diameter of 100 mm and a height of 63 mm is loaded diametrically with a
constant deformation rate of 0.85 mm/s until collapse. Bottom (b): A force
displacement curve is measured. The Marshall quotient Qm is determined as the
quotient of the Marshall stability Pm and the Marshall flow, Fm , i.e.
Qm = Pm / Fm .

optimise the binder content of the mixture. The method yields the bitumen
content intervals in which the following properties meet the requirements:
. Marshall stability
. Marshall flow
. Marshall quotient
. voids content
. voids in mineral aggregate filled with bitumen.
The Marshall test is used to determine these properties, cf. figure 8. The
design bitumen content is determined as the median of the common
interval, in which these properties meet the requirements. The mineral
aggregate composition of an asphalt mixture is not actually designed, but
satisfies the grading requirements. The mixture design requirements for

Sec. 3

Discussion and conclusions

33

paving asphalt mixtures are summarised in table 5. The requirements with


respect to the constituents are summarised in table 6.

3 Discussion and conclusions


The current pavement design methodology consists of a pavement design
method, a material selection method, and a mixture design method. The
pavement design method is focused on the pavements bearing capacity.
The method uses two design criteria to control the bearing capacity. These
criteria require respectively the stiffness (complex modulus) and the
fatigue-life of the material. No design criteria are in effect to control the
surface characteristics. Two functional requirements are in effect:
i the pavement evenness, realised during paving, is not a design property;
ii the skid resistance; this is controlled by the properties of the aggregate
constituents in the top pavement layer.
The current material selection method is an empirical method, which
prescribes the use of standardised asphalt mixtures for a given application
(top layer, binder layer, or base layer). The mixtures are developed
empirically, and are described (defined and specified) by composition.
Mechanical asphalt mixture properties are controlled by mixture
composition. Mechanical asphalt mixture properties play no role in the
selection of the mixtures.
The current mixture design method is composition-related. The mixture
design properties, i.e. the Marshall properties, are empirical properties,
which serve merely to find the optimum bitumen content of the mixture.
The fact that the required Marshall properties depend on the traffic class
does not change the essence of the composition-relatedness of the mixture
design method.
The current pavement design methodology is not performance related for
the following reasons:
1 it is not possible to judge the pavements performance, i.e. its costeffectiveness and risk of failure, since the surface characteristics and the
long-term performance are not considered;
2 the design properties, needed for a pavement design, are normally not
available, for do not come available from the mixture design method, or
another method;
3 in the material selection method, only prescribed mixture types can be
selected for a given application (base layer, binder layer, top layer); it is
not possible to solve the material selection problem by optimisation of the
pavements performance;

34

The current pavement design methodology

Ch. 2

Table 5. Mixture design requirements for asphalt mixtures used


by Rijkswaterstaat (CROW 2000).
Mixture

porous asphalt
0/16 tc 2, 3, 4, 5
0/11 tc 2, 3, 4, 5
stone mastic asphalt
0/11 type 1 tc 2
0/11 type 2 tc 2, 3, 4, 5
0/6 tc 2
dense graded
asphalt concrete
0/16 tc 2
0/16 tc 3
0/16 tc 4
0/11 tc 2
0/11 tc 3
0/11 tc 4
open graded
asphalt concrete
0/22 tc 2
0/22 tc 3
0/22 tc 4
0/22 tc 5
0/16 type 2 tc 2
0/16 type 2 tc 3
0/16 type 3 tc 2
0/16 type 3 tc 3
0/16 type 3 tc 4
0/16 type 3 tc 5
0/11 tc 2
0/11 tc 3
crushed gravel asphalt
concrete
0/22 tc 4
0/22 tc 5
0/16 tc 4
0/16 tc 5
gravel asphalt concrete
0/32 tc 2
0/32 tc 3
0/16 type 1 tc 2
0/16 type 2 tc 2
0/16 type 2 tc 3

Marshall Marshall Marshall bitumen


voids
stability
flow
quotient content content
(N)
(mm)
(N/mm) (% m/m) (% v/v)

VFB
(% v/v)

4.5
4.5

20.0
20.0

7.0
7.0
8.0

6500
7000
7500
6500
7000
7500

2.0 - 5.0
2.0 - 4.0
2.0 - 4.0
2.0 - 5.0
2.0 - 4.0
2.0 - 4.0

2000
2500
3000
2000
2500
3000

6.2 - 6.6
6.2 - 6.6
6.0 - 6.4
6.4 - 6.8
6.4 - 6.8
6.2 - 6.6

4.0
4.0
6.0
4.0
4.0
6.0

87
80
80
87
83
80

5500
6500
7000
7500
5500
6500
5500
6500
7000
7500
5500
6500

2.0 - 4.0
2.0 - 4.0
2.0 - 4.0
2.0 - 4.0
2.0 - 4.0
2.0 - 4.0
2.0 - 4.0
2.0 - 4.0
2.0 - 4.0
2.0 - 4.0
2.0 - 4.0
2.0 - 4.0

2000
2500
3000
3000
2000
2500
2000
2500
3000
3000
2000
2500

4.2 - 5.0
4.2 - 5.0
4.0 - 5.0
4.0 - 5.0
4.6 - 5.4
4.6 - 5.4
5.0 - 5.8
5.0 - 5.8
4.8 - 5.8
4.8 - 5.8
4.6 - 5.4
4.6 - 5.4

7.0
7.0
7.0
7.0
7.0
7.0
7.0
7.0
7.0
7.0
7.0
7.0

77
75
72
70
77
75
77
75
72
70
77
75

6000
7000
6000
7000

1.5 - 3.0
1.5 - 3.0
1.5 - 3.0
1.5 - 3.0

3000
3500
3000
3500

4.0-5.0
4.0-5.0
4.0-5.0
4.0-5.0

7.0
7.0
7.0
7.0

50 - 68
50 - 65
50 - 68
50 - 65

4500
5000
4500
4500
5000

1.5 - 3.0
1.5 - 3.0
1.5 - 3.0
1.5 - 3.0
1.5 - 3.0

2000
2500
2000
2000
2500

4.5 - 5.5
4.0 - 5.0
4.5 - 5.5
4.5 - 5.5
4.0 - 5.0

7.0
7.0
7.0
7.0
7.0

50 - 75
50 - 72
50 - 75
50 - 75
50 - 72

VFB = voids in mineral aggregate filled with bitumen,


tc = traffic class, see table 1.

Sec. 3

Discussion and conclusions

35

Table 6. Requirements with respect to mixture constituents of asphalt mixtures


used by Rijkswaterstaat (CROW 2000).
Mixture

emulsion asphalt
concrete
0/3 0/6 0/8 tc 2, 3, 4, 5
porous asphalt
tc 2, 3, 4, 5

stone

crushed
stone

sand

filler

bitumen

medium strength1
with
Ca(OH)2 addition

70/100
70/100

stone mastic asphalt


crushed
0/11 type 1 tc 2
weak lime
0/11 type 2 tc 2, 3, 4, 5 stone sand C
0/6 tc 2
sand D
dense asphalt concrete
tc 2, 3
crushed sand B
weak, or
tc 4
stone sand B medium strength
dense asphalt concrete
with asphalt granulate
crushed crushed
weak, or
tc 2,3
tc 4
stone
sand medium strength
open asphalt concrete
tc 2, 3

open asphalt concrete


tc 4, 5
open asphalt concrete
with asphalt granulate
tc 2,3
tc 4,5
crushed gravel asphalt
concrete
tc 2

crushed
stone

-2

70/100
70/100
70/100

70/100
40/60

40/60
or
70/100

crushed sand B
weak, or
medium strength
stone

40/60

weak, or
medium strength

anti binder
segregation
additive

70/100
40/60

weak, or
medium strength

crushed
stone

additive

40/100
40/60

granulated old
asphalt,
max. 50%

granulated old
asphalt,
max. 50%

weak
40/60 or
crushed
stone
70/100
tc 3, 4, 5
crushed
weak
40/60
stone
crushed gravel asphalt concrete with asphalt granulate
tc 2
weak
40/100
old asphalt,
tc 3, 4, 5
weak
40/60
max. 50%
gravel asphalt concrete
tc 2
weak
40/603
tc 3, 4, 5
weak
40/60
gravel asphalt concrete with asphalt granulate
tc 2
weak
40/100
old asphalt,
tc 3, 4, 5
weak
40/60
max. 50%
1
filler strength is defined by voids Rigden, or the Bitumen number (SVC 1982)
2
sand B is required for 0/16 type 3 tc 2, 3, 4, 5.
3
also 80/100 bitumen is permitted.
Sand A : a natural sand, or mixture of natural sands. Sand B : at least 75% crushed sand.
Sand C : at least 50% crushed sand. Sand D : at least 50% natural sand.

36

The current pavement design methodology

Ch. 2

4 it is not possible to judge an asphalt mixture by its performance


properties related to the pavements main functions, bearing capacity,
surface characteristics, and long-term performance; the Marshall mixture
design properties serve to determine the (read: an arbitrary) asphalt
mixtures optimum bitumen content, however, are not relevant to the
pavement design, or the material selection; i.e. are not used and not
suitable to optimise the pavements performance;
5 it is not possible to accept a newly developed (type of) asphalt mixture,
if the composition deviates from known standardised asphalt mixtures,
and if that causes reason to doubt the associated risk of failure or costeffectiveness in a given (standard) application.
Possible improvements
1 The pavement design method could be modified,
i to permit a judgement of a pavement design on its three main functions,
bearing capacity, surface characteristics, and long-term performance;
ii to improve the control of the decline of the surface characteristics;
iii to perform a cost-effectiveness judgement and a risk analysis of the
pavement design.
2 The asphalt mixture design method could be modified,
i to make available asphalt mixture properties needed in pavement design,
and to solve the associated material selection problem;
ii to permit judgement of asphalt mixtures performance properties
relevant to the pavements main functions;
iii to perform a cost-effectiveness judgement of an asphalt mixture in a
given application (pavement design);
iv to facilitate acceptation of newly developed asphalt mixtures.

3
Characterisation of viscoelastic and
viscoplastic behaviour of asphalt
mixture
1 Aim
The aim of the analysis presented in this chapter is to investigate the
experimental evidence of the complex modulus and the linear viscoelastic
creep susceptibility in the four point bending test, and the viscoplastic
creep susceptibility in the uniaxial creep test and triaxial creep test, to
come to a judgement as to whether or not it is justified to have confidence
that these test methods are suitable for use in a practical context to
characterise the viscoelastic/viscoplastic behaviour. The ultimate aim is to
be able to judge the cost-effectiveness and the risk of failure of the asphalt
mixture when applied in a pavement. The tests considered are the four
point bending test as frequency sweep test and the creep test with and
without confinement stress.

2 Methodology
The method to come to the above judgement comprises a study of the
complex modulus, the linear viscoelastic creep susceptibility, and the
viscoplastic creep susceptibility with the purpose to investigate whether or
not these properties can be obtained as true material properties.
In general, the stress strain behaviour of asphalt mixtures is nonlinearly
elasto-viscoplastic, cf. Intermezzo 1. However, it can be shown that the
stress strain behaviour of asphalt mixture can be approximately linearly

38

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

Ch. 3

viscoelastic under specific conditions. This can be based on the validity of


the time temperature superposition principle, and on the reproducibility of
the complex modulus using different specimen geometries. The latter was
shown in an international interlaboratory investigation (Partl and Francken
1997). If the real behaviour of a given asphalt mixture shows negligible
deviations from the linearly viscoelastic behaviour, then, from a practical
viewpoint, the behaviour of that asphalt mixture might be considered
linearly viscoelastic. Based on the validity of the time temperature
superposition principle and the reproducibility of the complex modulus
using different specimen geometries, a physical meaning can be attributed
to linear viscoelastic asphalt mixture properties. Subsequently, a physical
meaning (predictive value) can be attributed to the viscoplastic creep
susceptibility, based on the correspondence of nonlinear viscoplastic creep
and linear viscoelastic creep.

3 Theory
3.1 Linear viscoelastic creep susceptibility of asphalt mixture
Let us assume linear viscoelastic creep. This is described by the following
equation,
(t )
J (t ) =
= J 0 + J1 t m
(1)
[ ]
where J ( t ) is the creep compliance as a function of the time, t , ( t ) is
the creep strain, [ ] represents a stress invariant, J 0 is the instantaneous
elastic compliance at t = 0, J1 is the permanent creep compliance at 1 s,
and m is a material constant. In triaxial creep, the following stress
invariants can be defined,
q = 1 3
(2a)
+ 23
p= 1
(2b)
3
where q is the deviatoric stress, p is the volumetric stress, 1 is the
maximum axial stress, and 3 is the maximum radial stress. Note, that
both 1 and 3 are compressive stresses, which, for convenience, have
been defined with a positive sign. From equation (1), with the aid of the
Laplace transform, the following expressions can be obtained for the real
and imaginary parts of the complex creep compliance, J * , respectively
(Schapery 1974):
m
J ( ) = J 0 + J 1 (1 + m) m cos
(3a)

m
J ( ) = J1 (1 + m) m sin
(3b)

2
where J ( ) is the real part of the complex compliance as function of the
angular frequency, , J ( ) is the imaginary part of the complex

Sec. 3

Theory

39

Intermezzo 1
The stiffness modulus of asphalt mixture is stress dependent, i.e. differs in
tension and in compression (Monismith 1962, 1966). Therefore, the stiffness
modulus and the complex modulus vary as function of the height of the bending
beam in the four point bending test. Thus, the complex modulus of an asphalt
mixture in the four point bending test, or any other frequency sweep test, is not
really a fundamental material property of an asphalt mixture. Rather it is a
property of the specimen.
Owing to the difference between the stiffness modulus in tension and in
compression, the neutral plane of the bending beam is not continuously halfway
the height of the beam, but oscillates between two extreme positions. Owing to
the fact that the stiffness modulus in tension, S + , is smaller than the stiffness
modulus in compression, S , the neutral zone is a zone under continuous
tension. Therefore, the neutral zone is expected to experience tensile creep. If
the applied strain amplitude is small, and if the duration of the test is not too
long, this creep is negligible. If the creep is negligible, then the actual
viscoplasticity of the material of the bending beam remains undetected.
Seemingly, the material behaves linearly viscoelastic. Owing to S being
greater than S + there is relatively abundant tensile stress and strain. Possibly,
this explains why m in equation (1) as obtained in four point bending may be
relatively large in comparison to ~z in equation (8) as obtained in compressive
creep.

modulus, (1 + m) represents the Gamma-function with argument


(1 + m) , and J 0 , J1 , and m is the same parameter as in equation (1). The
complex creep compliance is defined as
J * = J + iJ
(4)
where i = 1 . Substitution of equations (3) into equation (4) yields
J * = J 0 + J 1 (1 + m) m e i m / 2
(5)
Taking logarithms on both sides, and differentiating with respect to ln
yields, for a sufficiently long time, i.e. J 0 << J 1 t m ,
(ln J * )
= m
(6)
(ln )
According to equation (1), m represents the slope of the linear viscoelastic
creep curve on log-log scale. Therefore, m can be considered the linear
viscoelastic creep susceptibility. Note, that, in general, J 0 is negligible
with respect to J1 . Therefore, according to equation (6), m represents the
slope of the complex modulus master curve on log-log scale. Thus, the
linear viscoelastic creep susceptibility can be obtained from the complex
modulus master curve, which, itself, is a characterisation of the dynamic

40

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

Ch. 3

linear viscoelastic stress strain behaviour. The time temperature


superposition principle is a fundamental principle of physics. Linear
viscoelasticity as a phenomenon is a form of this principle. In
mathematical form, the principle can be formulated by means of the
following two equations1:

(t ) = J (t )d ( )

(7a)

(t ) = S (t )d ( )

(7b)

where (t ) is the strain, J (t ) is the compliance, (t ) is the stress, S (t ) is


the stiffness modulus, and is the reference time. Let us define a
fundamental constitutive model as a set of mathematical equations which
describes a fundamental principle of physics. Then equations (7) represent
a fundamental constitutive model. It can be shown that the complex
modulus master curve, i.e. its continuity, is a consequence of the validity
of this principle, cf. Ch. 5, Intermezzo 1. In fact, the complex modulus
master curve is described by equation (7), and also by equation (6).
Equation (6) can be derived mathematically from equation (1), as shown.
Thus, linear viscoelastic creep as described by equation (1), and dynamic
linear viscoelastic stress strain behaviour as described by equation (6), are
both described by the same constitutive model, equation (7). Let us define
a constitutive model as a fundamental constitutive model if it represents a
fundamental principle of physics. Then, since the time temperature
superposition principle is a fundamental principle of physics, equation (7)
can be considered a fundamental constitutive model (only for a
homogeneous continuum). Since equation (7) implies equation (6), which,
in turn, can be derived mathematically from equation (1), equations (1)
(6) together can be considered a fundamental constitutive model.
Therefore, the parameters of this model, J 0 , J1 , and m , can be considered
fundamental properties of a linear viscoelastic continuum. This implies
that also the constitutive model variables, complex compliance, J * , and
the complex modulus, S * = 1 / J * , can be considered fundamental
properties of a linear viscoelastic continuum. The linearity of the model
implies that the constitutive model parameters, J 0 , J1 , and m , the
complex compliance, J * , and the complex modulus, S * , are independent
of the stress.
Equation (7) shows that linear viscoelastic stress strain behaviour is time
dependent, and that the creep compliance and the stiffness modulus are
time dependent properties. This implies that the shape of waveform of the
1

This is the form in which the integrals were introduced in the theory of the
viscoelastic after-effect by Boltzmann (1874). For a more detailed description
the reader is referred to Ch. 5.

Sec. 3

Theory

41

applied stress or strain influences the properties. To gain insight in the


consequences of this time dependence, the Burgers equation was solved
for different shapes of waveform of applied stress, cf. appendix 1. In
appendix 1, it is shown for a Burgers material, that the response strain,
( t ) , to an applied stress, ( t ) , depends on the shape of the waveform of
the applied stress. If the creep term of ( t ) is defined as the term which
causes accumulated creep strain as function of the time, creep (t ) , then the
creep strain rate, &creep , also depends on the shape of the waveform of the
applied stress.
In general, an asphalt mixture exhibits nonlinear stress strain behaviour.
Two reasons can be given for this:
1 Nonlinearity of stress strain behaviour as function of the stress is an
intrinsic property of an unbound confined grain skeleton2.
2 Nonlinearity of stress strain behaviour as a function of the time and as a
function of the stress is an intrinsic property of the bituminous binder3.
This nonlinearity is transferred to the asphalt mixture.
However, asphalt mixtures can exhibit approximate linearly viscoelastic
behaviour under special conditions4. In principle, if a complex modulus
master curve can be constructed, then the time temperature superposition
principle is valid. It means that the stress strain behaviour is linearly
viscoelastic, and implies that there is a linear domain where constitutive
model parameters, J 0 , J1 , and m , and constitutive model variables, the
complex compliance, J * , and the complex modulus, S * , are independent
of the stress.
A true material property is defined as a property, which can be reproduced
independent of the specimen geometry (size and shape), and is not
influenced by the measurement itself.
Strictly, S * and m of an asphalt mixture are not true material properties,
cf. Intermezzo 1. However, it was shown by Partl and Francken (1997),
that the complex modulus, S * , is reasonably reproducible using different
specimen geometries. Therefore, by approximation, S * of asphalt mixture
can be considered a true material property. If S * of asphalt mixture is
obtained as a true material property, then, by equation (6), also the slope
of the complex modulus master curve on log-log scale, m , can be
considered a true material property.

An unbound grain skeleton confined in a box shows resistance to deformation


under compressive stress, but lacks resistance to deformation under tensile
stress.
3
Bitumen is a thixotropic and pseudoplastic fluid.
4
Strain controlled frequency sweep test, small applied strain amplitude, short
test duration.

42

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

Ch. 3

If a rheological model consisting of springs and dashpots, e.g. a Burgers


model, cf. appendix 1, is used to model the creep of a linear viscoelastic
material, then it is found that the material is linear as a function of the
stress and as a function of the time. If the creep of a linear viscoelastic
material is linear as a function of the time, then m in equation (1) is equal
to 1. This linearity as a function of the time is a consequence of the use of
Newtons viscosity law to define the dashpot of the rheological model. In
general, however, a linear viscoelastic material is only linear as a function
of the stress, which means that m may differ from unity. It is suggested
here to consider that the linear viscoelastic creep susceptibility, m , for an
asphalt mixture, might be determined from the complex modulus master
curve.

3.2 Viscoplastic creep susceptibility of asphalt mixture


The creep of asphalt mixture may be described by
~
J (t ) = J 0 + J 1 t z
(8)
where J ( t ) is the creep compliance as a function of the time, t , and J 0 ,
J1 , and ~
z are constants. Based on equation (8), it is suggested here to
z the viscoplastic creep susceptibility, in analogy to the linear
consider ~
viscoelastic creep susceptibility, m , in equation (1).
Equation (8) is merely a suitable mathematical equation to describe the
creep. A mathematical equation which is merely descriptive, i.e. which is
not derived from a fundamental principle of physics, is generally called a
phenomenological equation. Also other phenomenological creep models
may be found to describe different types of creep behaviour, cf. 6.
Equation (8) is special because of the analogy to the linear viscoelastic
creep law, equation (1). This analogy provides a basis for the physical
z in equation (8) for asphalt mixtures. This
interpretation of the parameter ~
is elaborated in the remainder of this chapter.
Let us assume that the mathematical definition of J ( t ) in equation (8) is
the same as that of J ( t ) in equation (1). Thus, a parameter of a
phenomenological equation can be a fundamental property. This does not
mean that a fundamental property is a true material property. While S *
and m of asphalt mixture can be obtained as true material properties from
a practical viewpoint, J ( t ) in equation (8) is found to depend on the
specimen height, cf. 8.4. It is made plausible in 9, that the dependence of
J ( t ) of an asphalt mixture on the specimen height is caused by the
material structure of the asphalt mixture, and is therefore to be considered
an intrinsic property of the mixture, cf. 9. This means, that the dependence
on the specimen height cannot be avoided or eliminated.

4 Experimental details
Four test methods are investigated, cf. 4.1. The influences of the stress

Sec. 5

Results

43

Figure 1. Four point bending test.

condition, the shape of the waveform of the applied stress, and the
specimen size are investigated, cf. 4.2. Also the influence of the mixture
composition is investigated, cf. 4.3.

4.1 Test methods


The following four test methods are investigated:
1 the dynamic four point bending test as frequency sweep test (4PB)
2 the static uniaxial compression creep test (static UCC)
3 the dynamic uniaxial compression creep test (dynamic UCC)
4 the dynamic triaxial compression creep test (TCC).
The tests were carried out according to the normal practice at the time of
the experiments, laid down in documented test procedures. The four point
bending test is used as a frequency sweep test, to determine the master
curves of the complex modulus and the loss angle. The creep test is used
to determine the resistance to permanent deformation. Originally, only the
static UCC was used for asphalt concrete mixtures. Later, the dynamic
UCC was recommended for polymer modified asphalt mixtures (Valkering
et al. 1990). Again, later, the dynamic TCC was recommended for stone
skeleton mixtures (Molenaar et al. 1995).
4.2 Test set-up and testing conditions
4.2.1 Four point bending test
Figure 1 is a representation of the test set-up. It represents a prismatic
beam, 50 x 50 x 450 mm, subjected to four point bending with supposedly
free rotation and translation at all load and reaction points (it can be

44

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

Ch. 3

Figure 1 (Contd). Schematic representation of the four point bending test.

doubted whether this is indeed realised). The bending is realised by a


sinusoidal movement of the centre bearing, in vertical direction. The
vertical position of the end-bearings is fixed. The applied displacement
was generated by means of a Schenck Hydropuls servo-hydraulic testing
device. A displacement transducer for the measurement of the deflection,
not shown in figure 1, is positioned at the middle of the upper surface of
the test specimen, where the deflection is at maximum. The measuring
range of the displacement transducer was 1 mm, and its resolution,
according to the specifications, 1 m. The measuring range of the load cell
for the measurement of the force was 2000 N. The resonance frequency of
the load cell and the coupled moving mass is at least 10 times higher than
the test frequency. The load-cell is positioned on the hydraulic jack, not
shown in figure 1. The applied deflection amplitude corresponds to a
maximum strain amplitude of 50 m/m. Test temperatures vary. In part,
the master curves were constructed from series of measurements at -10C,
0C, 10C, and 20C, and in part from series measurements at 0C, 10C,
20C, and 40C. The frequency was varied between 1 Hz and 60 Hz.

4.2.2 Static uniaxial compression creep test


A cylindrical specimen is subjected to a static uniaxial load. The principle
of the test is similar to that shown in figure 2 for the triaxial creep test,
except that no radial stress is applied. The applied stress was generated by
means of a Schenck Hydropuls servo-hydraulic testing device. The axial
deformation was measured as the change in the distance between the
loading plates. The radial deformation was measured using two
diametrically placed displacement transducers, situated halfway the height
of the specimen. The Schenck Hydropuls apparatus permitted tests on
specimens up to 100 mm in height. Measurements were performed at two
temperatures: 40C and 50C. Two applied stresses were used: 800 N,
corresponding to 0.1 MPa, and 1600 N, corresponding to 0.2 MPa, cf.
table 1. The friction between the specimen and the loading plate was

Sec. 5

Results

45

Figure 2. Triaxial cell of triaxial creep test.

reduced according to the standard procedure at the time of the


experiments. This procedure was as follows: The specimen end-faces were
polished plane parallel, and subsequently greased with talcum powder and
glycerine. This procedure was developed by Bolk (1980).

4.2.3 Dynamic uniaxial compression creep test


The principle of the test is similar to that shown in figure 2 for the triaxial
creep test, except that no radial stress is applied. Measurements were
performed at two temperatures: 40C and 50C. Two applied stress
amplitudes were used: 800 N, corresponding to 0.1 MPa, and 1600 N,
corresponding to 0.2 MPa. Different shapes of waveform of the applied
stress were used, cf. table 1. To investigate the departure from static
loading, a rest-time of 1.8 s was introduced after a loading time of,
respectively, 600 s, 10 s, 1 s, 0.2 s, and 0.05 s. Subsequently, the loading
time was held constant at 0.2 s, and the following rest-times were used:
0.2 s, 0.5 s, 1.0 s, and 5.0 s (cf. table 1 on page 45). The friction between
the specimen and the loading plate was reduced according to the standard
procedure at the time of the experiments, cf. 4.2.2.
4.2.4 Dynamic triaxial compression creep test
The test set-up is represented in figure 2. A Schenck Hydropuls servohydraulic testing device permitted tests on specimens up to 100 mm in
height. At the Delft University of Technology, a pneumatically controlled

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

46

A) displacement
transducer actuator
B) load cell
C) axial displacement
transducer (one of two)
D) radial displacement
transducer (one of
three)
E) air supply for
confinement pressure
F) control unit
confinement pressure
G) closure rings to
separate specimen from
water
H) water supply
I) water outlet
J) water for
confinement pressure
K) triaxial cell
L) frame
M) test sample

A
L

E
F
K

B
C
D
J

Ch. 3

G
M
G
H
I

Figure 2 (Contd). Schematic representation of the triaxial creep test.

test device, made by the Australian IPC, permitted tests to be performed


on specimens up to 200 mm in height having the conventional height to
diameter ratio of 2. Measurements were performed at 50C. Various
combinations of axial stress amplitudes and radial confinement stress were
applied, indicated in table 1. The maximum applied axial stress was 0.9
MPa, and the maximum radial stress was 0.3 MPa. Two shapes of waveform were used; a block-wave with a loading time of 0.2 s, and a rest-time
of 0.8 s hence a frequency of 1 Hz and a sinusoidal waveform, also
with a frequency of 1 Hz. The same block-wave was used on the hydraulic
device and the pneumatic device. It is noted, that the hydraulically
controlled waveform corresponds accurately to the desired waveform,
whereas the pneumatically controlled waveform deviates slightly from the
desired waveform. This is illustrated in figure 3.
The combinations of applied axial stress and radial stress in table 1 are
shown in figure 4. The friction between the specimen and the loading plate
was reduced according to the standard procedure at the time of the
experiments, cf. 4.2.2.

Results

applied stress
(normalised)

Sec. 5

47

0
0

Figure 3. Shape of waveform of the


applied stress, with hydraulic load
control and pneumatic load control.

0,2

0,4

0,6 0,8
time (s)

1,2

hydraulically controlled block-w ave


pneumatically controlled block-w ave

Table 1. Testing conditions applied in the different creep tests.


Specimen: 60 x 100 mm height x diameter.
load
loading restAxial temp.
time
signal
time
load
(s)
(s)
(MPa) (C)
static creep
0.2
40
constant

0
dynamic creep
0.1
40
block wave
600
1.8
10
1.8
1
1.8
0.2
1.8
0.05
1.8
dynamic creep
0.2
40
block wave
600
1.8
uniaxial
10
1.8
1
1.8
0.2
1.8
0.05
1.8
dynamic creep
0.1
50
block wave
600
1.8
uniaxial
10
1.8
1
1.8
0.2
1.8
0.05
1.8
dynamic creep
0.2
50
block wave
600
1.8
uniaxial
10
1.8
1
1.8
0.2
1.8
0.05
1.8
dynamic creep
0.1
40
block wave
0.2
0.2
uniaxial
0.2
0.5
0.2
1.0
0.2
1.8
0.2
5.0
The minimum axial force in the unloaded situation was 20 N.

period

freq.

601.8
11.8
2.8
2.0
1.85
601.8
11.8
2.8
2.0
1.85
601.8
11.8
2.8
2.0
1.85
601.8
11.8
2.8
2.0
1.85
0.4
0.7
1.2
2.0
5.2

(Hz)
0.00166
0.0847
0.357
0.5
0.54
0.0017
0.085
0.36
0.5
0.54
0.0017
0.085
0.36
0.5
0.54
0.0017
0.085
0.36
0.5
0.54
2.5
1.43
0.83
0.5
0.19

48

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

Ch. 3

Table 1 (Contd). Testing conditions applied in the different creep


tests. Specimen: 60 x 100 mm height x diameter.
freq.
load
period
Axial temp.
signal
load
(Hz)
(s)
(MPa) (C)
dynamic creep
0.1
40
sinus
10
0.1
uniaxial
1
1
0.1
10
dynamic creep
0.1
50
sinus
10
0.1
uniaxial
1
1
0.1
10
dynamic creep
0.2
40
sinus
10
0.1
uniaxial
1
1
0.1
10
dynamic creep
0.2
50
sinus
10
0.1
uniaxial
1
1
0.1
10
The minimum axial force in the unloaded situation was 20 N.

deviatoric stress, q (MPa)

1.6
axial stress
1.4

0,9

1.2

0,8

radial stress
0,1
0,2

0,3

0,7

0,6
0.8

0,5

0.6

0,4
0,3

0.4

0,2
0.2

0,1

0
0

0.2

0.4

0.6

0.8

volumetric stress, p (MPa)

Figure 4. Combinations of applied axial stress and radial stress


in the q - p -space.

4.2.5 Modified friction reduction system


Recently, a new friction reduction system was introduced, cf. appendix 2,
to control the friction between the specimen and the loading plate. The
first experiences with this new system indicate that the creep is strongly
influenced by the friction (reduction) between the specimen and the
loading plate. The data discussed in the present investigation were
obtained using the old friction reduction method, described in 4.2.2. A
limited comparative investigation was performed, to compare results
obtained with different friction reduction systems, cf. appendix 2.

Sec. 5

Results

Table 1 (Contd). Testing conditions applied in the different creep tests.


radial
load
axial
Temp
p
q
load
signal
load
.
(MPa) (MPa)
(C)
specimen height: 60 mm
dynamic creep
50 block-wave
0.1
0
0.033
0.1
triaxial
hydraulic
0.3
0.1
0.167
0.2
control
0.5
0.1
0.233
0.4
0.7
0.1
0.300
0.6
0.9
0.1
0.367
0.8
dynamic creep
50 block-wave
0.3
0.2
0.233
0.1
triaxial
hydraulic
0.5
0.2
0.300
0.3
control
0.7
0.2
0.367
0.5
0.9
0.2
0.433
0.7
dynamic creep
50 block-wave
0.4
0.3
0.333
0.1
triaxial
hydraulic
0.5
0.3
0.367
0.2
control
0.6
0.3
0.4
0.3
0.7
0.3
0.433
0.4
0.8
0.3
0.467
0.5
0.9
0.3
0.500
0.6
specimen height: 100 mm
dynamic creep
50 block-wave
0.7
0.1
0.300
0.6
triaxial
hydraulic
0.9
0.1
0.367
0.8
specimen height: 200 mm
dynamic creep
50 block-wave
0.3
0.03
0.120 0.27
triaxial
pneumatic
0.6
0.05
0.233 0.55
control
0.6
0.10
0.266 0.50
0.6
0.15
0.300 0.45
0.75
0.15
0.350 0.60
0.3
0.20
0.233 0.10
0.6
0.25
0.367 0.35
0.9
0.30
0.500 0.60
The minimum axial force in the unloaded situation was 20 N.

49

q/ p

3.00
1.20
1.72
2.00
2.18
0.43
1.00
1.36
1.63
0.30
0.54
0.75
0.93
1.07
1.20
2.00
2.18
2.25
2.36
1.88
1.50
1.71
0.43
0.95
1.20

4.3 Materials
An overview of the asphalt mixtures used, and the investigations are given
in table 2.
Table 2, section A. The mixtures GAC 0/22 M1..M3 were investigated as a
part of an investigation into the recyclability of polymer modified porous
asphalt for the Direction East Netherlands of Rijkswaterstaat, in
connection to a large maintenance operation on motorway A12
Grijsoord/Waterberg (1997). The RHEI took the opportunity to investigate
the possibility to increase the recycling percentage of polymer modified
asphalt granulate from the 20%, recommended by CROW (1996), to 40%.

50

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

Ch. 3

Table 2. Overview of mixtures and investigated tensile properties.


Sec- type of asphalt mixture
tion
A
laboratory manufactured
three crushed gravel asphalt concrete 0/22 mixtures with
40% asphalt granulate from reclaimed polymer modified
porous asphalt, and respectively 4.0%, 4.5% and 5.0%
80/100 bitumen; in the text, the mixtures are indicated
respectively by M1, M2 and M3.
B
laboratory manufactured
five crushed gravel asphalt concrete 0/22 mixtures,
respectively with 0%, 20%, 40%, 60%, and 80% reclaimed
gravel asphalt concrete, in the text indicated respectively by
G1, G2, G3, G4, and G5.

Investigation

S * and m .
S * as function of

laboratory manufactured
polymer modified porous asphalt 4/11, with bitumen
composed of 97% Styrelf 13-80 and 3% Evathane.
porous asphalt 11/16 with 80/100 bitumen.
polymer modified porous asphalt 11/16, with polymer
modified bitumen compound
polymer modified guss-asphalt 2/8, with bitumen composed
of 96% Styrelf 13-80 and 4% Evathane.
laboratory manufactured and plant-mixed
5 projects: road trials
crushed gravel asphalt concrete 0/22 with 4.5% 70/100
crushed gravel asphalt concrete 0/22 with 4.0% 70/100
crushed gravel asphalt concrete 0/22 with 5.0% 70/100
crushed gravel asphalt concrete 0/22, plant-mixed, slabcompacted at the plant
crushed gravel asphalt concrete 0/22, plant-mixed, laid in the
pavement
laboratory manufactured
porous asphalt 0/16 with 4.0% 70/100 bitumen
porous asphalt 0/16 with 5.0% 70/100 bitumen
porous asphalt 0/11 with 4.0% 70/100 bitumen
porous asphalt 0/11 with 5.0% 70/100 bitumen
porous asphalt 0/16 with 4.0% polymer modified bitumen
porous asphalt 0/16 with 5.0% polymer modified bitumen
porous asphalt 0/11 with 4.0% polymer modified bitumen
porous asphalt 0/11 with 5.0% polymer modified bitumen
plant mixed
dense asphalt concrete 0/16 with 6.0% 80/100 bitumen
laboratory manufactured
dense asphalt concrete 0/16 with 6.2% elastomer modified
bitumen
dense asphalt concrete 0/16 with 6.2% plastomer modified
bitumen
laboratory manufactured
dense asphalt concrete 0/16 with 6.1% 45/60 bitumen
porous asphalt 0/16 with 4.1% 80/100 bitumen

S * and m .
S * as function of
frequency and
temperature, and
mixture composition

S * and m .
S * as function of
frequency and
temperature, and
mixture composition

frequency and
temperature, and
mixture composition

S * and m .
S * as function of
frequency and
temperature, and
mixture composition

S * and m .
S * as function of
frequency and
temperature, and
mixture composition

J (t ) and related
constitutive model
parameters
J (t ) and related
constitutive model
parameters
J (t ) and related
constitutive model
parameters

Sec. 4

Experimental details

51

Table 2, section B. The mixtures GAC 0/22 R1, G1..G5 were investigated
as a part of an investigation into the possibility of adding reclaimed gravel
asphalt concrete to crushed gravel asphalt concrete, for the Direction
South-Holland of Rijkswaterstaat. The investigation was performed in
connection to the reconstruction of motorway A16 near Groenix van
Zoelenlaan/Langenweg (1998/1999). In this reconstruction, an estimated
400,000 tons of asphalt granulate became available for re-use, of which
100,000 tons of dense asphalt concrete and open asphalt concrete, and
300,000 tons of gravel asphalt concrete.
Table 2, section C. The polymer modified porous asphalt mixture PA 4/11
was first applied in 1994, in the Merwede bridge-deck, and in 1996 in the
right lane porous asphalt layer of the Middachten bridge-deck. The binder
used consisted of 97% Styrelf 13-80 and 3% Evathane.
A reference porous asphalt, PA 11/16, containing pure 80/100 bitumen,
and the same mixture, densely filled with a polymer modified compound.
This is indicated in the text as PAM 11/16.
A polymer modified gussasphalt, GAM 2/8, used by the Bouwdienst of
Rijkswaterstaat, for bridge-decks. The binder used consisted of 96%
Styrelf 13-80 and 4% Evathane.
Table 2, section D. Laboratory manufactured and slab-compacted, and
plant-mixed and in the road compacted crushed gravel asphalt concrete,
prepared for 5 projects (road trials). Per project, the following mixtures are
compared:
1 the mixture according to the Marshall mixture design method (MD);
2 the MD mixture minus 0.5% bitumen (MD -);
3 the MD mixture plus 0.5% bitumen (MD +);
4 the mixture from the plant, slab-compacted at the plant (PLA);
5 the mixture from the pavement (PAV).
The following amounts of reclaimed asphalt were used: road trial A: 50%
m/m, road trial B: 40% m/m, road trial C: 40% m/m, road trial D: 45%
m/m, and road trial E: 0% m/m.
Table 2, section E. Four porous asphalt mixtures PA 0/16 and PA 0/11
were tested to investigate the influence of the mixture composition
(grading and bitumen content) on the complex modulus:
P1: PA 0/16 with 4.0% 70/100 bitumen;
P2: PA 0/16 with 5.0% 70/100 bitumen;
P3: PA 0/11 with 4.0% 70/100 bitumen;
P4: PA 0/11 with 5.0% 70/100 bitumen;
PM1: PA 0/16 with 4.0% elastomer modified bitumen;
PM2: PA 0/16 with 5.0% plastomer modified bitumen;
PM3: PA 0/11 with 4.0% elastomer modified bitumen;
PM4: PA 0/11 with 5.0% plastomer modified bitumen.
Table 2, section F. A dense asphalt concrete mixture DAC 0/16 was tested
to compare the creep in the static uniaxial creep test, the dynamic uniaxial

52

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

Ch. 3

creep test, and the dynamic triaxial creep test:


DAC 0/16 with 6.0% 80/100 bitumen.
Table 2, section G. Two polymer modified dense asphalt concrete
mixtures were tested to compare the creep in the static uniaxial creep test
and the dynamic uniaxial creep test:
DAC 0/16 E with 6.2% elastomer modified bitumen
DAC 0/16 P with 6.2% plastomer modified bitumen.
Table 2, section H. A dense asphalt concrete mixture, DAC 0/16, and a
porous asphalt mixture, PA 0/16, were tested to compare the creep in the
triaxial creep test of these mixtures:
DAC 0/16 with 6.1% 45/60 bitumen
PX1: PA 0/16 with 4.1% 80/100 bitumen
The DAC 0/16 mixture is considered representative for asphalt concrete
mixtures, and the PA 0/16 mixture is representative for stone skeleton
mixtures.

5 Results of frequency sweep tests Complex modulus


Figure 5 shows an example of a result of a four point bending test as
frequency sweep test and the subsequent analysis. Similar results were
obtained for the other mixtures of table 2. Figure 5a shows the stiffness
master curves, i.e. the complex modulus as function of the reduced
frequency, of three different specimens. The reduced frequency, f r , is
defined as the product of the frequency, f , and the shift factor, aT . The
shift factor the Arrhenius type, is defined as
1
1

aT = exp K
(9)
T
T
R

where K is a constant, T is the temperature, and TR is the reference
temperature. Figure 5b shows the average complex modulus of the three
specimens. Figuur 5b is used for the determination of the following
parameters:
*
. S min
: lower table value of the master curve (can be negative5)
*
. S max
: upper table value of the master curve (glass modulus)
. ln (aT f m ) : the x-axis value of the inflection point of the master curve.
Figure 5c shows the loss angle master curves of three different specimens.
Figure 5d shows the average loss angle of the three specimens. Figure 5e
shows the Black-diagram: the loss angle as function of the logarithm of the
complex modulus. A Black-diagram is useful for two reasons:
1 It shows that the linearity of the stress strain behaviour was maintained
during the measurements, i.e. if the curve is continuous. If the curve is
discontinuous, it means that the time temperature superposition principle
was invalidated, probably by nonlinear stress strain behaviour (not
5

For a better fit.

Results of frequency sweep tests Complex modulus

35000

35000

30000

30000

25000

25000

S* (MPa)

S* (MPa)

Sec. 5

20000
15000

20000
15000

10000

10000

5000

5000
0

0
-5

10

-5

15

50
45
40
35
30
25
20
15
10
5
0
0

15

10

15

-5

10

15

ln (aT*f)
y = 0.0011x2 - 0.3392x + 6.3531
7
6
5

ln aT

loss angle (degrees)

10

50
45
40
35
30
25
20
15
10
5
0

ln (aT*f)

50
45
40
35
30
25
20
15
10
5
0
3.50

ln (aT*f)

loss angle (degrees)

loss angle (degrees)

ln (aT*f)

-5

53

4
3
2
1
0

3.75

4.00

log S* (MPa)

4.25

4.50

10

15

temperature (C)

20

Figure 5. Analysis of the complex modulus of an asphalt mixture. Top left (a):
Measured complex modulus of three samples (specimens). Top right (b):
Averaged complex modulus of (a). Mid left (c): Measured loss angle of three
samples (specimens). Mid right (d): Averaged loss angle of (c). Bottom left (e):
Black-curve: averaged loss angle from (d) versus averaged complex modulus
from (b). Bottom right (d): Temperature shift-factor, cf. equation (9).

54

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

Ch. 3

necessarily a material property; e.g. also improper specimen alignment


might be a cause).
2 The curve gives a direct indication at which S * the behaviour is
relatively solid-like (elastic), respectively is relatively fluid-like (viscous);
if the loss angle is smaller than 45, then the stress strain behaviour is
relatively elastic, and if is greater than 45, then the stress strain
behaviour is relatively viscous. From this viewpoint, the Black-diagram
can be used as an aid to determine a suitable creep temperature.
Figure 5f shows the temperature shift factor as function of the temperature.
The temperature shift factor is needed to reconstruct the original measured
data as function of the temperature from the master curve and vice versa.
Figure 5g shows the transformed stiffness master curve according to the
equation
S * = S m* + S m* tanh (c1 ln f / f m )
(10a)
S 0* + S *
(10b)
S =
2
In equation (10a), S * is the complex modulus, S m* is the complex
modulus in the inflection point of the master curve, f is the frequency,
f m is the frequency where S * is equal to S m* , c1 is a constant, and tanh is
the tangent hyperbolicus function,
e x ex
tanh x = x
(10c)
e + e x
S * is equal to S 0* for f 0 , and S * is equal to S * for f . Equation
(10a) can be rewritten as:
y S ( x ) = arctanh [( S * S m* ) / S m* ] = c1 x
(11a)
where
x = ln ( f / f m )
(11b)
According to equation (11) the values of y S ( x ) are on a straight line as a
function of x , with slope c1 , cf. figure 5g. By means of iterative
regression analysis the constant c1 in equation (11) can be determined,
while the frequency, f m , and the constant K are calculated, and the sum
of the squared residues is minimised. Figure 5h compares the measured
master curve and the fitted curve. Figure 5i shows the variation coefficient
of the three specimens of figure 5a as function of the reduced frequency,
and figure 5j shows the percent deviation of the fitted complex modulus
with respect to the measured modulus as a function of x .
The slope of the stiffness master curve on log-log scale, m , can be
obtained by fitting the tangent hyperbolicus function,
*
m

ln S * = ln [ S m* (1 + tanh (c1 x) )]

Differentiating,

(12)

arctanh [(S*-Sm*)/Sm*]

Sec. 5

Results of frequency sweep tests Complex modulus

35000

y = 0.1947x + 2E-16
R2 = 0.9954

1.5

30000

25000

0.5

20000

15000

-0.5

10000

-1

5000

-1.5

-2

-10
-10

-5

10

12

(S*calc. - S*meas.) /
S*meas100%

10
8
6
4
2
0
-5

10

-5

S*=Sm*[1+tanh(c1*x)]

log (f / fm)
variation coefficient of S* (%)

55

S* avg

30
25
20
15
10
5
0
-5
-10
-15
-20
-10

15

10

-5

10

ln (f / fm)

ln (aT.f)

Figure 5 (Contd). Top left (g): Fit of the stiffness master curve to the tangent
hyperbolicus function, y S ( x) = arctanh[( S * S m* ) / S m* ] = c1 x . Top right (h):
Comparison of the measured and the fitted stiffness master curve. Bottom left
(i): variation coefficient of three curves in (a). Bottom right (j): percent
deviation of fitted versus measured average complex modulus.

d ln S *
d
=
ln (1 + tanh (c1 x) ) = c1 (1 tanh (c1 x) )
(13a)
dx
dx
Equation (13a) expresses how the slope of the complex modulus master
curve on log-log scale varies as function of the frequency,
f 0 (or x )
=> m 2c1
(13b)
In the inflection point,
f = f m (or x = 0)
=> m f = c1
(13c)
m =

f (or x )

=> m 0

(13d)

Let us define:

m0 = lim m( f ) = 2 c1
f 0

(13e)

where m0 is the limiting maximum value of m for f 0 . In tables 36,


m0 is given together with m -values for testing conditions used regularly

56

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

Ch. 3

Table 3. Values of K and m of 3 crushed gravel asphalt concrete mixtures, CGAC,


3 porous asphalt mixtures, PA, and one Guss-asphalt mixture, GA, cf. table 2, D.
m 40o C ,0.5 Hz m 40o C ,1 Hz m50o C ,0.5 Hz m50o C ,1 Hz
K
m0
(K)
0.302
0.303
0.295
CGAC 0/22 M1 26448 0.308
0.298
0.325
0.326
0.320
CGAC 0/22 M2 25724 0.330
0.322
0.336
0.337
0.330
CGAC 0/22 M3 25731 0.339
0.332
0.309
0.309
22468 0.312
0.305
PA 4/11
0.305
0.241
0.242
27496 0.246
0.238
PA 11/16 MM
0.238
0.340
0.340
30935 0.270
0.339
PA 11/16 MR
0.339
0.247
0.248
27977 0.250
0.244
GA 2/8 M
0.244
Table 4. Values of K and m of 5 crushed gravel asphalt concrete mixtures, CGAC,
with respectively 0%, 20%, 40%, 60%, and 80% reclaimed asphalt R.A.),
cf. table 2, section D.
m 40o C ,0.5 Hz m 40o C ,1 Hz m50o C ,0.5 Hz m50o C ,1 Hz
K
m0
(K)
0.498
0.499
0.491
CGAC 0% R.A.
22806 0.502
0.494
0.428
0.430
0.419
CGAC 20% R.A. 23047 0.433
0.423
0.392
0.394
0.384
0.387
CGAC 40% R.A. 24065 0.398
0.388
0.389
0.380
0.383
CGAC 60% R.A. 23197 0.394
0.361
0.362
0.351
0.355
CGAC 80% R.A. 23966 0.368
Table 5. m50o C ,1 Hz -values of 25 crushed gravel asphalt concrete
mixtures, CGAC, cf. table 2, section D.
Project
MD MD
MD +
0.317
0.334
0.321
A
0.299
0.270
0.294
B
0.439
0.380
0.400
C
0.334
0.331
0.335
D
0.363
0.362
0.342
E

PLA
0.360
0.350
0.420
0.370
0.400

PAV
0.387
0.371
0.481
0.371
0.414

Table 6. m -values of 25 crushed gravel asphalt concrete mixtures, CGAC,


cf. table 2, section D.
CGAC
mean
stdev
min
max
m0
0.365
0.047
0.276
0.482
m40o C ,0.5 Hz
0.359
0.048
0.266
0.479
m 40o C ,1 Hz
0.357
0.049
0.264
0.478
m50o C ,0.5 Hz
0.362
0.048
0.271
0.481
m50o C ,1 Hz
0.362
0.048
0.270
0.481

Sec. 5

Results of frequency sweep tests Complex modulus

57

lin. viscoelastic creep susceptibility

0.6
0.5
MDMD
MD+
PLA
PAV

0.4
0.3
0.2
0.1
0
A

C
project

Figure 6. Linear viscoelastic creep susceptibility, m50o C ,1 Hz , per project (road


trial) A..E, and per mixture MD-, MD, MD+, PLA, and PAV, cf. table 2,
section D. Each bar represents the average of three tests. MD-: mixture design
mixture minus 0,5% bitumen, MD: mixture design mixture, MD+: mixture
design mixture plus 0,5% bitumen, PLA: plant-mixed mixture, slab-compacted
at the site, PAV: plant-mixed mixture, compacted in the pavement.

in the (triaxial) creep test. The following combinations were applied:


[40C, 0.5 Hz], [40C, 1 Hz], [50C, 0.5 Hz], and [50C, 1 Hz]. Each data
point represents the average of three tests (specimens). The following
observations can be made with tables 36:
Asphalt concrete: m0 varies roughly between 0.25 and 0.50. The
representative mean is approximately 0.36.
Asphalt concrete with reclaimed asphalt: m0 decreases with increasing
percentage of reclaimed asphalt. This is ascribed to the reclaimed asphalt
being polymer modified. It is known that polymer modification of bitumen
reduces the slope of complex modulus master curve, i.e. reduces the
thermal susceptibility (Molenaar 1991, Hagos 2002).
Porous asphalt, unmodified (PA 4/11): m0 corresponds closely to that for
asphalt concrete, CGAC M1.
Porous asphalt, polymer modified: the reduced value of m0 with respect
to the value for unmodified porous asphalt can be attributed to the
polymer.
Guss-asphalt, polymer modified: m0 corresponds to that for polymer
modified porous asphalt concrete, PA 11/16 MM.
From figure 6 and table 5, it is seen that there is no clear dependence of
the m -values on the bitumen content, i.e. not within the tolerance of
0.5% in mixture design. It is seen further that the m -value is practically
constant within this tolerance. Figure 6 and table 5 show further that the

58

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

Ch. 3

m -values of the PLA mixtures and the PAV mixtures may differ from the
MD mixture. The circumstances are too complex to ascribe the differences
to a particular cause. They may be related to the mixture composition (the
fraction reclaimed asphalt), or production and paving conditions.
Additional statistical data for the 25 mixtures are given in table 6. It can be
seen in tables 36, that m has practically attained its maximum value, m0 ,
for the combinations [40C, 0.5 Hz], [40C, 1 Hz], [50C, 0.5 Hz], and
[50C, 1 Hz]. This is discussed further in 10.6.

6 Results of static creep tests in the absence of confinement


Dependence of the creep of asphalt concrete on time,
temperature, stress, and mixture composition
The purpose of the investigation described in 6, 7, and 8 was to gain
insight in the dependence of the creep of asphalt concrete on time, stress,
temperature, and mixture composition. In this section, the results of static
creep tests in the absence of confinement are discussed. The time
dependence is shown at constant temperature and stress. Measurements
were performed at two temperatures, 40C and 50C. Different stresses
were applied to investigate the stress dependence. The stress dependence
is shown at constant time, or frequency, and temperature. The tests were
performed on three asphalt concrete mixtures; DAC 0/16, unmodified,
DAC 0/16 E, elastomer modified, and DAC 0/16 P, plastomer modified,
cf. table 2, sections F..H.
To permit comparison of the investigations in 6, 7, and 8, the results are
presented in terms of the creep compliance, J (t ) . The creep compliance is
obtained by dividing the creep strain by a stress invariant,
J (t ) = (t ) /[ ]
(14)
where (t ) is the total permanent creep strain, and [ ] is a suitable stress
invariant. In principle, any stress invariant can be used. The three stress
invariants used in this study are the deviatoric stress, q , the volumetric
stress, p , cf. equation (2), and the quotient q / p . The preferred stress
invariant was p rather than q or the quotient q / p , to avoid division by
zero ( q = 0 if 1 = 3 ), cf. equation (1). Thus, [ ] = p . The following
creep models were used,
J (t ) = J 1 + ln t z
(15)
~
~
J (t ) = J 0 + J 1 t z
(16)
J (t ) = J e (ln t ) z
(17)
( z(
J (t ) = J 0 + J 1 t + C (et 1)
(18)
where J 0 represents the instantaneous elastic creep
( compliance at t = 0 s,
~
J 1 in equation (15), J 1 in equation (16), and J 1 in equation (18) are the

Sec. 6

Results of static creep tests in the absence of confinement

59

creep compliance at t = 1 s 6. J e in equation (17) is the creep compliance


at e s, where e is the base of the natural logarithm, e = 2.71828. C is a
constant. In equation (15), z is the slope of J (t ) as a function of ln t in a
z is the slope of J ( t ) as function of t in a
semi-ln plot. In equation (16), ~
log-log plot, or equivalently, the slope of ln J (t ) as function of ln t
plotted in a linear plot. In equation (17), z is the slope of J (t ) as function
of ln t in a ln-ln plot, or equivalently, the slope of ln J (t ) as function of ln
ln t in a linear plot. The creep according to equation (18) is not discussed
in detail.
The values of the constants of equation (15) and equation (17) depend on
the logarithm that is used. The notation in terms of the natural logarithm is
preferred, because the physical interpretation of the exponents of the
different creep models, z , ~
z , and z , in relation to m of equation (1) is
considered. The notation in terms of the natural logarithm is also preferred
in mathematical operations that involve the constants of the equations.
Substituting
ln x = 2.3026 log x
(19)
equations (15) and (17) are rewritten, respectively, as
J (t ) = J 1 + 2.3026 log t z = J 1 + z log t
(20)
J (t ) = J e ( 2.3026 log t ) z = J 10 (log t ) z
(21)
J 1 in equation (20) represents the creep compliance at 1 s. In a semi-log
plot of J (t ) as function of log t , the slope of J (t ) as function of log t is
equal to 2.3026 z . Let us define z as
z = 2.3026 z
(22)
J 10 in equation (21) represents the creep compliance at 10 s,
J 10 = J e ( 2.3026) z

(23)

In equation (21), z is the slope of J (t ) as function of log t in a log-log


plot, or the slope of log J (t ) as function of log log t in a linear plot. A list
of creep models for the various materials and testing conditions is given in
appendix 3.

6.1 Dense graded asphalt concrete 0/16, unmodified


Figure 7a shows the time dependence in nine creep tests (nine specimens)
of the same DAC 0/16 mixture, at constant applied stress, 0.2 MPa, and
constant temperature, 40C. The creep models fitted to the curves shown
in figure 7a are given in appendix 3, table 1. Figure 7a gives an impression of the normal scatter. The average value of the creep compliance is
shown in figure 7b. Also the fitted equation (15) is shown. Figure 7b
shows that the creep is nonlinear as function of the time, since the
6

J 0 is normally negligible. Therefore, J 1 is the creep compliance at t = 1 s.

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

60

DAC 0/16, 40C


0.1

0.1

compliance J(t) (MPa^-1)

compliance, J(t), (MPa^-1)

DAC 0/16, 40C

Ch. 3

0.01

J(t) = 0.0063 ln (t) + 0.0161


R2 = 0.9785

0.01
1

10

100

1000 10000

time (s)

10

100

1000 10000

time (s)

DAC 0/16, 40C


compliance, J(t) (MPa^-1)

0.1

0.01
100

J(t) = 0.033 t0.0843


R2 = 0.9933

1000

10000

time (s)

Figure 7. Static creep of dense graded asphalt concrete, DAC 0/16.


Temperature: 40C. Applied stress: 0.2 MPa. Specimen: 60 x 100 mm height x
diameter. Load control: hydraulic. Top left (a): J (t ) versus t ; both plotted
logarithmically. Results of nine tests are shown to give an impression of the
normal scatter. Top right (b): Equation (15) fitted to the average value of the
nine tests of fig. (a). Bottom (c): The average value of the nine tests of fig. (a);
for t > 100 s the creep is described by the power function, equation (16).

exponent of the fitted equation, z is smaller than 1. The equation permits


a reasonable fit for t > 10 s. J 1 is equal to 0.0161, whereas the measured
creep compliance at 1 s is equal to 0.0105. Thus, J 1 deviates 53.3% from
the measured creep compliance at 1 s. In fact, in the time-interval between
1 s and 10 s, the creep curve is not described by equation (15), but by
equation (16)7. This raises the difficulty, in this case, that J 1 cannot be
used as a constitutive model parameter to characterise the material. If
importance is to be attributed to J 1 as a constitutive model parameter, then
7

This can be based on results presented later, cf. 7.37.4, and the discussion in
7.6 (the coefficient of the creep model cannot be negative).

Sec. 6

Results of static creep tests in the absence of confinement

61

it may be useful to fit equation (16) to the measured creep curve in the
time-interval for t < 10 s. However, in principle, greater importance is to
be attributed to the exponent of equation (15), z , because z describes the
greater part of the creep curve.
For a sufficiently long time, the curve becomes quasi-linear on log-log
scale. Equation (16) might be fitted to this quasi-linear portion of the creep
curve, cf. figure 7c. Figure 7c shows the mean value of the creep
~
compliance, J ( t ) , from figure 7a, for t > 100 s. Note, that J 1 in figure 7c
is not related to the measured creep compliance at 1s. The result in figure
7c shows that the static creep of asphalt mixture is nonlinear as a function
of the time, since 0 < ~
z < 1. The results are discussed further in 6.3.

6.2 Dense graded asphalt concrete 0/16, polymer modified


The same dense graded asphalt concrete mixture used for the investigation
described in 6.1 was manufactured using polymer modified binder instead
of the conventional binder to investigate the influence of polymer
modification of asphalt mixture. Two binders were used: an elastomer
modified binder and a plastomer modified binder. The elastomer modified
binder contained an elastomer of the type styrene-butadiene-styrene blockcopolymer (SBS), approximately 5% by mass, in a 200 pen base bitumen.
The plastomer modified binder contained a plastomer of the type ethylenevinylacetate block-copolymer (EVA), approximately 5% by mass, in a
70/100 base bitumen.
Figure 8 shows the time dependence in different creep tests (specimens) of
the similar DAC 0/16 mixture, for different combinations of temperature
and applied axial stress. Figure 8 shows that the creep is non-linear as a
function of the time, since the slope of the creep on log-log scale is smaller
than 1, and also gives an impression of the normal scatter. Figures 8b/d
show the same results but averaged per combination of temperature and
applied axial stress. Equation (15) yields a reasonable fit for t > 10 s. An
example is shown in figure 8d, where equation (15) is fitted to the upper
creep curve, to give a value of z equal to 0.0137 in a semi-ln-plot, or
0.01372.3026 = 0.0315 in a semi-10log-plot. The creep models fitted to
the averaged creep curves, averaged per testing condition shown in figures
8b and 8d, are given in appendix 3, table 2. Figure 9 shows the averaged
coefficient and the averaged exponent of equation (15). The results show
that the static creep of polymer modified asphalt is nonlinear as a function
of the time, with 0 < z < 1. In table 7, J 1 and z are listed for the various
testing conditions. Figure 9 and table 7 also show the temperature and
stress dependence of the constitutive model parameters J 1 and z . As for
the temperature dependence, figure 9 and table 7 show that J 1 increases as
the temperature increases, while z decreases.
As for the stress dependence, table 7 shows that z decreases as the applied

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

62

Ch. 3

DAC 0/16, elastomer modified, 40C and 50C

compliance, J(t), MPa^-1

1
40C/0.1 MPa

40C/0.1 MPa

40C/0.1 MPa

40C/0.2 MPa

40C/0.2 MPa

40C/0.2 MPa

50C/0.1 MPa

50C/0.1 MPa

50C/0.1 MPa

50C/0.2 MPa

50C/0.2 MPa

50C/0.2 MPa

0.1

0.01
1

10

100

1000

10000

time (s)

DAC 0/16, elastomer modif. 40C and 50C


40C/0.1 MPa
40C/0.2 MPa
50C/0.1 MPa
50C/0.2 MPa

compliance J(t) (MPa^-1)

0.1

0.01
1

10

100
time (s)

1000

10000

Figure 8. Static creep of dense graded asphalt concrete, DAC 0/16, polymer
modified. J (t ) versus the time, t ; both plotted logarithmically. Temperatures:
40C and 50C. Applied stress: 0.1 MPa and 0.2 MPa. Specimen: 60 x 100 mm
height x diameter. Load control: hydraulic. Top (a): DAC 0/16, elastomer
modified. Each curve represents a single test (specimen). Per combination of
temperature and applied stress, three tests are shown to give an impression of
the normal scatter. Bottom (b). As figure (a). Each curve represents the average
of three tests of fig. 8a.

Sec. 6

Results of static creep tests in the absence of confinement

63

DAC 0/16, plastomer modified, 40C and 50C

compliance, J(t), MPa^-1

40C/0.1 MPa

40C/0.1 MPa

40C/0.1 MPa

40C/0.2 MPa

40C/0.2 MPa

40C/0.2 MPa

50C/0.1 MPa

50C/0.1 MPa

50C/0.1 MPa

50C/0.2 MPa

50C/0.2 MPa

50C/0.2 MPa

0.1

0.01
1

10

100

1000

10000

time (s)

DAC 0/16, plastomer modified 40C and 50C


compliance J(t) (MPa^-1)

40C/0.1 MPa
40C/0.2 MPa
50C/0.1 MPa
50C/0.2 MPa

J(t) = 0.0137 Ln (t) + 0.0398


R2 = 0.9575
0.1

0.01
1

10

100

1000

10000

time (s)

Figure 8 (Contd). Top (c): As figure 8a, DAC 0/16, plastomer modified.
Bottom (d): As figure 8b, DAC 0/16, plastomer modified.

stress is increased. Unfortunately, the stress dependence of J 1 cannot be


interpreted, owing to the probability that the value of J 1 for 40C/0.1 MPa
in table 7 is an outlier.

6.3 Comparison of static creep of DAC 0/16, unmodified and DAC


0/16, polymer modified
Let us compare the static creep of DAC 0/16, cf. 6.1, and DAC 0/16
polymer modified, cf. 6.2. The creep model for DAC 0/16, unmodified in

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

constant of eq. (15)

500

400
300
200
100

500

constant of eq. (15)

64

0.1 MPa

400

0.2 MPa

300
200
100
0

0
30

40

50

30

60

150
100
50

0.1 MPa
0
30

40

50

temperature (C)

60

exponent of eq. (15), z.10^4

200

40

50

60

temperature (C)

temperature (C)

exponent of eq. (15), z.10^4

Ch. 3

200

0.2 MPa

150
100
50
0
30

40

50

60

temperature (C)

Figure 9. Static creep DAC 0/16, elastomer modified (E) and plastomer
modified (P). Coefficient and exponent of equation (15). Top left (a): Coefficient
of equation (15), applied stress: 0.1 MPa. Top right (b) Coefficient of equation
(15), applied stress: 0.2 MPa. Bottom left (c): Exponent of equation (15), z 104,
applied stress: 0.1 MPa. Bottom right (d): Exponent of equation (15), z 104,
applied stress 0.2 MPa.
Table 7. DAC 0/16, polymer modified. Static creep. Constitutive
model parameters J 1 , and z of eq. (15), and z of eq. (20), if J (t )
is in MPa-1.
J 1 104
z
z
DAC 0/16, elastomer modified
0.0181
0.0079
213
40C/0.1 MPa
0.0138
0.0060
382
50C/0.1 MPa
0.0097
0.0042
209
40C/0.2 MPa
0.0092
0.0040
311
50C/0.2 MPa
DAC 0/16, plastomer modified
0.0378
0.0164
9
40C/0.1 MPa
0.0315
0.0137
398
50C/0.1 MPa
0.0214
0.0093
191
40C/0.2 MPa
0.0175
0.0076
293
50C/0.2 MPa

Sec. 6

Results of static creep tests in the absence of confinement

65

figure 7b,
DAC 0/16 unmodified, 40C/0.2 MPa:

J (t ) = 0.0161 + 0.0063 ln t
J (3,600) = 0.068 MPa-1
compares to the following two models for DAC 0/16, elastomer modified
(E) and DAC 0/16, plastomer modified (P), cf. table 7, respectively,
DAC 0/16 E, 40C/0.2 MPa:
J (t ) = 0.0209 + 0.0042 ln t ;
J (3,600) = 0.055 MPa-1
DAC 0/16 P, 40C/0.2 MPa:
J (t ) = 0.0191 + 0.0093 ln t ;
J (3,600) = 0.095 MPa-1
where J (3,600) is the final creep compliance at 3,600 s. Note, the ranking
of J (3,600) . J (3,600) is lowest for the elastomer modified mixture, and
is lower for the unmodified mixture than for the plastomer modified
mixture. This ranking is not in agreement with the resistance to permanent
deformation in the pavement, where polymer modified asphalt mixtures
show normally improved resistance to permanent deformation. It was also
unexpected that the final creep compliance, J (3,600) , of the plastomer
modified mixture was higher than that of the elastomer modified mixture.
This was expected, because the plastomer modified binder contained a
harder base bitumen (penetration 70/100) than that of the elastomer
modified bitumen (penetration 200), and on the experience that asphalt
mixtures containing the particular plastomer modified bitumen normally
show a greater resistance to permanent deformation in empirical rutting
tests. Valkering et al. (1990) reported similar unexpected results in the
static creep test. They argued that the dynamic creep test is to be preferred
rather than the static creep test. The matter is discussed further in 7.6.

7 Results of dynamic creep tests in the absence of confinement


Dependence of the creep of asphalt concrete on time,
temperature, stress, and mixture composition
The purpose of the investigation described in 6, 7, and 8 was to gain
insight in the dependence of the creep of asphalt concrete on time, stress,
temperature, and mixture composition (Molenaar et al. 1995). In this
section, the results of dynamic creep tests in the absence of confinement
are discussed. Different waveforms of the applied stress were used to
investigate the time dependence. The time dependence is shown at
constant applied stress, in case of the static creep test, or at constant
applied stress amplitude, in case of the dynamic creep test. Measurements
were performed at two temperatures, 40C and 50C. Different stress
amplitudes were applied to investigate the stress dependence. The stress
dependence is shown at constant time, or frequency, and temperature. The
tests were performed on three asphalt concrete mixtures; DAC 0/16,
unmodified, DAC 0/16 E, elastomer modified, and DAC 0/16 P, plastomer
modified, cf. table 2, sections F..H.

66

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

Ch. 3

600 s
1

load signal

1: 600 s/1.8 s
2: 10 s/1.8 s
3: 1 s/1.8 s

4: 0.2 s/1.8 s
5: 0.05 s/1.8 s

10

20

30600

40

time (s)

Figure 10. Different waveforms used. Loading times, t l : 600 s,


10 s, 1 s, 0.2 s, and 0.05 s. Rest-time, t r : constant 1.8 s.

7.1 Dense graded asphalt concrete 0/16, unmodified, using a blockwaveform of applied stress
7.1.1 Time dependence of creep properties - constant rest-time
The departure from static loading, or the transient to normal dynamic
loading, was investigated by introducing a rest-time, t r , of 1.8 s after the
following loading times, t l : 600 s, 10 s, 1 s, 0.2 s, cf. figure 10. This was
done with two applied stress amplitudes, 0.1 MPa and 0.2 MPa, and two
temperatures, 40C and 50C. In figure 11, the creep compliance is shown
as a function of the total time, at 40C. In figure 12, the creep compliance
is shown as a function of the total time, at 50C. Each curve in figures 1112 represents the average of three tests. The corresponding creep models
are listed in appendix 3, table 3.
If the results in figures 11 - 12 are compared, then it seems that the creep
is enhanced as the rest-time pulse of 1.8 s is repeated faster. That is: the
figures show that the creep curve shifts to the left, i.e. to lower time value,
if the loading time, t l , decreases from 600 s to 0.2 s, at constant rest-time
t r . The figures show further that the creep curve shifts back to the right,
i.e. to higher time value, if t l decreases further from 0.2 s to 0.05 s. This is
a form of time dependence of creep, where the creep depends on the shape
of the waveform of the applied stress. This is discussed later in this section
in more detail, cf. figure 18.
Let us, as an example, consider the creep curve 1 s/1.8 s in figure 11a.

Sec. 7 Results of dynamic creep tests in the absence of confinement

67

Equation (20) cannot be fitted to this curve, cf. figure 13a. Also equation
(16) cannot be fitted, since this represents a linear log-log relationship,
whereas the creep curve in figure 13a clearly is not. However, equation
(21) can be fitted to J ( t ) versus log t in a log-log plot, cf. figure 13b.
Note, that the x -axis of figure 13a represents the time, t , plotted
logarithmically, and that the x -axis of figure 13b represents log t plotted
logarithmically (i.e. log log t on normal scale). The advantage of the latter
representation is that a single creep model can fit the whole creep curve.
Otherwise, equation (15) or equation (16) could be fitted only to parts of
the creep curve to be selected arbitrarily. Differentiating equation (17),
d J (t )
= J e z (ln t ) z 1
(24)
d ln t
If we define
(t ) = J e z (ln t ) z 1
(25)
then (t ) is the time dependent analogue of z in equation (15),
d J (t )
= z
(26)
d ln t
Note that z is constant. Substitution of equation (19) and equation (23)
into equation (25) yields
J z
(t ) = 10 (log t ) z 1
(27)
2.3026
(t ) is the time dependent analogue of z in equation (20). Taking the
logarithm of equation (17),
ln J (t ) = ln J e + ln (ln t ) z = ln J e + z ln ln t
(28)
Substituting x = ln t ,
ln J (t ) = ln J e + ln x z = ln J e + z ln x
(29)
Differentiating,
d ln J (t )
d ln J (t )
z
z
=
=
=
(30)
d ln t
dx
x
ln t
If we define
z
~ (t ) =
(31)
ln t
then ~ (t ) is the time dependent analogue of ~z in equation (16),
d ln J (t ) ~
= z
(32)
d ln t
Note that ~z is constant. Substitution of equation (19) and equation (23)
into equation (30) yields
z
~ (t ) =
(33)
2.3026 log t

68

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

compliance J(t) (MPa^-1)

Ch. 3

DAC 0/16, 40C


0.1 MPa

0.1
600 s/1.8 s
10 s/1.8 s
1 s/1.8 s
0.2 s/1.8 s
0.05 s/1.8 s

0.01

0.001
1

10

100

1000

10000

time (s)

compliance J(t) (MPa^-1)

DAC 0/16, 40C


0.2 MPa

0.1

600 s/1.8 s
10 s/1.8 s
1 s/1.8 s
0.2 s/1.8 s
0.05 s/1.8 s

0.01

0.001
1

10

100

1000

10000

time (s)

Figure 11. Dynamic creep of dense graded asphalt concrete, DAC 0/16. Whole
creep curves J (t ) versus t ; both plotted logarithmically. Each curve represents
the average of three tests as a function of the total time, i.e. loading time plus
rest-time. The waveform of the applied load was a block-wave defined by a
loading time and a rest-time, cf. fig. 10. Different loading times: 600 s, 10 s, 1 s,
0.2 s, 0.05 s. Constant rest-time: 1.8 s. Temperature: 40C. Specimen: 60 x 100
mm height x diameter. Load control: hydraulic. Top (a): Applied stress
amplitude: 0.1 MPa. Bottom (b): Applied stress amplitude: 0.2 MPa.
Example. In figure 13b, the creep is described by equation (21). Let us compare
this creep to the creep described by equation (15), an example of which is shown
in figure 7b. J 10 is equal to 0.0028, and z is equal to 2.9543. By equation (27),
J z
0.0028 2.9543
(10) = 10
=
0.0036
(34a)
2.3026
2.3026

compliance J(t) (MPa^-1)

Sec. 7 Results of dynamic creep tests in the absence of confinement

DAC 0/16, 50C


0.1 MPa

0.1

600 s/1.8 s
10 s/1.8 s
1 s/1.8 s
0.2 s/1.8 s
0.05 s/1.8 s

0.01

69

0.001
1

10

100

1000

10000

time (s)

compliance J(t) (MPa^-1)

DAC 0/16, 50C


0.2 MPa

0.1
600 s/1.8 s
10 s/1.8 s
1 s/1.8 s
0.2 s/1.8 s
0.05 s/1.8 s

0.01

0.001
1

10

100

1000

10000

time (s)

Figure 12. As figure 11, 50C.


Note the order of magnitude, which is similar to that of the example in figure 7b.
For t = 100 s,
J z
0.0028 2.9543
(10) = 10 (log t ) z 1 =
21.9543 0.0139
(34b)
2.3026
2.3026
The time dependent analogue of z in equation (27), (t ) , increases with
increasing log t . Hence, the creep in figure 13b according to equation (21) runs
faster than the creep described by equation (20), according to which the slope of
J (t ) versus log t is constant as function of log t . This can also be seen by
comparison of the creep on the same time scale; i.e. if figure 13a and figure 7b
are compared. Let us now compare the creep in figure 13b, described by
equation (21), and the creep described by equation (16). By equation (33),

70

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

compliance, J(t) (MPa^-1)

Ch. 3

1 s/1,8 s measured

0.1

0.01

0.001
1

10

100

1000

10000

compliance, J(t) (MPa^-1)

time (s)

1 s/1,8 s measured

0.1
J(t) = 0.0028 t2.9543
R2 = 0.9941
0.01

0.001
0.1

10

log time (s)

Figure 13. Top (a): J (t ) versus t ; both plotted logarithmically. Equation (15)
cannot be fitted to the creep curve 1 s/1.8 s of fig. 11a. Bottom (b): The creep
compliance of figure (a). J (t ) versus log t ; both plotted logarithmically
(equivalent to log J ( t ) versus log log t on linear scales). Equation (21) fitted to
the creep of figure 13a.
z
z
= 1.28,
= 0.64,
and
~ (100) =
2.3026
2.3026 2
z
= 0.32
(34c)
~ (10,000) =
2.3026 4
The time dependent analogue of ~z in equation (33), ~ (t ) , slope of log J (t )
versus log log t , decreases with increasing log t . Note the order of magnitude of
~ (t ) , which is greater than that of ~z in figure 7c. Hence, the creep in figure 13b

~ (10) =

is relatively fast as compared to the creep in figure 7c.

It is important that the creep can be characterised by a constitutive model,


i.e. by constitutive model parameters. For a representative and reliable
characterisation, whole creep curves must be fitted, rather than arbitrary
selected parts. In addition, it is important that model parameters reflect

Sec. 7 Results of dynamic creep tests in the absence of confinement

71

measured data. Thus, the creep compliance at 1 s, J 1 , should reflect the


measured creep compliance at 1 s, or if applicable, J 10 should reflect the
measured creep compliance at 10 s. If not, J 1 and J 10 are not useful as
material property.
In figures 1415, J (t ) , from figures 1112, is plotted versus log t ; both
plotted logarithmically. In figures 1415 the creep curves seem to form a
wedge with the sharp end on the right-hand side. It can be shown that
this wedge is caused by the log-transformation of the abscissa, from log
(time) to log (log (time)), if the untransformed creep curves are parallel
curves, cf. 7.7. It seems that creep curves of a given asphalt mixture do
cross sometimes, cf. figures 7, 8, and 1112, and similar results presented
later, presumably owing to bad specimens, but as a rule do not cross.
The results seem to indicate, that, if the material quality is constant, the
creep curves do not cross when the waveform of the applied load changes,
but shift parallel (horizontally along the time axis). It is shown in 7.7, that
in that case the wedge of creep curves forms upon log-transformation, i.e.
by plotting the data as in figures 1415. Conversely, from the fact that a
wedge is obtained in a plot of log J (t ) versus log log t , it follows that the
untransformed creep curves are (read: should be) parallel creep curves.
From the fact that a wedge is obtained it follows further that J 10 and z are
dependent; i.e. if J 10 increases, then z decreases, and vice versa. This is
also discussed further in 7.7.
Figure 16 shows J 10 of equation (21) from table 8 as function of the
loading time, tl . Figure 16 shows that J 10 increases with increasing tl
between tl = 0.05 s and tl = 0.2 s, and decreases with increasing tl
between tl = 0.2 s and tl = 600 s. This behaviour is found with all of the
four applied testing conditions, [40C, 0.1 MPa], [40C, 0.2 MPa], [50C,
0.1 MPa], and [50C, 0.2 MPa].
Figure 17 shows z of equation (21) from table 8 as a function of the
loading-time, tl . Figure 17 shows that z decreases with increasing tl
between tl = 0.05 s and tl = 0.2 s, and increases with increasing tl
between tl = 0.2 s and tl = 600 s. This behaviour was found with three of
the four applied testing conditions, [40C, 0.2 MPa], [50C, 0.1 MPa], and
[50C, 0.2 MPa]. The result for the condition [40C, 0.1 MPa] deviates
from the results for the other conditions.
In table 8, J 10 , z , (t ) and ~ (t ) are listed for the various testing
conditions of the curves shown in figures 1415. In all cases in table 8, the
applied creep model is equation (21). z of equation (21) can be compared
to z of equation (20), or ~z of equation (16). The relationships between
these parameters are defined by means of the parameters (t ) and ~ (t ) .
(t ) is the time dependent analogue of z in equation (20) or z in
equation (15). ~ (t ) is the time dependent analogue of ~z . Later, cf. 13,

72

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

DAC 0/16, 40C


0.1 MPa

compliance J(t) (MPa^-1)

Ch. 3

0.1
600 s/1.8 s
10 s/1.8 s
1 s/1.8 s
0.2 s/1.8 s
0.05 s/1.8 s

0.01

0.001
0.1

10

log time (s)


y = 0.0027x2.7115
R2 = 0.9954; 10 s/1.8 s

y = 0.0029x2.2579
2
R = 0.9985; 600 s/1.8 s

y = 0.0115x2.3369
R = 0.9979; 0.2 s/1.8 s

y = 0.0028x2.9543
R = 0.9941; 1 s/1.8 s
2

y = 0.0024x3.21
R2 = 0.9949; 0.05 s/1.8 s

compliance J(t) (MPa^-1)

DAC 0/16, 40C


0.2 MPa

0.1
600 s/1.8 s
10 s/1.8 s
1 s/1.8 s
0.2 s/1.8 s
0.05 s/1.8 s

0.01

0.001
0.1

10

log time (s)


y = 2E-05x5.0293
R = 0.9997; 600 s/1.8 s
2

y = 0.0013x3.2306
R = 0.9985; 10 s/1.8 s
2

y = 0.0066x2.294
R2 = 0.9917

y = 0.0054x2.5592
R = 0.9929; 1 s/1.8 s
2

y = 0.0032x2.6239
R = 0.9985; 0.05 s/1.8 s
2

Figure 14. Dynamic creep of dense graded asphalt concrete, DAC 0/16; data of
fig. 11. J (t ) versus log t ; both plotted logarithmically. In the regression
equations, y represents J ( t ) , and x represents log t , cf. equation (21). Top (a):
Curves of fig. 11a. Bottom (b): Curves of fig. 11b.

Sec. 7 Results of dynamic creep tests in the absence of confinement

compliance J(t) (MPa^-1)

73

DAC 0/16, 50C


0.1 MPa

0.1
600 s/1.8 s
10 s/1.8 s
1 s/1.8 s
0.2 s/1.8 s
0.05 s/1.8 s

0.01

0.001
0.1

10

log time (s)


y = 0.0002x4.0136
R2 = 0.9991; 600 s/1.8 s

y = 0.005x2.9456
2
R = 0.997; 10 s/1.8 s

y = 0.0226x2.0893
R = 0.9875; 0.2 s/1.8 s

y = 0.0149x2.3447
R2 = 0.9911; 1 s/1.8 s

y = 0.0047x3.0735
R = 0.9994; 0.05 s/1.8 s

compliance J(t) (MPa^-1)

DAC 0/16, 50C


0.2 MPa

0.1
600 s/1.8 s
10 s/1.8 s
1 s/1.8 s
0.2 s/1.8 s
0.05 s/1.8 s

0.01

0.001
0.1

10

log time (s)


y = 4E-05x5.3661
R = 0.9999; 600 s/1.8 s
2

y = 0.0017x3.4764
R = 0.9987; 10 s/1.8 s
2

y = 0.0126x2.3001
R = 0.9897; 0.2 s/1.8 s
2

y = 0.0109x2.0719
R = 0.9999; 1 s/1.8 s
2

y = 0.0032x2.552
R = 0.9996; 0.05 s/1.8 s
2

Figure 15. Dynamic creep of dense graded asphalt concrete, DAC 0/16; data of
fig. 12. J (t ) versus log t ; both plotted logarithmically. In the regression
equations, y represents J ( t ) , and x represents log t , cf. equation (21). Top (a):
Curves of fig. 12a. Bottom (b): Curves of fig. 12b.

74

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

Ch. 3

Table 8. DAC 0/16. Dynamic creep. Block-wave load with constant rest-time,
1.8 s1. Constitutive model parameters J 10 , and z of eq. (21), if J (t ) is in
MPa-1; (10) of eq. (27), and ~ (10) of eq. (33). (10) = J 10 z /2.3026;
~ (10) = z 2.3026.
J 10 10 4
z
(10)
~ (10)
Loading time/rest-time
40C/0.1 MPa
600 s/1.8 s
29
2.2579
0.0028
0.98
10 s/1.8 s
27
2.7115
0.0032
1.18
1 s/1.8 s
28
2.9543
0.0036
1.28
0.2 s/1.8 s
115
2.3369
0.0117
1.01
0.05 s/1.8 s
24
3.2100
0.0033
1.39
40C/0.2 MPa
600 s/1.8 s
0.2
5.0293
0.00004
2.18
10 s/1.8 s
13
3.2306
0.0018
1.40
1 s/1.8 s
54
2.5592
0.0060
1.11
0.2 s/1.8 s
66
2.2940
0.0066
1.00
0.05 s/1.8 s
32
2.6239
0.0036
1.13
50C/0.1 MPa
600 s/1.8 s
2
4.0136
0.0003
1.74
10 s/1.8 s
50
2.9456
0.0064
1.28
1 s/1.8 s
149
2.3447
0.0152
1.02
0.2 s/1.8 s
226
2.0893
0.0205
0.91
0.05 s/1.8 s
47
3.0735
0.0063
1.33
50C/0.2 MPa
600 s/1.8 s
0.4
5.3661
0.00009
2.33
10 s/1.8 s
17
3.4764
0.0026
1.51
1 s/1.8 s
109
2.0719
0.0098
0.90
0.2 s/1.8 s
126
2.3001
0.0126
1.00
0.05 s/1.8 s
32
2.5520
0.0036
1.11
1
~
(t ) is the time dependent analogue of z in eq. (20). (t ) is the time
dependent analogue of ~
z in eq. (16). These parameters are used later to be
able to compare different creep behaviours of different asphalt mixtures.

(t ) and ~ (t ) are used to attribute a physical meaning to z and z . The


time dependence of J 10 and z is illustrated schematically in figure 18.
Figure 18 shows, that, if the creep curve shifts to the left, i.e. to lower time
value, then J 10 increases, and z decreases. Conversely, if the creep curve
shifts back to the right, i.e. to higher time value, then J 10 decreases, and z
increases. Thus, the time dependence of J 10 and z is shown. To explain
the time dependence, one must explain the shift of the creep curve along
the time axis. Different time dependencies were found in this study,
depending on the shape of the waveform. In the present case, the creep
curve shifts to the left, if the loading time decreases from 600 s to 0.2 s at
constant rest-time, and the curve shifts back to the right if the loading time
decreases further from 0.2 s to 0.05 s at constant rest-time. This behaviour

Sec. 7 Results of dynamic creep tests in the absence of confinement

parameter J1010^4

250

75

40C/0.1 MPa
40C/0.2 MPa
50C/0.1 MPa
50C/0.2 MPa

200

150

100

50

0
0.05 s/1.8 s

0.2 s/1.8 s

1 s/1.8 s

10 s/1.8 s

600 s/1.8 s

loading-time/rest-time

Figure 16. Coefficient J 10 of equation (21) as function of the loading time.


6

40C/0.1 MPa

exponent z of eq. (21)

40C/0.2 MPa

5
4

50C/0.1 MPa
50C/0.2 MPa

3
2
1
0
0.05 s/1.8 s

0.2 s/1.8 s

1 s/1.8 s

10 s/1.8 s

600 s/1.8 s

loading-time/rest-time

Figure 17. Exponent z of equation (21) as function of the loading time.

cannot be explained on the basis of the time temperature superposition


principle (if that were valid, which it is not because the behaviour is
nonlinear). Based on that principle, the mixture is expected to respond
with greater stiffness if the loading time decreases at constant rest-time.
Then, the period time of the load signal decreases, and the frequency
increases8. Accordingly, the creep compliance is expected to decrease, i.e.
8

It is assumed that the notion frequency applies to any periodical function.


Hence, for a block-wave, defined by a loading time and a rest-time, the blockwave frequency increases if the loading time decreases at constant rest-time, or
if the rest-time decreases at constant loading time.

76

compliance, J(t) (MPa^-1)

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

Ch. 3

J 10 , z

0.1

t l , t r = constant

t l , t r = constant
0.01
J 10 , z
0.001
1

10

100

1000

10000

100000

time (s)

Figure 18. Schematic illustration of the time dependence of J 10 and z . If the


loading time, t l , decreases from 600 s to 0.2 s at constant rest-time, t r , then the
creep curve shifts to the left, i.e. to lower time value; J 10 increases and z
decreases. If t l decreases further from 0.2 s to 0.05 s at constant t r , then the
creep curve shifts back to the right, i.e. to higher time value; J 10 increases and
z decreases.

the creep curve is expected to shift to higher time value. Since the result is
obviously not explained by the time temperature superposition principle, it
is likely to be attributable to the response of the grain skeleton in stress
controlled compression. Possibly, the shift to the left as a function of
increasing frequency is to be attributed to the aggressiveness of the
block-wave, which prevents the grain skeleton from becoming
immobilised by the internal friction. Evidently, at some point, the creep
compliance cannot continue to increase as the loading time decreases at
constant rest-time, since there will be insufficient applied load left to keep
the creep going. This probably explains the turning point, where the shift
turns back to higher time value, if the loading time decreases further from
0.2 s to 0.05 s. It is probable that the loading time of the turning point
depends on the material composition, cf. 7.2, figure 30.
7.1.2 Time dependence of creep properties - constant loading-time
The loading time was held constant: 0.2 s, and the rest-time was increased:
0.2 s, 0.5 s, 1 s, 1.8 s, 5 s. The temperature was 40C. The applied stress
amplitude was constant, 0.2 MPa. Therefore, only the time dependence of
the constitutive model parameters can be considered. In figure 19a, J (t ) is
shown as a function of the time, t . Each curve represents an average of
three tests. A list of creep models is given in appendix 3, table 4.
The result in figure 19a is roughly similar to the results in figures 1112.
The creep curve shifts to the right, i.e. to higher time value, if the rest-time

Sec. 7 Results of dynamic creep tests in the absence of confinement

77

increases at constant loading time, i.e. if the frequency decreases. In terms


of block-wave frequency, this trend is similar to that in figures 1112. If
the creep curves are plotted as log J (t ) versus log log t , cf. figure 19b,
then the typical wedge mentioned previously is formed, from which it
follows that the original creep curves in figure 19a are (read: should be)
parallel creep curves. The dependence on the time of J 10 and z$ is
illustrated schematically in figure 20. In terms of block-wave frequency, it
is qualitatively similar to the trend in figure 18.
In figure 19b, J (t ) from figure 19a is shown as a function of log t . The
result in figure 19b is roughly similar to the results in figures 14 - 15. In
figure 21, the coefficient J 10 of equation (21) is given as a function of the
rest-time, t r . In figure 22, the exponent z$ of equation (21) is given as a
function of the rest-time, t r . Figure 21 shows J 10 to decrease as a function
of increasing rest-time. Figure 22 shows z$ to increase as a function of
increasing rest-time. The results in figure 21 and figure 16 are
complentary. Similarly, the results in figure 22 and figure 17 are
complementary. Figures 1617 and figures 2122 show that the
dependence of the creep on the loading time and the rest-time is complex.
Table 9. DAC 0/16. Dynamic creep. Block-wave load with constant loading
time, 0.2 s. Constitutive model parameters J 10 , and z of eq. (21), if J (t ) is
in MPa-1; (10) of eq. (27), and ~ (10) of eq. (33). (10) = J 10 z /2.3026;
~ (10) = z 2.3026.
J 10 10 4
z
(10)
~ (10)
loading time/rest-time
40C/0.2 MPa
3.03
0.0150
1.3171
262
0.2 s/0.2 s
3.40
0.0123
1.4775
191
0.2 s/0.5 s
3.96
0.0129
1.7211
173
0.2 s/1.0 s
5.42
0.0064
2.3530
63
0.2 s/1.8 s
6.06
0.0045
2.6339
39
0.2 s/5.0 s

In table 9, the model parameters, J 10 , z , (t ) and ~ (t ) are listed for the


various testing conditions. Note the dependence of J 10 and z ; if J 10
decreases, then z increases, and vice versa. An explanation for this was
given in the previous section.
7.1.3 Temperature dependence and stress dependence of creep properties
If figures 11a and 12a are compared, it can be seen that the creep curve
shifts to the left, i.e. to lower time value, as the temperature increases,
causing the creep compliance to increase for a given time. A similar shift
can be observed upon comparison of figures 11b and 12b. This shift
corresponds to an increase of J 10 with increasing temperature, which is
confirmed in figure 16.

78

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

compliance J(t) (MPa^-1)

Ch. 3

DAC 0/16, 40C


0.2 MPa

0.1
0.2 s/0.2 s
0.2 s/0.5 s
0.2 s/1.0 s
0.2 s/1.8 s
0.2 s/5.0 s

0.01

0.001
1

10

100

1000

10000

100000

time (s)

compliance J(t) (MPa^-1)

DAC 0/16, 40C, 0.2 MPa

0.1

0.2 s/0.2 s
0.2 s/0.5 s
0.2 s/1.0 s
0.2 s/1.8 s
0.2 s/5.0 s

0.01

0.001
0.1

10

log time (s)


y = 0.0262x1.3171
R = 0.9901; 0.2 s/0.2 s
2

y = 0.0191x1.4775
R = 0.9969; 0.2 s/0.5 s
2

y = 0.0063x2.353
R = 0.9844; 0.2 s/1.8 s
2

y = 0.0173x1.7211
R = 0.9975; 0.2 s/1.0 s
2

y = 0.0039x2.6339
R2 = 0.996; 0.2 s/5.0 s

Figure 19. Dynamic creep of dense graded asphalt concrete, DAC 0/16. Whole
creep curves. The waveform of the applied load was a block-wave defined by a
loading time and a rest-time, cf. fig. 3. Constant loading time: 0.2 s. Different
rest-times: 0.2 s, 0.5 s, 1.0 s, 1.8 s, 5.0 s. Temperature: 40C. Applied stress
amplitude: 0.2 MPa. Specimen: 60 x 100 mm height x diameter. Load control:
hydraulic. Each creep curve represents the average of three tests. Top (a): J (t )
versus the time, t ; both plotted logarithmically. Bottom (b): J (t ) versus log t ;
both plotted logarithmically. In the regression equations, y represents J ( t ) , and
x represents log t , cf. equation (21).

compliance, J(t) (MPa^-1)

Sec. 7 Results of dynamic creep tests in the absence of confinement

79

0.1

t l = constant, t r
J 10 , z

0.01

0.001
1

10

100

1000

10000

100000

time (s)

Figure 20. Schematic illustration of the time dependence of J 10 and z . If the


rest-time, t r , decreases from 0.2 s to 5.0 s at constant loading time, t l , then the
creep curve shifts to the right, i.e. to higher time value; J 10 decreases and z
increases.
3

exponent of eq. (21), z

coefficient of eq. (21),


J1010^4

350

y = 115.85x-0.6231
R2 = 0.9045

300
250
200
150
100
50

2,5
2
1,5
1
0,5
0

0
0

rest-time

Figure 21. Coefficient J 10 of eq. (21)


as function of the rest-time. y (x)
represents J 10 as function of the
rest-time, t r .

rest-time (s)

Figure 22. Exponent z of eq. (21) as


function of the rest-time.

If figures 11a and 11b are compared, it can be seen that the creep curve
shifts to the right, i.e. to higher time value, as the applied stress amplitude
increases, causing the creep compliance to decrease for a given time. A
similar shift can be observed upon comparison of figures 12a and 12b.
This shift corresponds to a decrease of J 10 with increasing applied stress
amplitude. This is confirmed by eight out of ten cases in figure 16; i.e.
except the following two cases:
1 first group from left, block-wave: 0.05 s/1.8 s, the first two bars;
2 third group from left, block-wave: 1 s/1.8 s, the first two bars.

80

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

Ch. 3

Furthermore, as noticed previously, z is expected to decrease as J 10


increases, and vice versa. This dependence between z and J 10 is caused,
as also noted previously, by the parallelism of the creep curves in figures
1112 and the log-transformation of the log (time) data when equation
(21) is used.
Not in all cases the real changes in J 10 and z as a function of the
temperature or the stress are in agreement with expected changes. In those
cases it seems the scatter on the data overshadows the influences of the
temperature respectively the stress.
7.2 Dense graded asphalt concrete 0/16, polymer modified, using a blockwaveform of applied stress
7.2.1 Time dependence of creep properties
The same dense graded asphalt concrete mixture used for the investigation
described in 7.1 was manufactured using polymer modified binder instead
of the conventional binder to investigate the influence of polymer
modification of asphalt mixture. Two binders were used: an elastomer
modified binder and a plastomer modified binder. The elastomer modified
binder contained an elastomer of the type styrene-butadiene-styrene blockcopolymer (SBS), approximately 5% by mass, in a 200 pen base bitumen.
The plastomer modified binder contained a plastomer of the type ethylenevinylacetate block-copolymer (EVA), approximately 5% by mass, in a
70/100 base bitumen.
The investigations described in 7.1 were in part repeated for the polymer
modified mixtures. Different block-waves of applied stress were used:
different loading times, 600 s, 1 s, and 0.2 s, at constant rest-time, 1.8 s. A
list of creep models is given in appendix 3, table 5.
Figure 23 shows J ( t ) versus the time, t , on log-log scales. Each curve
represents an average of three tests. It seems that in figure 23 two creep
domains can be distinguished, one creep domain in a lower time-interval,
in which the creep is of the type of equation (16), and a second creep
domain in a higher time-interval, in which the creep is of the type of
equation (15), or the type of equation (16). If it is attempted to fit equation
(15) to the whole creep curve, then the fitted relation is found to show a
misfit similar to the example in figure 13a. A disadvantage of
distinguishing two creep domains is that two constitutive models are
needed to describe the creep.
Unfortunately, for the curves designated by 1 s/1.8 s and 0.2 s/1.8 s, no
measurements were sampled between the first and the tenth load
repetition. Therefore, these creep curves are not known in detail between
the first and the second data point in figure 23.

Sec. 7 Results of dynamic creep tests in the absence of confinement

DAC 0/16, elastomer,


40C, 0.1 MPa

0.1

0.01

600 s/1.8 s
1 s/1.8 s
0.2 s/1.8 s

0.001

compliance J(t) (MPa^-1)

compliance J(t) (MPa^-1)

DAC 0/16, elastomer,


40C, 0.2 MPa

0.1

600 s/1.8 s

0.01

1 s/1.8 s
0.2 s/1.8 s

0.001
1

10

100

1000 10000

10

100

DAC 0/16, plastomer,


40C, 0.1 MPa

0.1

600 s/1.8 s

0.01

1 s/1.8 s
0.2 s/1.8 s

0.001

compliance J(t) (MPa^-1)

1000 10000

time (s)

time (s)

compliance J(t) (MPa^-1)

81

DAC 0/16, plastomer,


40C, 0.2 MPa

0,1

600 s/1.8 s

0,01

1 s/1.8 s
0.2 s/1.8 s

0,001
1

10

100

1000 10000

time (s)

10

100

1000 10000

time (s)

Figure 23. Dynamic creep of dense graded asphalt concrete, DAC 0/16, polymer
modified. J (t ) versus the time, t , both plotted logarithmically. The waveform of
the applied load was a block-wave, given by a loading time and a rest-time, cf.
fig. 10. Different loading times: 600 s, 1 s, and 0.2 s. Constant rest-time: 1.8 s.
Temperature: 40C. Specimen 60 x 100 mm height x diameter Load control:
hydraulic. Top left (a): DAC 0/16, elastomer modified. Applied stress amplitude:
0.1 MPa. Top right (b) As fig. (a), applied stress amplitude: 0.2 MPa. Bottom
left (c): DAC 0/16, plastomer modified. Applied stress amplitude: 0.1 MPa.
Bottom right (d): As fig. (c), applied stress amplitude: 0.2 MPa. Continue on
next page.

82

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

DAC 0/16, elastomer,


50C, 0.1 MPa

0,1

600 s/1.8 s
1 s/1.8 s

compliance J(t) (MPa^-1)

compliance J(t) (MPa^-1)

Ch. 3

DAC 0/16, elastomer,


50C, 0.2 MPa
600 s/1.8 s
1 s/1.8 s
0.2 s/1.8 s

0.1

0.2 s/1.8 s

0,01

0.01
1

10

100

1000 10000

10

time (s)

DAC 0/16, plastomer,


50C, 0.1 MPa

0.1

600 s/1.8 s

0.01

1000 10000

1 s/1.8 s
0.2 s/1.8 s

0.001

DAC 0/16, plastomer,


50C, 0.2 MPa

compliance J(t) (MPa^-1)

compliance J(t) (MPa^-1)

100

time (s)

0.1

0.01

600 s/1.8 s
1 s/1.8 s
0.2 s/1.8 s

0.001
1

10

100

time (s)

1000 10000

10

100

1000 10000

time (s)

Figure 23 (Contd). Top left (e): As figure 23a, 50C. Top right (f): As figure
23b, 50C. Bottom left (g): As figure 23c, 50C. Bottom right (h): As figure 23d,
50C.

Sec. 7 Results of dynamic creep tests in the absence of confinement

83

Notice, that the creep curve shifts to the left, i.e. to lower time value, as
the loading time decreases from 600 s to 1 s at constant rest-time, and that
the creep curve shifts back to the right, i.e. to higher time value, as the
loading time decreases further from 1 s to 0.2 s at constant rest-time. This
time dependence of the creep curve, i.e. its dependence on the blockwaveform is similar as in figure 18; however, the loading time where the
direction of the shift changes, differs. Possibly, this can be attributed to the
mixture composition.
Figures 24 show J ( t ) versus log t , on log-log scale. It seems that the
curves are linear between the first and the second data point. However,
this is uncertain, because no measurements were sampled between the first
and the tenth load repetition.
If the first data point, which corresponds to the first load repetition, is
omitted, then a reasonably reliable characterisation of the creep behaviour
is still possible.
Equation (15) was fitted to the creep curves designated by 1 s/1.8 s and 0.2
s/1.8 s, in figure 23, omitting the first data point. Equation (16) was fitted
to the creep curves designated by 600 s/1.8 s. The results are summarised
in figures 25.
Equation (21) was also fitted to the creep curves designated by 1 s/1.8 s
and 0.2 s/1.8 s in figure 23, omitting the first data point. The results are
discussed later in this section, cf. figure 29.
Let us first discuss the results obtained using equations (15) and (16), cf.
figure 25. Figure 26 shows J1 of equation (15) as function of the loadingtime, t l . Figure 27 shows z of equation (15) as function of the loadingtime, t l . Note, as previously, that J1 of equation (15), is wrongly
predicted for the cases in figure 26, because equation (15) does not fit well
to the observed creep at low time value. It means that the J1 -values from
the regression analysis do not reflect the measured creep compliance at 1
s. For example, negative J1 -values may be obtained, cf. figure 26, but
measured J1 -values cannot be negative.
Thus, although the fit of equation (15) to the observed creep is reasonable
with an explained variance greater than 99% (because the first data point
was omitted9), it is not functional in a practical context, since J1 does not
reflect the measured creep compliance at 1 s. To avoid this, J 10 could be
determined using equation (21).

The data point that corresponds to the first load repetition (respectively after 1
+ 1.8 = 2.8 s in case of the 1 s/1.8 s block-waveform, and after 0.2 + 1.8 = 2.0 s
in case of the 0.2 s/1.8 s block-waveform), was omitted.

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

84

DAC 0/16, elastomer,


40C, 0.1 MPa

0.1

0.01

600 s/1.8 s
1 s/1.8 s
0.2 s/1.8 s

0.001
0.1

10

0.1

0.01

0.2 s/1.8 s

0.001
0.1

10

0.1

600 s/1.8 s
1 s/1.8 s

DAC 0/16, plastomer,


40C, 0.2 MPa

compliance J(t) (MPa^-1)

compliance J(t) (MPa^-1)

log time (s)

DAC 0/16, plastomer,


40C, 0.1 MPa

0.01

600 s/1.8 s
1 s/1.8 s

log time (s)

DAC 0/16, elastomer,


40C, 0.2 MPa

compliance J(t) (MPa^-1)

compliance J(t) (MPa^-1)

Ch. 3

0.1

0.01

600 s/1.8 s
1 s/1.8 s
0.2 s/1.8 s

0.2 s/1.8 s

0.001

0.001
0.1

log time (s)

10

0.1

10

log time (s)

Figure 24. Dynamic creep of dense graded asphalt concrete, DAC 0/16, polymer
modified. J (t ) versus log time, both plotted logarithmically. The waveform of
the applied load was a block-wave, given by a loading time and a rest-time, cf.
fig. 10. Different loading times: 600 s, 1 s, and 0.2 s. Constant rest-time: 1.8 s.
Temperature: 40C. Specimen 60 x 100 mm height x diameter. Load control:
hydraulic. Top left (a): DAC 0/16, elastomer modified. Applied stress amplitude:
0.1 MPa. Top right (b) As fig. (a), applied stress amplitude: 0.2 MPa. Bottom
left (c): DAC 0/16, plastomer modified. Applied stress amplitude: 0.1 MPa.
Bottom right (d): As fig. (c), applied stress amplitude: 0.2 MPa.

Sec. 7 Results of dynamic creep tests in the absence of confinement

DAC 0/16, elastomer,


50C, 0.1 MPa

0.1

0.01

600 s/1.8 s
1 s/1.8 s
0.2 s/1.8 s

0.001

DAC 0/16, elastomer,


50C, 0.2 MPa

compliance J(t) (MPa^-1)

compliance J(t) (MPa^-1)

0.1

0.01

600 s/1.8 s
1 s/1.8 s
0.2 s/1.8 s

0.001
0.1

10

0.1

log time (s)

10

600 s/1.8 s
1 s/1.8 s

DAC 0/16, plastomer,


50C, 0.2 MPa

compliance J(t) (MPa^-1)

0.1

0.01

log time (s)

DAC 0/16, plastomer,


50C, 0.1 MPa

compliance J(t) (MPa^-1)

85

0.1

0.01

600 s/1.8 s
1 s/1.8 s
0.2 s/1,8 s

0.2 s/1.8 s

0.001

0.001
0.1

log time (s)

10

0.1

10

log time (s)

Figure 24 (Contd). Top left (e): As figure 24a, 50C. Top right (f): As figure
24b, 50C. Bottom left (g): As figure 24c, 50C. Bottom right (h): As figure 24d,
50C.

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

86

compliance J(t) (MPa^-1)

Ch. 3

DAC 0/16, elastomer, 40C, 0.1 MPa


600 s/1.8 s
1 s/1.8 s
0.2 s/1.8 s

0.1

0.01
1

10

100

1000

10000

time (s)
y = 0.0371x0.0927
R = 0.9983; 600 s/1.8 s
2

y = 0.0139Ln(x) + 0.0103
R2 = 0.9903; 1 s/1.8 s

compliance J(t) (MPa^-1)

y = 0.0115Ln(x) - 0.0125
R2 = 0.9976; 0,2 s/1,8 s

DAC 0/16, elastomer, 40C, 0.2 MPa


600 s/1.8 s
1 s/1.8 s
0.2 s/1.8 s

0.1

0.01
1

10

100

1000

10000

time (s)
y = 0.0233x0.0948
R = 0.9997; 600 s/1.8 s
2

y = 0.009 Ln (x) + 0.0106


R2 = 0.9976; 1 s/1.8 s

y = 0.0079 Ln (x) - 0.0027


R2 = 0.9947; 0.2 s/1.8 s

Figure 25. Dynamic creep of dense graded asphalt concrete, DAC 0/16, polymer
modified. J (t ) versus the time, t , both plotted logarithmically. In the regression
equations y (x), y represents J (t ) , and x represents the time. The waveform of
the applied load was a block-wave, given by a loading time and a rest-time, cf.
fig. 10. Different loading times: 600 s, 1 s, and 0.2 s. Constant rest-time: 1.8 s.
Temperature: 40C. Specimen 60 x 100 mm height x diameter. Load control:
hydraulic. Top (a): DAC 0/16, elastomer modified. Applied stress amplitude: 0.1
MPa. Bottom (b) As fig. (a), applied stress amplitude: 0.2 MPa.

Sec. 7 Results of dynamic creep tests in the absence of confinement

compliance J(t) (MPa^-1)

87

DAC 0/16, plastomer, 40C, 0.1 MPa

0.1

0.01

600 s/1.8 s
1 s/1.8 s
0.2 s/1.8 s

0.001
1

10

100

1000

10000

time (s)
0.1162

y = 0.046 x
y = 0.0242 Ln (x) - 0.049 y = 0.0248 Ln (x) - 0.0768
R2 = 0.9965; 600 s/1.8 s R2 = 0.9954; 1s/1.8 s
R2 = 0.975; 0.2 s/1.8 s

compliance J(t) (MPa^-1)

DAC 0/16, plastomer, 40C, 0.2 MPa

0.1
600 s/1.8 s
1 s/1.8 s
0.2 s/1.8 s
0.01
1

10

100

1000

10000

time (s)
y = 0.0552x0.0807
R2 = 0.999; 600 s/1.8 s

y = 0,0154 Ln (x) - 0,0053


R2 = 0.9942; 1 s/1.8 s

y = 0.0124 Ln (x) - 0.0242


R2 = 0.9988; 0.2 s/1.8 s

Figure 25 (Contd). Top (c): As figure 25a, DAC 0/16, plastomer modified.
Bottom (d): As figure 25b, DAC 0/16, plastomer modified.

88

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

Ch. 3

DAC 0/16, elastomer, 50C, 0.1 MPa

compliance J(t) (MPa^-1)

600 s/1.8 s
1 s/1.8 s
0.2 s/1.8 s

0.1

0.01
1

10

100

1000

10000

time (s)
y = 0.0105 Ln (x) + 0.0229
y = 0.0422x0.0956
2
R2 = 0.9964
R = 0.9978; 600 s/1.8 s

y = 0.0116Ln(x) + 0.006
R2 = 0.9893; 0.2 s/1.8 s

DAC 0/16, elastomer 50C, 0.2 MPa

compliance J(t) (MPa^-1)

600 s/1.8 s
1 s/1.8 s
0.2 s/1.8 s

0.1

0.01
1

10

100

1000

10000

time (s)
y = 0.0246x0.1044
R = 0.9889; 600 s/1.8 s
2

y = 0.0086Ln(x) + 0.0155
R2 = 0.9997; 1 s/1.8 s

y = 0.0095Ln(x) + 0.009
R2 = 0.9959; 0.2 s/1.8 s

Figure 25. Top (e): As figure 25a, 50C. Bottom (f): As figure 25b, 50C.

Sec. 7 Results of dynamic creep tests in the absence of confinement

compliance J(t) (MPa^-1)

89

DAC 0/16, plastomer, 50C, 0,1 MPa


600 s/1.8 s
1 s/1.8 s
0.2 s/1.8 s

0.1

0.01
1

10

100

1000

10000

time (s)
y = 0.0403x0.1289
R = 0.9985; 600 s/1.8 s
2

y = 0.0284 Ln (x) - 0.0448


R2 = 0.9909; 1 s/1.8 s

compliance J(t) (MPa^-1)

y = 0.027 Ln (x) - 0.0593


R2 = 0.9969; 0.2 s/1.8 s

DAC 0/16, plastomer, 50C, 0,2 MPa


600 s/1.8 s
1 s/1.8 s
0.2 s/1.8 s

0.1

0.01
1

10

100

1000

10000

time (s)
y = 0.0415x0.1048
R = 0.9943; 600 s/1.8 s
2

y = 0.0175 Ln (x) + 0.0084


R2 = 0.993; 1 s/1.8 s

y = 0.0165 Ln (x) - 0.0303


R2 = 0.9841; 0.2 s/1.8 s

Figure 25 (Contd). Top (g): As figure 25c, 50C. Bottom (h): As figure 25d,
50C.

90

400
200
0
-200
-400
-600
-800
0

0.2

0.4

0.6

0.8

1.2

loading-time, tl (s)

coefficient of eq.(15), J110^4

coefficient of eq.(15), J110^4

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

Ch. 3

400
200
0
-200
-400
-600
-800
0

0.2

0.4

0.6

0.8

loading-time, tl (s)

1.2

40C/0.1 MPa

40C/0.2 MPa

40C/0.1 MPa

40C/0.2 MPa

50C/0.1 MPa

50C/0.2 MPa

50C/0.1 MPa

50C/0.2 MPa

300
250
200
150
100
50
0
0

0.2

0.4

0.6

0.8

1.2

loading-time, tl (s)
40C/0.1 MPa
50C/0.1 MPa

40C/0.2 MPa
50C/0.2 MPa

exponent of eq. (15), z10^4

exponent of eq. (15), z10^4

Figure 26. Coefficient J 1 of equation (15) as function of the loading time.


Left (a): Elastomer modified DAC 0/16. Right (b): Plastomer modified DAC
0/16.
300
250
200
150
100
50
0
0

0.2

0.4

0.6

0.8

loading-time, tl (s)

40C/0.1 MPa
50C/0.1 MPa

1.2

40C/0.2 MPa
50C/0.2 MPa

Figure 27. Exponent z of equation (15) as function of the loading time.


Left (a): Elastomer modified DAC 0/16. Right (b): Plastomer modified DAC
0/16.

Sec. 7 Results of dynamic creep tests in the absence of confinement

DAC 0/16, elastomer, 40C, 0.1 MPa

compliance J(t) (MPa^-1)

91

J(t)2 = 0.039 t0.9175


R2 = 0.9921

J(t)1 = 0.0832 Ln (t) + 0.0184


R2 = 0.9982
1 s/1.8 s
0.2 s/1.8 s

0.1

J(t)3 = 0.0649 Ln (t) - 0.0017


R2 = 0.9865
J(t)4 = 0.0158 t1.2988
R2 = 0.9943

0.01
1

log time (s), logarithimically plotted

10

Figure 28. Example: Logarithmic creep functions regression equations, J (t )1


and J (t )3 compared to power creep functions regression equations, J (t )2
and J (t )4 . The logarithmic function, e.g. J (t )1 shows a monotonically
decreasing slope as a function of log time in a log-log plot; the power function,
e.g. J (t )2 , shows a constant slope as function of log time in a log-log plot.

An example is shown in figure 28. Figure 28 shows data of DAC 0/16 E


(elastomer modified), tested at 40C, and 0.1 MPa stress amplitude, cf.
figure 25a (the first data point of the original data, cf. figure 24, was
omitted). In figure 28, two creep curves are shown, for the waveform
1 s/1.8 s, and the wave form 0.2 s/1.8 s (i.e. loading-time/rest-time). J (t )
is shown as function of log t , both plotted logarithmically. Equation (15)
and equation (21) are both fitted to each curve. Using equation (15),
J (t ) = J e + ln (ln t ) z = J e + z ln ln t
(35)
where J e is the compliance at e s, e is the base of the natural logarithm,
2.71828. Since J e cannot be measured, we will rewrite equation (35) in
terms of the 10log. Using equation (19), equation (35) can be rewritten as:
J (t ) = J e + z 2.3026 log (2.3026 log t )
J (t ) = J e + z 2.3026 (log 2.3026 + log log t )
J (t ) = J 10 + z 2.3026 log log t
(36a)
where J 10 = J e + z 2.3026 log 2.3026 , or
J 10 = J e + 0.834 z
(36b)
z = z 2.3026
(36c)
Equation (36a) is a logarithmic function. A logarithmic function can be
represented by a straight line in a semi-log plot. Equation (36a) yields a
straight line in a plot of J ( t ) versus log log t . However, in a log log plot,

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

92

0.2 s/1.8 s

y = 0.039x
R2 = 0.9921
0.1

y = 0.0158x1.2988
R2 = 0.9943

compliance J(t) (MPa^-1)

1 s/1.8 s
0.9175

0.01

DAC 0/16, elastomer,


40C, 0.2 MPa
1 s/1.8 s
0.2 s/1.8 s
0.8432

y = 0.029x
2
0.1 R = 0.9989

y = 0.0152x1.1163
R2 = 0.9914

0.01
1

compliance J(t) (MPa^-1)

DAC 0/16, elastomer,


40C, 0.1 MPa

10

log time (s)

DAC 0/16, plastomer,


40C, 0,1 MPa
1 s/1.8 s

y = 0.0191x1.6306
R2 = 0.9901

0.2 s/1.8 s

0.1

y = 0.0065x2.3385
R2 = 0.9958

0.01

compliance J(t) (MPa^-1)

compliance J(t) (MPa^-1)

Ch. 3

10

log time (s)

DAC 0/16, plastomer,


40C, 0.2 MPa
1 s/1.8 s

y = 0.0297x1.1104
R2 = 0.9898

0.2 s/1.8 s

0.1

y = 0.0102x1.5998
R2 = 0.9982

0.01
1

log time (s)

10

log time (s)

10

Figure 29. As figure 24 if the first data point is omitted. J (t ) versus log t , J (t )
and log t both plotted logarithmically. In regression equations, y represents,
J (t ) and x represents log t . For testing conditions, cf. fig. (29). Temperature:
40C. Top left (a): DAC 0/16, elastomer modified. Applied stress amplitude: 0.1
MPa. Top right (b) As fig. (a), applied stress amplitude: 0.2 MPa. Bottom left
(c): DAC 0/16, plastomer modified. Applied stress amplitude: 0.1 MPa. Bottom
right (d): As fig. (c), applied stress amplitude: 0.2 MPa.

it represents a curve with a monotonically decreasing slope, cf. figure 28.


In figure 28, clearly, the data points are on a curve with monotonically
decreasing slope. Since the first data point, at t = 1 s, was omitted, the
second data point is the new first data point. This is the compliance at
t = 10 s, J 10 . Recall equation (21),
J (t ) = J10 (log t ) z log J (t ) = log J10 + z log log t
(37)
Equation (37) is a power function. A power function can be represented

Sec. 7 Results of dynamic creep tests in the absence of confinement

DAC 0/16, elastomer,


50C, 0.1 MPa

0.2 s/1.8 s

y = 0.0421x
R2 = 0.9983
0.1

y = 0.03x0.9621
R2 = 0.9884

DAC 0/16, elastomer,


50C, 0.2 MPa
1 s/1.8 s
0.2 s/1.8 s

0.1

y = 0.0323x0.7725
R2 = 0.9992

y = 0.0286x0.875
R2 = 0.9968

0.01

0.01
1

compliance J(t) (MPa^-1)

compliance J(t) (MPa^-1)

1 s/1.8 s
0.7498

log time (s)

10

DAC 0/16, plastomer,


50C, 0.1 MPa

y = 0.0341x1.3382
R2 = 0.9973
0.1

y = 0.0186x1.7122
R2 = 0.9988
1 s/1.8 s

compliance J(t) (MPa^-1)

compliance J(t) (MPa^-1)

93

log time (s)

10

DAC 0/16, plastomer,


50C, 0.2 MPa
1 s/1.8 s
0.9015

y = 0.048x
R2 = 0.9941

0.2 s/1.8 s

0.1

y = 0.0163x1.4653
R2 = 0.9968

0.2 s/1.8 s

0.01
1

log time (s)

0.01
10

log time (s)

10

Figure 29 (Contd). Top left (e): As figure 29a, 50C. Top right (f): As fig. 29b,
50C. Bottom left (g): As figure 29c, 50C. Bottom right (h): As fig. 29d, 50C.

by a straight line in a log-log plot, cf. figure 28. Of the data in figures 25,
those in figure 25a show the greatest curvature (curve 1 s/1.8 s). Fitting
equation (37) yields an explained variance satisfactorily greater than 99%.
Therefore, the power function, equation (37), can be used to fit all cases in
figures 25, so that a maximum explained variance is obtained. The results
are shown in figures 29, and in tables 1011. In tables 1011, J 10 , z ,
(10) , and ~ (10) are listed for the various testing conditions.
Notice that the creep curves form wedges in figures 29. It was
mentioned previously that this means that the untransformed curves are
(read: should be) parallel creep curves, cf. 7.7. Figure 30 illustrates
schematically the dependence of J 10 and z on the block-wave form.

94

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

Ch. 3

Table 10. DAC 0/16, elastomer modified. Dynamic creep. Block-wave loading with
constant rest-time, 1.8 s. Constitutive model parameters J 10 , and z of eq. (21), if
J (t ) is in MPa-1; (10) of eq. (27), and ~ (10) of eq. (33). (10) = J 10 z /2.3026;
~ (10) = z 2.3026.
J 10 10 4
z
(10)
~ (10)
Loading-time/rest-time
40C/0.1 MPa
0.2 s/1.8 s
158
1.2988
0.0089
0.56
1.0 s/1.8 s
390
0.9715
0.0165
0.42
40C/0,2 Mpa
40C/0.2 MPa
0.2 s/1.8 s
152
1.1163
0.0074
0.48
1.0 s/1.8 s
290
0.8432
0.0106
0.37
50C/0.1 MPa
0.2 s/1.8 s
300
0.9621
0.0125
0.42
1.0 s/1.8 s
421
0.7498
0.0137
0.33
50C/0.2 MPa
0.2 s/1.8 s
286
0.8750
0.0109
0.38
1.0 s/1.8 s
323
0.7725
0.0108
0.34
Table 11. DAC 0/16, plastomer modified. Dynamic creep. Block-wave loading with
constant rest-time, 1.8 s. Constitutive model parameters J 10 , and z of eq. (21), if
J (t ) is in MPa-1; (10) of eq. (27), and ~ (10) of eq. (33). (10) = J 10 z /2.3026;
~
(10) = z 2.3026.
J 10 10 4
z
(10)
~ (10)
Loading-time/rest-time
40C/0.1 MPa
0.2 s/1.8 s
65
2.3385
0.0066
1.02
1.0 s/1.8 s
191
1.6306
0.0135
0.71
40C/0.2 MPa
0.2 s/1.8 s
102
1.5998
0.0071
0.69
1.0 s/1.8 s
297
1.1104
0.0143
0.48
50C/0.1 MPa
0.2 s/1.8 s
186
1.7122
0.0138
0.74
1.0 s/1.8 s
341
1.3382
0.0198
0.58
50C/0.2 MPa
0.2 s/1.8 s
163
1.4653
0.0104
0.64
1.0 s/1.8 s
480
0.9015
0.0188
0.39

Figure 30 shows, that, if the creep curve shifts to the left, i.e. to lower time
value, then J 10 increases, and z decreases. Conversely, if the creep curve
shifts back to the right, i.e. to higher time value, then J 10 decreases, and z
increases. Thus, the time dependence of J 10 and z is shown. To explain
the time dependence, one must explain the shift of the creep curve along
the time axis. Different time dependencies were found in this study,
depending on the shape of the waveform. In the present case, the creep
curve shifts to the left, if the loading time decreases from 600 s to 1 s at
constant rest-time, and the curve shifts back to the right if the loading time

compliance, J(t) (MPa^-1)

Sec. 7 Results of dynamic creep tests in the absence of confinement

95

J 10 , z

0.1

t l , t r = constant

t l , t r = constant
0.01
J 10 , z
0.001
1

10

100

1000

10000

100000

time (s)

Figure 30. Schematic illustration of the time dependence of J 10 and z . If the


loading time, t l , decreases from 600 s to 1 s at constant rest-time, t r , 1.8 s, then
the creep curve shifts to the left, i.e. to lower time value; J 10 increases and z
decreases. If t l decreases further from 1 s to 0.2 s at constant t r , then the creep
curve shifts back to the right, i.e. to higher time value; J 10 increases and z
decreases.

decreases further from 1 s to 0.2 s at constant rest-time. This behaviour


cannot be explained on the basis of the time temperature superposition
principle (if that were valid, which it is not because the behaviour is
nonlinear). Based on that principle, the mixture is expected to respond
with greater stiffness if the loading time decreases at constant rest-time.
Then, the period time of the load signal decreases, and the frequency
increases10. Accordingly, the creep compliance is expected to decrease,
i.e. the creep curve is expected to shift to higher time value. Since the
result is obviously not explained by the time temperature superposition
principle, it is likely to be attributable to the response of the grain skeleton
in stress controlled compression. Possibly, the shift to the left as a
function of increasing frequency is to be attributed to the aggressiveness
of the block-wave, which prevents the grain skeleton from becoming
immobilised by the internal friction. Evidently, at some point, the creep
compliance cannot continue to increase as the loading time decreases at
constant rest-time, since there will be insufficient applied load left to keep
the creep going. This probably explains the turning point, where the shift
turns back to higher time value, if the loading time decreases further from
0.2 s to 0.05 s. It is probable that the loading time of the turning point
depends on the material composition, cf. 7.1, figure 18.
10

It is assumed that the notion frequency applies to any periodical function.


Hence, for a block-wave, defined by a loading time and a rest-time, the blockwave frequency increases if the loading time decreases at constant rest-time, or if
the rest-time decreases at constant loading time.

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

96

parameter J1010^4

600
500
400

Ch. 3

40C/0.1 MPa
40C/0.2 MPa
50C/0.1 MPa
50C/0.2 MPa

300
200
100
0
0.2 s/1.8 s

1 s/1.8 s

0.2 s/1.8 s

1 s/1.8 s

loading-time/rest-time
DAC 0/16 P

DAC 0/16 E

Figure 31. Coefficient J 10 of equation (21) as function of the loading-time.


Left (a): DAC 0/16 E: with elastomer. Right (b): DAC 0/16 P: with plastomer.

exponent z of eq. (21)

2.5

40C/0.1 MPa
40C/0.2 MPa

50C/0.1 MPa
50C/0.2 MPa

1.5
1
0.5
0
0.2 s/1.8 s

1 s/1.8 s

0.2 s/1.8 s

1 s/1.8 s

loading-time/rest-time
DAC 0/16 E

DAC 0/16 P

Figure 32. Coefficient z$ of equation (21) as function of the loading-time.


Left (a): DAC 0/16 E: with elastomer. Right (b): DAC 0/16 P: with plastomer.

Figure 31 shows J 10 of equation (21) as function of the loading time, t l .


In figure 31, J 10 , for 0.2 s t l 1 s (constant rest-time: 1.8 s), ranges
between 0.0152 and 0.0421 for DAC 0/16 E (elastomer) and between
0.0065 and 0.0480 for DAC 0/16 P (plastomer). Those values compare to
values between 0.0028 and 0.0226 for the unmodified mixture, cf. figure
16. Thus, comparison of figure 31 and figure 16 shows that J 10 of the
polymer modified DAC 0/16 is significantly greater than the
corresponding J 10 of the unmodified DAC 0/16 mixture.
Figure 32 shows the corresponding z of equation (21) as function of the

Sec. 7 Results of dynamic creep tests in the absence of confinement

97

loading time, t l . In figure 32, z , ranges between 0.7 and 1.3 for DAC
0/16 E and between 0.9 and 2.4 for DAC 0/16 P. Those values compare to
values between 2 and 3 for the unmodified mixture, cf. figure 17. Thus,
comparison of figure 32 and figure 17 shows that z of the polymer
modified DAC 0/16 is significantly smaller than the corresponding z of
the unmodified DAC 0/16 mixture.

7.2.2 Temperature dependence and stress dependence of creep properties


If figures 29a..d (40C) and the respective figures 29e..h (50C) are
compared, it can be seen that the creep curve shifts to the left, i.e. to lower
time value, as the temperature increases, causing the creep compliance to
increase for a given time. This shift corresponds to an increase of J 10 with
increasing temperature, which is confirmed in figure 31.
If figures 29a,c,e,g and the respective figures 29b,d,f,h are compared, it
can be seen that the creep curve shifts to the right, i.e. to higher time
value, as the applied stress amplitude increases, causing the creep
compliance to decrease for a given time. This shift corresponds to a
decrease of J 10 with increasing applied stress amplitude. This is
confirmed by five out of eight cases in figure 31; i.e. except the following
three cases:
1 third group from left, block-wave: 0.2 s/1.8 s, the first two bars;
2 fourth group from left, block-wave: 1 s/1.8 s, bars 1 and 2;
3 fourth group from left, block-wave: 1 s/1.8 s, bars 3 and 4.
As noticed previously, z is expected to decrease as J 10 increases, and
vice versa. This dependence between z and J 10 is caused, as also noted
previously, by the parallelism of the creep curves in figure 25 and the logtransformation of the log (time) data when equation (21) is used.
Not in all cases the real changes in J 10 and z as a function of the
temperature or the stress are in agreement with expected changes. In those
cases it seems the scatter on the data overshadows the influences of the
temperature respectively the stress.
7.3 Dense graded asphalt concrete 0/16, unmodified, using a sinusoidal
applied stress
7.3.1 Time dependence of creep properties
The influence of the shape of the waveform of the applied stress was
investigated further by the application of a sinusoidal applied stress at two
temperatures, 40C and 50C, and two applied stress amplitudes: 0.1 MPa
and 0.2 MPa. The applied frequencies were: 0.1 Hz, 1 Hz, and 10 Hz. A
list of creep models is given in appendix 3, table 6.
Figure 33 shows J (t ) versus the time, t , on log-log scale, for DAC 0/16
using a sinusoidal load. From figure 33 it seems that two creep domains

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

98

compliance, J(t) (MPa^-1)

Ch. 3

DAC 0/16, 40C


sinusoidal load

0.1

0.1 Hz/0.1 MPa


1 Hz/0.1 MPa
10 Hz/0.1 MPa
0.1 Hz/0.2 MPa
1 Hz/0.2 MPa
10 Hz/0.2 MPa

0.01

0.001
1

10

100

1000

10000

time (s)

compliance, J(t) (MPa^-1)

DAC 0/16, 50C


sinusoidal load

0.1
0.1 Hz/0.1 MPa
1 Hz/0.1 MPa
10 Hz/0.1 MPa
0.01

0.1 Hz/0.2 MPa


1 Hz/0.2 MPa
10 Hz/0.2 MPa

0.001
1

10

100

1000

10000

time (s)

Figure 33. Dynamic creep of dense graded asphalt concrete, DAC 0/16. J (t )
versus the time, t ; both plotted logarithmically. Applied load: sinus. Different
frequencies: 0.1 Hz, 1.0 Hz, 10 Hz. Different applied stress amplitudes: 0.1
MPa, 0.2 MPa. Each creep curve represents the average of three tests. Load
control: hydraulic. Top (a): Temperature: 40C. Bottom (b): Temperature 50C.

Sec. 7 Results of dynamic creep tests in the absence of confinement

99

compliance, J(t) (MPa^-1)

can be indicated, except for the testing condition [0.1 Hz/0.2 MPa]. For
the latter, perhaps three creep domains may have to be distinguished;
however, this curve is not considered further. Figure 33 shows one creep
domain in the time-interval between 1 s and 50 s, in which the creep is of
the type of equation (16), and a second creep domain in the time-interval
between 100 s and 10.000 s, in which the creep is of the type of equation
(16), or of the type of equation (15). If it is attempted to fit equation (15)
to the whole creep curve, then the fitted equation is found to show a misfit
similar to the example in figure 13a. A disadvantage of distinguishing two
(or more) creep domains is that two (respectively more) constitutive
models are needed to describe the creep.
Figure 33 shows that the creep curve shifts to the left, i.e. to lower time
value, as the frequency decreases from 10 Hz to 0.1 Hz. The time
dependence of J 10 and z is illustrated schematically in figure 34.
1

J 10 , z

0.1

f
0.01

J 10 , z
0.001
1

10

100

1000

10000

100000

time (s)

Figure 34. Schematic illustration of the time (frequency) dependence of J 10 and


z . If the frequency decreases from 10 Hz to 0.1 Hz, then the creep curve shifts to
the left, i.e. to lower time value; J 10 increases and z decreases. If the frequency
increases from 0.1 Hz to 10 Hz, then the creep curve shifts to the right, i.e. to
higher time value; J 10 decreases and z increases.

Figure 34 illustrates that J 10 increases, and z decreases if the creep curve


shifts to the left, i.e. to lower time value, when the frequency decreases.
Conversely, J 10 decreases, and z increases if the creep curve shifts to the
right, i.e. to higher time value, when the frequency increases. Thus, the
time dependence of J 10 and z is shown. To explain the time dependence,
one must explain the shift of the creep curve along the time axis. In the
present case, the creep curve shifts to the left, if the frequency decreases
from 10 Hz to 0.1 Hz, or equivalently, to the right if the frequency
increases from 0.1 Hz to 10 Hz. This behaviour can be explained on the
basis of the time temperature superposition principle (if that were valid,

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

100

DAC 0/16, 40C


sinusoidal load

compliance, J(t) (MPa^-1)

Ch. 3

0.1 Hz/0.1 MPa


1 Hz/0.1 MPa
10 Hz/0.1 MPa
0.1 Hz/0.2 MPa
1 Hz/0.2 MPa
10 Hz/0.2 MPa

0.1

0.01

0.001
1

10

log time (s)

y = 0.0513x1.1837
R = 0.9996; 0.1 Hz/0.1 MPa

y = 0.0276x1.3008
R = 0.9977; 0.1 Hz/0.2 MPa

y = 0,0187x1.4184
R2 = 0.9992; 1 Hz/0.1 MPa

y = 0.0128x1.6693
R = 0.9935; 1Hz/0.2 MPa
2

y = 0.0072x1.7473
R = 0.9994; 10 Hz/0.1 MPa

y = 0.0063x1.6383
R = 0.9999; 10 Hz/0.2 MPa

compliance, J(t) (MPa^-1)

DAC 0/16, 50C


sinusoidal load
0.1 Hz/0.1 MPa
1 Hz/0.1 MPa
10 Hz/0.1 MPa
0.1 Hz/0.2 MPa
1 Hz/0.2 MPa
10 Hz/0.2 MPa

0.1

0.01
1

10

log time (s)


1.6121

1.6238

y = 0.0293x
R2 = 1; 0.1 Hz/0.1 MPa

y = 0.0345x
R2 = 1; 0.1 Hz/0.2 MPa

y = 0.0384x1.2841
R = 1; 1 Hz/0.1 MPa

y = 0.0222x1.4044
R = 1; 0.1 Hz/0.2 MPa
2

y = 0.0168x1.6168
R = 0.9991; 10 Hz/0.1 MPa
2

y = 0.0122x1.5314
R = 0.9997; 10 Hz/0.2 MPa
2

Figure 35. J (t ) versus log t from fig. 33; both plotted logarithmically. Applied
load: sinus. Different frequencies: 0.1 Hz, 1.0 Hz, 10 Hz. Different applied
stress amplitudes: 0.1 MPa, 0.2 MPa. Each creep curve represents the average
of three tests. Load control: hydraulic. In the regression equations, y represents
J ( t ) , and x represents log t , cf. equation (21). Top (a): Temperature: 40C.
Bottom (b): Temperature 50C.

Sec. 7 Results of dynamic creep tests in the absence of confinement 101

which it is not because the behaviour is nonlinear). Based on that


principle, the mixture is expected to respond with lesser stiffness as the
frequency decreases. Accordingly, the creep compliance is expected to
increase, i.e. the creep curve is expected to shift to a lower time value.
Although the experimental results confirm this expectation, it must be
noted, as also indicated previously, that the time temperature
superposition principle is not applicable to creep of asphalt mixture, since
the creep is nonlinear.
Figure 35 shows J (t ) versus log t , both plotted logarithmically. Figure
35 shows that J (t ) versus log t is practically linear in a log-log plot.
That is: J (t ) is logarithmically linear until 10,000 s for the following
testing conditions: [1 Hz, 0.1 MPa], [1 Hz, 0.2 MPa], [10 Hz, 0.1 MPa],
and [10 Hz, 0.2 MPa], and is logarithmically linear until approximately
600 s for the following testing conditions: [0.1 Hz, 0.1 MPa], cf. figure
35b, and [0.1 Hz, 0.2 MPa], cf. figure 35a/b.
Table 12. DAC 0/16. Dynamic creep. Sinusoidal load. Different frequencies.
Constitutive model parameters J 10 , and z of equation (21), if J (t ) is in
MPa-1; (10) of eq. (27), and ~ (10) of eq. (33). (10) = J 10 z /2.3026;
~ (10) = z 2.3026.
J 10 10 4
z
(10)
~ (10)
frequency (Hz)
40C/0.1 MPa
0.1
513
1.1837
0.0264
2.73
1.0
187
1.4184
0.0115
3.27
10
72
1.7473
0.0055
4.02
40C/0.2 MPa
0.1
276
1.3008
0.0156
3.00
1.0
128
1.6693
0.0093
3.84
10
63
1.6383
0.0045
3.77
50C/0.1 MPa
0.1
293
1.6238
0.0207
3.74
1.0
384
1.2841
0.0214
2.96
10
168
1.6168
0.0118
3.72
50C/0.2 MPa
0.1
345
1.6121
0.0242
3.71
1.0
222
1.4044
0.0135
3.23
10
122
1.5314
0.0081
3.53

Table 12 shows J 10 , z , (t ) and ~ (t ) for the various testing


conditions. Figure 36 shows J 10 of equation (21) as a function of the
frequency. J 10 decreases as a function of increasing frequency. Figure 37
shows z of equation (21) as a function of the frequency. z seems to
increase (at 40C), or to be more or less constant (at 50C) as a function of
increasing frequency. The result in figure 36 is complementary to the
result in figure 16. The result in figure 37 is complementary to the results
in figure 17.

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

102

40C/0.1 MPa

600

2,5

50C/0.2 MPa

400
300
200
100
0

exponent z of eq. (21)

parameter J1010^4

50C/0.1 MPa

40C/0.1 MPa
40C/0.2 MPa

40C/0.2 MPa
500

Ch. 3

50C/0.1 MPa
50C/0.2 MPa

1,5
1
0,5
0

0.1 Hz

1 Hz

10 Hz

frequency

Figure 36. Coefficient J 10 of eq. (21)


as function of the frequency.

0.1 Hz

1 Hz

10 Hz

frequency

Figure 37. Exponent z of eq. (21)


as function of the frequency.

7.3.2 Temperature and stress dependence of creep properties


If figures 33a and 33b are compared, it can be seen that the creep curve
shifts to the left, i.e. to lower time value, as the temperature increases,
causing the creep compliance to increase for a given time. This shift
corresponds to an increase of J 10 with increasing temperature, which is
confirmed in figure 36 (except by J 10 for 40C, 0.1 MPa).
If figures 33a and 33b are compared, it can be seen that the creep curve
shifts to the right, i.e. to higher time value, as the applied stress amplitude
increases, causing the creep compliance to decrease for a given time. This
shift corresponds to a decrease of J 10 with increasing applied stress
amplitude. This is confirmed by five out of six cases in figure 36; i.e.
except the following case:
. first group from left, the two bars for 50C.
As noticed previously, z is expected to decrease as J 10 increases, and
vice versa. This dependence between z and J 10 is caused, as also noted
previously, by the parallelism of the creep curves in figure 33 and the logtransformation of the log (time) data when equation (21) is used.
Not in all cases the real changes in J 10 and z as a function of the
temperature or the stress are in agreement with expected changes. In those
cases it seems the scatter on the data overshadows the influences of the
temperature respectively the stress.

Sec. 7 Results of dynamic creep tests in the absence of confinement 103

J(t) (MPa^-1)

7.4 Dense graded asphalt concrete 0/16, polymer modified, using a


sinusoidal applied stress
7.4.1 Time dependence of creep properties
The influence of the shape of the wave form of the applied stress was
investigated further by the application of a sinusoidal applied stress to the
polymer modified mixtures at two temperatures, 40C and 50C, and two
applied stress amplitudes: 0.1 MPa and 0.2 MPa. The applied frequencies
were: 0.1 Hz, 1 Hz, and 10 Hz. A list of creep models is given in appendix
3, table 7.
Figure 39 shows J (t ) versus the time, t , on log-log scale, for DAC 0/16
with elastomer, and DAC 0/16 with plastomer. Each curve represents the
average of three tests. It seems that two creep domains in figure 39 can be
indicated, one in a lower time-interval, in which the creep is of the type of
equation (16), and a second in a higher time-interval, in which the creep is
of the type of equation (15), or of the type of equation (16). As a result, it
is in general not possible to fit equation (15) or equation (16) to the whole
creep curve. If it is attempted to fit equation (15) to the whole creep curve,
then the fitted equation shows a misfit similar to the example in figure
13a. A disadvantage of distinguishing two creep domains is that two
constitutive models are needed to describe the creep. No measurements
were sampled between the first and the tenth load repetition. Therefore,
the creep curves are not known in greater detail between the first and the
second data point of each curve in figure 39. A reasonable characterisation
of the creep is still possible if
1
DAC 0/16, plastomer, 40C
the first data point, corresponsinusoidal load
ding to the first load repetition,
is omitted. It was found that
0.1
equation (15) is in general not
suitable to fit the creep curve.
An example is shown in figure
0.01
38. Once more the reason was
that the measured creep
1 Hz/0,1 MPa
compliance at 1 s, represented
0.001
by J 1 , was wrongly predicted. It
1
10
100
1000 10000
is undesirable to have a
time (s)
constitutive model parameter
which does not reflect a Figure 38. Misfit of equation (15) to the
measured value. Therefore, measured creep curve. The constitutive
equation (21) was used. The model parameter J 1 wrongly predicts the
results are shown in tables 13 creep compliance at 1 s.
14 and figures 4042. Tables
1314 are complementary to tables 1011. The creep curve in figure 39
shifts to the left, i.e. to lower time value, if the frequency decreases. This
dependence of the creep curve on the frequency is similar to that shown in

104

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

Ch. 3

Table 13. DAC 0/16 E: elastomer modified. Dynamic creep. Sinusoidal load.
Constitutive model parameters J 10 , and z of eq. (21), if J (t ) is in MPa-1;
(10) of eq. (27), and ~ (10) of eq. (33). (10) = J 10 z /2.3026;
~ (10) = z 2.3026.
J 10 104
z
(10)
~ (10)
frequency (Hz)
40C/0.1 MPa
0.1
501
0.5564
0.0121
1.28
1
335
0.8645
0.0126
1.99
10
108
1.4259
0.0067
3.28
40C/0.2 MPa
0.1
404
0.5066
0.0089
1.17
1
274
0.6744
0.0080
1.55
10
107
1.3009
0.0060
3.00
50C/0.1 MPa
0.1
599
0.5111
0.0133
1.18
1
558
0.656
0.0159
1.51
10
266
1.0586
0.0052
2.44
40C/0.2 MPa
0.1
545
0.4542
0.0108
1.05
1
437
0.5672
0.0108
1.31
10
236
0.8089
0.0083
1.86
Table 14. As table 13. DAC 0/16 P: plastomer modified.
J 10 104
z
(10)
frequency (Hz)
40C/0.1 MPa
0.1
487
0.8017
0.0170
1
201
1.5771
0.0138
10
74
1.9969
0.0064
40C/0.2 MPa
0.1
575
0.6493
0.0162
1
196
1.0840
0.0092
10
81
1.7933
0.0063
50C/0.1 MPa
0.1
884
0.7697
0.0295
1
429
1.1247
0.0209
10
126
1.6083
0.0088
50C/0.1 MPa
0.1
1
266
1.1206
0.0129
10
153
1.5123
0.0100

~ (10)
1.85
3.63
4.60
1.50
2.50
4.13
1.77
2.59
3.70
2.58
3.48

Sec. 7 Results of dynamic creep tests in the absence of confinement 105

DAC 0/16, elastomer, 40C


sinusoidal load

compliance, J(t) (MPa^-1)

compliance, J(t) (MPa^-1)

0.1

0.01

0.001
1

10

100

1000

10000

0.1

0.01

0.001
1

time (s)
0.1 Hz/0.1 MPa
10 Hz/0.1 MPa
1 Hz/0.2 MPa

10

100

1000

10000

time (s)

1 Hz/0.1 MPa
0.1 Hz/0.2 MPa
10 Hz/0.2 MPa

0.1 Hz/0.1 MPa


10 Hz/0.1 MPa
1 Hz/0.2 MPa

DAC 0/16, elastomer, 50C


sinusoidal load

compliance, J(t) (MPa^-1)

compliance, J(t) (MPa^-1)

DAC 0/16, plastomer, 40C


sinusoidal load

0.1

0.01

0.001

1 Hz/0.1 MPa
0.1 Hz/0.2 MPa
10 Hz/0.2 MPa

DAC 0/16, plastomer, 50C


sinusoidal load

0.1

0.01

0.001
1

10

100

1000

10000

10

0.1 Hz/0.1 MPa


10 Hz/0.1 MPa
1 Hz/0.2 MPa

100

1000

10000

time (s)

time (s)
1 Hz/0.1 MPa

0.1 Hz/0.1 MPa

1 Hz/0.1 MPa

0.1 Hz/0.2 MPa

10 Hz/0.1 MPa

0.1 Hz/0.2 MPa

10 Hz/0.2 MPa

1 Hz/0.2 MPa

10 Hz/0.2 MPa

Figure 39. Dynamic creep of polymer modified dense graded asphalt concrete,
DAC 0/16. Whole creep curves. J (t ) versus time, t , both logarithmically plotted.
Waveform of applied load: sinus. Different frequencies: 0.1 Hz, 1.0 Hz, and 10
Hz. Different temperatures: 40C and 50C. Different applied stress amplitudes:
0.1 MPa and 0.2 MPa. Specimen: 60 x 100 mm height x diameter. Load control:
hydraulic. Each curve represents the average of three tests. Top left (a): DAC
0/16, elastomer modified, 40C. Top right (b): DAC 0/16, plastomer modified,
40C. Bottom left (c): DAC 0/16, elastomer modified, 50C. Bottom right (d):
DAC 0/16, plastomer modified, 50C.

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

compliance, J(t) (MPa^-1)

106

Ch. 3

DAC 0/16, elastomer, 40C


sinusoidal load

0.1

0.1 Hz/0.1 MPa


1 Hz/0.1 MPa
10 Hz/0.1 MPa
0.1 Hz/0.2 MPa

0.01

1 Hz/0.2 MPa
10 Hz/0.2 MPa

0.001
1

10

log time (s)


0.5564

0.5066

y = 0.0404x
R2 = 0.9998; 0.1 Hz/0.2 MPa

y = 0.0501x
R2 = 0.9978; 0.1 Hz/0.1 MPa

y = 0.0274x0.6744
R = 0.9963; 1 Hz/0.2 MPa

y = 0.0335x0.8645
R2 = 0.99; 1 Hz/0.1 MPa

y = 0.0108x1.4259
R2 = 0.9966; 10 Hz/0.1 MPa

y = 0.0107x1.3009
R = 0.9912; 10 Hz/0.2 MPa
2

compliance, J(t) (MPa^-1)

DAC 0/16, elastomer, 50C


sinusoidal load
0.1 Hz/0.1 MPa
1 Hz/0.1 MPa

0.1

10 Hz/0.1 MPa
0.1 Hz/0.2 MPa
1 Hz/0.2 MPa
10 Hz/0.2 MPa

0.01
1
y = 0.0599x0.5111
R2 = 0.9985; 0.1 Hz/0.1 MPa
y = 0.0558x0.656
2
R = 0.9913; 1 Hz/0.1 MPa
y = 0.0266x1.0586
R = 0.9883; 10 Hz/0.1 MPa
2

10

log time (s)

y = 0.0545x0.4542
R = 0.9918; 0.1 Hz/0.2 MPa
2

y = 0.0437x0.5675
R2 = 0.9996; 1 Hz/0.2 MPa
y = 0.0236x0.8089
R = 0.9856; 10 Hz/0.2 MPa
2

Figure 40. J (t ) versus log t from figure 39a/c; both plotted logarithmically.
Applied load: sinus. Different frequencies: 0.1 Hz, 1.0 Hz, 10 Hz. DAC 0/16
elastomer modified. Specimen 60 x 100 mm height x diameter. Load control:
hydraulic. Top (a): 40C. Bottom (b): 50C.

Sec. 7 Results of dynamic creep tests in the absence of confinement 107

compliance, J(t) (MPa^-1)

DAC 0/16, plastomer, 40C


sinusoidal load
0.1 Hz/0.1 MPa

0.1

1 Hz/0.1 MPa
10 Hz/0.1 MPa
0.1 Hz/0.2 MPa

0.01

1 Hz/0.2 MPa
10 Hz/0.2 MPa

0.001
1
y = 0.0487x0.8017
R2 = 0.9995; 0.1 Hz/0.1 MPa

0.6493

y = 0.0575x
R2 = 0.9983; 0.1 Hz/0.2 MPa

y = 0.0201x1.5771
R = 0.9969; 1 Hz/0.1 MPa

y = 0.0196x1.084
R2 = 0.9997; 1 Hz/0.2 MPa

y = 0.0074x1.9969
R2 = 0.9966; 10 Hz/0.1 MPa

y = 0.0081x1.7933
R = 0.9972; 10 Hz/0.2 MPa
2

comliance, J(t) (MPa^-1)

10

log time (s)

DAC 0/16, plastomer, 50C


sinusoidal load
0.1 Hz/0.1 MPa
1 Hz/0.1 MPa
10 Hz/0.1 MPa

0.1

0.1 Hz/0.2 MPa


1 Hz/0.2 MPa
10 Hz/0.2 MPa

0.01
1
y = 0.0884x0.7697
R2 = 0.9933; 0.1 Hz/0.1 MPa
y = 0.0429x1.1247
R2 = 0.991; 1 Hz/0.1 MPa
y = 0.0126x1.6083
R2 = 0.9966; 10 Hz/0.1 MPa

10

log time (s)

0.1 Hz/0.2 MPa: not fitted


y = 0.0266x1.1206
R = 0.9844; 1 Hz/0.2 MPa
2

y = 0.0153x1.5123
R2 = 0.9932; 10 Hz/0.2 MPa

Figure 40 (Contd). As figure 40a/b, DAC 0/16, plastomer modified.


Top (c): 40C. Bottom (d): 50C.

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

parameter J1010^4

108

1000
900
800
700
600
500
400
300
200
100
0

Ch. 3

40C/0.1 MPa
40C/0.2 MPa
50C/0.1 MPa
50C/0.2 MPa

0.1 Hz

1 Hz

10 Hz

0.1 Hz

1 Hz

10 Hz

loading-time/rest-time

DAC 0/16 E

DAC 0/16 P

Figure 41. Coefficient J 10 of equation (16) as function of the frequency.


Left (a): DAC 0/16 E: with elastomer. Right (b): DAC 0/16 P: with plastomer.
Figure 42. Coefficient z$ of equation (16) as function of the frequency.

exponent z of eq. (21)

2.5

40C/0.1 MPa
40C/0.2 MPa

50C/0.1 MPa
50C/0.2 MPa

1.5
1
0.5
0
0.1 Hz

1 Hz

10 Hz

0.1 Hz

1 Hz

10 Hz

loading-time/rest-time

DAC 0/16 E

DAC 0/16 P

Left (a): DAC 0/16 E: with elastomer. Right (b): DAC 0/16 P: with plastomer.

figure 34. Figure 40 shows J ( t ) as a function of log t , plotted on log-log


scales. Figure 41 shows J 10 of equation (21) as function of the frequency,
and figure 42 shows z as a function of the frequency. J 10 decreases as a
function of increasing frequency, and z increases as a function of
increasing frequency. The results in figures 41 and 42 are in qualitative
agreement with similar results for the unmodified DAC 0/16 mixture
shown, respectively, in figures 36 and 37. The dependence on the
frequency of J 10 and z is a consequence of the time dependence of the
creep curve, cf. figure 34 and 7.7.

Sec. 7 Results of dynamic creep tests in the absence of confinement 109

7.4.2 Temperature and stress dependence of creep properties


If figures 39a and 39c, respectively figures 39b and 39d are compared, it
can be seen that the creep curve shifts to the left, i.e. to lower time value,
as the temperature increases, causing the creep compliance to increase for
a given time. This shift corresponds to an increase of J 10 with increasing
temperature, which is confirmed in figure 41.
If figures 39a and 39c, respectively figures 39b and 39d are compared, it
can be seen that the creep curve shifts to the right, i.e. to higher time
value, as the applied stress amplitude increases, causing the creep
compliance to decrease for a given time. This shift corresponds to a
decrease of J 10 with increasing applied stress amplitude. This is
confirmed by eight out of eleven cases in figure 36; i.e. except the
following three cases:
1 fourth group from left, the two bars for 40C;
2 sixth group from left, bars 1 and 2;
3 sixth group from left, bars 3 and 4.
As noticed previously, z is expected to decrease as J 10 increases, and
vice versa. This dependence between z and J 10 is caused, as also noted
previously, by the parallelism of the creep curves in figure 39 and the logtransformation of the log (time) data when equation (21) is used.
Not in all cases the real changes in J 10 and z as a function of the
temperature or the stress are in agreement with expected changes. In those
cases it seems the scatter on the data overshadows the influences of the
temperature respectively the stress.
7.5 Block-wave form of applied stress versus sinusoidal applied stress
7.5.1 Time dependence of creep properties
Let us assume, as previously, that the notion frequency applies to any
periodical function. Then, the frequency decreases if the loading time, t l ,
of a block-wave increases whilst the rest-time, t r , is constant. The
frequency decreases also if the rest-time of a block-wave increases whilst
the loading time is constant. The 1 s/1.8 s block-wave has a period of 2.8
s, hence a frequency of 0.357 Hz (for convenience, this is abbreviated to
0.3 Hz in the following). The 0.2 s/1.8 s block-wave has a period of 2 s,
hence a frequency of 0.5 Hz.
In figure 43a/b, J 10 and z of the 0.3 Hz and 0.5 Hz block-wave loads are
compared to J 10 and z of the 0.1 Hz and 1 Hz sinusoidal loads, cf. tables
8 and 12. Figure 43c/d is the analogue of figure 43a/b for DAC 0/16 E
(elastomer modified), cf. tables 10 and 13. Figure 43e/f is the analogue of
figure 43a/b for DAC 0/16 P (plastomer modified), cf. tables 11 and 14.
In figure 43a, J 10 increases with increasing frequency of the block-wave,
whereas J 10 decreases with increasing frequency of the sinusoidal wave.
In figures 43c and 43e, J 10 decreases with increasing frequency of the

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

1000

J1010^4

800
600

40C/0.1 MPa
40C/0.2 MPa
50C/0.1 MPa
50C/0.2 MPa

3.5

DAC 0/16

3
2.5

exponent z

DAC 0/16

J1010^4

1000
900
800
700
600
500
400
300
200
100
0

Ch. 3

40C/0.1 MPa
40C/0.2 MPa
50C/0.1 MPa
50C/0.2 MPa

1.5
1

0.5
0
block block sinus sinus 1
Hz
0.3 Hz 0.5 Hz 0.1 Hz 1Hz

block block sinus sinus 1


Hz
0.3 Hz 0.5 Hz 0.1 Hz 1 Hz

waveform/frequency
wave-form/frequency

wave-form/frequency
waveform/frequency

40C/0.1 MPa DAC 0/16 E


40C/0.2 MPa
50C/0.1 MPa
50C/0.2 MPa

3.5

DAC 0/16 E

3
2.5

exponent z

110

40C/0.1 MPa
40C/0.2 MPa
50C/0.1 MPa
50C/0.2 MPa

1.5

400
200

0.5
0

0
block block sinus sinus 1
Hz
0.3 Hz 0.5 Hz 0.1 Hz 1Hz

block block sinus sinus 1


Hz
0.3 Hz 0.5 Hz 0.1 Hz 1Hz

wave-form/frequency
waveform/frequency

J1010^4

800

40C/0.1 MPa
40C/0.2 MPa
50C/0.1 MPa
50C/0.2 MPa

DAC 0/16 P

600

3.5
3
2.5

exponent z

1000

wave-form/frequency
waveform/frequency

DAC 0/16 P

40C/0.1 MPa
40C/0.2 MPa
50C/0.1 MPa
50C/0.2 MPa

1.5

400
200

0.5
0

0
block block sinus sinus 1
Hz
0.3 Hz 0.5 Hz 0.1 Hz 1Hz

block block sinus sinus 1


Hz
0.3 Hz 0.5 Hz 0.1 Hz 1 Hz

wave-form/frequency
waveform/frequency

wave-form/frequency
waveform/frequency

Figure 43. J 10 and z as obtained for different waveforms of applied stress.


Different temperatures, 40C and 50C. Different stress amplitudes, 0.1 MPa
and 0.2 MPa. Two types of block-wave: 1 s/1.8 s loading time/rest-time
(frequency: 0.357 Hz, abbreviated as 0.3 Hz; 0.2 s/1.8 s (frequency: 0.5 Hz).
Two sinusoidal waves: 0.1 Hz and 1 Hz. Top left (a): J 10 for DAC 0/16,
unmodified. Top right (b): z for DAC 0/16, unmodified. Mid left (c): J 10 for
DAC 0/16 E (elastomer modified). Mid right (d): z for DAC 0/16 E. Bottom left
(e): J 10 for DAC 0/16 P (plastomer modified). Bottom right (d): z for DAC 0/16 P.

Sec. 7 Results of dynamic creep tests in the absence of confinement 111

block-wave, and decreases with increasing frequency of the sinusoidal


wave.
In figure 43b, z decreases with increasing frequency of the block-wave.
In the case of the sinusoidal wave it is not clear how z depends on the
frequency.
In figures 43d and 43f, z increases with increasing frequency of the
block-wave, and increases with increasing frequency of the sinusoidal
wave.
The following general trends can be observed irrespective of whether the
applied load is a block-wave or a sinus,
1 J 10 decreases as a function of increasing frequency;
2 z increases as a function of increasing frequency;
3 The unmodified mixtures show relatively low values of J 10 and
relatively large values of z in comparison to the polymer modified
mixtures.
The various cases show that J 10 and z differ for the different blockwaveforms and sinusoidal waveforms. The experimental evidence
supports the idea that the shape of the waveform of the applied stress
influences the creep.

7.5.2 Temperature and stress dependence of creep properties


Figure 43 also shows the temperature dependence of J 10 and z in the
temperature interval between 40C and 50C, and the stress dependence of
J 10 and z in the applied stress interval between 0.1 MPa and 0.2 MPa.
As for the temperature dependence, J 10 is greater than at 50C than at
40C (in all cases but one). This is an expected result, which can be based
on the fact that the compliance is known to increase as a function of
increasing temperature. In contrast, z is usually smaller at 50C than at
40C. There is no explanation for this experimental observation, except
this: The parallelism of the creep curves implies that z decreases if J 10
increases, and vice versa. In as much this explains the results, the
explanation is elaborated in 7.7. In some cases, it seems the scatter on the
data overshadows the dependence on the temperature in the interval
between 40C and 50C.
As for the stress dependence, J 10 is lower at 0.1 MPa than at 0.2 MPa
applied stress in 17 out of 23 cases in figure 43a/c/e. A probable
explanation is that the mixture responds with greater stiffness at the
greater stress amplitude.
In 18 out of 23 cases, a lower J 10 corresponds to a greater z in figure
43b/d/f. In some cases, it seems the scatter on the data overshadows the
temperature dependence in the interval between 0.1 MPa and 0.2 MPa.

112

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

Ch. 3

7.6 Comparison of dynamic creep to static creep


Results for unmodified DAC 0/16 in static creep, figure 7, may be
compared to results in dynamic creep, figures 1112 (block-wave loading)
and figure 33 (sinusoidal loading). Similarly, results for polymer modified
DAC 0/16 in static creep, figure 8, may be compared to results in dynamic
creep, figure 25 (block-wave loading) and figure 39 (sinusoidal loading).
Let us compare the results in figure 7a to those in figures 11-12. In figure
7a, J (10) varies between 0.01MPa-1 and 0.06 MPa-1, and the final creep
compliance varies roughly between 0.05 MPa-1 and 0.09 MPa-1. If in
figures 1112 the results for the waveform designated by 600 s/1.8 s are
not considered, then J (10) varies roughly between 0.001 MPa-1 and 0.02
MPa-1, and the final creep compliance varies roughly between 0.09 MPa-1
and 0.5 MPa-1. Thus, a crude comparison of the results for dynamic and
static creep shows that the creep curves are steeper in dynamic creep.
Fitting equation (15) to the creep of figure 7b and the creep of the 600
s/1.8 s block-wave of figures 1112, yielded the results of table 15.
Table 15. The logarithmic creep model, equation (15), fitted to the static
creep of figure 7b, and to the creep with the 600 s/1.8 s signal of
figures 1112.
R2
test condition
creep model
static
40C/0.2 MPa
J (t ) = 0.0161 + 0.0063 ln t
0.979
dynamic, 600 s/1.8 s
J (t ) = 0.0643 + 0.0141 ln t
40C/0.1 MPa
0.996
40C/0.2 MPa
J (t ) = 0.0577 + 0.0089 ln t
0.993
50C/0.1 MPa
0.997
J (t ) = 0.1181 + 0.0194 ln t
50C/0.2 MPa
J (t ) = 0.1455 + 0.0221 ln t
0.992

Choosing the logarithmic model, to fit the creep curve in figure 7b, we
obtain
J 1 = 0.0161
static case, 40C/0.2 MPa:
dynamic case, 40C/0.2 MPa:
J 1 = 0.0577
In the dynamic case, J 1 is predicted wrongly by equation (15), since the
creep compliance cannot be negative at 1 s. The explanation for this is,
that the creep is actually described by two different creep models: by
equation (16) for t < 10 s, and by equation (15) for t > 10 s. Therefore, a
simple comparison using equation (15) was not possible.
Using equation (21), the creep functions given in table 16 were obtained,
which are shown in figure 44. In figure 44/table 16, J 10 is a lot greater in
the static case than in the dynamic case, and z is considerably smaller in
the static case as compared to the dynamic case. It seems that a large value
of z does not necessarily indicate a large creep compliance, J (t ) . It
seems, that a large value of z is usually associated to a low value of J 10 ,

Sec. 7 Results of dynamic creep tests in the absence of confinement 113

Comparison of static creep and dynamic creep


with 600 s/1.8 s load signal

compliance J(t) (MPa^-1)

static: 40C/0.2 MPa


dynamic: 40C/0.1 MPa

0.1

dynamic: 40C/0.2 MPa


dynamic: 50C/0.1 MPa
0.01

dynamic: 50C/0.2 MPa

0.001
1

log time (s)

10

Figure 44. Comparison of static creep in figure 7b to dynamic creep with


the 600 s/1.8 s block-wave load of figures 1112.
Table 16. The power law creep model, equation (21), fitted to the
static creep of figure 7b, and to the creep with the 600 s/1.8 s blockwave of figures 11-12 (dynamic creep constants from table 8).
test condition
creep model: J (t ) = J 10 (log t ) z
R2
static
40C/0.2 MPa
J (t ) = 0.0307 (log t ) 0.6061
0.997
dynamic
0.999
40C/0.1 MPa
J (t ) = 0.0029 (log t ) 2.2579
1.000
40C/0.2 MPa
J (t ) = 0.00002 (log t ) 5.0293
4.0136
0.999
50C/0.1 MPa
J (t ) = 0.0002 (log t )
1.000
50C/0.2 MPa
J (t ) = 0.00004 (log t ) 5.3661

and vice versa, cf. table 8, figure 18, table 9, figure 20, tables 10-11,
figure 30, figure 34, table 12, and tables 13-14.
Table 16 clearly shows that the creep curves are steeper in dynamic creep.
Similarly, the results for the polymer modified mixtures may be
compared. The regression equation shown in figure 8d, J (t ) = 0.0137ln
(t) + 0.0398 is also described by J (t ) = 0.0691(log t)0.6043, R 2 = 0.995.
Comparison of the exponent, z = 0.6043, to the z -values in tables 13 and
14 shows, for the investigated mixtures, that the creep curve is steeper in
dynamic creep than in static creep for frequencies greater than 0.1 Hz.
Recall figure 43; comparison of figures 43a/b, 43c/d, and 43e/f shows that
J 10 is relatively low for the unmodified mixture as compared to the
polymer modified mixtures, whereas z is relatively large for the
unmodified mixture as compared to the polymer modified mixtures.
However, if the creep curves in figure 25 (dynamic creep) are compared to
those in figure 8 (static creep), then the creep compliance, J (t ) , does not

114

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

Ch. 3

seem to differ greatly, that is for for t > 10 s. Thus, typically, the creep
curve is steeper in dynamic creep than in static creep, but this does not
mean that the creep compliance, J (t ) , is greater in dynamic creep than in
static creep other conditions being equal.
It seems that the shape of the waveform of the applied stress also
influences the creep during the first 10 s. It seems that this initial creep is
relatively strong in static creep, and in dynamic creep at low frequency
using a sinusoidal load. That is: it seems as if much of the creep
capacity of the specimen (read: the grain skeleton) is used during the first
10 seconds, and that in that case less capacity is left for the creep after the
first 10 seconds. If this is true, as seems to be the case considering that this
result is reproducible, then it seems important to use a waveform of
applied stress that is representative for the waveform the material
experiences in the pavement.
The steeper creep curve in dynamic creep was a familiar result, and was
therefore expected. There are also some unexpected peculiarities in the
results shown in figures 7, 8, 11-12, and 25. An unexpected result was that
the polymer modified mixtures in static creep, figure 8, show on average a
greater final creep compliance, than the unmodified mixture in figure 7a.
This was unexpected, because it was known from experience that polymer
modified mixtures show improved resistance to permanent deformation
(rutting in the pavement). Also unexpected was that the plastomer
modified mixture shows a greater final creep compliance, figure 8c/d, than
the elastomer modified mixture, figure 8a/b. This was unexpected,
because the plastomer modified binder contained a harder base bitumen.
These unexpected results may be explained by the scatter on the data,
since this is considerable. That is, if the scatter bands in figures 7a and
8a..d are considered, then it seems there is a fair chance that two different
mixtures need not rank reproducibly if an average test result is based on
only three test repetitions.
The results obtained with the waveform 600 s/1.8 s shown in figures 11
12 are peculiar, because the creep compliance, J (t ) , needs a longer time
to exceed the value of 0.01 MPa 1; compare figures 1112 to figure 7. In
figure 7b (static creep), J (t ) reaches the value of 0.01 MPa-1 at
approximately 1 s, whereas in figures 11b and 12a/b (dynamic creep),
J (t ) attains the value of 0.01 MPa-1 after more than 100 s. The result for
the polymer modified mixtures in figure 25 is in qualitative agreement
with the results for the unmodified mixtures in figures 1112. However, in
figure 25, the creep curves obtained with the waveform 600 s/1.8 s are
situated closer to the curves obtained with other waveforms.

Sec. 7 Results of dynamic creep tests in the absence of confinement 115

7.7 The interrelatedness of J 10 and z of equation (21) and the parallelism


of the creep curves
As mentioned previously, cf. page 71, it seems that creep curves of the
same material do cross sometimes, presumably owing to bad specimens,
but as a rule do not cross. The results seem to indicate, that, if the material
quality is constant, the creep curves do not cross when the waveform of
the applied load changes, but shift parallel (horizontally along the time
axis). This is investigated in the following.
Let us assume that the creep curves in figures 1112, 19a, 23, 33, and 39
are (read: should be) parallel curves, and that deviations (crossings of
curves) can be attributed to the scatter, which, in turn, can be attributed to
the material heterogeneity.
As an example, recall the creep model from figure 13b,
(38)
1 s/1.8 s: J (t ) = 0.0028 (log t ) 2.95
This model is shown in figure 45a as a function of the time, plotted logarithmically (i.e. as function of log t on normal scale), and together with
the same curve shifted to (t / 3) , and the same curve shifted to (3 t ) . In
figure 45a, the curves are parallel. In figure 45b, the same curves are
shown as a function of log (time) plotted logarithmically (i.e. as function
of log log t on normal scale). In figure 45b, line B is a straight line, and
lines A and C are curvilinear. This is caused by the log-transformation.
Notice the wedge in figure 45b, which forms owing to the log-transformation of the parallel creep curves in figure 45a. Recall equation (21),
J (t ) = J 10 (log t ) z
(39a)
Equation (39a) can be written as
log J (t ) = log J 10 + z log (log t )
(39b)
Equation (39b) represents a straight line with slope z if J (t ) is plotted
versus log t on log-log scale (line B in figure 45b). For t = 10 s, J (t ) =
J 10 . Let us choose curve B as the creep master curve, so t is the time of
the creep master curve. Let us now write
J (t ) = J 10 (log t ) z
(40a)
t = aT t
(40b)
For curve A in figure 45, aT = 1/3, and for curve C, aT = 3. Figure 45a
shows that J (t ) = J (t ) , for different times, respectively t and t . For
example, J (10) is equal to 0.0028, and J A (10 / 3) and J C (30) are also
equal to 0.0028. Thus, if J (t ) = J (t ) and aT = 1/3, curve A in figure 45
can be obtained from curve B. Note, that t is the time of curve A.
Equation (40a) can be written as
log J (t ) = log J 10 + z log (log aT t )
(41)

Equation (41) represents a straight line with slope z if J (t ) is plotted


versus log aT t on log-log scale. Note, that curve A is plotted versus log t

116

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

compliance, J(t) (MPa^-1)

compliance, J(t) (MPa^-1)

Ch. 3

A
B
C

0.1

0.01

A
B
0.1

0.01

0.001
100000

time, t (s)

10000

1000

100

10

0.001

0.1

10

log time, t (s)

Figure 45. Left (a): The creep model from figure 13b, J (t ) = 0.0028 (log t ) 2.95 ,
shown as function of the time, plotted logarithmically, as curve B. Also shown
are the same curve shifted to (t / 3) , curve A, and the same curve shifted to
(3 t ) , curve C. Right (b): The same models shown as function of log (time),
plotted logarithmically.

J (t )
line B

t:

3.3

10

30

log t:

0.523

1.477

log log t:

-0.282

0.169

Figure 46. Illustration of the log-transformation. On t-scale, the abscissa differ


by a constant factor; on log t-scale, the
abscissa differ by a constant distance; on
log log t-scale, the abscissa are not
equidistant.

in figure 45b, i.e. in the time of curve B, and not versus log aT t , and
therefore deviates from the straight line. The curvature of curve A and
curve C is negligible in the domain where J (t ) J 10 ; in this example,
J 10 = 0.0028. J 10 and z can be determined by means of regression
analysis using equation (21). z might also be estimated as follows:
z A =

log J (10,000) log J (10)


log 0.1682 log 0.0028
=
= 2.147
log log (aT 10,000) log log (aT 10) log 3.5228 log 0.5228

Sec. 7 Results of dynamic creep tests in the absence of confinement 117

log J (10,000) log J (10)


log 0.1682 log 0.0028
=
= 2.9543
log log (10,000) log log (10)
log 4 log 1
log J (10,000) log J (10)
log 0.1682 log 0.0028
z C =
=
= 3.693
log log (10,000 / aT ) log log (10 / aT )
log 4.477 log 1.477
z B =

Thus, the creep models for lines A..C in figure 45b in the domain J (t )
0.0028 are:
J (t ) = 0.0099 ( log t )2.147
curve A:
curve B:
J (t ) = 0.0028 ( log t )2.954
curve C:
J (t ) = 0.0009 ( log t )3.693
The time dependence of the creep curve is shown by a parallel shift of the
creep curve along the time axis, cf. figure 45a. Semi-log transformation
yields a wedge of creep curves as in figure 45b. Similar wedges can be
seen in figures 1415, 19b, 29, 35, and 40. A parallel shift of the creep
curve to the left, i.e. to a lower time value causes J 10 to increase and z to
decrease. Conversely, a parallel shift of the creep curve to the right, i.e. to
a higher time value causes J 10 to decrease and z to increase.
A relatively large J 10 indicates that the asphalt mixture creeps relatively
fast initially, i.e. before t = 10 s. A relatively low z indicates that the
asphalt mixture creeps relatively slowly later, i.e. for t > 10 s. Therefore,
it seems that a greater importance must be attributed to the exponent of the
creep model, which means that for the purpose of material selection in
pavement design, a low creep susceptibility, z , is to be preferred. A large
z does not necessarily indicate a large final creep compliance, J (10,000)
in the creep test, cf. figure 43, since a large z is usually associated to a
low J 10 .
7.8 Change of volume of the specimen
The radial deformation was measured to be able to determine an estimate
of volumetric strain of the test specimen. The volumetric strain was
calculated using
V / V0 = 1 + 2 3
(42)
where V is the specimen volume, V0 is the initial specimen volume, 1 is
the axial deformation, and 3 is the radial deformation. There is a net
specimen volume increase as soon as the ratio 3 /| 1 | starts to become
greater than 0.5. The radial deformation was measured by means of a
displacement transducer at half the height of the specimen. Table 17
summarises values of the axial and radial deformations obtained for
asphalt mixtures and testing conditions. The values in table 17 are
indicative. The reason is that the specimens showed barrelling. This means
that the radial deformation was not constant over the specimen height.
Consequently, the measured radial deformation is the maximum value.
The volume change is overestimated if it is assumed that the measured

118

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

Ch. 3

Table 17. Axial deformation, 1 , radial deformation, 3 , ratio of radial


deformation and axial deformation, 3 /| 1 |, and volumetric strain, V = 1 +
2 3 , obtained for stone mastic asphalt, SMA, dense graded asphalt concrete
unmodified, DAC/U, dense graded asphalt concrete, elastomer modified,
DAC/E, and dense graded asphalt concrete, plastomer modified, DAC/P, after
7,200 s loading (total testing time) under different testing conditions.
Temperatures: 40C, 50C. Applied stress: 0.1 MPa, 0.2 MPa, frequencies in
sinusoidal loading: 0.1 Hz, 1 Hz, or 10 Hz. Loading time, t l : 0.05, 0.2, 1.0, or
600 s. In all cases the rest-time was constant, 1.8 s. Creep tests performed in the
absence of confinement.
SMA
sinusoidal load
block-wave load
f
1
3 3 /| 1 | V
tl
- 1
3 3 /| 1 | V
(Hz) (%) (%)
(%) (s) (%) (%)
(%)
0.1
4.52 3.67 1.23 +2.8 0.05 2.86 1.54 0.54 +0.2
40C/
2.74 1.78 0.65 +0.8 0.2 3.34 2.49 0.75 +1.6
1
0.1 MPa
10
1.55 1.14 0.74 +0.7 1.0 4.36 3.22 0.74 +2.1
0.1
6.38 6.23 0.98 +6.1 0.05 5.64 4.81 0.85 +4.0
40C/
4.12 3.34 0.81 +2.6 0.2 6.49 5.80 0.89 +5.1
1
0.2 MPa
10
2.10 2.07 0.99 +2.0 1.0 6.62 5.56 0.84 +4.5
0.1
5.71 5.86 1.03 +6.0 0.05 4.04 3.29 0.81 +2.5
50C/
3.59 2.68 0.75 +1.8 0.2 8.41 6.71 0.80 +5.0
1
0.1 MPa
10
1.59 1.24 0.78 +0.9 1.0 8.08 6.87 0.85 +5.7
0.1
6.37 6.82 1.07 +7.3 0.05 7.29 6.76 0.93 +6.2
50C/
5.73 5.43 0.95 +5.1 0.2 7.41 6.40 0.86 +5.4
1
0.2 MPa
10
2.37 1.98 0.84 +1.6 1.0 8.06 7.45 0.92 +6.8
DAC/U

sinusoidal load
f
1
3 3 /| 1 |
(Hz) (%) (%)
0.1
3.10 2.57 0.83
40C/
1.31 0.83 0.64
1
0.1 MPa
0.77 0.53 0.69
10
0.1
12.5 7.66 0.61
40C/
2.58 1.96 0.76
1
0.2 MPa
10
1.16 0.94 0.81
0.1
7.00 7.57 1.08
50C/
2.85 2.41 0.84
1
0.1 MPa
10
1.54 1.01 0.66
0.1
6.98 7.23 1.04
50C/
7.64 7.42 0.97
1
0.2 MPa
10
2.51 2.35 0.95
This table continues on the next page.

block-wave load

V
(%)
+2.0
+0.4
+0.3
+2.8
+1.3
+0.7
+8.1
+2.0
+0.5
+7.4
+7.4
+2.4

tl
(s)
0.2
1.0
600
0.2
1.0
600
0.2
1.0
600
0.2
1.0
600

- 1

(%)
2.54
1.71
0.65
3.53
4.60
0.46
4.65
4.80
0.56
6.85
6.90
1.10

3 3 /| 1 |
(%)
1.60 0.63
2.00 1.17
0.51 0.78
2.85 0.81
3.85 0.84
0.49 1.07
4.25 0.91
5.05 1.05
0.52 0.93
5.40 0.79
6.40 0.93
1.14 1.04

V
(%)
+0.6
+2.3
+0.4
+2.2
+3.1
+0.5
+3.9
+5.3
+0.5
+4.0
5.9
+1.2

Sec. 7 Results of dynamic creep tests in the absence of confinement 119

Table 17 Contd.
DAC/E

40C/
0.1 MPa
40C/
0.2 MPa
50C/
0.1 MPa
50C/
0.2 MPa

f
(Hz)
0.1
1
10
0.1
1
10
0.1
1
10
0.1
1
10

sinusoidal load
1
3 3 /| 1 |
(%) (%)
1.07 0.64 0.60
1.06 0.44 0.42
0.73 0.31 0.42
1.63 0.77 0.47
1.37 0.74 0.54
1.20 0.67 0.56
1.22 0.70 0.57
1.36 0.63 0.46
0.73 0.31 0.42
2.08 1.31 0.63
1.92 0.89 0.46
1.39 0.76 0.55

V
(%)
+0.2
+0.2
+0.1
-0.1
+0.1
+0.1
+0.2
-0.1
-0.1
+0.5
-0.1
+0.1

tl
(s)
0.2
1.0
600
0.2
1.0
600
0.2
1.0
600
0.2
1.0
600

f
(Hz)
0.1
1
10
0.1
1
10
0.1
1
10
0.1
1
10

sinusoidal load
1
3 3 /| 1 |
(%) (%)
1.45 0.80 0.55
1.68 1.06 0.63
1.13 0.53 0.47
2.85 1.76 0.62
1.72 1.08 0.63
1.77 0.84 0.47
2.63 1.64 0.62
1.95 1.10 0.56
1.07 0.66 0.62
4.91 4.68 0.95
2.63 1.85 0.70
2.29 1.24 0.54

V
(%)
+0.2
+0.4
-0.1
+0.7
+0.4
-0.1
+0.7
+0.3
+0.3
+4.5
+1.1
+0.2

tl
(s)
0.2
1.0
600
0.2
1.0
600
0.2
1.0
600
0.2
1.0
600

DAC/P

40C/
0.1 MPa
40C/
0.2 MPa
50C/
0.1 MPa
50C/
0.2 MPa

block-wave load
- 1

(%)
0.94
1.32
0.92
1.42
1.93
1.17
1.14
1.24
1.06
1.98
2.02
1.38

3 3 /| 1 |
(%)
0.25 0.27
0.52 0.39
0.32 0.35
0.64 0.45
0.71 0.37
0.66 0.56
0.45 0.39
0.52 0.42
0.54 0.51
1.04 0.53
0.86 0.43
0.92 0.67

V
(%)
-0.4
-0.3
-0.3
-0.1
-0.5
+0.2
-0.2
-0.2
+0.0
+0.1
-0.3
+0.5

block-wave load
- 1

(%)
1.65
1.75
1.38
1.89
2.79
2.41
2.12
2.47
1.42
3.14
3.92
2.36

3 3 /| 1 |
(%)
0.92 0.56
0.59 0.34
0.78 0.57
1.18 0.62
1.61 0.58
1.37 0.57
1.11 0.52
1.53 0.62
0.74 0.52
2.24 0.71
2.17 0.56
1.93 0.82

V
(%)
+0.1
-0.6
+0.2
+0.5
+0.4
+0.3
+0.1
+0.6
+0.1
+1.3
+0.4
+1.5

maximum radial deformation is constant over the specimen height. The


V -values can be used to have a relative indication of the volume change.
7.9 Barrelling
Figure 47 shows the radial deformation as function of the loading time.
The left most three bars belong to the static test in which the specimen was
loaded during 3,600 s. Figure 47 shows the radial deformation on the
upper section, the middle section, and the lower section of the test
specimen. Figure 47 illustrates that barrelling occurred during the creep
test in the absence of confinement (using the friction reduction system that
was used at the time the tests were performed), and that it was at
maximum for a loading time of 1 s (rest-time: 1.8 s).

120

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

Ch. 3

DAC 0/16, 40C/0.2 MPa


radial deformation (%)

6
5

upper part
middle part

lower part

3
2
1
0

3600 s

600 s

10 s

1s

0.2 s

0.05 s

loading time (s)

Figure 47. Barrelling of dense graded asphalt concrete in


the static creep test (extreme left) and in the dynamic
creep test in the absence of confinement.

8 Results of dynamic creep tests in the presence of confinement


Dependence of creep properties on time, stress, specimen height,
and mixture composition
The purpose of the investigation described in 6, 7, and 8 was to gain
insight in the dependence of the creep of asphalt concrete on time, stress,
and mixture composition (Molenaar and Molenaar 2000). In this section,
also the influence of the specimen height is investigated. The results of
dynamic creep tests in the presence of confinement are discussed.
Different waveforms of the applied stress were used to investigate the
time dependence. The time dependence is shown at constant applied stress
amplitude. Measurements were performed at two temperatures, 40C and
50C. Different stress amplitudes were applied to investigate the stress
dependence. The stress dependence is shown at constant time, or
frequency, and temperature. Three specimens were used to investigate the
influence of the specimen height; a 60 x 100 mm height x diameter
specimen, a 100 x 100 mm height x diameter specimen, and a 200 x 100
mm height x diameter specimen. Two asphalt mixtures were used to
investigate the influence of the mixture composition; dense graded asphalt
concrete, DAC 0/16, and porous asphalt, PA 0/16. The mixtures contained
pure, unmodified bitumen, and are indicated in table 2, sections F..H. All
tests were performed at 50C.
8.1 Creep susceptibility of dense graded asphalt concrete Transition
of creep according to eq. (15) to creep according to eq. (16)
In figure 48, the creep compliance, J ( t ) = p ( t ) / p , using the 200 x 100
mm height x diameter specimen is shown as function of the time. Each

Sec. 8

Results of creep tests in the presence of confinement

121

curve represents an average of three tests. The combinations of applied


axial stress, 1 , and applied radial stress, 3 , are indicated in the legend
as 1 / 3 . From figure 48 and table 18, it seems that between q / p =
0.43 and q / p = 0.95 the creep transforms from the behaviour according
to equation (15) to the behaviour according to equation (16). The actual
transition point (q / p) c is not known, and the values of J 1 , ~
z and z in
the transition point are not known. However, let us assume, as an example,
there is a transition point, (q / p) c , and that, in this example, it is given by
(q / p) c = 0.43. Let us assume that the creep is described by equation (15)
for q / p (q / p) c , and that the creep is described by equation (16) for
q / p > (q / p) c . Moreover, let us assume that the slope of equation (15)
in a log-log plot must reach that of equation (16) at a given time, t , before
the creep can transform. Thus, in the transition point, the following
condition might be assumed,
d ln J (t ) ~
d
=z =
ln (J 1 + z ln t )
(43)
d ln t
d ln t
Differentiating,
1
d
~
(J 1 + z ln t )
z=

J 1 + z ln t d ln t
Hence,
z
~
z=
(44)
J 1 + z ln t
z~
z J1
1 J
ln t = ~ 1 =
(45)
z
z
z~
z

Using the following creep models from table 18,


J (t ) = 0.0010 t 0.304
(46)
J (t ) = 0.0001 + 0.0004 ln t
(47)
~
and z = 0.304, J 1 = 0.0001, and z = 0.0004, it follows that
0.0004 0.304 0.0001
t = exp
(48)
= 20.9 s
0.304 00004

If the transition point is in (q / p) c = 0.43, then equation (44), predicts the


slope on log-log scale of the creep according to equation (15), in this
example, at t = 20.9 s. Thus it has been made plausible that z of equation
(15) and ~z of equation (16) are related in the transition point.
Remark
If equation (43) is used with the 10log, then care must be taken to use the
correct transformation to the 10log. The correct transformation is the
following:
z J1
1 J1 z ~
=
ln t = ~
(49a)
z
z
z~
z

122

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

Ch. 3

z
~
z J1
2
.
3026
(49b)
2.3026 log t =
z
~
z
2.3026
z
~
z J1
log t = 2.3026 ~
(correct formula)
(49c)
z z
A scaling error is introduced if the relationship is written directly using the
10
log,
d log J (t ) ~
d
=z =
log ( J 1 + z log t )
49d)
d log t
d log t

yielding,
z~
z J1
log t =
z~
z

(incorrect formula)

(49e)

8.2 Creep susceptibility of dense graded asphalt concrete and porous


asphalt Time and stress dependence
Figure 4950 show the creep compliance, J ( t ) = p ( t ) / p , as obtained
using the 60 x 100 mm specimen as function of the time. Figure 49 shows
the results for dense asphalt concrete, DAC 0/16, unmodified, and figure
50 shows the results for porous asphalt, PA 0/16, unmodified. Each curve
represents an average of three tests. The applied axial stress, 1 , is given
in the legend in MPa. In figures 4950 it can be seen that the creep obeys
the creep law according to equation (16) only if the radial stress is equal to
0 MPa. These cases are indicated by means of square shaded markers. In
all other cases shown, the creep obeys the creep law according to equation
(15). A complete list of models is given in appendix 3.

Results of creep tests in the presence of confinement

DAC 0/16, block-wave, 1 Hz

1.000

0.100

0.010

0.001
1

10

100

1000 10000

123

DAC 0/16, block-wave, 1 Hz

0,5

exponent z of eq. (16)

compliance J(t) (MPa^-1)

Sec. 8

0,45

z = 0.0713 q/p + 0.2353

0,4

0,35

time (s)
0,9/0,3 MPa
0,3/0,2 MPa
0,3/0,03 MPa
0,6/0,05 MPa

0,6/0.25 MPa
0,6/0,1 MPa
0,75/0,15 MPa
0,6/0,15 MPa

0,3
0

0,5

1,5

2,5

q/p

Figure 48. Dynamic creep of dense graded asphalt concrete, DAC 0/16. Left (a):
Each curve represents the average of three tests. The waveform of the applied
load was a block-wave with a loading time of 0.2 s and a rest-time, of 0.8 s.
Load control: pneumatic. Temperature: 50C. Specimen: 200 x 100 mm height x
diameter. In the legend, combinations of axial stress and radial stress in MPa
are given; see also table 18. Right (b): Exponent ~z of eq. (16) as function of the
quotient of the deviatoric stress, q, and the volumetric stress, p.
Table 18. Creep models belonging to the creep curves in figure 48.
1 / 3
creep model
q/ p
R2
99.9
0.6/0.05
J (t ) = 0.0032 t 0.4255
2.36
0.3673
99.7
0.3/0.03
2.25
J (t ) = 0.0040 t
99.4
0.6/0.1
1.88
J (t ) = 0.0029 t 0.3648
0.3750
99.7
0.75/0.15
1.71
J (t ) = 0.0020 t
0.3363
99.6
0.6/0.15
1.5
J (t ) = 0.0020 t
0.3192
99.1
0.9/0.3
1.2
J (t ) = 0.0010 t
0.3040
99.4
0.6/0.25
0.95
J (t ) = 0.0010 t
99.1
0.3/0.2
0.43
J (t ) = 0.0001 + 0.0004 ln t

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

DAC, block-wave, 1 Hz

compliance J(t) (MPa^-1)

1.000

y = 0.0841x0.2171
R2 = 0.9784

0.100

compliance J(t) (MPa^-1)

124

DAC, block-wave, 1 Hz

1.000

y = 0.0063 Ln (x) + 0.0055


R2 = 0.9997

0.100

0.010

0.010

y = 0.0097 Ln (x) + 0.0098


R2 = 0.9984

y = 0.0006 Ln (x) + 0.0006


R2 = 0.9902

0.001

0.001
1

10

100

1000

10000

10

0.50

0.70

0.30

0.90

Figure 49. Creep compliance of dense


asphalt
concrete,
DAC
0/16,
unmodified, in the triaxial creep test.
J (t ) versus the time, t , plotted
logarithmically. Each creep curve
represents the average of three tests. In
y (x) ,
y
regression
equations
represents J (t ) , and x represents the
time, t . Waveform of applied load: 1
Hz block-wave, loading time: 0.2 s,
rest-time: 0.8 s. Applied axial stress
amplitude: see legend, values in MPa.
Temperature: 50C. Specimen: 60 x
100 mm height x diameter. Load
control: hydraulic. Top left (a):
Applied radial stress (MPa): 3 = 0
for the shaded markers, and 3 = 0.1
for the open markers. Top right (b):
Radial stress: 0.2 MPa. Bottom (c):
Radial stress: 0.3 MPa.

1.000

compliance J(t) (MPa^-1)

0.30

100

1000

10000

time (s)

time (s)
0.10

Ch. 3

0.50

0.70

0.90

DAC, block-wave, 1 Hz

y = 0.0036 Ln (x) + 0.011


R2 = 0.9954
0.100

0.010

y = 0.0004 Ln (x) + 0.0057


R2 = 0.9963
0.001
1

10

100

1000

10000

0.80

0.90

time (s)
0.40

0.50

0.60

0.70

Sec. 8

Results of creep tests in the presence of confinement

1.000

1.000

DAC, sinus, 1 Hz

125

DAC, sinus 1 Hz

y = 0.1024x
R2 = 0.9891

0.100

0.010

y = 0.0063 Ln (x) + 0.0259


R2 = 0.9992

compliance J(t) (MPa^-1)

compliance J(t) (MPa^-1)

0.1939

y = 0.0032 Ln (x) + 0.0155


R2 = 0.9954

0.100

0.010
y = 0.0005 Ln (x) + 0.0111
R2 = 0.9409
0.001

0.001

10

100

1000

10

10000

0.50

0.70

0.30

0.90

Figure 49 (Contd). Creep compliance of dense asphalt concrete, DAC


0/16, unmodified, in the triaxial creep
test. J (t ) versus the time, t , plotted
logarithmically. Each creep curve
represents the average of three tests.
In regression equations y ( x) , y
represents J (t ) , and x represents the
time, t . Waveform of the applied
load: 1 Hz sinus. Applied axial stress
amplitude: see legend, values in MPa.
Temperature: 50C. Specimen: 60 x
100 mm height x diameter. Load
control: hydraulic. Top left (d):
Applied radial stress (MPa): 3 = 0
for the shaded markers, and 3 = 0.1
for the open markers. Top right (e):
Radial stress: 0.2 MPa. Bottom (f):
Radial stress: 0.3 MPa.

0.50

1,000

compliance J(t) (MPa^-1)

0.30

1000 10000

time (s)

time (s)
0.10

100

0.70

0.90

DAC, sinus 1 Hz

y = 0.0011 Ln (x) + 0.0117


R2 = 0.9927

0,100

0,010

y = 0,0009Ln(x) + 0,008
R2 = 0,9655
0,001
1

10

100

1000

10000

time (s)
0,40

0,50

0,70

0,90

126

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture
PA, block-wave, 1 Hz

1.000

Ch. 3

PA, block-wave, 1 Hz

y = 0.0448x0.5511
R2 = 0.9847
0.100

0.010
y = 0.007Ln(x) + 0.0453
R2 = 0.9794

compliance J(t) (MPa^-1)

compliance J(t) (MPa^-1)

1.000

y = 0.0026 Ln (x) + 0.0208


R2 = 0.9981

0.100

0.010

y = 0.0018 Ln (x) + 0.0026


R2 = 0.9994
0.001
10

0.001
10

100

1000

100

10000

1000

10000

time (s)

time (s)
0.30

0.50

0.70

0.90

Figure 50. Creep compliance of


porous asphalt, PA 0/16, unmodified,
in the triaxial creep test. J (t ) versus
the time, t , plotted logarithmically.
Each creep curve represents the
average of three tests. In regression
equations y ( x) , y represents J (t ) ,
and x represents the time, t .
Waveform of applied load: 1 Hz
block-wave, loading time: 0.2 s, resttime: 0.8 s. Applied axial stress
amplitude: see legend, values in MPa.
Temperature: 50C. Specimen: 60 x
100 mm height x diameter. Load
control: hydraulic. Top left (a):
Applied radial stress (MPa): 3 = 0
for the shaded markers, and 3 = 0.1
for the open markers. Top right (b):
Radial stress: 0.2 MPa. Bottom (c):
Radial stress: 0.3 MPa.

0.30

0.50

1.000

compliance J(t) (MPa^-1)

0.10

0.70

0.90

PA, block-wave, 1 Hz

y = 0.0014 Ln (x) + 0.0164


R2 = 0.9988

0.100

0.010
y = 0.0001 Ln (x) + 0.0072
R2 = 0.8122
0.001
10

100

1000

10000

time (s)

0.40

0.50

0.70

0.90

Sec. 8

Results of creep tests in the presence of confinement

1.000

PA, sinus, 1 Hz

127

PA, sinus, 1 Hz

y = 0.0903x0.3755
R2 = 0.9966
0.100

0.010
y = 0.0021 Ln (x) + 0.0157
R2 = 0.9637

compliance J(t) (MPa^-1)

compliance J(t) (MPa^-1)

1.000

y = 0.0023 Ln (x) + 0.0141


R2 = 0.9992

0.100

0.010
y = 0.0004 Ln (x) + 0.0074
R2 = 0.9895

0.001
1

10

100

1000

10000

0.001

time (s)

10

100

1000

10000

time (s)
0.30

0.50

0.70

0.90

Figure 50 (Contd). Creep compliance of porous asphalt, PA 0/16,


unmodified, in the triaxial creep test.
J (t ) versus the time, t , plotted
logarithmically. Each creep curve
represents the average of three tests.
In regression equations, y ( x) y
represents J (t ) , and x represents
the time, t . Waveform of applied
load: 1 Hz sinusoidal load. Applied
axial stress amplitude: see legend,
values in MPa. Temperature: 50C.
Specimen: 60 x 100 mm height x
diameter. Load control: hydraulic.
Top left (d): Applied radial stress
(MPa): 3 = 0 for the shaded
markers, and 3 = 0.1 for the open
markers. Top right (e): Radial
stress: 0.2 MPa. Bottom (f): Radial
stress: 0.3 MPa.

0.30

0.50

1.000

0.70

0.90

PA, sinus, 1 Hz

compliance J(t) (MPa^-1)

0.10

0.100
y = 0.0013 Ln (x) + 0.0095
R2 = 0.999

0.010
y = 0.0004 Ln (x) + 0.0072
R2 = 0.855

0.001
1

10

100

1000

10000

time (s)
0.40

0.50

0.70

0.90

128

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

exponent, z10^4

150

y = 29.312x1.6396
R2 = 0.976

100
50

0
0.00

p=0.37

1.00
p=0.17
p=0.4

1000

exponent, z10^4

DAC 0/16, block-wave, 1 Hz

200

q/p 2.00

p=0.23
p=0.43

p=0.3
p=0.47

p=0.33
p=0.5

DAC 0/16, block-wave, 1 Hz

100

10

y = 29.312x1.6396
R2 = 0.976

1
0.10

3.00

Ch. 3

1.00

10.00

q/p
p=0.37

p=0.17
p=0.4

p=0.23
p=0.43

p=0.3
p=0.47

p=0.33
p=0.5

Figure 51. Exponent z of equation (15) as function of the quotient of the deviatoric stress, q , and the volumetric stress, p . The legend shows the volumetric
stress in MPa. DAC 0/16, block-wave form of applied stress. Hydraulic load
control. (a): On linear scale. Right (b): On log-log scale. In the regression
equations, y(x), y represents z104, and x represents q/p.
DAC 0/16, sinus, 1 Hz

80
60
40

y = 18.374x1.514
R2 = 0.9157

20
0
0.00

1.00

p=0.37

p=0.17
p=0.43

q/p

2.00

p=0.23

3.00

DAC 0/16, sinus, 1 Hz

100

exponent, z10^4

exponent, z10^4

100

10
y = 18,374x1,514
R2 = 0,9157
1
0,10

p=0.3
p=0.3

1,00
p=0.17
p=0.37

q/p

10,00

p=0.23
p=0.43

Figure 52. Exponent z of equation (15) as function of the quotient of the deviatoric stress, q , and the volumetric stress, p . The legend shows the volumetric
stress in MPa. DAC 0/16. Sinusoidal applied stress. Hydraulic load control. Left
(a): On linear scale. Right (b): On log log scale. In the regression equations,
y(x), y represents z104, and x represents q/p.

Figures 51-54 show the exponent of equation (15), z , plotted as function


of q / p . z is given on metric scale and on log-log scale to show that
z ( q / p ) can be described by a power function. In figure 54a, for PA
0/16, sinus, possibly the exponential function provides a better fit9.
Let us consider the functions z ( q / p ) in figures 51-54,
(50)
DAC 0/16, block-wave: z = 29.312 (q / p )1.6396
1.5140
DAC 0/16, sinus:
z = 18.374 ( q / p)
(51)
9

It is known that porous asphalt is susceptible to creep in the absence of


confinement pressure. Therefore, it seems improbable that the power function
correctly predicts the exponent if q / p approaches the value 3 (i.e. 3 = 0 ).

Sec. 8

Results of creep tests in the presence of confinement

120

PA 0/16, block, 1 Hz

exponent, z10^4

exponent, z10^4

100
1.7515

80

y = 17.035x
R2 = 0.8654

40
20

1.00

p=0.33

p=0.17
p=0.37

PA 0/16, block, 1 Hz

100

60

0
0.00

1000

129

q/p

2.00

p=0.23
p=0.43

10
1.7515

y = 17.035x
R2 = 0.8654
1
0.10

3.00
p=0.3
p=0.5

p=0.33

1.00
p=0.17
p=0.37

q/p

p=0.23
p=0.43

10.00
p=0.3
p=0.5

Figure 53. Exponent z of equation (15) as function of the quotient of the deviatoric stress, q , and the volumetric stress, p . The legend shows the volumetric
stress in MPa. DAC 0/16, block-wave form of applied stress. Hydraulic load
control. (a): On linear scales. Right (b): On log-log scales. In the regression
equations, y(x), y represents z104, and x represents q/p.
PA 0/16, sinus, 1 Hz
y = 2.578e1.4067x
R2 = 0.9587

150
100
50
0
0.00

y = 13.41x1.3299
R2 = 0.9166

1.00

2.00

3.00

q/p
p=0.37

p=0.17

p=0.23

p=0.3

p=0.33

p=0.4

p=0.43

p=0.47

p=0.5

PA 0/16, sinus, 1 Hz

100

exponent, z10^4

exponent, z10^4

200

10

y = 13.41x1.3299
R2 = 0.9166
1
0.10

p=0.37

1.00
p=0.17
p=0.4

p=0.23
p=0.43

q/p
p=0.3
p=0.47

10.00
p=0.33
p=0.5

Figure 54. Exponent z of equation (15) as function of the quotient of the deviatoric stress, q , and the volumetric stress, p . The legend shows the volumetric
stress in MPa. DAC 0/16. Sinusoidal applied stress. Hydraulic load control. Left
(a): On linear scale. Right (b): On log log scale. In the regression equations,
y(x), y represents z104, and x represents q/p.

PA 0/16, block-wave: z = 17.035(q / p )1.7515


(52)
PA 0/16, sinus:
z = 2.578exp (1.4067 q / p)
(53)
If figure 51a, equation (50) and figure 53a, equation (52) are compared, it
can be seen that z for PA 0/16 is smaller than z for DAC 0/16 in the
range between q / p = 0, i.e. 1 = 3 , and q / p = 3, i.e. 3 = 0.
Comparing figure 52a, equation (51) and figure 54a, equation (53), z
seems to become greater for PA 0/16 for q / p 2.2. In other words, if the
radial stress is small, then PA 0/16 is more susceptible to creep (permanent
deformation, or rutting in the pavement) than DAC 0/16. A minimal radial
stress (confinement pressure) is needed to let PA 0/16 show greater

130

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

Ch. 3

DAC 0/16, block-wave, 1 Hz


z=100

0.80

97

0.70
q=3p

q (MPa)

0.60

85

z=50
36
z=35

45

0.50

25

72

0.40

37
z=15

32

0.30
54

0.20

14

15

z=5

0.10
0.00
0.00

z=70
63

4
0.10

0.20

0.30

0.40

0.50

p (MPa)

DAC 0/16, sinus, 1 Hz


z=65

0.80

63

0.70
q=3p

q (MPa)

0.60

z=15

28
41

0.40

z=30

56

0.50

11

z=10

15

0.30
41

0.20

z=5

0.10
0.00
0.00

z=40
32

0.10

0.20

0.30

0.40

0.50

p (MPa)

Figure 55. Contour lines of constant z 104 in a diagram of the deviatoric stress
q versus the volumetric stress p . Values next to markers indicate measured
average values of z for the given stress condition ( p, q ) . Values of z 104 near
the dashed lines are indicative, owing to the scatter on the measured values of
z 104, cf. figures 49-52. Specimen height: 60 mm. Top (a): DAC 0/16, blockwave load, 1 Hz. Bottom (b): DAC 0/16, sinusoidal load, 1 Hz. The line q = 3 p
represents the states of stress accessible in simple creep, i.e. in the absence of
confinement.

resistance to creep than DAC 0/16, which then corresponds to a greater


resistance to permanent deformation in the pavement, in conformity with
the behaviour observed in the pavement.
In figure 55, the dashed lines represent contour lines of constant z 104. If
indeed z can be described as a function of q / p , as indicated by figures
51-54, then contour lines of constant z must be straight lines through the
origin in a diagram of q versus p , similar to the dashed straight lines in
figure 55. The lines must be straight lines, because only a straight line in a
q - p -diagram can represent an ensemble of stress states [ p, q ] such that

Sec. 8

Results of creep tests in the presence of confinement

131

PA 0/16, block-wave, 1 Hz
z=75
70

0.80
0.70

q=3p

q (MPa)

0.60

64

z=30
15
z=20

27

0.50

14

48

0.40

z=10

18

0.30
37

0.20

z=5

10

0.10
0.00
0.00

z=50
26

1
0.10

0.20

0.30

0.40

0.50

p (MPa)

PA 0/16, sinus, 1 Hz
z=70

0.80

65

0.70
q=3p

q (MPa)

0.60

23

33

z=15

17
11

27

0.40

z=20
13

0.50

z=10

11

0.30
21

0.20

z=5

0.10
0.00
0.00

z=30

4
0.10

0.20

0.30

0.40

0.50

p (MPa)

Figure 55 (Contd). Top (c): PA 0/16, block-wave load, 1 Hz. Bottom (d): PA
0/16, sinusoidal load, 1 Hz.

is constant; this is required if z is a function of q / p . Figure 55 seems to


confirm this expectation, although it must be admitted that the
confirmation could have been better, had the scatter on the data in figures
5154 been smaller.
8.3 Change of volume of the specimen
The change of volume of the specimen was determined by measuring the
specimen volume before and after the test. The specimen volume was
determined from measurements of the height and the diameter of the
specimen using Vernier callipers. The height was determined as the
average of four equidistant measurements. The diameter was determined
as the average of four or five equidistant measurements. The results are
shown in figure 56. It can be observed that the volume change is usually
negative for q / p < 2.2, whereas there seems to be a fair chance that the
volume change is positive for q / p > 2.2

132

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

Ch. 3

volume change (%)

volume change of specimen


DAC 0/16
1.5
1
0.5
0
-0.5
-1
-1.5
-2
-2.5
0

q/p
200 mm block P

60 mm block H

60 mm sinus

volume change (%)

volume change of specimen


PA 0/16
5
4
3
2
1
0
-1
-2
-3
0

0.5

1.5

2.5

q/p
200 mm block P

60 mm block H

60 mm sinus

Figure 56. Volume change of the specimen in the triaxial creep test. 200 mm
block P: specimen height: 200 mm, load signal: 1 Hz block-wave (loading time
0.2 s, rest-time 0.8 s), load control: pneumatic; 60 mm block H: specimen
height: 60 mm, load signal: 1 Hz block-wave (loading time 0.2 s, rest-time 0.8
s), load control: hydraulic; 60 mm sinus: specimen height: 60 mm, load signal:
1 Hz sinusoidal, load control: hydraulic. Top (a): Dense graded asphalt
concrete, DAC 0/16. Bottom (b) Porous asphalt, PA 0/16.

8.4 Dependence on the specimen height


There is experimental evidence which shows that the creep properties
depend on the height of the specimen.
With the 200 x 100 mm specimen, figure 48 and table 18, different models
were found to describe the creep curves, depending on the testing
conditions. For q / p > 0.95, the creep is described by equation (16), cf.
table 18. Figure 48b shows that ~z ( q / p ) follows a linear relationship for
q / p > 0.95.
With the 60 x 100 mm specimen, the creep curves were found to be
described by the logarithmic creep model according to equation (15), cf.

Sec. 8

Results of creep tests in the presence of confinement

0.100

133

compliance J(t) (MPa^-1)

DAC, block 1 Hz, 50C

0.7MPa (60 mm)


0.9 MPa (60 mm)
0.7 MPa (100 mm)
0.9 MPa (100 mm)

0.010
1

10

100

1000

10000

time (s)

Figure 57. Comparison of the creep compliance obtained with


different specimen heights other conditions being equal. Radial
stress: 0.1 MPa. In the legend, the axial stress is indicated.

figures 4950. Figures 5154 show the exponent, z , of equation (15) as


function of q / p . Figures 5154 show that z ( q / p ) follows a power law
relationship.
Notice, that the relationship z ( q / p ) differs entirely for the 200 x 100 mm
specimen and the 60 x 100 mm specimen. So, it seems, the creep
properties depend strongly on the specimen geometry. Geometrydependent mechanical properties are not predictive for the mechanical
behaviour in a different geometry. That is: geometry-dependent
mechanical properties cannot be considered true (geometry-independent)
material properties. If creep properties are indeed geometry-dependent,
they cannot be considered predictive for the creep behaviour in the
pavement.
In figure 57, the creep compliance, J ( t ) , as function of the time, is
compared for the 60 x 100 mm height x diameter specimen and the 100 x
100 mm height x diameter specimen. Each curve represents the average of
three tests. The applied axial stress, 1 , is indicated, respectively 0.7 MPa
and 0.9 MPa. Clearly, the compliance of the 100 x 100 mm specimen is
greater than that of the 60 x 100 mm specimen, other conditions being
equal. The specimens were manufactured in the same manner10, and tested
in the same apparatus under the same testing conditions (hydraulic load
control).

10

Comparison of the average densities of the slabs from which the 60 mm


specimens and the slabs from which the 100 mm specimen revealed no
significant difference in compaction.

134

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

Ch. 3

9 Results and interpretation of finite element computations of the


creep of a heterogeneous visoelasto-viscoplastic creep specimen
Finite element computations of the creep of a heterogeneous specimen
were conducted, using the FEMMASSE finite element code owned by
Intron (Verburg et al. 1995). The analysis was conducted because the
stress state of the creep specimen is three-dimensional, and the stress strain
behaviour of the material is nonlinear. Under those conditions, the
property determined from the test is expected to depend on the stress
distribution in the material, which, in turn, depends on the specimen
geometry. This makes it probable that the property determined from the
test is not a true material property. The analysis and the outcome are
described in the following.
The material was assumed to be a two-dimensional heterogeneous
granular material containing air voids. It is shown in figure 58. This
structure was created artificially, by hand. The air voids were created by
elimination of elements from the finite element mesh. The mineral
aggregates were assumed to be infinitely rigid, which means they can only
translate or rotate. Note, that this artificial structure differs from a real
dense graded asphalt concrete structure. This can be explained as follows.
An asphalt mixture can be thought to constitute a granular phase
embedded in a matrix phase, called mastic. The granular phase can be a
mineral aggregate skeleton, or can be coarse mineral aggregate grains
floating in the matrix. The matrix also contains small mineral aggregate
particles, of 2 mm and smaller, i.e. sand and filler. In the case of a
skeleton asphalt mixture, these particles may also be a part of the
skeleton, but are considered to be a part of the binder (mastic). In creating
the artificial structure of figure 58, it was a question how many of these
small particles had to be introduced in the structure as part of the granular
phase. In principle, the model described in the following can be extended
to three dimensions, but a difficulty is to define a realistic asphalt mixture
structure.
In figure 59a, the stress distribution is given for a given load step. Figure
59a shows that mainly the horizontal layers of the mastic between the
aggregates particles (black) transfer the applied compressive stress. The
stresses in the vertical layers (grey) are tensile, and are caused by
dilation. Figure 59b shows the corresponding displacements. Typically,
figure 59b shows mainly vertical displacements and negligible lateral
displacements of the aggregate particles.
The model was used to predict the simple static creep of dense asphalt
concrete observed experimentally, and shown in figure 60. Figure 60
shows creep curves obtained for two applied stresses, 0.1 MPa and 0.2
MPa. Figure 60 shows that the creep is strongly nonlinear, since a twice as
large stress causes an approximately 20% greater creep strain. This non-

Sec. 9

Results and interpretation of finite element computations


of the creep of a heterogeneous creep specimen

135

Figure 58. Two dimensional artificial


granular structure representing an
asphalt mixture.

Figure 59. Left (a): Typical stress distribution of principal stresses in the creep
test specimen. Indicated are compressive stresses (black) and tensile stresses
(grey). Right (b): Original position of the grains and position after loading.

linear behaviour can be attributed to the displacement of mastic in the


more or less horizontal layers between the aggregate particles in lateral
directions. The effect of this is that less mastic in the horizontal layers
between rigid particles is left to deform. Although this cannot be based on
evidence, it is plausible, because the model predicts negligible lateral
displacements of the aggregate particles (figure 59b). That is: it seems the
observed creep is localised mainly in the horizontal layers of mastic. To
make this plausible, the material parameters of the model were first fitted
to obtain close agreement with the result of the static test for 0.1 MPa.
Then the creep was computed for 0.2 MPa of applied stress. As expected,
this was found to be twice as large, cf. figure 61a, because the material
model (cf. appendix 4) used was a linear viscoelastic model. Subsequent
computations were made in which the elasticities and viscosities of the
material model were reduced proportionally, until the difference between
the curves diminished to 20%, in agreement with the experimental result.

136

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

Ch. 3

1.4

creep strain (%)

1.2
1
0.8
0.6
0.4
0.2
0
0

2.5

0,8

1.5
1
0.5

2000 4000 6000 8000

time (s)

creep strain (%)

creep strain (%)

Figure 60. Creep strain of a specimen


of 100 mm height after 1 hour loading
and 1 hour unloading, at 0.1 MPa
and 0.2 MPa applied axial stress.

0,6
0,4
0,2

0
0

2000 4000 6000 8000

time (s)

0
0

5000

10000

time (s)

Figure 61. Left (a): Computed creep of a specimen of 100 mm height, loaded at
0.1 MPa and 0.2 MPa. Right (b): Creep of the specimen of figure (a), computed
with reduced elasticities and viscosities of the material model so as to obtain the
same nonlinearity (difference of 20% between the curves shown) as the
experimental result in figure 60.

The result thus obtained is shown in figure 61b. Assuming that only the
bituminous matrix can deform (because aggregate particles are infinitely
rigid), a comparison of figure 61b and figure 60 permits an estimate of the
total thickness of the horizontal bituminous layers between the
aggregate particles. The vertical thickness reduction of the bituminous
layers of approximately 65%, cf. figure 61b, results in a macroscopic creep
of the asphalt mixture of approximately 1%, cf. figure 60. Hence, the
effective thickness of the bituminous layers in the loading direction was
approximately (1/65)100% = 1.5% of the specimen height.

Sec. 10

Modelling the time dependence of linear viscoelastic


stress strain behaviour

137

Thus, the two-dimensional finite element model described here explains


nonlinear creep of the heterogeneous material by the relative movement of
the two phases of the material. This shows, that the creep of a
heterogenous composite material differs from that of a linearly viscoelastic
continuum. The physical meaning of J (t ) in equation (8) differs from that
of J (t ) in equation (1), although the mathematical form of the definition
is the same. J (t ) in equation (1) is a true material property of a linearly
viscoelastic continuum, whereas J (t ) in equation (8), applied to asphalt
mixture, is influenced by the geometry of the grain skeleton. J (t ) of the
heterogeneous material could still be a geometry-independent (true)
material property, provided the stress state of the material were
homogeneous, cf. Ch. 5, figure 7, i.e. if the specimen were infinitely large.
Because of the finiteness of the specimen geometry, the threedimensionality of the state of stress, and the stress dependence of the stress
strain behaviour, J (t ) cannot be considered a material property. It is to be
considered a specimen property.

10 Modelling the time dependence of linear viscoelastic stress


strain behaviour
A further complication of the observed viscoelastic and viscoplastic stress
strain behaviour described in 5, 6, 7, and 8 represents the time dependence
of the behaviour. The time dependent behaviour is a complication since an
asphalt mixture in a pavement experiences shapes of waveform of loading
induced by passing wheel-loads that differ from those used in the
laboratory. Therefore, the question arises if (and how) the predictive value
of characterising properties determined in the laboratory depends on the
shape of the waveform of the applied load11. A consequence of the time
dependence of asphalt mixture properties is that they have at best relative
(qualitative) predictive value for the behaviour of the material in the
pavement.
It is probable that also a viscoplastic property depends on the shape of the
waveform of the applied load. The experimental evidence presented in 5
8 indicates that the elasto-viscoplastic stress strain behaviour of asphalt
mixtures depends on the shape of the waveform of the applied stress.
However, the experimental evidence does not by itself have the power to
prove that the different shapes of waveform of applied stress used caused
the observed differences in time dependent stress strain behaviour. The

11

The applied load is an applied stress in a stress controlled test, or an applied


strain in a strain controlled test. Here, only the response to an applied stress is
considered.

138

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

Ch. 3

reason is the scatter on the data, which is often considerable (caused in


part by the material heterogeneity and in part by the manufacturing
conditions, e.g. compaction), easily overshadows the influence of the
shape of the waveform. The following analysis was performed to further
make plausible that viscoelastic properties depend on the shape of the
waveform of the applied load. The analysis was performed using the
Burgers model. The Burgers model is a rheological model composed of
springs and dashpots. It is a linear model and is shown in figure 62.

E1

1
2
E2
Figure 62. The Burgers model is a rheological model composed of two springs, E1
and E2 , and two dashpots, 1 and 2.

A rheological model such as the Burgers model or a similar rheological


model is necessary if one wishes to model the influence of the shape of the
waveform on the response strain to an applied stress. The Burgers model
represents a simplification of an asphalt mixture. As such, it is not capable
of predicting the behaviour of real asphalt mixture quantitatively correctly.
10.1 Linear viscoelastic stress strain behaviour in the Burgers model
The Burgers model, cf. figure 62, is a suitable rheological model to obtain
a rough understanding of the relationships between linear viscoelastic
creep and dynamic linear viscoelastic stress strain behaviour. The stress
strain behaviour is described by the boundary value problem consisting of
the differential equation,
1 1
E

E1&& + 1 & = && + [ ] & +
(54a)
2
1 2
where
1
1
1
(54b)
[ ] =
+
+
1 21 2
1 = 1 / E1
(54c)
21 = 2 / E1
(54d)
2 = 2 / E 2
(54e)

Sec. 10

Modelling the time dependence of linear viscoelastic


stress strain behaviour

139

and the boundary conditions,


( 0)
1
(54f)
=

E1
&(0)
1
1
=
+
(54g)

1 2
Note, that (0) 0; this is the instantaneous elastic strain upon loading at
t = 0.

10.2 Static creep


For a constant applied stress,
(t )
(55)
=1

the boundary value problem equation (54) may be solved for ( t ) , to give
& (0)
t
1
(t ) (0 )
(1 e t / )
=
+
+ 2

(56a)

E11
E

1 1
The transient term, e t / , describes the compliance if the stress is applied
at t = 0 . Let us, for convenience, assume that the relaxation time, 2 , is
short, so that this term is negligible. Then equation (56a) reduces to
1
(t )
t
t
1
1
1
1
1
=
(56b)
=
+
+ 2 +

+
+
$
E1
E11
E11
E1
E 2 1
1 2
2

Rewriting,
E1 E 21
$
S (t ) =
=
(t ) ( E1 + E 2 )1 + E1 E 2 t
If E1 E 2 t >> ( E1 + E 2 )1 , then
S (t )

(56c)

(56d)
t
From equation (56b), the static creep strain rate is
d $
& =
=
(57)
dt 1
Note, that equation (56b) can be written as
J (t ) = J 0 + J 1 t m
(58)
with m = 1. In equation (58), J 0 represents the instantaneous elastic
compliance at t = 0.
10.3 Dynamic creep using a block-wave form of applied stress
The influence of the shape of the waveform of the applied stress can be
made plausible, by solving the differential equation of the rheological
model, to see that the response strain as a function of the time, ( t ) , and
the creep strain rate, &creep , depend on the shape of the wave-form of the
applied stress. For details the reader is referred to appendix 1. Then, ( t )

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

140

1.5

1.5

Ch. 3

1
0.5
0.5
0

0
0

1
2
1 + sin x x
constant

in units

x (x pi)

1.5

x (x)(x pi)

halfsine x

1
0.5
0
0

block-width = pi

x x((x)
pi)

Figure 63. Different shapes of waveform of applied stress. (a) Left: constant
stress and haversine stress. (b) Right: halfsine stress. (c) Low: unidirectional
block-wave; the loading-time, or block-width is defined as 2 .

is found to be of the following general form:


( t ) = integration constant + creep term + transient terms + dynamic terms

(59)

The dynamic terms consist of sums of sines and cosines, which do not
contribute to the cumulative creep strain. The transient terms are
proportional to exp ( t / ) , where t is the time and is the relaxation
time. These terms vanish as t . To evaluate the creep strain rate, it
suffices to consider the creep term. For example, for the Burgers model,
for the waveform shapes shown in figure 63, the creep strain rates are
given in table 19. The creep strain rate may be defined as the slope of the
line tangent to the minima (or maxima) of & (t ) . In table 19, E1 , 1 refer to
the model parameters of the Maxwell element of the Burgers model,
1 = 1 / E1 . For the block-wave shown in figure 63c, = / 2 . The
same solution applies to a block-wave with a loading time of 0.2 s and a
rest-time of 0.8 s. Then 2 = 0.2 s, for a frequency of 1 s-1, hence
= / 5 . Thus, the creep strain rate depends on the shape of the
waveform of the applied stress. According to equation (60d) in table 19,
the creep strain rate for the block-wave differs from that for the half sine
of equation (60c), by a factor of , equal to half the loading time.
Depending on the loading time, the creep strain rate for the block-wave
can be greater or smaller than that for the half sine. For example, for the
block-wave of figure 63c, = / 2 , and the creep strain rate is greater

Sec. 10

Modelling the time dependence of linear viscoelastic


stress strain behaviour

141

Table 19. Creep strain rate for different shapes of waveform of the applied stress in
the Burgers model.
shape of waveform
creep strain
rate

(
t
)
constant

=1
(60a)
&creep =
$
E1 1
(t )
haversine

= 1 + sin ( t )
(60b)
&creep =
$
E1 1
half sine
1
(t ) 1 1
2 cos(2i 2 t / T )
(60c)
= + sin(2 t / T )
& creep =
$

2
i = 1 (2i 1) (2i + 1)
E1 1
unidirectional (t )
sin (i )
2

&creep =
(60d)

( 1) i
cos (2i t / T )
=
+

block-wave
$

i =1
i

E1 1

than that for the half sine, by a factor of / 2 . For a block-wave with a
loading time of 0.2 s and a rest-time of 0.8 s, the creep strain rate is
smaller than that for the half sine, by a factor of / 5 . For the block-wave
of figure 63c, the creep strain rate is half that for the sinusoidal applied
stress of equation (60b), since = / 2 . Thus, equation (57) for static
creep can be generalised for dynamic creep,
d
k
&creep =
=
(61)
dt
1
where k is a constant, cf. table 19. Note, that for a Burgers material the
creep strain rate is the same in the static case and in the dynamic case
using a sinusoidal load.
10.4 Dynamic viscoelastic stress strain behaviour, using a sinusoidal
waveform of applied stress
The dynamic stress strain behaviour of the Burgers model under a
sinusoidal applied stress is described by the following equations (Findley
et al. 1976):
1 2 (12 2 / E 2 ) [1 12 2 / ( E 1 E 2 ) ]
E =
(62a)
2 2 + [1 1 2 2 / ( E 1 E 2 )] 2
(12 / E 2 ) 3 + 1 [1 1 2 2 / ( E 1 E 2 )]
E =
(62b)
2 2 + [1 12 2 / ( E 1 E 2 )] 2
=

E1 E 2
E2
From equation (62b) it is readily seen that
for 0 , E * E 1
for , E E E1
From equation (63a), ln E = ln 1 + ln , hence
*

(62c)

(63a)
(63b)

142

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

Ch. 3

d ln E * d ln E
=
=1
(63c)
d ln
d ln
Thus, the slope of the complex modulus as a function of the frequency on
log-log scale varies from 1 for 0 , to 0 for .
Equation (63a) expresses that, in the low frequency domain, E > E , the
complex modulus (master curve) is determined by 1 . Equation (63b)
expresses that, in the high frequency domain, E > E , the complex
modulus is determined by E1 .
An example is shown in figure 64, where E1 = 13,000 MPa, 1 = 2 =
700 MPa.s, and E 2 takes different values depending on the relaxation
time of the Voight-Kelvin element of the model. This relaxation time is
given by 2 = 2 / E 2 . In figures 64a..d the relaxation time is respectively
equal to 0.1 s, 1 s, 10 s, and 100 s. To obtain these relaxation times, the
required E 2 -values are respectively 7,000 MPa, 700 MPa, 70 MPa, and 7
MPa. The delayed elasticity is controlled by the parameters E 2 and 2 . It
influences the master curve only at intermediary frequencies. This can be
seen by comparison of figures 64a..d, where E1 , 1 , and 2 are constant,
and only E 2 varies. Figure 64 shows that the complex modulus, E * ,
which is approximately equal to the loss modulus, E , is hardly
influenced by delayed elasticity, i.e. by 2 and E 2 , in the low frequency
domain12. The reason that E in the low frequency domain is hardly
influenced by 2 and E 2 , is that these parameters occur in coefficients of
3 in equation (62b) which decrease rapidly if 0 . Under those
conditions, the energy represented by E * , i.e. E , is practically
dissipative, i.e. irrecoverable, if the angular frequency is sufficiently low.
for 0 ,

12

Phillips and Robertus (1995, 1996, 1997, 1999) pointed out that the delayed
elasticity influences the creep, and that the complex modulus, S * , may wrongly
predict the creep. The present analysis shows that this can be true, if the
frequency applied in the creep test is in the intermediary range where the
behaviour is influenced by the delayed elasticity. However, in this study it was
found that the temperatures and frequencies at which rutting occurs in the
pavement are normally not in the intermediary frequency domain of the complex
modulus master curve (assuming the reference temperature is 20C), but in the
low frequency domain.

Modelling the time dependence of linear viscoelastic


stress strain behaviour

100000

100000

10000

10000

143

1000
1000
100
100

stiffness modulus, E', E",


E* (MPa)

1000

100

E' pred.
E* pred.

10000

E" pred.

0.1

angular frequency, rad/s


E' pred.
E* pred.

10

10
0.01

10000

1000

100

10

0.1

0.01

10

E" pred.

100000

100000

10000

10000

1000

E' pred.
E* pred.

E" pred.

10000

1000

10

10
100

10000

1000

100

10

0.1

0.01

10

100

100

0.1

1000

0.01

stiffness modulus, E', E",


E* (MPa)

Sec. 10

angular frequency, rad/s


E' pred.
E* pred.

E" pred.

Figure 64. Effect of delayed elasticity in the Burgers model. E 1 = 13.000 MPa, 1
= 700 MPa.s, 2 = 1 . Top left (a): 2 = 0,1 s E 2 = 7000 MPa. Top right (b):
2 = 1 s E 2 = 700 MPa. Bottom left (c): 2 = 10 E 2 = 70 MPa. Bottom
right (d): 2 = 100 E 2 = 7 MPa.

10.5 Complex modulus and linear viscoelastic creep susceptibility


of asphalt mixture
For a hypothetical homogeneous asphalt mixture, one might assume the
following rheological model instead of equation (63),
for 0 , S * S k1 m
(64a)
(64b)
for , S * S k 2
From equation (64a):
d ln S * d ln S
=
=m
(64c)
for 0 ,
d ln
d ln
where k1 is the analogue of 1 of the Burgers model, cf. equation (63a),
and k 2 , the glass modulus, is the analogue of E1 of the Burgers model, cf.
equation (63b). Recall equation (1),
J (t ) =

= J 0 + J1 t m
[ ]

(65)

according to which the slope of the master curve is equal to m ( m < 1),

144

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

Ch. 3

cf. equation (6). Recall that equation (56b) is of the form of equation (58),
i.e. equation (65) with m = 1. Generalising equation (61) yields
d k
k m
&creep =
=
=>
creep = C +
t ( m = 1)
(66)
dt
1
1
m k m 1
&creep =
t
( m < 1)
(67a)
k1
where C is an integration constant, and k1 replaces 1 . From equation
(65),
&creep = m [ ] J 1 t m1
(67b)

Equations (67a) and (67b) are different expressions for the same creep
strain rate. Combining equations (67a) and (67b),
[ ] J 1 = k / k1
(67c)
Equation (67a) expresses that the creep strain rate depends on the slope of
the complex modulus master curve on log log scale, m , and on the shape
of the waveform applied in the creep test, k ; see also table 19.
10.6 Discussion
A matter of discussion has been the question to which extent an asphalt
mixtures complex modulus characterises its resistance to permanent
deformation (rutting in the pavement). Another question that has to be
addressed then is to which extent the complex modulus may wrongly
predict an asphalt mixtures resistance to permanent deformation if the
mixture exhibits delayed elasticity.
On the one hand, there are different limitations associated to the complex
modulus:
1 the complex modulus characterises a mixtures linear viscoelastic
behaviour under strain controlled conditions, whereas permanent
deformation (creep) is always nonlinear and stress controlled;
2 the complex modulus characterises only the dynamic component of the
stress strain behaviour, whereas permanent deformation is normally
considered as the cumulative summation of dynamic and static strain
components;
3 the resistance to permanent deformation depends importantly on the
mixtures structure, i.e. its grain skeleton, and on the presence of a
confinement pressure. The response to loading of the grain skeleton
differs strongly under strain controlled conditions and stress controlled
conditions. The conditions under which the complex modulus is
determined are not representative for the conditions under which the
material develops permanent deformation.
On the other hand, in the physical model used to characterise the material,
the material is a homogeneous linearly viscoelastic continuum, and the
stress strain behaviour is governed by the time temperature superposition

Sec. 10

Modelling the time dependence of linear viscoelastic


stress strain behaviour

145

principle, expressed by equation (7), irrespective of whether the loading


mode is strain controlled or stress controlled. Therefore, an additional
characterisation is required if this principle is violated, i.e. for a material
exhibiting nonlinear stress strain behaviour under stress controlled
conditions. In this study, this additional characterisation is the creep
behaviour in the (triaxial) creep test, characterised by any one of the creep
models, equations (15)-(17), depending on the testing conditions.
In conclusion at this point, an asphalt mixtures complex modulus
characterises its resistance to permanent deformation, but is as such an
incomplete characterisation of the resistance to permanent deformation.
Considering that the materials stress strain behaviour is stress dependent
and time dependent, the testing conditions in the triaxial test are relatively
simple as compared to the loading conditions in the pavement. Since it is
probable that an asphalt mixture in the pavement undergoes a greater
variety in time stress paths than can be realised in the (triaxial) creep test,
a characterisation of the nonlinear viscoplastic behaviour in the (triaxial)
creep test is also likely to be incomplete as a characterisation of the
behaviour of the material in the pavement.
The slope of the complex modulus as a function of the frequency on loglog scale varies from m0 for 0 , to 0 for , cf. 5. It was shown
in 5, cf. tables 3, 4, and 6, that, for the creep conditions used in this study,
[40C, 0.5 Hz], [40C, 1 Hz], [50C, 0.5 Hz], [50C, 1 Hz], the condition
m m0 is fulfilled. Under those conditions, cf. equation (62), S * S .
Under the conditions applied in the triaxial creep test, the frequency seems
sufficiently low so that the influence of delayed elasticity is negligible.
Theoretically, an asphalt mixture might perform better than expected
based on its characterised behaviour, if the mixtures delayed elasticity
develops, i.e. in the intermediary temperature and frequency domain of the
master curve. This possibility can be ignored if temperatures where rutting
develops in the pavement are sufficiently high so that the delayed
elasticity is negligible.
10.7 Limitations of the linear viscoelastic model
It is recognised that the model assumes the material to be homogeneous.
This may be an oversimplification of a granular bituminous material. It is
recognised that the expectation according to equation (61) is strongly
simplified, and that it is incorrect to assume implicitly (as we have done)
that 1 is constant for the various conditions in the experiments. This is
explained further in the following.
The binder, bitumen, is thixotropic and pseudoplastic. A fluid is
thixotropic, if the viscosity decreases as a function of the time upon
loading (shearing). A fluid is pseudoplastic, if the viscosity decreases as a

146

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

Ch. 3

function of increasing shear rate, & . For the purpose of the following
considerations, we may assume that & can be replaced by & .
The pseudoplastic viscosity can be defined according to the equation
& k = /
(68)
Equation (68) is known as the pseudoplasticity equation. In rheological
modelling, the dashpot is modelled as a Newtonian viscosity law, i.e.
k = 1 . For a pseudoplastic material, k < 1 . A consequence of k = 1 is that
the static creep according to equation (56) and the dynamic creep
according to equation (60), see table 19, are proportional with time, which
implies that m = 1 in equation (58) for static creep.
The thixotropy and pseudoplasticity of the binder will be transferred to the
asphalt mixture. Therefore, in general (read: under stress controlled
conditions), 1 cannot be considered a constant, not as function of the
time, and not as function of the creep rate.
However, if the dynamic properties E ( ) and E ( ) of equation (60)
are fitted to the corresponding experimental properties S ( ) and S ( ) ,
then a set of values of E1 , 1 , E 2 and 2 is obtained, which describes the
dynamic time dependent behaviour under strain controlled conditions, in a
wide frequency domain. An example is shown in figure 64. Note, that this
applies to the strain controlled case, where the stress is caused by a
sinusoidal applied strain. Thus, the value of 1 differs for the stress
controlled case and the strain controlled case.

11 Discussion I Dependence of creep properties on the specimen


geometry (height)
Two results from previous sections are recalled here. The first result, in
figure 65, is for a 60 x 100 mm height x diameter specimen, see also
figure 51. The second result, figure 66 and table 20, is for a 200 x 100 mm
height x diameter specimen, see also figure 48 and table 18. For both
results the material was dense graded asphalt concrete. The testtemperature was 50C, and the applied load was a block-wave with a
loading-time of 0.2 s, and a rest-time of 0.8 s. It is assumed here that not
only the size of the 60 x 100 height x diameter specimen and the 200 x
100 height x diameter specimen differs, but also the shape, because the
height : diameter ratio of the specimens differs.
As for the 60 x 100 mm specimen, figure 65, the creep model according to
the following equation describes the creep curves
J (t ) = J 1 + ln t z
(69)
where J (t ) is the creep compliance as function of the time, and J 1 is the
creep compliance at 1 s, and z is the viscoplastic creep susceptibility.
Figure 65 shows the exponent, z , of equation (69) as function of q / p .
Figure 65 shows that z ( q / p ) is a power law relationship.

exponent, z10^4

200
150

Discussion I Dependence of creep properties on


the specimen geometry (height)
DAC 0/16, block-wave, 1 Hz

1000

exponent, z10^4

Sec. 11

y = 29.312x1.6396
R2 = 0.976

100
50
0
0.00

1.00

2.00
q/p

3.00

p=0.37

p=0.17
p=0.4

p=0.23
p=0.43

p=0.3
p=0.47

p=0.33
p=0.5

DAC 0/16, block-wave, 1 Hz

100

10

y = 29.312x1.6396
R2 = 0.976

1
0.10

4.00

147

p=0.37

1.00
q/p
p=0.17
p=0.4

p=0.23
p=0.43

10.00
p=0.3
p=0.47

p=0.33
p=0.5

Figure 65a. Exponent z of equation (69) as function of the quotient of the


deviatoric stress, q , and the volumetric stress, p . The legend shows the
volumetric stress in MPa. DAC 0/16, block-wave form of applied stress.
Hydraulic load control. Specimen: 60 x 100 mm height x diameter. (a): On
linear scale. Right (b): On log-log scale.

exponent, z10^4

100
80

DAC 0/16, sinus, 1 Hz


y = 18.374x1.514
R2 = 0.9157

60
40
20
0
0.00

p=0.37

1.00

2.00
3.00
4.00
q/p
p=0.17
p=0.23
p=0.3
p=0.43

DAC 0/16, sinus, 1 Hz

100

exponent, z10^4

120

10
y = 18.374x1.514
R2 = 0.9157
1
0.10
p=0.3

1.00
q/p
p=0.17
p=0.37

10.00
p=0.23
p=0.43

Figure 65b. Exponent z of eq. (69) as function of the quotient of the deviatoric
stress, q , and the volumetric stress, p . The legend shows the volumetric stress
in MPa. DAC 0/16. Sinusoidal applied stress. Hydraulic load control. Specimen:
60 x 100 mm height x diameter. Left (a): On linear scale. Right (b): On log log
scale.

As for the 200 x 100 mm specimen, figure 66 and table 20, the creep
curves are described by different creep models, depending on the testing
conditions). In figure 66, a transition can be observed from the logarithmic
model, equation (69), to the power law model,
~ ~
J (t ) = J 0 + J 1 t z
(70)
where J 0 is the instantaneous elastic compliance at 0 s (negligible), and
~
J 1 is the creep compliance at 1 s, and ~z is the viscoplastic creep
susceptibility. This transition takes place as the ratio of the deviatoric
stress and the volumetric stress increases, somewhere between q / p =
0.43 and 0.95. Figure 66 shows that ~
z ( q / p ) beyond 0.95 is a linear

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

DAC 0/16, block-wave, 1 Hz

1.000

0.100

0.010

0.001
1

10

100

1000 10000

Ch. 3

DAC 0/16, block-wave, 1 Hz

0,5

exponent z of eq. (16)

compliance J(t) (MPa^-1)

148

0,45

z = 0.0713 q/p + 0.2353

0,4

0,35

time (s)
0,9/0,3 MPa
0,3/0,2 MPa
0,3/0,03 MPa
0,6/0,05 MPa

0,6/0.25 MPa
0,6/0,1 MPa
0,75/0,15 MPa
0,6/0,15 MPa

0,3
0

0,5

1,5

2,5

q/p

Figure 66. Dynamic creep of dense graded asphalt concrete, DAC 0/16. Left (a):
Each curve represents the average of three tests. The waveform of the applied
load was a block-wave with a loading time of 0.2 s and a rest-time, of 0.8 s.
Load control: pneumatic. Temperature: 50C. Specimen: 200 x 100 mm height x
diameter. In the legend, combinations of axial stress and radial stress in MPa
are given; see also table 20. Right (b): Exponent ~z of eq. (16) as function of the
quotient of the deviatoric stress, q, and the volumetric stress, p.
Table 20. Creep models belonging to the creep curves in figure 66.
1 / 3
creep model
q/ p
R2
99.9
0.6/0.05
J (t ) = 0.0032 t 0.4255
2.36
0.3673
99.7
0.3/0.03
2.25
J (t ) = 0.0040 t
0.3648
99.4
0.6/0.1
1.88
J (t ) = 0.0029 t
0.3750
99.7
0.75/0.15
1.71
J (t ) = 0.0020 t
99.6
0.6/0.15
1.5
J (t ) = 0.0020 t 0.3363
0.3192
99.1
0.9/0.3
1.2
J (t ) = 0.0010 t
0.3040
99.4
0.6/0.25
0.95
J (t ) = 0.0010 t
99.1
0.3/0.2
0.43
J (t ) = 0.0001 + 0.0004 ln t

relationship. It is noticed, that this relationship differs entirely from that


for the 60 x 100 mm specimen of figure 65. So, it seems, the creep
properties of asphalt mixture depend strongly on the specimen geometry.
If that is the case, then those properties cannot be considered predictive
for the behaviour in a geometry which differs from the one in which they
were determined, e.g. the pavement. That is, the creep properties of
asphalt mixture cannot be considered geometry-independent (true)
material properties.
Weaker experimental evidence of the dependence of the creep properties
on the specimen geometry is shown in figure 57. A possible explanation
for the dependence of creep properties on the specimen height in the
absence of confinement was given in 9.

Sec. 12

Discussion II Relatedness of dynamic viscoelastic and


elasto-viscoplastic (creep) properties of asphalt mixture

149

12 Discussion II Relatedness of dynamic viscoelastic and elastoviscoplastic (creep) properties of asphalt mixture
12.1 Relatedness of complex modulus and creep properties
It was shown that the complex modulus and the creep properties are
closely related properties. For a homogeneous linear viscoelastic
continuum the linear dynamic viscoelasticity (i.e. strain controlled stress
strain behaviour) and the linear viscoelastic creep (i.e. stress controlled
stress strain behaviour) are two types of viscoelastic behaviour that are
governed by the same constitutive model, i.e. the time temperature
superposition principle, equation (7).
An asphalt mixture is capable of exhibiting linearly viscoelastic behaviour
under strain controlled conditions. Under stress controlled conditions, an
asphalt mixture exhibits nonlinear stress strain behaviour. Therefore, it is
probable that the complex modulus and the creep properties of asphalt
mixture are related. It implies that the complex modulus characterises an
asphalt mixtures resistance to creep or permanent deformation. There are,
however, a couple of limitations connected to the complex modulus as a
characterisation of the creep susceptibility,
. it does not account for the influence of nonlinear stress strain behaviour;
i.e. increased stiffness under compression;
. it does not account for the influence of confinement.
It was argued in 10.6 that a characterisation of the resistance to permanent
deformation on the basis of the complex modulus is incomplete, because
the complex modulus characterises linear viscoelastic behaviour. It was
argued further in 10.6 that a characterisation of the resistance to
permanent deformation on the basis of the creep properties in the triaxial
creep test is also incomplete, because the testing conditions in the triaxial
test are relatively simple as compared to the loading conditions in the
pavement. Therefore, it is plausible that the complex modulus and the
creep properties characterise the resistance to permanent deformation in
different domains of stress strain behaviour. An advantage of the complex
modulus is that it is approximately geometry-independent.
12.2 Creep of asphalt mixture in the absence of confinement
The creep properties of asphalt mixtures were investigated for their
dependence on the time, the temperature, the applied stress, and the
mixture composition.
Dependence of the creep on the time. Experimental evidence was
presented, cf. 6, 7, 8, on the basis of which it is plausible that the creep of
asphalt mixture depends on the shape of the waveform of the applied
stress. The experimental evidence by itself does not have the power of
proof. Therefore the influence of the shape of the waveform was further

150

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

Ch. 3

made plausible by modelling the time dependent behaviour of the Burgers


model for different shapes of waveform of applied stress.
It is probable that the creep curves that were found for different shapes of
waveform in the absence of confinement are parallel curves. That is, the
effect of the time dependence of the creep is that the creep curve shifts
horizontally along the time axis. The dependence of the shift on the shape
of the waveform of the applied stress can be complex. It cannot be
predicted on the basis of the time temperature superposition principle.
In the absence of confinement, a suitable creep model is equation (21),
because it permits a fit to whole creep curves, meaning that the fitted
creep parameters of this model reflect measured values of the creep
compliance. The parameters J 10 and z of equation (21), were found to be
suitable characterising properties. It was shown that the parallelism of the
creep curves implies that J 10 and z are interrelated. That is: if J 10
increases, then z decreases, and vice versa. The effect of a shift of the
creep curve to the left, i.e. to lower time value, is that J 10 increases and z
decreases. Conversely, the effect of a shift to the right, i.e. to higher time
value, is that J 10 decreases and z increases.
Dependence of the creep on the temperature. The experimental evidence
shows that an increase of the temperature causes J 10 to increase. The
expected decrease of z , based on the interrelatedness of J 10 and z , was
confirmed by the experimental data.
Dependence of the creep on the applied stress. The experimental evidence
shows that an increase of the applied stress amplitude causes J 10 to
decrease. The expected increase of z , based on the interrelatedness of J 10
and z , was confirmed by the experimental data.
Dependence of the creep on the mixture composition. The results must not
be generalised, since only three mixtures were investigated. Despite that,
the observed trends are representative, which is known based on
subsequent experience. The experimental evidence indicates that J 10 is to
be considered an indicator for the creep during the first 10 s, and that z is
an indicator for the creep after the first 10 s. In agreement with what can
be observed in pavements, polymer modified asphalt mixtures show
improved resistance to permanent deformation, i.e. a reduced creep
susceptibility, z .

12.3 Creep of asphalt mixture in the presence of confinement


A more complete characterisation of the creep susceptibility and the
resistance to permanent deformation includes the influence of
confinement. The creep properties of asphalt mixtures were investigated
for their depen-dence on the time, the applied stress, and the mixture
composition, 8.

Sec. 13

Discussion III Physical meaningfulness and predictive


value of viscoelastic and elasto-viscoplastic properties

151

Dependence of the creep on the time. Only two waveforms were used: a 1
Hz block-wave of applied stress and a 1 Hz sinusoidal applied stress. It
appears that with those waveforms the influence on the creep is not
significant. That is: there is probably an influence, but that influence is
small compared to the potential influence of the scatter. The scatter can be
significant, cf. figures 7 and 8. The influences of the shape of the
waveform and the scatter on the data were not investigated in great detail.
Two creep models were found to describe the time dependent behaviour;
the logarithmic model, equation (15), or the power law model, equation
(16), depending on the testing conditions.
Dependence of the creep on the applied stress. The experimental evidence
shows that confinement has a very significant influence on the creep
properties, in particular the exponent of the creep model, respectively z of
z of equation (16). For convenience, let us use Z to
equation (15), or ~
indicate z , ~
z , or z . The experimental evidence shows that the influence
of confinement can be characterised by Z as a function of the ratio of the
deviatoric stress, q , and the volumetric stress, p . Z ( q / p )-relationships
for dense graded asphalt concrete and porous asphalt were found to be
logarithmically linear and to depend on the shape of the waveform of the
applied stress and the material composition.
Dependence of the creep on the mixture composition. The results must not
be generalised, since only two mixtures were investigated, a dense graded
asphalt concrete mixture and a porous asphalt mixture. Despite that, the
observed trends are representative, which is known based on subsequent
experience. The experimental evidence shows that Z ( q / p ) on log-log
scale is reduced for porous asphalt as compared to dense graded asphalt.
The reduction is attributed to the grain skeleton, i.e. its resistance to
deformation in the presence of confinement.

13 Discussion III - Physical meaningfulness and predictive value


of viscoelastic and elasto-viscoplastic properties of asphalt
mixture
Many properties of asphalt mixture currently used for asphalt pavement
design and asphalt mixture design (type testing) are arbitrary empirical
properties, which do not fit in any theory of physics. A limitation of such
properties is that they are defined in a framework of practical experience,
and lose applicability outside that framework. Arbitrarily defined
empirical properties of asphalt mixture do not have predictive value for
the behaviour of the material in the pavement. To solve this problem,
physically meaningful properties have to be defined.
Physically meaningful asphalt mixture properties are predictive for the
behaviour of the material in the pavement. Physically meaningful
properties are properties of a physical model that describe the material

152

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

Ch. 3

behaviour. For physical properties to have predictive value it is necessary


that they are geometry-independent (true) material properties.
As already argued in 3, equation (7) represents a fundamental principle of
physics, based on which the variables and parameters of equations (1)-(7)
can be considered fundamental physical properties of a linearly
viscoelastic continuum. Applied to asphalt mixture, although not a linearly
viscoelastic continuum, the complex modulus is approximately geometryindependent, therefore can be considered approximately a true material
property. Therefore, it should be predictive for the behaviour of the
mixture in the pavement.
Basically, a variable or parameter of a constitutive model equation that
cannot be obtained as a true material property is almost equally empirical
as an arbitrarily defined empirical property. Creep properties of asphalt
mixture depend on the specimen height, and can therefore not be
considered true material properties.
A physical meaning14 is attributable to ~
z in equation (16), based on the
mathematical correspondence of equation (16) and equation (1), the fact
that equation (1) is a fundamental constitutive equation, and the fact that
its variables and parameters are obtained as true material properties. A
physical meaning is attributable to z in equation (15), based on the
relationship to ~
z in equation (16) in the transition point, (q / p) c , cf.
equation (45). A physical meaning is attributable to z , based on the
relationship to z , defined by (t ) , cf. equation (25), and the relationship
to ~
z defined by ~ (t ) , cf. equation (31).
Predictive value of physically meaningful properties is lost in a number of
ways. Firstly, a physical model that is an oversimplification of the reality
does not permit accurate predictions. Apart from that, two major
influences on the viscoelastic and elasto-viscoplastic properties of asphalt
mixture were investigated in this chapter:
. the influence of the specimen geometry (specimen height);
. the influence of the time.
Two main observations, which arise from this study, are:
1 The dependence on the specimen geometry of a creep property is an
intrinsic property of a granular bituminous material. This means, that this
dependence cannot be avoided or eliminated, cf. 8.4, 9, and 11.
2 The stress strain behaviour depends on the shape of the waveform of the
applied stress. The shape of the waveform of the stress induced in the
pavement by a passing wheel-load differs from the shapes of waveform
used in the laboratory, cf. 10.

14

This is stated here based on the notion of physical meaningfulness discussed in


Ch. 5, 3.2, and further elaborated in Ch 5, 5 and 6.

Sec. 14

Conclusions

153

Owing to the dependence on the specimen geometry, an asphalt mixtures


creep properties cannot be considered true material properties. A
geometry-dependent mechanical property is not predictive for the
mechanical behaviour in a different geometry. Therefore, creep properties
of asphalt mixture are not predictive for a geometry that differs from the
geometry in which they were determined, e.g. the pavement.
Even if an asphalt mixture property can be considered a true material
property from a practical standpoint the complex modulus can be
considered as such then the predictive value is limited, if, as in the case
of a viscoelastic or elasto-viscoplastic property, it depends on the time, i.e.
on the shape of the waveform of the applied stress (or strain).
Therefore at best mechanical asphalt mixture properties have relative
(qualitative) predictive value rather than absolute (quantitative) predictive
value for the behaviour of the material in the pavement.

14 Conclusions
1 Viscoelastic properties and creep properties of asphalt mixture as
determined in the laboratory lack quantitative predictive value for the
behaviour of the material in the pavement. Main reasons are the
dependence of the stress strain behaviour on the specimen geometry and
the time, i.e. the shape of the waveform of the applied stress.
2 It is plausible that the dependence of viscoelastic properties and creep
properties of asphalt mixture on the specimen height is an intrinsic
property, and can therefore not be avoided or eliminated.
3 The viscoelastic properties are fairly reproducible using different
specimen geometries, and can therefore, from a practical standpoint, be
considered true material properties.
4 A physical meaning can be attributed to phenomenological creep
models, based on the correspondence to the fundamental viscoelastic
model.
5 The predictive value of viscoplastic asphalt mixture properties for the
behaviour of the material in the pavement is not entirely lost as a result of
the time dependence and the dependence on the specimen geometry. A
qualitative predictive value can be retained based on a characterisation
based on constitutive models. By this it is meant that viscoplastic
properties have a qualitative predictive value for the behaviour in the
pavement (meaning that it can be compared to the corresponding property
of a similar standardised asphalt mixture).

154

Characterisation of viscoelastic and viscoplastic


behaviour of asphalt mixture

Ch. 3

6 The viscoelastic creep susceptibility of asphalt mixture as obtained from


the complex modulus master curve in the low frequency domain can be
used as an estimator of the viscoplastic creep susceptibility, and the
resistance to permanent deformation.
7 The triaxial creep test provides additional information, in addition to the
information from the frequency sweep test, about the resistance to
permanent deformation of an asphalt mixture. The triaxial creep test
provides an indication of the contribution of the mixture composition and
granular/bituminous structure to the resistance to permanent deformation,
depending on the state of stress and the shape of the waveform of the
applied stress
8 Since viscoplastic properties depend on the specimen geometry and the
geometry of the load application, it is important for the reproducibility of
the properties in the laboratory that test methods are standardised.

4
Characterisation of the resistance to
crack-growth and the resistance to
fracture of asphalt mixture
1 Aim
The aim of the analysis in this chapter is to investigate the experimental
evidence of the resistance to crack-growth and fracture in different tests, to
come to a judgement as to whether or not it is justified to have confidence
that the simpler test methods are suitable for use in a practical context to
characterise the resistance to cracking. The ultimate aim is to be able to
judge the cost-effectiveness and the risk of failure and of the asphalt
mixture when applied in a pavement. The tests considered are two
dynamic crack-growth tests, a static crack-growth test, a fracture
toughness test, and three different tensile tests.

2 Methodology
The method to come to the above judgement comprises a study of the
applicability of Paris Law and the stress intensity factor with the purpose
to investigate whether or not the constants of the Paris equation, the
fracture toughness and the tensile strength of an asphalt mixture can be
obtained as true material properties.
The methodology is illustrated by figure 1, and explained in the following.
For details, the reader is referred to 3. In a dynamic crack-growth test, a

156

Resistance to crack-growth and fracture

Ch. 4

LINEAR ELASTIC FRACTURE MECHANICS

dynamic
crack-growth test
stable crack-growth

Paris Law
da
= A( K ) n
dN
K c = K max at amax

static creep
crack-growth test
stable crack-growth

Paris Law
da
= A( K ) n
dt

monotonic test
constant rate: & = constant
unstable crack-growth

fracture toughness test


K Ic = ac

Figure 1. Overview of possible characterisations of resistance to cracking in


linear elastic fracture mechanics. Left: Dynamic crack-growth behaviour obeys
Paris Law (fatigue crack-growth); properties: constants A and n of Paris Law,
and critical stress intensity factor. Middle: Static crack-growth behaviour obeys
Paris Law (creep crack-growth); properties: constants A and n of Paris Law.
Right: Monotonic (constant deformation rate) test, to characterise the resistance
to fracture; properties: fracture toughness, tensile strength. For an explanation
of symbols, cf. 3.

periodical load is applied to the specimen. The specimen contains a


prefabricated crack. In a constant rate test, the specimen is subjected to a
deformation at a constant applied deformation rate. Two types of constant
rate tests were used: a fracture toughness test, to determine the fracture
toughness, and different tensile tests to determine the tensile strength. The
specimen for the determination of the fracture toughness is equipped with
a prefabricated notch. The specimen for the determination of the tensile
strength is without a notch. A special test is the static crack-growth test, in
which a static load is applied to the specimen equipped with a
prefabricated crack.
Figure 1 illustrates that a dynamic crack-growth test may be used for the
characterisation of the crack-growth behaviour of the material, and that a
constant rate test may be used for the characterisation of the resistance to
fracture. A dynamic crack-growth test is designed to let a crack grow in a
stable, controlled manner. A fracture toughness test is designed to let a
crack grow in an unstable manner, called fracture.
Dynamic crack-growth behaviour can be characterised by means of Paris
Law. Dynamic crack-growth properties are the constants A and n of the
Paris equation. In addition, the critical stress intensity factor, Kc, can be
obtained from the stress and the crack-length at the moment where the
residual fracture in a crack-growth test occurs. The critical stress intensity
factor can be expected to correspond to the fracture toughness, provided

Sec. 2

Methodology

157

the residual fracture is linearly elastic, and the specimen fulfils the ASTM
specimen size requirements for the stress state of the specimen to be plane
strain.
Fracture can be characterised by means of the critical stress intensity
factor. The stress intensity factor describes the stress field at the crack-tip.
The critical stress intensity factor is the stress at which fracture occurs.
Therefore, this is also considered to be the residual tensile strength of the
cracked specimen (however, in MPam instead of in MPa). It is known
that the critical stress intensity factor depends on the specimen thickness.
For sufficient thickness, the critical stress intensity factor becomes
independent of the specimen thickness. It is then called the plane strain
critical stress intensity factor, also called fracture toughness. It is common
to consider the fracture toughness a material property, which characterises
the resistance of the material to fracture.
In principle, the dynamic crack-growth properties, i.e. the constants of the
Paris equation, can be related to the fracture toughness. The relation can
be shown, if the critical stress intensity factor from the dynamic crackgrowth test fulfils the conditions for the determination of a valid fracture
toughness. Crack-growth tests are laborious and expensive, and therefore
not suitable for routine purposes, such as asphalt mixture design. One
might consider to use the fracture toughness instead of the constants of the
Paris equation, to characterise the resistance to cracking, since the fracture
toughness can be determined in a simple and much more efficient way,
more suitable for routine purposes. An additional benefit of the fracture
toughness is that it can be used as an engineering property. This is
illustrated in figure 2. According to the definition, K Ic = ac , K Ic
can be represented as a residual tensile strength hyperbole in a plot of
versus ac . Figure 2 shows that for a given design stress, d , the
tolerable flaw size, 2 a d , is increased if the critical stress intensity factor,
K Ic , is increased.
The conditions which have to be fulfilled in order to obtain a valid fracture
toughness are very strict. The conditions are:
1 the stress strain behaviour must be linearly elastic;
2 the stress state of the specimen must be plane strain.
These conditions are difficult to fulfil, in particular at temperatures above
approximately 10C. Therefore, one might consider using the tensile
strength instead of the fracture toughness. The fracture toughness and the
tensile strength differ by a constant factor in the absence of strain
hardening, i.e. if the temperature is sufficiently low. Thus, in principle, the
fracture toughness and the tensile strength are related. Based on the
relationship between the fracture toughness and the tensile strength, the

Resistance to crack-growth and fracture

158

Ch. 4

stress

Figure 2. Engineering interpretation of the critical stress


intensity factor. d represents the design stress, 2 a d is the
tolerable flaw size, curves K c1 and K c 2 represent the critical stress intensity factor of material 1 and material 2.

latter might be useful to characterise the resistance to cracking of the


material if a valid fracture toughness cannot be obtained. It was found
experimentally, using different specimen geometries, that the tensile
strength was not obtained independent of the specimen geometry. This can
be attributed to the nonlinearity of the stress strain behaviour.

3 Theory
3.1 The stress intensity factor
In the linear elastic model, the biaxial state of stress of an infinite thin
plate with a central ellipsoidal crack of length 2 a , is described by the
following equations:
a
3

(1a)
cos 1 sin sin
x =

2r
2
2
2

y =

a
3

cos 1 + sin sin

2r
2
2
2

(1b)

a
3

(1c)
sin cos cos
2r
2
2
2
where x and y represent the stress in the x-direction, respectively the
stress in the y-direction, xy is the shear stress, is the applied stress (or
the stress at infinite distance from the crack, r is radial distance of the
location ( x , y ) with respect to the crack-tip, and is the argument of the
radius vector, r . This is illustrated in figure 3.

xy =

Sec. 3

Theory

159

xy
x

Figure 3. A biaxially loaded infinite plate containing a slit crack.

Equations (1) can be rewritten in the following generalised form:


ij = K f ij (r )

(2)

where ij is the stress tensor, f ij (r ) is a geometrical function, and

K = a
(3)
is the stress intensity factor. For specimens of finite size approximate
solutions for the stress similar to equations (1) were obtained (Ewalds and
Wanhill 1984, Broek 1986), where K is written as
K = a f (a / W )
(4)
where f ( a / W ) is the geometry factor. For the centre-cracked tensile
(CCT) specimen with a central ellipsoidal crack of length 2a , figure 4a
(Murakami 1986),
P
a
K = 0 a f (a / W ) =
(5)
a sec
BW
W
where 0 is the applied stress, a is the crack-length, f ( a / W ) is the
geometry factor, P is the applied force, B is the specimen thickness, and
W is the specimen width. The error of the geometry factor is less than 1%
for 2a / W = 0.8 . For the single edge notched four point bending
specimen, figure 4b (Murakami 1986),
3Pl
K = 0 a f (a / W ) =
a f (a / W )
(6a)
BW 2

160

Resistance to crack-growth and fracture

P/2
W

Ch. 4

P/2

2a
P

W
B

Figure 4. Schematic representations of specimen geometries used. Top left (a):


Centre-cracked tensile specimen. Top right (b): Single edge notched four point
bending specimen. Below (c): Notched semic-circular bending specimen.
2

a
a
a
a
(6b)
f (a / W ) = 1.122 1.40 + 7.33 13.08 + 14.0
W
W
W
W
where P is the applied force, l is the arm of the bending moment, B is
the specimen thickness, and W is the specimen width. The error of the
geometry factor is less than 0.2% for a / W 0.6 . For the notched semicircular bending specimen, figure 4c (Krans 1995):

K = max a f (a / W )
max = 4.263 Pmax /( D B)

(7a)
(7b)
2

a
a
a
a
a
a
f = 0.623 + 29.29 171.2 + 457.1 561.2 + 265.54
W
W
W
W
W
W

(7c)
where max is the maximum stress, Pmax is the maximum applied force, D
is the specimen diameter, and B is the specimen thickness. Note, that
equations (7b) and (7c) apply to the case where the support span is equal
to 0.8 times the specimen diameter.
Fracture, i.e. unstable crack-growth, occurs if the stress, i.e. the stress
intensity factor, reaches a critical value. This is called the critical stress

Sec. 3

Theory

stress intensity factor (MPam)

plane stress

161

transition

plane strain

250

200

150

100

fracture toughness

50

0
1

10

100

specimen thickness (mm)

Figure 5. Dependence of the stress intensity factor on the specimen thickness.


[From: Ewalds and Wanhill 1984].

intensity factor, K c . It was found (for metals) that the critical stress
intensity factor depends on the thickness of the specimen, figure 5. Figure
5 shows that for a given thickness, K c tends to a constant value. This is
called the plane strain stress intensity factor, or fracture toughness, K Ic .
Since K Ic is independent of the specimen thickness, it is commonly
considered a true material property.
3.2 ASTM minimum size requirements
It is known that the specimen thickness has to be large with respect to the
size of the plastic zone otherwise plane stress will develop. It is known
from the literature, that the size of the plastic zone ahead of the crack-tip is
proportional to K Ic2 / ys2 , where ys is the yield strength, the generalised
minimum size requirement must be B K Ic2 / ys2 . It is normally
assumed that a valid plane strain K Ic is obtained if 2.5.

3.3 Fatigue crack-growth and creep crack-growth


Fatigue crack-growth is the process in which a crack originates and grows
under the influence of an alternating stress. Fatigue can cause fracture
even when the alternating stress is small, much smaller than the tensile
strength. The fracture mechanical characterisation of the crack-growth
behaviour of a material can be based on vast experience largely for metals
(Ewalds and Wanhill 1984, Broek 1986). It is usual to represent the crackgrowth behaviour in the form of a graph of the crack-growth rate, da / dN ,
versus the variation of the stress intensity factor, K , where
K = K max K min , as shown in figure 6. The da / dN - K -curve shows

162

Resistance to crack-growth and fracture

Ch. 4

10

region III

region II
7

log da/dN

da
= A( K ) n
dN

region I
1

Figure 6. The fatigue crack-growth rate


curve da / dN - K .

0
2,5

3,5

4,5

5,5

6,5

7,5

log K

three regions of crack-growth behaviour. In region II, which is of interest


to the present study, there is a linear log-log relationship between the
crack-growth rate, da / dN , and K , described by the Paris equation,
da
= A( K ) n
(8a)
dN
where a is the crack-length, N is the number of load repetitions,
K = K max K min , and A and n are constants. It was found that the
da / dN - K -relationship can be independent of the specimen geometry.
That is an indication that the constants A and n of equation (8a) can be
obtained as true material properties. Schapery was able to show that the
Paris equation can be derived from elementary principles (1973, 1975).
For crack-growth in a linear viscoelastic medium, he derived the following
expressions for the constants A and n :
(1 2 ) J 1

A=
6 R2 I 2
2

1/ m

[ w(t )]n dt

(8b)

n = 2/m
(8c)
n = 2 (1+ 1 / m)
(8d)
where I 1.5 , R is the tensile strength, is Poissons constant, J 1 is
the creep compliance at 1 s, is the fracture energy per unit fracture
surface, w(t ) is the normalised shape of the wave-form, and m is the
linear viscoelastic creep susceptibility, defined by
J (t ) = J 0 + J 1t m
(9)
where J (t ) is the creep compliance as function of the time, t , and J 0 is a
constant. The exponent of the Paris equation, n , is given by equation (8c)
or (8d), depending on the properties of the fracture process zone.

Sec. 3

Theory

163

Normally, the above equations are used to describe fatigue crack-growth,


i.e. frequency dependent crack-growth. However, it is known that crackgrowth can also take place under constant applied stress. This is called
creep crack-growth and can be described by the following equation:
da
da
= f
= A ( K ) n
(10)
dt
dN
More generally, the notion creep crack-growth is used to indicate that the
crack-growth is controlled by the time during which the (creep) stress is
present. Thus, different mechanisms of crack-growth appear to be
described by a Paris type of equation.
3.4 Indirect methods of characterising the resistance to crack-growth
Based on equation (8), it is expected that the resistance to crack-growth
can be characterised based on the properties upon which A and n depend:
i.e. the complex modulus, the creep compliance, the tensile strength and
the fracture energy. An advantage of such an indirect characterisation is
that simpler, more practical tests can be used. The complex modulus and
the creep compliance were discussed in the previous chapter. The fracture
toughness and the tensile strength are considered in the present chapter. To
investigate the possibilities of obtaining the fracture toughness and the
tensile strength, tests were performed using the semi-circular bending test,
the uniaxial tensile test and the indirect tensile test. The fracture energy
was not studied.
3.5 Definitions of tensile strength
The tensile strength is defined in a uniaxial state of stress as the maximum
force per unit original cross-section in which the force operates. The
tensile strength is calculated assuming the material is homogeneous and
linearly elastic.
In the uniaxial tensile test, the tensile stress is uniformly distributed,
provided the specimen is slender. Then, the uniaxial tensile strength can be
calculated as:
P
UT = max
(11)

where UT is the uniaxial tensile strength, Pmax is the maximum applied


force, and is the specimen cross section.
In the indirect tensile test, the tensile stress in the x -direction
perpendicular to the vertical is known to be approximately uniform.
Therefore, the indirect tensile strength can be calculated as (Hondros
1959):
2 Pmax
(12)
IT =
DB
where IT is the indirect tensile strength, Pmax is the maximum applied

164

Resistance to crack-growth and fracture

Ch. 4

force, D is the specimen diameter, and B is the specimen thickness.


In the semi-circular bending test, the tensile stress is inhomogeneously
distributed. The maximum tensile stress occurs where the bending moment
is largest, normally at the middle of the specimen base, unless a greater
stress concentration develops elsewhere in the specimen owing to material
imperfections. The maximum stress at the specimen base is expected to
give an estimate of the tensile strength of the material. Assuming linear
elastic behaviour, it is calculated as (Van de Ven et al. 1997):
4.263 Pmax
BT =
(13)
DB
where BT is the bending tensile strength.

4 Experimental details
Six test methods were investigated, cf. 4.1. The influences of the stress
condition, the shape of the waveform of the applied stress, and the
specimen geometry (size and shape) were investigated, cf. 4.2. Also the
influence of the mixture composition was investigated, cf. 4.3.
4.1 Test methods
The following six test methods were investigated:
1 a dynamic crack-growth test, using a centre-cracked tensile specimen
(CCT);
2 a dynamic crack-growth test, using a single edge notched four point
bending specimen (4PB);
3 the SCB-test, using a notched half-cylindrical specimen (SCB/N);
4 the uniaxial tensile test (UTT), using a cylindrical specimen;
5 the semi-circular bending (SCB) test, using an un-notched half-cylindrical specimen (SCB/U);
6 the indirect tensile test (ITT), using a cylindrical specimen.
The CCT-test and the 4PB-test were used to investigate the applicability of
Paris Law and the linear elastic stress intensity factor to asphalt mixtures.
The SCB-test using the notched specimen was used to determine the
fracture toughness. Three tensile tests were used to determine the tensile
strength: the UT-test, the SCB-test with un-notched specimen, and the ITtest.
The CCT-specimen and the 4PB-specimen were selected based on a
literature study. It was found that the following specimen geometries had
been used:
. bending beam on elastic base (Majidzadeh et al. 1971, Molenaar 1983);
. double edge notched uniaxial tensile specimen (Molenaar 1983, Jacobs
1995);
. three point bending beam (Tangella et al. 1990);
. four point bending beam (Krans et al. 1993).

Sec. 4

Experimental details

165

All of these specimens have various advantages and disadvantages. With


the beam on an elastic base, the crack-length is calculated from the
compliance. To do that, it is necessary to know the friction between the
beam and the base. Since this is not known and has to be estimated, an
error in the calculation of the crack-length is unavoidable. Despite that,
Majidzadeh was able to determine the constants A and n of the Paris
equation, finding values of the exponent of the Paris equation, n , between
3 and 4 for different mixtures at temperatures between 0C and 25C.
In order to avoid the friction problem, Molenaar glued the asphalt beam to
the elastic base. Then, he found that the crack grew over a certain distance,
but stopped when the crack-tip approached the compression zone, caused
by bending of the specimen. For that reason, Molenaar judged this test as
less suitable for crack-growth studies.
In the double edge notched uniaxial tensile specimen, a bending moment
may arise because two cracks grow from the two notches on opposite sides
of the specimen, which, however, may not grow equally fast. As a result, it
is more difficult to calculate K . In this specimen the crack-length is
calculated from the crack opening displacement, C.O.D., (Jacobs 1995),
and from the compliance (Molenaar 1983).
In the three point bending beam, the crack may be found propagating in a
different plane than the vertical plane through the point of load
application. That causes an error in the calculation of the stress intensity
factor, K , also because the geometry factor, f ( a / W ) , varies significantly
as a function of the crack-length. Another disadvantage is that the beam
bends owing to creep. This is unavoidable because of the continuation of
the bending moment on the beam in the same direction. In principle, the
creep causes a violation of the condition of linear elasticity, which
possibly invalidates the applicability of the stress intensity factor. The
increasing curvature of the beam owing to creep causes the calculation of
K to be more difficult.
In the four point bending beam, the stress at the crack-tip is defined better,
but the geometry factor still changes significantly. The single edge
notched four point bending beam was preferred because of the better
defined stress at the crack-tip. But because of the likelihood of errors
caused by the bending moment and the variation of the geometry factor, it
was decided to perform also CCT-tests. There are the following
advantages of using the CCT-specimen:
. the stress intensity factor, K , is well defined, because the specimen is
under a uniaxial tensile stress1; therefore a bending moment remains
1

This is probably true if the test material is a metal; however, in the case of an
asphalt mixture this is questionable because the material is heterogeneous and
normally elasto-viscoplastic.

166

Resistance to crack-growth and fracture

Ch. 4

25
11

a = 30

Figure 7. The starter notch of the


CCT-specimen; measures in mm.

absent as long as the crack grows symmetrically;


. the crack-growth path is long, because the specimen is wide;
. the geometry factor f ( a / W ) does not change much as the crack grows.
Disadvantages of the CCT-specimen are:
. the specimen shape is relatively complicated;
. the starter notch is difficult to make;
. the specimen alignment is relatively complicated.
Note, that at the time of the experiments, the CCT-geometry had not been
used before in investigations into crack-growth of asphalt mixtures.
The uniaxial tensile test, the semi-circular bending test, and the indirect
tensile test are well-documented tests that were performed according to
RHEI test prescriptions.
4.2 Test set-up and testing conditions Influences of stress condition,
shape of wave-form of applied stress, and specimen geometry
4.2.1 Crack-growth test using the centre-cracked tensile (CCT) specimen
The dynamic crack-growth test, using a CCT-specimen, is described in the
ASTM-standard E 647-93 (ASTM 1993). The test is shown schematically
in figure 4a. The CCT-specimen is a prismatic slab, 230 x 390 mm (width
x height); two thicknesses were used: 30 mm and 60 mm. The starter
notch was made from a circular hole, figure 7. From this hole a normal
edge notch was sawn, 3 mm wide and about 14 mm deep, using a tungsten
carbide blade. Subsequently, a cut was made, about 0.5 mm, using a sharp
knife. In total the notch was 25 mm in length. The crack-growth rate was
not determined until the crack-length was 30 mm. The 5 mm prefatiguing was done to obtain a sharp crack-tip; the stress intensity factor
is defined for a sharp crack-tip. The end of the crack-growth path is
determined by the geometry factor, which is accurate up to 2a / W = 0.8 ,
provided the ligament is still elastic. Thus, the crack-growth path for

Sec. 4

Experimental details

R = 0.1
1.5

applied stress

applied stress

R = -0.5
1.5

0.5
0

-0.5

167

0.5

-0.5

time

time
R = 0.5

Figure 8. Schematic representation of


the applied stress for different R-values.
Top left (a): R = -0.5. Top right (b): R
= 0.1. Below (c): R = 0.5.

applied stress

1.5
1
0.5
0

-0.5

time

which K is calculable ranges between 30 mm and 90 mm from the centre


of the circular hole. The specimen was painted white and a millimetre
mesh was carved on it, to allow the crack-length to be determined by
visual observation. This method is demanding, but is believed to produce
reliable measurements of the crack-length2. The crack-length was
measured at the front-side and at the back-side of the CCT-specimen. The
average value was used as the crack-length. This method was found more
accurate than the method using a crack-foil. In particular, when a thick
specimen is used, different crack-lengths may be observed at the front-side
and the backside because the material is heterogeneous and the specimen
alignment may not be perfect. The slab was subjected to a sinusoidal
stress. Different R -values were used: 0.5, 0.1 and 0.5, illustrated by
figure 8. Note that R is defined here as
2

A different method uses crack-foils (Krans et al. 1993). This method was not
used in the present study, because it is very expensive, and because there was
uncertainty as to whether the length of the crack in the crack-foil corresponded
to the actual length of the crack in the specimen. An advantage of the method is
that the crack-length causes an electrical signal, permitting feedback, and
creating the possibility of performing tests with constant K , cf. 5.2.2.

168

Resistance to crack-growth and fracture

R = min / max = K min / K max

Ch. 4

(14)

where min is the minimum stress, max is the maximum stress, K min is
the minimum stress intensity factor, and K max is the maximum stress
intensity factor. Three frequencies were used: 1 Hz, 10 Hz, and 29.3 Hz.
The tests were performed at a temperature of 0C.
The ASTM-standard poses the following requirements:
. the stress state in the ligament must be elastic at the applied stress;
. no rounding off of the crack-tip is allowed;
. the ratio of the specimen length (height) to its width, L / W , must be
greater than 1.2,
L / W > 1 .2
(15a)
. the specimen thickness, B , must be between W / 20 and W / 4 ,
W / 20 < B < W / 4
(15b)
. the following condition must be satisfied:
(W 2 a ) 1.25 Pmax /( B ys )
(15c)
where ys is the yield strength of the material. Care was taken to follow
the ASTM-standard; however, this was not possible at all points, because
the standard was not made for asphalt mixtures. The asphalt specimen
deviates from the specimen according to the ASTM-standard in the
following aspects:
. the stress state of the ligament cannot be purely elastic, because an
asphalt mixture is a viscoelastic material;
. possible rounding off of the crack-tip cannot be verified; different models
assume the formation of micro-cracks or viscoplastic deformation at the
crack-tip; the formation of crazes as in synthetic organic construction
materials is unlikely because of the lower molecular weight of bitumen;
. the yield strength, ys , is not defined for a viscoelastic material, since
such a material deforms permanently at any stress, even the lowest, if there
is only sufficient time; the viscous deformation depends on the applied
stress, the temperature and the loading-time.
The crack must pass the aggregates in an asphalt mixture one by one. The
crack-growth rate may fluctuate increasingly as the maximum aggregate
size increases. In order to reduce the chances that individual large
aggregates cause a temporary stop of the crack, it is desirable to have a
large specimen thickness in comparison to the maximum aggregate size.
Therefore, two specimen thicknesses, 30 mm and 60 mm, were chosen in
combination to a maximum aggregate size of 2 mm. The material is
described in 4.3.1.
With the specimen dimensions as chosen, the ASTM condition L / W >
1.2 was satisfied ( L / W was actually greater than 1.5), and the ASTM
condition , W / 20 < B < W / 4 , was satisfied for the 30 mm slabs. For the
60 mm slabs, B / W was equal to 0.26; this was still acceptable because

Sec. 4

Experimental details

169

the ASTM-standard permits use of specimens for which B / W is as large


as 0.50, if the crack front stays straight.
Table 1. Investigation of ASTM minimum size requirement B 2.5 K Ic2 / ys2
CCT-specimen
4PB-specimen
ys (MPa)
5
10
5
10
K c (MPam)
0.97
0.97
1.47
1.47
2
2
B 2.5 K Ic / ys (mm)
B 94
B 24
B 216
B 54

In table 1, the requirement B 2.5 K Ic2 / ys2 is investigated. The yield


strength is not defined for a viscoelastic material. However, it is known
from fracture toughness tests and tensile tests, cf. 6 and 7, that plasticity is
negligible at 0C. Then it is reasonable to assume that the yield strength is
equal to the tensile strength. A representative value of the tensile strength
of asphalt concrete at 0C is 5 MPa, cf. 7, figures 36 and 38. Observed
values range between 0 MPa and 10 MPa, cf. 7. Therefore, a
representative maximum value is 10 MPa. A representative average value
of K c is 1 MPam for the CCT-specimen, cf. table 8, and 1.5 MPam for
the 4PB-specimen, cf. table 9. Table 1 shows, that the condition for plane
strain is not fulfilled for a tensile strength equal to 5 MPa (minimum
required CCT specimen thickness is 94 mm; minimum required 4PB
specimen thickness is 216 mm), but is fulfilled for a tensile strength equal
to 10 MPa (minimum required CCT specimen thickness is 24 mm;
minimum required 4PB specimen thickness is 54 mm).
Table 2. Investigation of ASTM minimum size requirement
W 2a 1.25 Pmax /( B ys ) .
CCT-specimen
B (mm)
30
60
W (mm)
230
230
2a (mm)
60
60
Pmax (N)
42000
42000
ys (MPa)
5
5
W 2a 1.25 Pmax /( B ys ) (mm)
170 <350 170 < 175

In table 2, the requirement W 2a 1.25 Pmax /( B ys ) is investigated.


Table 2 shows, that the condition is not fulfilled for a tensile strength
equal to 5 MPa ( W 2a = 170 mm is smaller instead of greater than
1.25 Pmax /( B ys ) , 350 mm for the 30 mm (thickness) specimen, and 175
for the 60 mm specimen. If the yield strength is assumed to be 10 MPa,
then the condition is fulfilled. Thus, applying the minimum size
requirements, the stress state of the specimen may not be purely plane

170

Resistance to crack-growth and fracture

Ch. 4

strain, cf. figure 5. This allows a possible explanation for the higher value
of K c of the 4PB-specimen as compared to the CCT-specimen: The
geometry dependence of the material properties can be ascribed to the
three-dimensional (non-uniform) stress state of the material, caused by the
materials heterogeneity, its stress dependent (nonlinear) stress strain
behaviour, and the extension of the plastic zone.
Summarising, the ASTM-requirements with respect to the specimen
dimensions can be satisfied (if the yield strength is greater than 5 MPa),
but not the ASTM-conditions with respect to the linear elastic behaviour.
This is also true in the case of organic construction material and concrete
if an estimate of the size of the plastic zone in front of the crack-tip cannot
be made.
4.2.2 Crack-growth test using the four point bending (4PB) specimen
The test is shown schematically in figure 4b. The specimen is a prismatic
beam, 450 x 60 mm (length x height, i.e. width); two thicknesses were
used: 30 mm and 60 mm. The starter notch was made using a thin, 0.5
mm, water-cooled saw blade. It was made 5 mm deep. The crack-growth
rate was not measured until the notch had grown 2 mm on both the frontside and the backside of the specimen. It was assumed that the crack-tip
was sharp then and suitable to calculate the stress intensity factor. The end
of the crack-growth path is determined by the geometry factor, which is
accurate up to a / W = 0.6 , provided the ligament is still elastic. Thus, the
crack-growth path, for which K is calculable, ranges between 7 mm and
36 mm from the lower surface of the specimen. The specimen was painted
white and a millimetre mesh was carved on it, to allow the crack-length to
be determined by visual observation. The crack-length was determined on
the front-side and the backside of the specimen. The average value was
used as the crack-length.
The beam was subjected to a sinusoidal stress. Different R -values were
used: 0.5, 0.1 and 0.5. Three frequencies were used: 1 Hz, 10 Hz, and
29.3 Hz. The tests were performed at a temperature of 0C.
4.2.3 Determination of the fracture toughness test using the semi-circular
bending (SCB) specimen
The test is shown schematically in figure 4c. The specimen is a halfcylindrical notched specimen in three point bending. The notch, 0.3 x 10
mm, was sawn at the middle of and perpendicular to the specimen base.
The mode I stress intensity factor can be calculated as (Lim et al. 1993):
K I = 0 a YI
(16a)
where

Sec. 4

Experimental details

171

P0
(16b)
DB
where P0 is the pertinent force, D is the specimen diameter, B is the
specimen thickness, and a is the crack-length, i.e. the length of the notch,
and YI is the normalised mode I stress intensity factor. From equation
(16a), it follows that YI can be defined as
KI
(16c)
YI =
0 a

0 =

YI was calculated for the SCB-specimen by Lim et al. These calculations


were found useful as equation (7c) can be used only if the ratio of the
support span and the specimen diameter is equal to 0.8.
6
5.5

Y-I

5
4.5
4
3.5

Figure 9. Normalised stress intensity


factor, YI , for a/r = 0.1 and 0.2 (a is
the notch length, r the specimen
radius). [Based on Lim 1993].

3
0.65

0.75

0.85

s/r

Y-I(a/r=0.1)

Y-I(a/r=0.2)

Figure 9 shows YI as a function of the ratio of the support span, 2 s , and


the specimen diameter, 2r , or s / r , for a / r = 0.1, and a / r = 0.2, based
on data given by Lim et al. For the parameters s / r and a / r of the
specimen geometries employed in the present study, YI was determined by
interpolation or extrapolation of the lines in figure 9. The values obtained
are shown in table 3. Thus, in equation (16a), the pertinent stress, 0 ,
Table 3. Normalised mode I stress intensity factors for the SCB
specimen
Specimen
diameter
(mm)
100
150
220

support
span
(mm)
80
130
200

s/r

a/r

YI

0.8
0.867
0.909

0.200
0.133
0.091

4.587
5.117
5.390

172

Resistance to crack-growth and fracture

Ch. 4

15

15
15

load, P

FRACTURE

load, P

load, P

FRACTURE

FRACTURE

0
0

15

displacement

0
0

15

displacement

15

displacement

Figure 10. Principal types of load-displacement diagram obtained during K Ic testing for metals according to the ASTM-standard.

the crack-length, a , and the normalised mode I stress intensity factor, YI ,,


are known (it is assumed that the crack-length at the moment of fracture is
equal to the length of the notch), so that K I can be calculated.
4.2.3.1 The ASTM-method for metals
Plots of load versus displacement may have different shapes. For metals,
according to the ASTM-standard E399-78A (ASTM 1979), the types of
diagram that can be obtained are shown schematically in figure 10.
Initially, the displacement increases linearly with the load, P . In many
cases there is either a gradually increasing nonlinearity, figure 10a, or
sudden crack extension and arrest, called pop-in, followed by nonlinearity, figure 10b. The nonlinearity is caused by plastic deformation and
stable crack-growth before fast fracture. If a material (metal) behaves
almost perfectly elastically, a diagram like that in figure 10c is obtained.
To establish whether a valid K Ic can be obtained from the test it is first
necessary to calculate a candidate value, K Q . This involves a graphical
construction of the load-displacement record. Then it must be determined
whether this K Q is consistent with the specimen size and material yield
strength according to the following equations,
a 2.5 (K Q / ys ) 2
(17a)
B 2.5 (K Q / ys ) 2

W 5 (K Q / ys )

(17b)
(17c)

If K Q meets these requirements, then K Q = K Ic . If not, the test is invalid


and the result may be used only to estimate the fracture toughness; it is not
an ASTM standard value.

Sec. 4

173

Experimental details

15

15
15

A
FRACTURE

load, P

FRACTURE

load, P

III

II

PS

15

displacement

PS
FRACTURE

0
0

load, P

PS

0
0

0 displacement

15

0 displacement

15

Figure 11. Determination of the pertinent load, P0 . P0 is equal to the maximum


load between the tangent line and the secant line. PS is the load at the intercept
of the secant line and the test record.

As shown in figure 10, the general types of force-displacement diagram


exhibit different degrees of nonlinearity. Various criteria to establish the
load corresponding to K Ic were considered. A 5% secant offset was
chosen to define K Ic as the fracture toughness. Although this definition is
apparently arbitrary, it turned out that for the standard ASTM specimens
the effects of plasticity and stable crack-growth were more or less
accounted for by assuming an effective length increase of 2%. The
procedure for determining the load corresponding to K Ic then consists of
drawing a secant line from the origin 0 with a slope 5% less than that of
the tangent 0A to the initial part of the test record, figure 11. The load PS
is the load at the intercept of the secant with the test record. The pertinent
force, P0 , is defined according to the following procedure. If the load at
every point on the force-displacement-curve which precedes PS is lower
than PS , then P0 is PS , figure 11 type I. If there is a maximum load
preceding PS that is larger than PS , then this maximum load is P0 , figure
11 types II and III. In order to prevent acceptance of a test record for a
specimen in which excessive stable crack-growth occurred, it is required at
the present stage that Pmax / P0 be less than 1.10. Hence,
Pmax
1.10
(17d)
P0
The value of this ratio is based on experience. After having determined
P0 , the value of K Q is calculated using the appropriate stress intensity
factor expression, equation (16), for the SCB-specimen. The quantity
2.5 ( K Q / ys ) 2 must then be found to be less than the thickness B , and
crack-length, a , of the specimen for K Q to be a valid K Ic .

174

Resistance to crack-growth and fracture

15

Ch. 4

15

15

PLASTIC
COLLAPSE

load, P

load, P

FRACTURE

0
0

15

displacement

load, P

STRAIN
HARDENING

15

displacement

15

displacement

Figure 12. Principal types of load-displacement diagram obtained for asphalt


mixture in the SCB-test as K Ic -test. Left (a): Elastic fracture at -10C or 1C.
Middle (b): Plastic collapse at 15C; no strain hardening. Right (c): Strain
hardening at 25C or higher temperature.

Figure 13. Method for the determination the pertinent force, P0 .

4.2.3.2 A modified ASTM-method for asphalt mixture


In the present study, also three principal types of load-displacement
diagram were observed for asphalt mixture. At low temperature (1C or
lower), the asphalt mixture behaves practically perfectly elastically i.e.
no plastic deformation is visible on the diagram, figure 12a3. At higher
temperatures, as shown in figures 12b and 12c, the general types of forcedisplacement-curves exhibit different degrees of nonlinearity. Figure 12b
3

A temperature of 1C was preferred rather than 0C to avoid ice formation on


and in the specimen.

Sec. 4

Experimental details

175

shows plastic stress strain behaviour without strain hardening. A real


example is shown in figure 13. Figure 12c shows plastic stress strain
behaviour and significant strain hardening.
The ASTM-procedure for the determination of K Ic described in the
previous section was slightly modified. To account for the heterogeneity
of asphalt specimens, a slightly more tolerant method to determine the
pertinent force was adopted, illustrated in figure 13. Figure 13 is typical
for a force-displacement-curve showing plastic collapse. First, the tangent
to the elastic part of the force displacement curve is drawn. Then, a secant
line, with a 10% lower slope, is drawn. The pertinent force, P0 , is
determined as the maximum force reached between these lines. The
fracture toughness is calculated using equation (16a). For convenience, the
calculated value, i.e. prior to the validity test according the ASTMstandard, is called the apparent fracture toughness, K IQ .
According to Lim et al. (1994), the minimum specimen size conditions,
equations (17), yield a conservative estimate of the minimum specimen
size required. Research referenced by Lim et al. indicates that a smaller
specimen size can be used to obtain a valid fracture toughness, because the
fracture process zone in the SCB specimen is relatively small.
The following specimen dimensions were used (diameter x thickness, in
mm):
100 x 25
100 x 50
100 x 75
150 x 25
150 x 50
150 x 75
220 x 25
220 x 50
220 x 75
The support span was equal to the specimen diameter minus 20 mm;
hence, with the diameters, 100 mm, 150 mm, and 220 mm the support
span was respectively: 80 mm, 130 mm, and 200 mm, see also table 3
(first and second column). The notch-length was constant, 10 mm; the
notch-width 0.35 mm. The test temperatures were -10C, 0C, 15C, and
25C, and the displacement rates 0.005 mm/s, 0.05 mm/s, and 0.5 mm/s.
The test programme in table 4 was conducted on unmodified dense graded
asphalt concrete, DAC 0/16. The test programme in table 5 was conducted
on polymer modified dense graded asphalt concrete, DAC 0/16, porous
asphalt 0/16, and stone mastic asphalt 0/11.
4.2.4. Three tensile tests
Three tensile tests were used for the determination of the tensile strength.
The aim was to determine the tensile strength using different specimen
geometries and to investigate whether the tensile strength can be obtained
as a true material property. The tests are shown schematically in figure 14.
The uniaxial tensile (UT) test was selected because of the stress state of

176

Resistance to crack-growth and fracture

Ch. 4

applied load

applied load

applied load

Figure 14. Top (a): uniaxial tensile test. Middle (b): semi-circular bending test
(three point bending test). Bottom (c): Indirect tensile test.

Sec. 4

Experimental details

177

Table 4. Test programme DAC 0/16, unmodified, for the determination of K IQ .


apparent fracture
indirect tensile
toughness
strength
100x25 150x25 220x25
specimen
dimensions
150x50
100x50 150x50 220x50
(D x B, mm)
100x75 150x75 220x75
temperature (C)
-10, 0, +15, +25
+15
displacement rate (mm/s)
0.005, 0.05 0.5
0.5
D = diameter, B = thickness.
Table 5. Test programme DAC 0/16, PA 0/16, SMA 0/11, elastomer modified,
and plastomer modified for the determination of K IQ .
Apparent fracture
indirect tensile
toughness
strength
specimen
150x50
dimensions
150x50
(D x B, mm)
temperature (C)
-10, +5, +15
+15
displacement rate (mm/s)
0.5
0.5
D = diameter, B = thickness.

the specimen, which, in principle, is uniaxial. However, this is true only


for a slender specimen and a homogeneous material. In reality, the test
material, an asphalt mixture, is heterogeneous. The effect of the specimen
geometry on the stress distribution has not yet been investigated in detail.
The semi-circular bending (SCB) test was selected because this has been
developed as a possible improvement of the indirect tensile (IT) test
(Krans et al. 1993, Van de Ven et al. 1997), the advantage being that the
localised plastic deformation near the point of load application is reduced
strongly (because a lower applied force is required).
The indirect tensile (IT) test was selected because this used to be the
traditional test to determine the tensile strength of asphalt mixtures. A
disadvantage of this test is that the observed tensile strength may be too
low owing to localised plastic deformation near the point of load
application. Each test was performed in fourfold for each testing
condition.
4.2.4.1 Uniaxial tensile (UT) test
The UT test was used to determine the uniaxial tensile strength. The
following testing conditions were used:
specimen height:
100 mm or 200 mm
specimen diameter:
50 mm
test temperature:
-5, 0, 15, and 25C
displacement rate:
various rates ranging between 0.01 mm/s
and 0.5 mm/s.

178

Resistance to crack-growth and fracture

Ch. 4

Table 6. Overview of mixtures and investigated tensile properties.


Sec- type of asphalt mixture
tion

investigation

laboratory manufactured
sand asphalt 0/2

laboratory manufactured
DAC 0/16 with 80/100 bitumen
DAC 0/16 with elastomer modified bitumen
DAC 0/16 with plastomer modified bitumen
PA 0/16 with elastomer modified bitumen
PA 0/16 with plastomer modified bitumen
SMA 0/11 with elastomer modified bitumen
SMA 0/11 with plastomer modified bitumen
laboratory manufactured
PA 0/16 with 4.0% 70/100 bitumen
PA 0/16 with 5.0% 70/100 bitumen
PA 0/11 with 4.0% 70/100 bitumen
PA 0/11 with 5.0% 70/100 bitumen
PA 0/16 with 4.0% elastomer modified bitumen
PA 0/16 with 5.0% elastomer modified bitumen
PA 0/11 with 4.0% elastomer modified bitumen
PA 0/11 with 5.0% elastomer modified bitumen
PA 0/16 with 4.0% plastomer modified bitumen
PA 0/16 with 5.0% plastomer modified bitumen
PA 0/11 with 4.0% plastomer modified bitumen
PA 0/11 with 5.0% plastomer modified bitumen
laboratory manufactured and plant-mixed
5 projects: road trials; 5 mixtures per project
selected mixture: crushed gravel asphalt concrete 0/22
1 mixture according to Marshall mixture design (MD),
with 4.5% 70/100
2 mixture MD minus 0.5% bitumen (MD-)
3 mixture MD plus 0.5% bitumen (MD+)
4 mixture, plant-mixed, slab-compacted at the plant
(PLA)
5 mixture, plant-mixed, laid in the pavement (PAV)
laboratory manufactured
gravel asphalt concrete 0/32 with 4.5% 45/60 bitumen
dense asphalt concrete 0/16 with 6.5% 80/100 bitumen

crack-growth
properties,
A , n , Kc
as function of
frequency and
applied stress
K Ic , ITS, BTSN,
as function of
specimen size,
deformation rate,
test temperature,
mixture composition

UTS, BTS, BTSN,


as function of
mixture composition

ITS, UTS, BTS,


BTSN, K Ic
as function of
test temperature,
mixture composition
and manufacturing
conditions

ITS as function of
specimen size,
loading strip width,
test temperature,
mixture composition
DAC: dense asphalt concrete, PA: porous asphalt, SMA: stone mastic asphalt, A :
constant of the Paris equation, n : exponent of the Paris equation, K c critical stress
intensity factor, K Ic : plane strain critical stress intensity factor, or fracture
toughness, BTS: bending tensile strength, BTSN: bending tensile strength of
notched specimen, ITS: indirect tensile strength, UTS: uniaxial tensile strength.

Sec. 4

Experimental details

179

4.2.4.2 Semi-circular bending (SCB) test


The SCB-test was used to determine the bending tensile strength. The
following testing conditions were used:
specimen diameter:
150 mm
specimen thickness height: 50 mm
test temperature:
-5, 0, 15, and 25C
displacement rate:
various rates ranging between 0.01 mm/s
0.10 mm/s

Two specimens were used: an un-notched specimen, and a notched


specimen. The notch, 0.35 x 10 mm, is at the middle of the specimen base
along the vertical.
4.2.4.3 Indirect tensile (IT) test
The indirect tensile test was used to determine the indirect tensile strength.
The following testing conditions were used:
specimen diameter:
97 mm, for dense asphalt concrete 0/16
148 mm, for gravel asphalt concrete 0/32
specimen thickness:
25, 50, and 75 mm
test temperature:
-5, 0, 10, and 20C
width of loading strip:
0.5, 1 inch
displacement rate:
0.85 mm/s
4.3 Materials Influence of the material composition
An overview of the asphalt mixtures used, and the properties determined
are given in table 6. Sections A, B, C and E in table 6 refer to laboratory
manufactured, slab-compacted mixtures. Section D in table 6 refers to 5
projects (road trials).
4.3.1 Sand asphalt tested in the CCT-tests and 4PB-tests
The test material for the CCT-tests and 4PB-tests was a sand asphalt with
a maximum aggregate size of 2 mm, cf. table 6, section A. The mixture
contained 75% crushed sand, 5% river sand, 20% Wigro filler, and 8.5%
45/60 bitumen. The average voids content of this mixture was 3.9%;
minimum: 3.5%, maximum 4.6%. The main reason for selecting this
material is the maximum grain size, only 2 mm, that made it a relatively
homogeneous material.
4.3.2 Asphalt mixtures tested in the SCB fracture toughness test
and indirect tensile test
Seven standard types of asphalt mixture were investigated: unmodified
dense asphalt concrete, DAC 0/16, elastomer modified DAC 0/16,
plastomer modified DAC 0/16, elastomer modified porous asphalt, PA

180

Resistance to crack-growth and fracture

Ch. 4

0/16, plastomer modified PA 0/16, elastomer modified stone mastic


asphalt, SMA 0/11, and plastomer modified SMA 0/11. The mixtures are
indicated in table 6, section B.
4.3.3 Asphalt mixtures used in the SCB tensile tests, uniaxial tensile test
and indirect tensile test
Section C in table 6 represents an investigation in which, for porous
asphalt, the uniaxial tensile strength, the bending tensile strength of the
un-notched SCB-specimen, and the bending tensile strength of the notched
SCB-specimen are compared. Two mixture types, with gradings 0/11 and
0/16, with respectively 4.0% bitumen and 5.0% bitumen were
investigated. The bitumens used were pure 70/100 bitumen, an elastomer
modified bitumen, and a plastomer modified bitumen.
Section D in table 6 refers to 5 projects (road trials), for which the tensile
strengths of the following mixtures were compared:
1 the mixture according to the Marshall mixture design method (MD)
2 the MD mixture minus 0.5% bitumen
3 the MD mixture plus 0.5% bitumen
4 the mixture from the plant, slab-compacted at the plant
5 the mixture from the road pavement.
Five projects (road trials) were realised, in which the following amounts of
reclaimed asphalt were used: road trial A: 50% m/m, road trial B: 40%
m/m, road trial C: 40% m/m, road trial D: 45% m/m, road trial E: 0%
m/m.
4.3.4 Asphalt concrete mixtures used in the indirect tensile test
The materials used were a dense asphalt concrete 0/16 mixture, DAC 0/16,
and a gravel asphalt concrete 0/32 mixture, GAC 0/32.

Sec. 5

181

Results and analysis of crack-growth tests

5 Results and analysis of crack-growth tests


5.1 Obtaining a graph of da / dN versus K from raw test data
In the CCT-test, the crack-length was determined on the front-side and on
the back-side of the specimen. Four readings per data record were
obtained. In table 7, these are indicated as crack-length front-left, cracklength front-right, crack-length back-left, and crack-length back-right.
Typically, approximately 80 data records per test were obtained.
Table 7. Scheme of raw data records as obtained from the CCT-test1.
a front left
a front right
a back left
a back right
record N
N1

a front left ,1

a front right ,1

a back left ,1

a back right ,1

N2

a front left , 2

a front right , 2

a back left , 2

a back right , 2

N3

a front left ,3

a front right ,3

a back left ,3

a back right ,3

Ni

a front left ,i

a front right ,i

a back left ,i

a back right ,i

Front-left refers to the front-side of the specimen, and to the left hand crack-tip of
the crack, cf. figure 6; front-right refers to the front-side of the specimen, and to the
right hand crack-tip of the crack; similarly, back-left and back-right refer to the
back-side of the specimen

The average of the four crack-length parameters was plotted versus the
number of load cycles, N . In the 4PB-test, two readings per data record
were obtained: the crack-length on the front-side and that on the back-side
of the specimen. Four methods for obtaining a graph of da / dN versus
K from the raw test data were investigated. The methods lead to
different results. The methods are described in the following.
Method 1. Calculation of da / dN according to the secant method
The crack-growth rate, da / dN is defined as (ASTM 1993)
a ai
da / dN = i +1
(18)
N i +1 N i
The variation of the stress intensity factor is defined by the equations
K = ( ) a f ( a / W )
(19a)
= max min
(19b)
K was calculated by substituting a = ( ai + ai +1 ) / 2 into equation (19a).
Figure 15a shows two a (N ) -curves from two tests on different specimens
tested under the same testing conditions. The corresponding data points
( K , da / dN ) are shown in figure 15b. Method 1 has two disadvantages:
1 the wide scatter band of da / dN versus K ;

182

Resistance to crack-growth and fracture

Ch. 4

120

crack-length, a (mm)

100
80
60
40

test #1
test #2

20
0
0

100000

200000

300000

400000

500000

600000

number of cycli, N

da/dN (m/cycle)

10

0.1

test #1
test #2
0.01
0.1

K (MPam)

Figure 15. Top (a): Raw data of the crack-length a versus the number
of load cycles, N . Bottom (b): Data points ( K , da / dN ) according
to method 1 (secant method).

2 if ai = ai +1 , then da / dN is equal to zero. Zero values cannot be plotted


in a log-log-plot of da / dN versus K . For the curve of test #1,
approximately 40%, and for the curve of test #2, approximately 30% of
the da / dN -values were found to be zero.
Therefore, the resulting da / dN versus K plot obtained by this method
might not be a good representation of the real da / dN ( K )-relationship.
Method 2. Polynomial fit of a versus N
A sixth order polynomial was fitted to the a (N ) -curve of figure 15a. The

Sec. 5

Results and analysis of crack-growth tests

183

120

crack-length, a (mm)

100
80
60
40

test #1
test #2

20
0
0

number of cycli, N:100,000

da/dN (m/cycle)

10

0.1

test #1
test #2
0.01
0.1

K (MPam)

Figure 16. Top (a): A sixth order polynomial fitted to the raw data of
figure 15a. Bottom (b): Data points ( K , da / dN ) according to
method 2 (polynomial method).

result is shown in figure 16a. The derivative of the polynomial is da / dN .


Since the polynomial is continuous, the derivative is also a continuous.
Hence, K can be determined in all points of the polynomial. K was
calculated by substituting a = ( ai + ai +1 ) / 2 into equation (19a). The
corresponding data points ( K , da / dN ) are shown in figure 16b. Method
2 has the following two disadvantages:
1 the predicted da / dN at greater K is too low; this is because the steep
end of the a (N ) -curve cannot be fitted properly; a similar effect was
obtained using a 9th order polynomial (results not shown).

Resistance to crack-growth and fracture

184

Ch. 4

da/dN (m/cycle)

10

0.1

test #1
test #2
0.01
0.1

K (MPam)

Figure 17. Result obtained according to method 3 (which involves a


smoothing of the a (N ) -curve). The result can be compared to that
in figure 14b, which shows the result according to method 1.

2 the assumption of an n-th order polynomial fit to a (N ) is a precondition


which influences the derivative da / dN . The effect may be that trends are
introduced which do not actually exist.
Therefore, probably, method 2 does not provide a good representation of
the real da / dN ( K )-relationship.

Method 3. Smoothing of a versus N


The a (N ) -curve was smoothed using the following formula.
ai = ( ai 3 + 2 ai 2 + 3 a i 1 + 4 ai + 3 ai +1 + 2 ai + 2 + ai +3 ) / 16

(20)

Thus a smoothed a (N ) -curve was obtained. Subsequently, da / dN was


calculated using equation (18). K was calculated by substituting
a = ( ai +1 + ai ) / 2 into equation (19a). The corresponding data points
( K , da / dN ) are shown in figure 17. Method 3 has two advantages with
respect to method 1:
1 there are no zero values of da / dN ;
2 the scatter on the da / dN -values is reduced (but still significant).

Method 4. Local averaging using a constant step


The method is illustrated by figure 18a. A minimum, constant step d in
the crack-length is defined by
a ai
da / dN = k
(21)
with k > i and a k ai d
Nk Ni
K was calculated by substituting a = ( ai +1 + ai ) / 2 into equation (19a).
The effect of the definition according to equation (21) is that da / dN , for

Sec. 5

Results and analysis of crack-growth tests

185

120

crack-length, a (mm)

100
80
60
40

20

test #1
0
0

100000

200000

300000

400000

500000

600000

number of cycli, N

da/dN (m/cycle)

10

test #1 d = 2
test #1 d = 4
test #1 d = 6

0.1

0.01
0.1

K (MPam)

Figure 18. Top (a): Local averaging of da / dN by means of a fixed


step d . Bottom (b): da / dN versus K according to method 4 (local
averaging method), respectively with d = 2, 4 and 6 mm.

low crack-growth rate (low K ), is averaged over a greater number of


load cycles. For high crack-growth rate high K ), the method turns over
into the secant method provided the step d is not too large. Figure 18b
shows that, for d = 2 mm, da / dN ( K ) oscillates rather strongly, and
that the oscillations diminish as d increases. Since the results obtained for
d equal to 4 resp. 6 mm do not differ much, it seems justified to conclude
that the result is not influenced much by the analysis method. Thus,
method 4 has two advantages in comparison to the previous methods:
1 the scatter of the da / dN ( K )-relationship is reduced to an acceptable
level, cf. figure 18c;

Resistance to crack-growth and fracture

186

Ch. 4

da/dN (m/cycle)

10

0.1

test #1
test #2
0.01
0.1

K (MPam)

Figure 18c. Result according to method 4 (local averaging using a


fixed step. The result can be compared to that in fig. 14b (method 1).

2 the trends in the raw data are not seriously influenced by the analysis
method.
Therefore, method 4 was adopted for the analyses of the crack-growth test
data discussed in the following sections. The step size d can be chosen to
represent a realistic extension of the crack: the largest structural elements
of sand asphalt are 2 mm in diameter. The step size was chosen equal to 4
mm; that is in principle as small as possible but nonetheless much greater
than the physical structure parts, the sand grains.
A possible disadvantage of method 4 may be that da / dN ( K ) is
somewhat overestimated for low K . This may happen when the a (N )
curve is flat, and many data points are needed to achieve the step size d ,
so that da / dN is averaged over many data points. The overestimate
diminishes as a (N ) becomes steeper and da / dN is averaged over lesser
data points. The effect may be that the da / dN ( K )-relationship shows
an initial more or less horizontal or slightly declining part. An example
is the curve for test #1 in figure 18c. This disadvantage is minor in
comparison to the scatter of the data points from test #1 in figure 15b and
figure 17, and therefore acceptable. Hence, the raw data are influenced by
the analysis method, but not seriously.

5.2 Crack-growth test using the CCT-specimen Influence of frequency,


applied stress, and specimen thickness
Three types of test using the CCT-specimen were performed:
. constant load amplitude tests;
. a constant- K -test;
. constant load tests (creep crack-growth tests).

Sec. 5

Results and analysis of crack-growth tests

187

Note, that a constant load amplitude test is a dynamic crack-growth test, a


constant- K test is a dynamic crack-growth test, and a constant load test
is a static creep test . All tests were performed at constant temperature,
0C (Kleemans 1994, Kleemans et al. 1997).
5.2.1 Constant load amplitude tests
Thirteen tests were performed at different frequencies, 1 Hz, 10 Hz, or
29.3 Hz, and R -values 0.5, 0.1, and 0.5. The specimen thicknesses used
were 30 mm and 60 mm. An overview of the crack-growth rate curves
a (N ) obtained is shown in figure 19. An overview of the testing
conditions and the main results is shown in table 8. Numbers between
brackets in the legend of figure 19 refer to the test number in table 8. It
can be seen from figure 19 that the a (N ) -curves show small disturbances
from the regular power function,
a( N ) = c N k
(22)
where c and k are constants. It is probable that the sand grains of the
material are often in the way of the propagating crack, and are capable of
temporarily halting the crack. Therefore, the disturbances are attributed to
the influence of the material heterogeneity, and similar small disturbances
of the da / dN - K -relation from the regular power function, Paris law,
equation (8a), are expected.
Figure 20 shows the corresponding da / dN - K -relationships. Figure 20a
compares results obtained from tests performed at 1 Hz and 29.3 Hz with
specimen thicknesses of 60 mm and 30 mm. In figure 20a, the da / dN K -relation for the 30 mm specimen, tested at 1 Hz and R = 0.1, is
somewhat lower than that for the 60 mm specimen tested under similar
conditions. The da / dN - K -relation for the 30 mm specimen, tested at
29.3 Hz and R = 0.1, is somewhat higher than that for the 60 mm
specimen. Figure 20b shows four da / dN - K -relations for specimens
tested at 10 Hz and R = 0.1, two for 30 mm thick specimens and two for
60 mm thick specimens. In figure 20b, the two da / dN - K -relations for
29.3 Hz and R = 0.5 are in the same band. These results seem to indicate
that there is no effect of the specimen thickness for these testing
conditions, and that differences between 60 mm thick specimens and 30
mm thick specimens tested under similar conditions have to be attributed
to the normal scatter caused by the material heterogeneity and the
measurement error.
Figures 19 and 20 show a significant influence of the frequency and the
applied stress. The da / dN - K -relation shifts to lower da / dN -values
with increasing frequency other conditions being constant. This frequency
dependence is discussed in more detail later in this section. The da / dN K -relation shifts to higher K with decreasing R -value (read:

Resistance to crack-growth and fracture

188

Ch. 4

100
1 Hz, R=0.1, 30 mm (1)

90

1 Hz, R=0.1, 60 mm (2)

crack-length, a (mm)

80

10 Hz, R=0.1, 30 mm (3)


10 Hz, R=0.1, 30 mm (4)

70

10 Hz, R=0.1, 60 mm (5)


60

10 Hz, R=0.1, 60 mm (6)


29,3 Hz, R=0.1, 30 mm (7)

50

29,3 Hz, R=0.1, 60 mm (8)


40

1 Hz, R=0.5, 60 mm (9)


29.3 Hz, R=0.5, 60 mm (10)

30

29.3 Hz, R=0.5, 30 mm (11)


20

1 Hz, R=-0.5, 60 mm (12)


29.3 Hz, R=-0.5, 60 mm (13)

10
0
1000

10000

100000

1000000

10000000

number of cycli, N

Figure 19. Crack-growth curves a (N ) obtained using the CCT-specimen at


different frequencies, 1 Hz, 10 Hz, and 29.3 Hz, applies stresses, R = -0.5, 0.1,
and 0.5, and specimen thicknesses, 30 mm, and 60 mm. Numbers between
brackets refer to the test numbers in table 8.

Table 8. Test conditions and results of CCT-tests.


Nr.

width thickness
(mm)

(mm)

freq.

(kN)

(Hz)

notch-

aR

length (mm)

(m/

(mm)

cyce)

R2

Kc
MPam

230.7

28.6

5.0

0.1

25

93

131

4.6

0.90

0.83

230.0

60.0

14.0

0.1

24

84

446

6.0

0.97

0.90

230.0

30.0

7.0

10

0.1

23

95

14

5.0

0.79

1.19

230.0

30.0

7.0

10

0.1

24

94

8.6

4.9

0.92

1.15

230.5

59.5

13.0

10

0.1

30

95

20

5.7

0.96

1.10

230.0

60.0

10.5

10

0.1

37

98

30

5.8

0.87

0.98

230.5

30.8

7.0

29.3

0.1

4.2

4.9

0.92

230.5

60.0

14.0

29.3

0.1

25

81

1.9

4.6

0.90

0.85

231.0

61.0

7.0

0.5

25

92

3700

4.7

0.91

0.95

10

230.8

60.7

7.0

29.3

0.5

25

90

31

4.0

0.96

0.91

11

231.0

31.0

3.5

29.3

0.5

24

91

120

4.9

0.87

0.91

12

230.5

58.6

21.0

-0.5

24

88

10

4.0

0.96

0.90

13

231.5

60.7

21.0

29.3

-0.5

25

92

1.3

5.4

0.93

0.95

Sec. 5

Results and analysis of crack-growth tests

100
100

189

1 Hz, R=0.1, 60 mm (2)

da/dN
da/dN(m/cycle)
(m/cycle)

Hz,R=0.1,
R=0.1,6030mm
mm(2)(1)
1 1Hz,
29.3
Hz,
R=0.1,
60
mm (8)
1 Hz, R=0.1, 30 mm (1)
10
10

29.3Hz,
Hz,R=0.1,
R=0.1,6030mm
mm(8)(7)
29,3
29,3 Hz, R=0.1, 30 mm (7)

1
1

0.1
0,1

0.01
0,01 0.1
0,1

K (MPam)
K (MPam)

100

da/dN (m/cycle)

10

1
1 Hz, R=0.5, 60 mm (9)
29.3 Hz, R=0.5, 60 mm (10)
29.3 Hz, R=0.5, 30 mm (11)

0.1

1 Hz, R=-0.5, 60 mm (12)


29.3 Hz, R=-0.5, 60 mm (13)
0.01
0.1

10

K (MPam)

Figure 20. Crack-growth relationships da / dN versus K at different


frequencies, f , applied stresses ( R -values) and specimen thicknesses.
Numbers between brackets refer to the test number in table 8. Top (a):
Comparison of da / dN ( K )-relationships from specimens of 60 mm
and 30 mm thickness for 1 and 29.3 Hz, and R = 0.1. Bottom (b): As
figure (a) but for R -values -0.5 and 0.5.

increasing stress variation, cf. figure 8). This is expected because K is


proportional to , cf. equation (19). From some of the tests, notably
test number 2 and test number 8 in figure 20a, the da / dN - K relationships show an initial horizontal part. Likely, this horizontal part,
if it occurs, is caused by the method of analysis, method 4. In method 4,
da / dN is determined by local averaging using a constant step size d ,
equal to 4 mm. The initial part of the crack-growth rate curve, a (N ) , can

Resistance to crack-growth and fracture

190

Ch. 4

100

crack-growth rate, f da/dN (m/s)

1 Hz, R=0.1, 30 mm (1)


1 Hz, R=0.1, 60 mm (2)

10

10 Hz, R=0.1, 30 mm (3)


10 Hz, R=0.1, 30 mm (4)
10 Hz, R=0.1, 60 mm (5)
10 Hz, R=0.1, 60 mm (6)

29.3 Hz, R=0.1, 30 mm (7)


29.3 Hz, R=0.1, 60 mm (8)
1 Hz, R=0.5, 60 mm (9)
29.3 Hz, R=0.5, 60 mm (10)

0.1

29.3 Hz, R=0.5, 30 mm (11)


1 Hz, R=-0.5, 60 mm (12)
29.3 Hz, R=-0.5, 60 mm (13)
0.01
0.1

10

K (MPam)

Figure 21. Crack-growth rate f da / dN as function of K for 1 Hz, 10 Hz and


29.3 Hz, and R = -0.5, 0.1 and 0.5.

be flat, which means that the crack progresses very slowly. Then, to
achieve the step size of 4 mm in method 4, a relatively large number of
data points may be needed. Consequently the crack-growth rate is
averaged over a relatively large number of data points. This caused an
initial horizontal part of the da / dN - K -relationship (i.e. at the lower
da / dN described by this relationship) in some of the tests. As long as this
initial horizontal part of the da / dN - K -relationship is within the
band-width of the remaining data points, it has little influence on the fit of
the relationship to the Paris equation. In a couple of cases, however,
notably test number 2, test number 8, and test number 9 in figure 20, the
initial horizontal part is long compared to the band-width of the
remaining data points. In those cases the determination of the constants of
the Paris equation was influenced. With the determination of the constants
A and n in table 8, the data points on the initial horizontal part outside
the bandwidth of the remaining data points were discarded in the
regression analysis.
Figure 21 shows f da / dN plotted versus K for all thirteen tests. The
originally separate da / dN - K -relationships for the different frequencies,
1 Hz, 10 Hz, and 29.3 Hz, cf. figure 20, have shifted into a single band for
each R -value. Figure 21 shows three bands of data, one for each R -value.

Sec. 5

Results and analysis of crack-growth tests

191

100

da/dN (m/cycle)

10

0.1
1 Hz, R=0.5, 60 mm (9)
1 Hz, R=-0.5, 60 mm (12)
0.01
0.1

10

K (MPam)

Figure 22. As figure 20b. The combination K = 0.3 and R = 0.5


and the combination K = 1 and R = 0.5. yield the same da / dN

Following Elber (1971) and Schijve (1977, 1981) a so-called crackclosure correction might be used, defined by
K eff = U ( R ) K
(23a)
where U (R ) represents the crack-closure function which depends on R .
The idea is that not K determines the crack-growth rate, but K eff , the
reason being that crack-closure may occur already at low positive R values.
At high R crack-closure is unlikely, since a large tensile stress is then
acting on the crack during the entire load cycle. For example, in figure
20b, which is reproduced here as figure 22, for 1 Hz, the combination of
K = 0.3 and R = 0.5, and the combination K = 1 and R = 0.5 yield
the same da / dN . K eff should be equal for these two combinations. If it
is assume that
U ( R) = a + b R + c R 2
(23b)
and further, that K eff = K if R = 1, or that U = 1 if R = 1, then it
follows that a + b + c = 1 . Substitution of the two combinations of K
and R yields: a = 0.18, b = 0.31, c = 0.51. Note, that the additional
assumption (i.e. K eff = K if R = 1) is not necessary. It implies only
that the ratio of the coefficients is fixed. Any U = kU , with k a
constant, has the same correlating capacity as Keff = U K (Ewalds et
al. 1983). Only if an absolute value of U has to be found, then an absolute
measurement of the opening or closure value of K must be conducted.
Figure 23 shows the data from figure 19 (all thirteen tests), plotted as
f da / dN versus K eff where
K eff = U ( R ) K
(24a)

Resistance to crack-growth and fracture

192

Ch. 4

crack-growth rate, f da/dN (m/s)

100

10

1 Hz, R=0.1, 30 mm (1)


1 Hz, R=0.1, 60 mm (2)
10 Hz, R=0.1, 30 mm (3)
10 Hz, R=0.1, 30 mm (4)
10 Hz, R=0.1, 60 mm (5)

10 Hz, R=0.1, 60 mm (6)


29.3 Hz, R=0.1, 30 mm (7)
29.3 Hz, R=0.1, 60 mm (8)
1 Hz, R=0.5, 60 mm (9)

0.1

29.3 Hz, R=0.5, 60 mm (10)


29.3 Hz, R=0.5, 30 mm (11)
1 Hz, R=-0.5, 60 mm (12)
29.3 Hz, R=-0.5, 60 mm (13)
0.01
0.01

0.1

Keff (MPam)

Figure 23. All results f da / dN for different frequencies, R -values, and


specimen thicknesses fall in the same bandwidth when plotted versus K eff .

U ( R ) = 018
. + 0.31 R + 0.51 R 2
(24b)
The originally separate bands for the three R -values in figure 21 have
shifted into a single data band in figure 23. According to Elber, K eff
accounts for the effect of R on the crack-growth rate. A good fit to the
data points is the solid line in figure 23, which represents the following
equation:
da
da
f
=
= 209000 Keff4.85
(25a)
dN
dt
with da / dt in m/s and K eff in MPam. To give an idea of the effects
of changes in the constants of the equation, also a second line is indicated,
the dashed line in figure 23:
da
da
f
=
= 100000 K eff4.57
(25b)
dN
dt
with da / dt in m/s and K eff in MPam. Notice the large effect on the
coefficient of equation (25) associated to the relatively small change of the
exponent. This shows that the coefficient cannot be known with precision
if there is a significant scatter on the da / dN K -relationship. The
procedure was repeated using the transformation f da / dN , cf. equation
(32), with = 1.3. This was not found to have a beneficial effect on the
explained variance.

Sec. 5

Results and analysis of crack-growth tests

193

Whether the above transformation of the data has something to do with


crack closure (giving it a physical meaning) or not, one might also
consider the pure mathematical meaning of the transformations. It is a
well-known mathematical method to eliminate the dependence of a
function on a given variable, i.e. to make it implicit, by forming a
composed variable, in order to reveal a dependence on another variable. In
the above analysis, the frequency dependence was made implicit, using
f da / dN , to reveal the dependence on the R -value. Subsequently, this
stress dependence was made implicit, using the transformation of K to
K eff . Equation (25) describes the crack-growth of all thirteen tests for
the different frequencies, 1 Hz, 10 Hz, and 29.3 Hz, R -values, -0.5, 0.1,
and 0.5, and specimen thicknesses, 30 mm and 60 mm.
In summary, the analysis makes plausible that Paris Law can be applied to
describe the crack-growth of the sand asphalt mixture. Within the range of
specimen thicknesses used, the parameters A and n of the Paris equation
do not seem to depend on the specimen thickness. Equation (25a) and
figure 23 suggest that the exponent n of the Paris equation is independent
of the frequency, at least within the scatter of the data in figure 23.

5.2.2 Constant K test


According to Paris Law, the crack growth per load repetition is constant
if K is constant as a function of the number of load repetitions. This
provides a possibility to verify that Paris Law yields a valid description of
the crack-growth behaviour of asphalt mixture. Therefore, a constant- K test was performed. The result is shown in figure 24. Figure 24a shows the
development of the crack-length as function of the number of load
repetitions, a ( N ) . The crack ran from 35 mm to 81 mm. The applied load
was adjusted every mm, i.e. every four points. Thus the applied load was
decreased from 13.5 kN to 6.23 kN. Figure 24b shows K as a function
of the crack-length, a . The average value is K 0.34 MPam. Note that
the geometry factor, f ( a / W ) corrects well for the finiteness of the
specimen. Figure 24c shows da / dN as a function of a . The parameters
da / dN and K are not perfectly constant. This is difficult to realise,
because the applied load must be reduced as the crack grows ( K depends
on and a ). Since the crack-length was measured visually, an electrical
feedback signal from the crack-length was lacking. Therefore, real-time
adjustment of K was not possible4. Besides, the accuracy of the cracklength measurement was limited by the fact that the material contains sand
grains of 2 mm. Given these difficulties, the result in figure 24 of this
constant K test suggests that Paris equation is a valid model to describe
the crack-growth behaviour.
Additional validation by means of an electrically controlled constant- K -test
using a crack-foil would have been interesting. Unfortunately this was prevented
by practical problems that arose with the use of crack-foils.
4

Resistance to crack-growth and fracture

194

Ch. 4

90

crack-length, a (mm)

80
70
60
50
40
30
20
10
0
0

100000

200000

300000

400000

500000

600000

700000

800000

number of load repetitions, N

0.365
0.36

K (MPam)

0.355
0.35
0.345
0.34
0.335
0.33
0.325
30

40

50

60

70

80

90

crack-length, a (mm)

0.2

da/dN (m/cycle)

0.18
0.16
0.14
0.12
0.1
0.08
0.06
0.04
0.02
0
30

40

50

60

70

80

90

crack-length a (mm)

Figure 24. Constant K test, CCT-specimen. Top (a): crack-growth curve,


a (N ) . Middle (b): Stress intensity factor, K , versus crack-length, a . Bottom
(c): crack- growth rate, d a / d N versus crack-length, a .

Sec. 5

Results and analysis of crack-growth tests

195

5.2.3 Creep crack-growth tests


Three static creep crack-growth tests were performed, two with a constant
load of 3 kN, one with a constant load of 5 kN. The crack-growth curves,
a (t ) , are shown in figure 25a. The a (t ) -curves were processed similar to
the a (N ) -curves using Method 4 in 5.1, to obtain a curve of da / dt as
function of K . The corresponding da / dt - K -curves are shown in figure
25b. A good fit to the data points is the solid line in figure 25b, which
represents the following equation:
da
= 110 K 3.6
(26)
dt
where the coefficient, A = 110 is in m/s, and K is in MPam. In
addition to the model according to equation (26), two other crack-growth
models are shown in figure 25b. For details concerning these models, the
reader is referred to 5.7.
5.3 Crack-growth test using the 4PB-specimen Influence of frequency
and specimen thickness
The 4PB-specimen was used to investigate if and to which extent results
from the CCT-specimen could be reproduced using a different specimen
geometry. With the 4PB-specimen, only constant load amplitude tests
were performed. All the tests were performed at constant temperature,
0C. Ten tests were performed at different frequencies, 1 Hz, 10 Hz, or
29.3 Hz, and specimen thicknesses, 30 mm and 60 mm. The R -value was
kept the same for all the tests, 0.1, for two reasons:
1 the test set-up, cf. figure 4b, was not suitable for use with negative R ;
2 creep: at any R -value greater than 1, the average stress is greater than
zero, cf. figure 8, and will cause the beam to bend gradually owing to
creep. Therefore, in principle, the condition of linear elasticity is not
fulfilled, meaning that an error in the calculation of the stress intensity
factor K is introduced. The greater the R -value and the average stress,
the greater, in principle, is the creep and the error in K .
Crack-growth curves, a ( N ) , similar to those in figure 19 for the CCTspecimen, and crack-growth rate relationships, da / dN versus K ,
similar to those in figure 20 for the CCT-specimen, were obtained. An
overview of the tests, the testing conditions, and the main results is given
in table 9.
5.4 Analysis of the parameters A and n of the Paris equation
According to equation (8), the parameters A and n of the Paris equation
are related. The relationship obtained using the CCT-specimen is shown in
figure 26a, and that obtained using the 4PB-specimen is shown in figure

Resistance to crack-growth and fracture

196

Ch. 4

120

crack-length, a (mm)

100
80
60
40
#1: P = 3 kN
#2: P = 3 kN

20

#3: P = 5 kN
0
0

20000

40000

60000

80000

100000

120000

140000

time (s)

100

da/dN (m/cycle)

10

1
#1: P = 3 kN
#2: P = 3 kN
#3: P = 5 kN

0.1

A=26
A=224
0.01
0.1

K (MPam)

Figure 25. Top (a): Crack-length a versus time t at applied loads, P , in the
static creep crack-growth test. Results of three tests, indicated by #1, #2, and #3
are indicated. Bottom (b): Corresponding crack-growth rate, da / dt versus the
stress intensity factor, K . Solid line: equation (26); also shown are
da / dt = 26 K 5.48 and da / dt = 224 K 5.48 , cf. 5.4.3.

26b. The A -values in table 8 were converted to A using


A = A / 1000 /( 1000 ) n
(27)
where A in mm/cycle is associated to K in MPamm, and A in
m/cycle is associated to K in MPam.
Regression analysis on the data from table 8 in figure 26a yields (not
shown in figure 26a):
log A = 1.367 n 2.214
R 2 = 0.44
(28)

Sec. 5

Results and analysis of crack-growth tests

197

Table 9. Test conditions and results of 4PB-tests.


Nr. height thickness
(mm)

(mm)

freq.

(N)

(Hz)

notch-

aR

length (mm)

(m/

(mm)

cyce)

R2

Kc
MPam

61.0

61.4

1125

0.1

39

39

4.2

0.89

1.50

60.9

60.9

1125

0.1

38

40

3.8

0.91

1.43

61.0

30.7

563

0.1

40

10

2.9

0.76

1.60

60.9

30.2

563

0.1

39.5

27

3.7

0.94

1.58

60.8

61.0

1125

10

0.1

37.5

3.2

3.2

0.95

1.39

60.5

31.3

563

10

0.1

37.5

5.5

3.9

0.96

1.38

60.9

60.6

1125

29.3

0.1

38

1.4

3.0

0.85

1.43

60.8

61.3

1125

29.3

0.1

38

1.2

2.7

0.70

1.43

60.7

30.7

563

29.3

0.1

38

3.6

4.1

0.88

1.43

10

60.8

30.4

563

29.3

0.1

39

1.7

3.9

0.97

1.53

Equation (27) is used to compare equation (28) to similar relationships


given by Molenaar (1983), Jacobs (1995), and Molenaar and Medani
(2000). Molenaar (1983) obtained the following relationship,
log A = 1.628 n 0.977 ,
R 2 = 0.884
(29)
Jacobs (1995) obtained the following relationship,
log A = 1.14 n 2.36
R 2 = 0.83
(30)
Medani and Molenaar (2000) proposed the following relationship,
log A = 0.308 n 0.739 n 0.273 log S 2.890
R 2 = 0.888
(31)
The relationships (29) and (30) are shown in figure 26a, respectively as
Mo 1983, and Ja 1995.
Figure 27a shows the values of the parameter A of the Paris equation for
the CCT-specimen from table 8 plotted as function of the frequency, f ,
for the different R -values. The average values for a given R -value are
connected by lines. Note, that the constant A is equal to the value of
da / dN (in m/cycle) at K = 1 MPam. Figure 27a shows that A
depends on the frequency and on the R -value, i.e. the stress5. Similarly,
the A -values from table 9 are plotted in figure 27b.
Figure 29a shows log A versus log f for R = 0.1 for the CCT-specimen.
Similarly, figure 29b shows log A versus log f for R = 0.1 for the 4PBspecimen. Figure 29a shows a complex relationship, which is fairly
approximated by a linear log-log relationship. Let us assume, cf. equation
(8b),
A = A0 t = A0 / f
(32)
mix

The constant A depends also on the temperature. The influence of the


temperature was not investigated.

Resistance to crack-growth and fracture

198

Ch. 4

-5

log A' (mm/cycle)

R = 0.1
R = 0.5
R = -0.5

CCT-specimen

-6
-7
-8
-9

Ja 1995

-10
-11

Mo 1983
-12
3

parameter n of Paris equation

log A (m/cycle)

R = 0.1

4PB-specimen

2.5
2
1.5
1
0.5
0
2

2.5

3.5

4.5

parameter n of Paris equation

Figure 26. Top (a): Constant of the Paris equation, A , as function of the
parameter n for R = -0.5, 0.1, and 0.5, as obtained with the CCTspecimen; data from table 8. A in mm/cycle, K in MPamm. Also
shown are Mo 1983, equation (29), and Ja 1995, equation (30). Bottom
(b): As figure (a), for R = 0.1, for the 4PB-specimen. A in m/cycle, K
in MPam. The solid line represents the regression line through the
data points.

Equation (32) can be used to make the frequency dependence implicit, by


plotting log A0 = log ( A f ) versus n , where A0 = A0 / 1000 /( 1000 ) n ,
cf. equation (27). This was done for = 1.3 (the slope of the regression
line in figure 29a), and for = 1. For convenience, the result for = 1 is
shown in figure 28b6.
The result for = 1.3 differed little from the result for = 1. The explained
variance for R = 0.1 was found somewhat greater (0.97) for = 1.3 as
compared to = 1, but the explained variance for R = 0.5 was found somewhat
smaller (0.89) as compared to = 1.
6

Sec. 5

10000

constant A of Paris' equation

199

Results and analysis of crack-growth tests

CCT-specimen

1000
R = -0.5
100

R = 0.1
R = 0.5
avg R=-0.5

10

avg R=0.1
avg R=0.5

0.1
0

10

20

30

40

frequency (Hz)

constant A of Paris equation

10000

4PB-specimen

1000
R = 0.1

100

avg R=0.1

10

0.1
0

10

20

30

40

frequency (Hz)

Figure 27. Top (a): Constant of the Paris equation, A , as function of


the frequency for different R , for the CCT-specimen; data from
table 8. A in m/cycle, K in MPam. Bottom (b): As figure (a), for
the 4PB-specimen; data from table 9. A in m/cycle, K in MPam.

Thus, the stress dependence is revealed: In figure 28b the log ( A f ) relationships for the different R -values are clearly separated, whereas in
figure 28a they are not.
If it were possible to separate the dependence of A on the frequency
completely by writing A as in equation (32),
A (n )
A = 0 0
(33)
f
then, at a first glance, one expects n0 not to depend on the frequency
otherwise A0 would still depend on the frequency (implicitly). When the
frequency was made implicit using the composed variable ( A f ), the
explained variance increased from 44% to 95%, leaving 5% variance

200

Resistance to crack-growth and fracture

Ch. 4

-5

R = 0.1
R = 0.5
R = -0.5

log A' (mm/cycle)

-6
-7
-8
-9

Ja 1995

-10
-11

Mo 1983
-12
3

parameter n of Paris equation

log (A'f) (mm/cycle/s)

-5

R = 0.1
R = 0.5
R = -0.5

y = -0,7899x - 2,8618
R2 = 0,9518

-6
y = -1,0339x - 3,2352
R2 = 0,9552

-7
-8

y = -1,0852x - 3,6595
R2 = 1

-9
-10
-11
3

3,5

4,5

5,5

6,5

parameter n of Paris equation

Figure 28. Top (a): As figure 26a. Bottom (b): Log ( A f ) as function of
the parameter n . A in mm/cycle, and K in MPamm. In the regression
equations y(x), y represents log ( A f ) , and x represents n .

unexplained. This might be implicit, inseparable, frequency dependence of


the parameters on which A depends, amongst which n . However, a part
of this 5% unexplained variance is attributable to bias on the data related
to, e.g. the material heterogeneity, and the testing conditions. Thus, it
seems that less than 5% is attributable to n ( f ); a practically negligible
amount. In table 10, the parameter n from the tests (table 8) does not seem
to depend on the R -value, and does not show a dependence on the
frequency. Plotting n versus the frequency (not shown), showed no
correlation at all. This weak or absent dependence of n on the frequency
does not seem to agree with the theory, since that predicts n to be
proportional to 1/ m while the dependence of m on the frequency is rather
strong, cf. Intermezzos 1 and 2, cf. pages 210..211. This is investigated in
more detail in the following.

Sec. 5

201

Results and analysis of crack-growth tests

R = 0.1

CCT-specimen
log A (um/cycle)

2.5
2
1.5
1
y = -1.2993x + 2.4341
R2 = 0.891

0.5
0
0

0.5

1.5

log frequency (Hz)

R = 0.1

4PB-specimen
log A (m/cycle)

2.5
2
y = -0.7954x + 1.4194
R2 = 0.8721

1.5
1
0.5
0
0

0.5

1.5

log frequency (Hz)

Figure 29. Top (a): Log A as function of log frequency for R = 0.1, for
the CCT-specimen; data from table 8. A in m/cycle, K in MPam.
In the regression equation y(x), y represents log A, and x represents log f.
Bottom (b): As figure (a), for the 4PB-specimen; data from table 9.
Table 10. Averaged n -values from table 8, as
function of the R -value and the frequency.
f (Hz)
n
R = 0.5
1
4.7
29.3
4.45
5.3
R = 0.1
1
5.35
10
4.75
29.3
R = 0.5
1
4
29.3
5.4

202

Resistance to crack-growth and fracture

Ch. 4

The following analysis is rather crude, owing to the fact that suitable
experimental data are not available. This implies that assumptions had to
made which may not be valid. Equation (8b) can be written as
t
(1 2 ) J 1
1

ln A = ln
ln
+
ln
[ w(t )]n dt
(34)
+
2 2

6 R I m
0
For the waveform
sin ( 2 f t ), 0 t 1 / 2 f
(35a)
w(t ) =
0 t 1/ f
0,
one obtains (Molenaar 1983),
t

log [ w(t )]n dt = a0 + a1 log (n)

(35b)

where the parameters a0 and a1 take the values given in table 11.
Table 11. Parameters a0 and a1 for the waveform according
to equation (35a) (Molenaar 1983).
f (Hz)
a0
a1
R2
2
-0.784
-0.419
0.998
4
-1.084
-0.419
0.998
8
-1.386
-0.419
0.998
16
-1.687
-0.419
0.998

From equation (35b),


t
a0
ln [ w(t )]n dt =
+ a1 ln ( n)
0.4343
0
Substitution of equation (35c) into equation (34) yields
k
2
ln A = k1 + a 0 ( f ) + 2 k 3 ln
m
m
where

k1 = ln
2 2
6 R I
a0 ( f ) = 2.3026 a0
(1 2 ) J 1

k 2 = ln
2

k 3 = | a1 |
From Ch. 3, eq. (13),

e c1 x e c1 x
m = c1 1 c1 x
e + e c1 x

where
x = ln ( f / f m )

(35c)

(36a)

(36b)
(36c)
(36d)
(36e)

(37a)

(37b)

Sec. 5

Results and analysis of crack-growth tests

203

Substitution of equation (37b) into equation (37a), using e ln x = x , yields,

( f / f m )c ( f / f m ) c

m = c1 1
c
c
+
(
f
/
f
)
(
f
/
f
)
m
m

( f / f m ) c (1 ( f / f m ) c /( f / f m ) c
= c1 1
( f / f m ) c (1 + ( f / f m ) c /( f / f m ) c

1 f m2 c / f 2 c

= c1 1
2c
2c
+
1
f
/
f
m

)
)

If f = f m , then m = c1 . If f >> f m , then m tends to

= c1 1
1 +

m 2 c1 = 2 c1 ( f m / f ) 2 c1

(38a)

where = ( f m / f ) 2 c1
Let us write
m = m ( f ) f
where
= 2c1

(38b)
(39a)
(39b)

= 2 c1 f m2 c

(39c)
As a qualitative estimate equation (39a) is useful to perform the following
analysis. Equation (37a) must be used for quantitative determination of m .
Substitution of equation (39a) into equation (36a) yields
1

ln A = k1 + a0 ( f ) +

k2

2f
+ ln

k3

(40a)


A = exp k1 + a0 ( f ) + 2 f
(40b)
2
f

Equation (40b) can be written as


Af
(41a)
A = k exp (k 2 f / )
f
where
A f = ( / 2) k exp (k1 + a 0 ( f ) )
(41b)
Let us write
(41c)
A f = A0 exp (a 0 ( f ))
where
A0 = ( / 2) k e k
(41d)
Using equation (36b),
( / 2) k
(41e)
A0 =
6 R2 I 2
where A0 is independent of the frequency. Eq. (41a) can be written as
k3

Resistance to crack-growth and fracture

204

ln ( A f k ) = ln A f + n f
where, cf. equation (39a),
n f = k2 f / = k2 / m

Ch. 4

(42a)

(42b)

Equation (42a) is of the form of equations (28)(30). A f depends on the


frequency, cf. equation (41c). Also n f depends on the frequency, cf.
equation (42b). The frequency dependence of A f and n f is shown in
table 12.
Table 12. Frequency dependence of A f and n f using a0 -values
from table 11 and m -values from Intermezzo 1 (asphalt concrete).
f (Hz)
a0
exp (a 0 )
1/ m
6.85
1
0.164
-0.784
2
0.082
-1.084
4
0.041
-1.386
8
13.2
10
0.0205
-1.687
16
18.5
29.3

The values of the parameters in table 11 are probably not valid to the
asphalt concrete mixture of Intermezzo 1, since the waveform according to
equation (35a) is a half sine ( R = 0), whereas the properties of the asphalt
concrete mixture of Intermezzo 1 were determined using a sinusoidal
applied strain (strain controlled test: R = 1). Let us assume, for the sake
of the analysis, that the complex modulus, had it been determined for R =
0.1, would have been approximately equal to the complex modulus for R
= 1. (Probably, this is untrue, since the stiffness modulus differs in
tension and compression; it is (much) greater under compression than
under tension). For equal n , the area under the normalised waveform,
t

[w(t )]

dt

can be roughly estimated to be a factor of three greater for a sinusoidal


load with R = 0 (and also still with R = 0.1) than for a half sine with R =
0. That could mean that the value of equation (35c) differs by a constant
factor for the sinus with R = 0.1 and the half sine with R = 0. Let us in
spite of these complexities assume the analysis is still valid qualitatively.
In table 12, the parameter exp (a 0 ) decreases by a factor of approximately
8 if the frequency increases from 2 Hz to 16 Hz. The parameter 1 / m
increases by a factor of almost 3 if the frequency increases from 1 Hz to 30
Hz. When the frequency dependence was made implicit by means of the
composed variable A f k , the explained variance increased from 44% to
95%, cf. equation (28) and figure 28b. If it is assumed that 100% of the
3

Sec. 5

Results and analysis of crack-growth tests

205

variance can be attributed to the frequency and to the terms in equation


(42a), then the net influence of the frequency on the right hand side of
equation (42a) is 5%. How is it possible that the net influence of the
frequency on the right hand side of equation (42a) is so small, while the
terms on that side of the equation depend strongly on the frequency? Let
us investigate this further.
Substitution of equation (41c) into equation (42) yields
(43a)
A f k = A0 exp ( a 0 ( f )) exp (k 2 / m )
Substituting equation (36d),
3

Af

k 3

(1 2 ) J 1

= A0 exp (a0 ( f ))
2

1/ m

(43b)
1/ m

(1 2 ) J 1

ln ( A f ) = ln A0 + a0 ( f ) + ln
(43c)
2

Equation (43c) corresponds to equation (42a). Let us assume that the net
dependence of the right-hand side of equation (42a) on the frequency can
be neglected, since this is only 5%. In other words, let us assume that the
right-hand side of equation (42a), or equivalently equation (43c), is
independent of the frequency. Then, the sum of the last two terms on the
right hand side is constant as a function of the frequency,
k 3

1/ m

(1 2 ) J 1

= a0 ( f ) + ln
2

where is a constant. Rewriting,

2 = (1 2 ) J 1 e m ( a ( f ))
0

(44a)

(44b)

Thus, it follows, that the right hand side of equation (42a) can be
independent of the frequency, if the fracture energy, , depends on the
frequency according to equation (44b). Rewriting equation (44b),
(1 2 ) J 1
+ ln e m ( a0 ( f ))
ln (1) = ln
(44c)
2

Using equation (36d) and equation (42), it follows that,


n f = k 2 / m = a 0 ( f )

(45)

Using equation (36c) and the values of a 0 in table 11, the following
regression equation was obtained,
a0 ( f ) = ln f 1.111
(46)
or, with equation (45),
n f = k 2 / m = + ln f + 1.111

(47)

The parameter 2 /[(1 ) J 1 ] , cf. equation (44b), is shown in figure 30


for different values of as a function of m . The influence of a 0 ( f )
relative to decreases slowly as increases, cf. equation (44b).
2

Resistance to crack-growth and fracture

206

Ch. 4

= 10

2/[(1^2).J1]

2,5

= 1
=0

=1
= 10

1,5
1
0,5
0
0

0,02

0,04

0,06

0,08

0,1

0,12

0,14

Figure 30. The fracture energy, 2 /[(1 2 ) J 1 ] , according to equation (44b),


for the asphalt concrete mixture of Intermezzo 1, for different values of the
constant , as a function of m . c1 = 0.1947, ln aT (0C) = 6.3546, ln ( aT f m )
= 5.030. The markers indicate the frequencies 2, 4, 8, 16 and 30 Hz (from the
right to the left).

Note, that m tends to zero if the frequency tends to infinity, and that m
tends to the maximum limiting value, m 0 = 2 c1 , if the frequency tends to
zero, cf. Ch. 3, eq. (13).
The parameter n f according to equation (47) is shown in figure 31, for
different values of as a function of the frequency, f . n f in equation
(47) is a negative number, since equation (42a) corresponds to equations
(28)(30). Therefore, let us assume that n f in equation (42a) is a negative
number. From figure 28b for R = 0.1,
(48)
ln ( Af ) = 1.034 n 3.235
In equation (48), n is a positive number. The coefficient of n is 1.034,
which produces equivalent | n f |-values 1.034 times greater, i.e. 3.4%, than
the n -values. If this difference is neglected, then the n -values for R = 0.1
in table 8 can be compared to | n f |-values in figure 31 7, i.e.
n |n f |

(49)

The average n -values from table 8 for R = 0.1 are given in table 13 (see
also table 10). If the n -values from table 13 are plotted in figure 31 (not
shown), i.e. as n f = n , then the curve obtained is fairly similar to the
curves shown in figure 31 (the n -value of 5.3 for 1 Hz being relatively
low) and could be characterised by approximately equal to 9. Thus,
7

For the sake of the analysis it is assumed here that the n -values for R = 0.1 in
table 8, for the sand asphalt mixture, can be compared to values obtained by
application of eq. (44b), using the waveform of eq. (35a) and the constants in
table 11.

Sec. 5

Results and analysis of crack-growth tests

207

20
=10

15

=5
=1

10

nf

=0
= 1

= 5
0

= 10

-5
-10
0

10

15

20

25

30

35

f (Hz)

Figure 31. n f = k 2 / m = + ln f + 1.111 as a function of the frequency for


different values of .
Table 13. Average n -values from table 8 for R = 0.1,
compared to values from figure 31 for = 9.
|nf |
f (Hz)
n
observed
from eq.(47)
average
for
values
= 9
7.89
5.3
R = 0.1
1
5.59
5.35
10
4.51
4.75
29.3

it seems that equation (49) is valid, in this case, for = 9. Note, that the
frequency dependence of | n f | in table 13 is fairly strong. Recall, that the
constancy of implies that the right-hand side of equation (42a) is
constant as a function of the frequency. If the n -values in table 13 are
considered, it seems possible that the value of 5.3 for 1 Hz is accidentally
low, and should have been greater. If that is true, then n decreases as a
function of increasing frequency. At a first glance, that does not seem to
agree with expectation based on the theory; m decreases as a function of
increasing frequency, and n is expected to increase accordingly, since n
= 2/ m . However, a decreasing n -value as a function of increasing
frequency is perfectly possible according to the theory, if equation (44b) is
fulfilled, from which it follows, using equation (49), that n depends on m
according to equation (45).
The | n f |-values from table 13 are compared to the m -values from
Intermezzo 1, and the corresponding values of 2/ m , in table 14. Note, that
assuming n to be equal to 2/ m , in agreement with equation (8c), yields n
differing greatly from | n f |. In figure 32, the | n f |-values from table 13 are
compared to the corresponding 1/ m -values.

Resistance to crack-growth and fracture

208

Ch. 4

Table 14. | n f |-values from table 13 for R = 0.1, compared to


corresponding values of m from Intermezzo 1 and values of 1/ m .
f (Hz)
|nf |
m
2/ m
from eq.(47)
from
for
Intermezzo 1
= 9
13.7
0.146
7.89
R = 0.1
1
26.2
0.076
5.59
10
37.2
0.054
4.51
29.3

10
8

|nf|

6
4

|nf| = 9.75 - 0.29 (1/m)


R2 = 0.975

2
0
0

10

15

20

1/m

Figure 32. The exponent of the Paris equation | n f | from table 14 as a function
of 1/ m for the asphalt concrete mixture described in Intermezzo 1.

The dependence of | n f | as a function of 1/ m is given by


| n f | = 9.75 0.29 (1 / m)

(50a)

or equivalently,
| n f | = 0.29 (33.6 1 / m)

(50b)

5.5 Discussion
It was found already by Molenaar (1983), and later by Jacobs (1995), that
the exponent of the Paris equation, n , of asphalt mixtures is not predicted
correctly by equations (8c) and (8d) using the slope of the master curve on
log-log scale, m . Unfortunately, for the sand asphalt mixture investigated
here, a master curve, hence m , was not available. Therefore, estimates of
n are made in Intermezzos 1 and 2, based on data for asphalt concrete and
polymer modified Guss-asphalt from chapter 3, in the expectation that the
n -values for the sand asphalt mixture should lie somewhere in between
those for the asphalt concrete mixture and the guss-asphalt mixture. In
Intermezzos 1 and 2, the earlier observations by Molenaar and Jacobs are
confirmed, since n -values calculated using equations (8c/d) are much
greater than the n -values obtained from the crack-growth tests.

Sec. 5

Results and analysis of crack-growth tests

209

The fact that for asphalt mixtures, n -values were found much lower than
predicted by the theory, equations (8c/d), was attributed to limitations of
the theory (Molenaar 1983) when applied to asphalt mixtures. The theory
was developed for so-called solid propellant, which is actually a particle
reinforced rubber material. An important difference in comparison to
asphalt mixtures is that the latter contain air voids, which act as crackinhibitors (crack stoppers). Therefore Medani and Molenaar (2000)
proposed to use a correction factor, to account for the influence of the air
voids,
nest = nmas / CF
(51a)
CF = 0.541 + 0.137 nmas 0.03524 Va

(51b)

where nest is the estimated value, nmas is the calculated value, using
equations (8a) and (8c/d), and m from the master curve. The estimated
(corrected) values, nest , are also given in Intermezzos 1 and 2. Indeed, the
corrected values are much smaller than the values predicted by equation
(8c), and the frequency dependence of nest is much weaker than that of n .
The analysis in 5.4 suggests another possible explanation. Schapery made
the following assumptions to derive equations (8c/d) (Schapery 1973,
1981b, Lee and Kim 1998):
= constant
n = 2/ m
(52a)
=>
= constant
= constant

(52b)
=> n = 2 (1 + 1 / m)
C . O . D . = constant
where is the fracture energy, is the distance over which the crack
travels in a time t , and C.O.D. is the crack opening displacement. In the
analysis in 5.4, it is made plausible that a different relationship for n as a
function of m is obtained, depending on the condition governing the
behaviour of the fracture process zone. This relationship, equation (50), is
obtained if it is assumed that in equation (44) is a constant. is a
constant which relates the fracture energy to the slope of the master curve,
cf. equation (44b). The constancy of permits the right-hand side of
equation (42a) to be constant as a function of the frequency, while A f and
n f depend on the frequency.
In Intermezzo 1 and Intermezzo 2, the n -values were estimated using the
following equations, cf. chapter 3. The temperature shift-factor is defined
by,
aT = exp [K (1 / T 1 / TR )]

(53)

where K is a constant, T is the temperature, and TR is the reference


temperature. The master curve is given by,

210

Resistance to crack-growth and fracture

S * = S m* + S m* tanh (c1 ln f / f m )

Ch. 4

(54a)

S 0* + S *
(54b)
2
where S 0* is the limiting value if f 0 at T = TR , and S * is the limiting
value if f at T = TR . The slope of the log-log master curve is given
by,
S m* =

d ln S *
d
=
(55a)
m =
ln (1 + tanh (c1 x) ) = c1 (1 tanh (c1 x) )
dx
dx
x = ln ( f / f m )
(55b)
Unfortunately, a master curve of the sand asphalt mixture used for this
study was not available. Therefore the master curves of two different
asphalt mixture types were used:
1 the asphalt concrete mixture, cf. Ch. 3, fig. 5.
2 the guss-asphalt mixture, cf. Ch. 3, table 3.
It is expected that the properties of sand asphalt lie somewhere between
the corresponding properties of those mixtures.
___________________________________________________________
Intermezzo 1: Asphalt concrete
Parameter n estimated using viscoelastic creep susceptibility, m
The parameter n is evaluated based on m from the master curve. For the
master curve of figure 5 in chapter 3, the constant K of equation (53) was found to
be equal to 25415 Kelvin; hence the shift-factor, ln aT , corresponding to 0C is
25415(1/273-1/293) = 6.3546. For the frequencies 1, 10 and 29.3 Hz, the shiftfactors are:
1 Hz:
ln (aT f ) = 6.3546
10 Hz: ln (aT f ) = 8.6572
29.3 Hz: ln (aT f ) = 9.7322
The constant c1 of equation (54a) was found to be equal to 0.1947, and the reduced
frequency of the inflection point, ln (aT f m ) , 5.030. The following values of
x = ln [(aT f ) /( aT f m )] , m according to equation (55a), and n according to
equation (8c) are obtained:
1 Hz:
x = 6.3546 5.030 = 1.325 => m = 0.146 => n = 13.7
10 Hz:
x = 8.6572 5.030 = 3.627 => m = 0.076 => n = 26.2
29.3 Hz: x = 9.7322 5.030 = 4.702 => m = 0.054 => n = 37.2
Using equation (51), with Va = 3.9%, the following values for nest are obtained:
1 Hz:
nest = 6.01
10 Hz:
nest = 6.56
29.3 Hz: nest = 6.76
These n -values are in reasonable agreement with the average values for sand
asphalt in the centre cracked tensile specimen, cf. table 8.
Using equation (54), S * , and using equation (31), yields A :

Sec. 5

Results and analysis of crack-growth tests

211

1 Hz:
x = 1.325 => S * = 21437 MPa => A = 13 m/cycle
10 Hz:
x = 3.627 => S * = 27529 MPa => A = 5.1 m/cycle
29.3 Hz: x = 4.702 => S * = 29505 MPa => A = 3.6 m/cycle
where A is in m/cycle, for a plot of da / dN versus K with K in MPam,
and S mix is in MPa. To obtain these values of A , the values obtained from equation
(31) have to be multiplied with 1000(1000)n. The predicted A -values are
somewhat on the downside of the A s in table 8; this is attributed to the fact that
the mixture considered is an asphalt concrete mixture, which has a relatively high
mineral aggregate content and a relatively low binder content, cf. Intermezzo 2.
___________________________________________________________________
Intermezzo 2: Polymer modified Guss-asphalt
Parameter n estimated using viscoelastic creep susceptibility, m
The parameter n is evaluated based on m from the master curve. The constant K
of equation (53) was found to be equal to 28188 Kelvin; hence the shift-factor,
ln aT , corresponding to 0C is 28188(1/273-1/293) = 7.0480. For the frequencies
1, 10 and 29.3 Hz, the shift-factors are:
1 Hz:
ln (aT f ) = 7.0480
10 Hz: ln (aT f ) = 9.3506
29.3 Hz: ln (aT f ) = 10.426
The constant c1 of equation (54a) was found to be equal to 0.1267, and the reduced
frequency of the inflection point, ln (aT f m ) , 9.088. The following values of
x = ln [(aT f ) /( aT f m )] , m according to equation (55a), and n according to
equation (8c) are obtained:
1 Hz:
x = 7.0480 9.088 = -2.040 => m = 0.159 => n = 12.6
10 Hz:
x = 9.3506 9.088 = 0.263 => m = 0.122 => n = 16.3
29.3 Hz: x = 10.426 9.088 = 1.338 => m = 0.105 => n = 19.0
Using equation (51), with Va = 3.0%, the following values for nest are obtained:
1 Hz:
nest = 5.83
10 Hz:
nest = 6.11
29.3 Hz: nest = 6.25
These n -values are in reasonable agreement with the average values for sand
asphalt in the centre cracked tensile specimen, cf. table 8.
Using equation (54), S * , and using equation (31), yields A :
1 Hz:
x = 1.325 => S * = 12789 MPa => A = 53 m/cycle
10 Hz:
x = 3.627 => S * = 17687 MPa => A = 23 m/cycle
29.3 Hz: x = 4.702 => S * = 19991 MPa => A = 16 m/cycle
where A is in m/cycle, for a plot of da / dN versus K with K in MPam,
and S mix is in MPa. To obtain these values of A , the values obtained from equation
(31) have to be multiplied with 1000(1000)n. The predicted A -values seem to be
in fair agreement with the A s in table 8; thus, it seems, that, when A is
considered, the guss-asphalt mixture is more alike the sand asphalt mixture, than the
asphalt concrete mixture of the previous example. This is probably explained by
their relatively fine material structure and relatively high binder content.

Resistance to crack-growth and fracture

212

Ch. 4

5.6 Prediction of dynamic crack-growth based on static creep crackgrowth


The transformation of the experimental data using f da / dN = da / dt
where f is the frequency, and t is the time, suggests that the crackgrowth rate per unit of time can be independent of the frequency; that is, if
the crack-growth rate depends on the time during which the load is present
rather than the number of load cycles per second. In the literature, the
associated crack-growth mechanism is called creep crack-growth, so as to
distinguish it from fatigue crack-growth.
If creep crack-growth is the only crack-growth mechanism, then equation
(26) should also predict the crack-growth in the constant load amplitude
test, i.e. the dynamic test. This was investigated as described in the
following.
K was determined at every time, t . K depends on the crack-length. The
crack-length develops as a function of the time. This implies that K is a
function of the time. The variation with time is given by equation (19),
K = ( ) a f (a / W )
(56a)
= max min
(56b)
where is the stress variation with time, cf. figure 8. K was assumed
to be given by K i during the i -th period, t i , figure 33. In figure 33, the
period time is 1 s. The period time, t i , was subdivided in 100 timeintervals t j , t j = t . Accordingly, K i is defined by 100 discrete
values, K ij . The crack increment during each time interval t j was
calculated as
da
a j =
t j = 110 K ij3.6 t
(57)
dt
The crack-length after each period, i , was calculated as
ai = ai 1 + a j
(58)
j

For each period a new K i -value was calculated based on the new cracklength ai . It was assumed that equation (26) is also valid for low K i values, cf. figure 25. It is reasonable to assume that the error introduced by
this assumption is small, since the influence of low K i on the crackgrowth is small, because da / dt is proportional to the power 3.6 of K .
The integration method was performed on the f da / dN ( K ) relationships for 1 Hz and the R -values 0.5, 0.1 and 0.5. For the R -value
of 0.5, it was assumed that there is no crack-growth during the
compressive part of the load cycle. The results are shown in figure 34.
To show that the integration is necessary, similar calculations were
performed taking K constant, equal to K max , and constant, equal to the
average stress intensity factor per cycle, K avg . When K was assumed
equal to K max , the crack-growth rate was overestimated. When K was

Sec. 5

213

Results and analysis of crack-growth tests

1,6

ai = ai 1 +

stress intensity factor, K

1,4

ai 1

1,2

ij

Ki
0,8

0,6
0,4

0,2
0
0

10

1 second
time

Figure 33. Illustration of the integration method for the calculation of the
dynamic stress intensity factor using the static stress intensity factor.

crack-growth rate, f da/dN (m/load cycle)

100

10

K max

K avg

1 Hz, R=0.1, 30 mm (1)

0.1
1 Hz, R=0.1, 60 mm (2)
1 Hz, R=0.5, 60 mm (9)
1 Hz, R=-0.5, 60 mm (12)
0.01
0.1

10

K (MPam)

Figure 34. Calculated f (da / dN ) as function of K for 1 Hz, and R = -0.5,


0.1, and 0.5; calculated by the integration method described in the text.

214

Resistance to crack-growth and fracture

Ch. 4

assumed constant equal to K avg , the crack-growth rate was


underestimated. This is an effect of the exponent in equation (26), since
this is greater than unity. This causes the effect of K to increase
nonlinearly as K increases; hence the K ij -values that are greater than
K avg have a nonlinearly greater effect than the K ij -values that are lower
than K avg . The calculation was performed for the three R -values, 0.1, 0.5,
and 0.5. The correspondence of the calculated crack-growth curves and
the observed curves seems to indicate that, in principle, dynamic creep
crack-growth might be predicted by static creep crack-growth; however,
the exponents of equations (25) and (26) differ. This is discussed further in
5.7. The correspondence of the calculated crack-growth curves and the
observed curves also seems to indicate that creep crack-growth is the
prevailing crack-growth mechanism also in the dynamic crack-growth
tests.
5.7 Discussion
The exponent of the Paris equation for the static creep crack-growth tests,
cf. equation (26), is smaller than the analogous exponent for the dynamic
crack-growth tests, cf. equation (25).
The question arises if the dynamic crack-growth theory is applicable to the
static case. The dynamic (complex) modulus, S * , and the static stiffness
modulus, S (t ) , are defined differently, cf. figure 35,
S * = (t ) / (t )
(59a)
S (t ) = 0 / cum (t )
(59b)
where (t ) is the dynamic stress as function of the time, t , (t ) is the
dynamic applied strain as function of the time, 0 is the static applied
stress, and cum (t ) is the cumulative strain as function of the time. As a
result, numerical values of S * and S (t ) differ orders of magnitude. For
example, typical values of S * range between say 5000 MPa and 15000
MPa, whereas S (t ) is typically of the order of 10 MPa. Note, that also in
dynamic creep, one may use the stiffness modulus, S (t ) , based on the
cumulative strain, cum (t ) . The corresponding m -values in static creep and
dynamic creep are different properties,
(ln S s (t ))
(60a)
static loading:
ms =
(ln t )
(ln S c (t ))
(60b)
dynamic loading: mc =
(ln t )
(ln S * )
dynamic loading: m =
(60c)
(ln )
where ms is associated to the stiffness modulus in the static case, S s (t ) ,
mc is associated to the stiffness modulus based on the cumulative strain in

Sec. 5

Results and analysis of crack-growth tests

215

creep compliance

(t ) + cum (t )

cum (t )

time

Figure 35. Schematic illustration of different definitions of stiffness


modulus in the static and the dynamic case.

the dynamic case, S c (t ) , and where m is associated to the complex


modulus. mc in equation (60b) corresponds to ~z in the power law creep
model, Ch. 3, eq. (15). Typical values of mc can be found e.g. in App. 3,
tables 8-12, and are found between 0.19 and 0.43. These values are of the
same order of magnitude as the limiting zero frequency m -value, m0 ,
for which a representative value for crushed gravel asphalt concrete was
found to be 0.365, cf. Ch. 3, table 6. The static m -value, ms , is likely to
be smaller than the dynamic m -values mc and m , other testing conditions
being equal. Accordingly, if the lower static ms -value is used in the
dynamic crack-growth theory to predict the exponent of the Paris
equation, n = 2/ m , then the associated exponent is greater than the
dynamic n -value. This contradicts what is observed experimentally.
One must be careful in making predictions of the behaviour in a tensile
creep crack-growth test, based on linear theories, since those theories are
based on the assumption that the material is a homogeneous continuum.
With an asphalt mixture, the behaviour of a grain skeleton may differ in a
strain controlled bending test (approximately 50% specimen volume in
compression and 50% specimen volume in tension) and a static creep test
(100% specimen volume in tension). Therefore, it seems improbable that
the m -value from a strain controlled dynamic bending test can be used to
predict the m -value in a stress controlled static creep test.
Despite all that, it is tempting to relate the static creep exponent to the
limiting maximum slope of the log-log complex modulus master curve, as
the frequency tends to zero, m0 , because this limiting maximum value is
expected to produce the lowest possible value of n ; in the stress controlled
case, n = 2/ m .

216

Resistance to crack-growth and fracture

Ch. 4

As mentioned previously, a representative value of m0 for the 25 asphalt


concrete mixtures of table 6 in chapter 3 is 0.365. Let us assume this m0 value predicts also the creep of the sand asphalt mixture as f 0 , i.e. in
the static creep test. The corresponding estimated n , equal to 2/ m , i.e.
5.48, overestimates the experimental static n , 3.6, cf. equation (26). Using
equation (51), with the voids content of the sand asphalt mixture, 3.9%,
the correction factor, CF , is found to be equal to 1.15, and the estimated
n -value 4.75, still greater than the experimental value.
It is not expected that the equations (29) and (30) produce correct A values, since, as mentioned previously, these equations do not correctly
account for the frequency dependence and the stress dependence of A .
Using equations (29) and (30) for the sake of completeness, the estimated
coefficient A is:
(61a)
from equation (29), A converted to A : 26 m/s
(61b)
from equation (30), A converted to A : 224 m/s
The corresponding creep crack-growth models are shown in figure 25.
5.8 Critical stress intensity factor
From the crack-length at rupture, a R , cf. table 8, measured afterwards on
the fracture surface, the critical stress intensity factor can be estimated
using equation (7),
[P /( BW )] a R f (a / W )
K c
K c = K max =
(62)
=
1 R
1 R
The values obtained are listed in table 8. The stress intensity factor, K ,
describes the stress field at the crack-tip, cf. equation (1) and figure 3. The
stress intensity factor represents the stress at the crack-tip, cf. equation (1).
The critical stress intensity factor represents the stress at the crack-tip at
which the crack-growth becomes unstable, and the specimen fractures.
Therefore, this is also commonly interpreted as the residual tensile
strength of the cracked specimen however, in MPam, instead of in
MPa. Based on experience with metals, cf. figure 5, the critical stress
intensity factor depends on the specimen thickness. For sufficient
specimen thickness, the critical stress intensity factor becomes
independent of the specimen thickness, in which case it is common to
consider it a material property.
The tensile mode of loading is commonly called mode I (to distinguish it
from the shearing mode, mode II, and the tearing mode, mode III; not
shown here). The so-called mode I plane strain critical stress intensity
factor is also called fracture toughness.
K c according to equation (62) represents the stress at the moment when
the remaining ligament fractures. It was discussed in the previous sections,
that the crack-growth properties of the sand asphalt mixture appear to be

Sec. 5

Results and analysis of crack-growth tests

217

independent of the specimen thickness. Equation (25a) represents a single


equation, which describes the crack-growth behaviour of tests performed
at three different frequencies, three different R -values, and two different
specimen thicknesses. Equation (25a) seems to indicate that the crackgrowth properties, the coefficient A and the exponent n of the Paris
equation, have been obtained independent of the specimen thickness.
Then, also K c in table 8 should be independent of the specimen thickness.
CCT-specimen: K c = 0.97 0.12 MPam
(63)
Therefore, the question arises: could K c from the dynamic crack-growth
test be considered the equivalent of the fracture toughness of the material?
This could be true if the ligament is linearly elastic, and if the properties of
material in the ligament have not changed by fatigue. It is probable that
these conditions are not fulfilled. In spite of that, the calculated K c might
still be a reasonable estimate of the original true K c of the material.
There is doubt to the validity of this assumption. On the one hand, the
properties obtained with the CCT-specimen do not seem to depend on the
specimen thickness. On the other hand, the 4PB specimen was used to
investigate to which extent results from the CCT-specimen are
reproducible using a different specimen geometry. If the results of the two
tests are compared, then the crack-growth properties of the material do not
appear to be reproducible using different specimen geometries. For
example, the relations between A and n , cf. figure 27, are totally
different, and also the coefficient A as function of the frequency differs in
the CCT-specimen and in the 4PB-specimen.
After the tests had been performed, it was clear that there is a major
disadvantage to the 4PB test as crack-growth test. The test set-up was not
suitable to be used with negative R -values. Therefore, the only useful R value was 0.1. Even with this value, creep is unavoidable, owing to the
testing conditions. The effect of creep is unidirectional bending of the
beam, which is the probable cause that the linear elastic condition at the
crack-tip is violated. This is required for application of Paris Law, and for
the stress intensity factor to be calculable. Therefore, it is probable that the
K c -values obtained with the 4PB-specimen, 1.47 0.08 MPam, are not
reliable. Thus, the 4PB-test does not seem to be particularly suitable to
perform crack-growth tests.
Also with the CCT-specimen there is still doubt to the validity of the
assumption that the obtained critical stress intensity factor is the true
fracture toughness of the material. Although the test results seem to
indicate that Paris Law is applicable, and that crack-growth properties do
not depend on specimen size, it was found that there is a stress
dependence, i.e. a dependence of the coefficient A of the Paris equation
on the R -value. Strictly, this means that the behaviour is not purely

218

Resistance to crack-growth and fracture

Ch. 4

linearly viscoelastic. However, similar to what was discussed in 5.7, it is


imaginable, and probable, that the dynamic component of the stress strain
behaviour is linearly viscoelastic on top of a static (creep) component
(which is nonlinear, but negligible), and that therefore linear elastic
fracture mechanics is still applicable8.
The K c -values in table 8 seem not to depend on the frequency. On the one
hand, this may seem peculiar, since K c is the residual tensile strength, the
tensile strength is related to the complex modulus (Heukelom 1966), and
the complex modulus is frequency dependent. On the other hand, the
fracture toughness is a monotonic test property (normally obtained from a
fracture toughness test performed at constant displacement rate). Dynamic
properties, and static or monotonic properties can be considered
uncoupled, as discussed in 5.7, illustrated by figure 35. Similarly, the
fracture toughness, and K c , might be considered uncoupled from the
dynamic crack-growth properties, A and n . Then, it seems there is no
need for K c to depend on the prehistory of the crack, hence on the
frequency under which that was created (provided the ligament is linearly
elastic, and its resistance to fracture not affected by fatigue).
In this study, the critical stress intensity factor, estimated by equation (62),
is an important property, because it is the only property based on which
crack-growth properties from so-called advanced dynamic crack-growth
tests and simple constant displacement rate tests can be linked. As
mentioned previously, the advanced crack-growth tests are laborious and
expensive, which is why they are less suitable for practical purposes, such
as asphalt mixture design. The critical stress intensity factor represents a
link to properties, which can be determined in simple tests for practical
purposes, i.e. the fracture toughness and the tensile strength. It provides
the opportunity to relate asphalt mixture properties from simple tests to
true material properties obtained from advanced tests (the parameters A
and n ). In addition, it provides the opportunity to attribute a physical
meaning to the properties from simple tests, to gain confidence that also
those properties have predictive value for the behaviour of the material in
the pavement.
Recall, that the aim of this study is to investigate if and to which extent
properties obtained from simple tests can be obtained as material
properties, i.e. with predictive value for the behaviour of the material in
the pavement. This is elaborated further in the following.
8

It was asked right from the start of the whole project whether this would likely
to be the case, or that the J -integral would have to be used instead of the stress
intensity factor, K . It was decided to investigate the applicability of the stress
intensity factor first, mainly for experimental reasons (e.g. the notch preparation
for J Ic -testing is more complicated than for K Ic -testing).

Sec. 6

Results and analysis of fracture toughness tests

219

6 Results and analysis of fracture toughness tests using the


SCB-specimen
The ensemble of tests performed with the CCT-specimen show that Paris
Law and based on that the stress intensity factor, hence also the critical
stress intensity factor, K c , are applicable to asphalt mixture, although
under strict conditions. Dynamic crack-growth tests to determine the
coefficient A and the exponent n of the Paris equation are laborious and
expensive, and therefore less suitable for routine purposes. An advantage
of K c , assuming this can be obtained as a true material property, is, that
this can also be obtained using a much simpler and therefore more
efficient method such as the SCB-test.
A further advantage of the SCB-test is that this test can be used for all
types of asphalt mixture including porous asphalt. Recall from 4.2.3.1 that
in order to obtain a valid fracture toughness, the specimen and crack must
fulfil the ASTM minimum size requirements. The values presented in the
following are candidate values, indicated as K IQ . Only if a candidate
value fulfils with the ASTM requirements, it can be considered a valid
fracture toughness, K Ic . For convenience, in the following, a candidate
value of the critical stress intensity factor is referred to as the apparent
fracture toughness. For convenience also, K IQ is expressed in MPamm
instead of MPam. Note, that 1 MPam is equal to 31.6 MPamm.

6.1 Dependence of the apparent fracture toughness on the specimen size


The apparent fracture toughness was calculated by means of equation (16)
using YI from table 3. Figure 36 shows the results obtained using three
specimen diameters, 100 mm, 150 mm, and 200 mm, and three specimen
thicknesses, 25 mm, 50 mm, and 75 mm. The tests were performed on
dense graded asphalt concrete 0/16, at four different temperatures, -10C ,
0C, +15C, and +25C. The displacement rate was constant, 0.05 mm/s.
Figure 36 shows the apparent fracture toughness, K IQ , as a function of the
temperature. For comparison, also the indirect tensile strength, ITS, is
shown. The ITS was calculated using equation (12). It can be seen in
figure 36, that K IQ (0C) is found between 25 MPamm and 35 MPamm,
and that K IQ (-10C) is found between 25 MPamm and 35 MPamm.
Note, that these values are of similar magnitude as the value of K c of 31.6
MPamm obtained with the CCT- specimen for the sand asphalt mixture,
cf. table 8.
6.2 Dependence of the apparent fracture toughness on the displacement
rate
Figure 37 shows results obtained using three specimen diameters, 100 mm,
150 mm, and 200 mm, and three specimen thicknesses, 25 mm, 50 mm,
and 75 mm. Figure 37 shows K IQ as a function of the displacement rate

Resistance to crack-growth and fracture

220

40

ITS (MPa), K-IQ (MPamm)

ITS (MPa), K-IQ (MPamm)

40

Ch. 4

35
30
25
20
15
10
5

35
30
25
20
15
10
5
0

0
-10

10

20

-10

30

temperature (C)
100x25 mm
220x25 mm

10

20

30

temperature (C)
100x50 mm
220x50 mm

150x25 mm
ITS (MPa)

150x50 mm
ITS (MPa)

ITS (MPa), K-IQ (MPamm)

40
35
30
25
20
15
10
5
0
-10

10

20

30

temperature (C)
100x75 mm
220x75 mm

150x75 mm
ITS (MPa)

Figure 36. Apparent fracture toughness, K IQ , and indirect tensile strength, ITS,
as a function of the temperature, for dense graded asphalt concrete 0/16.
Specimen diameters: 100, 150 and 220 mm. Displacement rate: 0.05 mm/s. Top
left (a): specimen thickness 25 mm. Top right (b): specimen thickness 50 mm.
Bottom (c): specimen thickness 75 mm.

Sec. 6

Results and analysis of fracture toughness tests

221

for dense graded asphalt concrete. The displacement rates applied were:
0.005 mm/s, 0.05 mm/s, and 0.5 mm/s. All test were performed at a
temperature of +15C.

6.3 Influence of the material composition


Figure 38 shows K IQ , as a function of the temperature, for respectively
unmodified dense asphalt concrete 0/16 (top left), polymer modified dense
asphalt concrete 0/16 (top right), polymer modified porous asphalt 0/16
(bottom left), and polymer modified stone mastic asphalt 0/11 (bottom
right). For comparison, also the ITS is shown.
By comparison of figure 38a and figure 38b it can be seen that the K IQ of
dense graded asphalt concrete are increased by polymer modification.
Figure 38c shows that the K IQ of the porous asphalt mixture are lower
than those of the unmodified dense graded asphalt concrete mixture
(figure 38a). This was expected because of the open structure of porous
asphalt.
Figure 38d shows that the K IQ of the stone mastic asphalt seem to be
somewhat higher than those of the dense graded asphalt concrete mixture.
Indicative K IQ -values for 0C and 15C from figure 38 are given in table
15.
Table 15. Indicative values of K IQ of various investigated asphalt
mixtures.
DAC unmodified
DAC elastomer
DAC plastomer
PA elastomer
PA plastomer
SMA elastomer
SMA plastomer

K IQ at 0C
25 30
27 40
27 40

K IQ at 15C
11
12.5 20
25 30

12 18
14 23

24
8 11

30 38
34 43

4
22-28

6.4 Analysis of the results


6.4.1 Dependence of the fracture toughness on the specimen size
Let us consider the pertinent stress, 0 , instead of K IQ , cf. equation (16),
since a c YI depends only on the specimen diameter. According to the
theory, K IQ is expected to increase with decreasing specimen thickness, as
a result of developing plane stress, cf. figure 5.
In figure 39, 0 is plotted versus the specimen thickness. Only the data for
10C and 0C are shown (the K IQ for these test temperatures can be
considered valid K Ic , cf. 6.4.3, table 20; therefore these are indicated as
K Ic ). To improve the visibility of the markers, the specimen thickness

Resistance to crack-growth and fracture

222

30

35

25

30

Ch. 4

K-IQ (MPamm)

K-IQ (MPamm)

25

20

20

15

15

10

10

0
-2.5

-2

-1.5

-1

log v (mm/s)

100x25 mm
220x25 mm

-0.5

0
-2.5

150x25 mm

-2

-1.5

-1

log v (mm/s)

100x50 mm
220x50 mm

-0.5

150x50 mm

30

K-IQ (MPamm)

25
20
15
10
5
0
-2.5

-2

-1.5

-1

log v (mm/s)

100x75 mm
220x75 mm

-0.5

150x75 mm

Figure 37. Apparent fracture toughness K IQ as a function of the displacement


rate, dense graded asphalt concrete 0/16. Temperature: +15C. Specimen
diameters: 100 mm, 150 mm, and 220 mm. Top left (a): specimen thickness 25
mm. Top right (b): specimen thickness 50 mm. Bottom (c): specimen thickness
75 mm.

Sec. 6

Results and analysis of fracture toughness tests

40

40

ITS (MPa), K-IQ (MPamm)

45

ITS (MPa), K-IQ (MPamm)

45

35
30
25
20
15
10
5

35
30
25
20
15
10
5
0

0
-20

-10

10

20

-20

30

K-I unmodified

-10

10

20

30

temperature (C)

temperature (C)

K-I elastomer
ITS elastomer

ITS unmodified

45

ITS (MPa), K-IQ (MPamm)

223

K-I plastomer
ITS plastomer

45

ITS (MPa), K-IQ (MPamm)

40
35
30
25
20
15
10
5
0

40
35
30
25
20
15
10
5
0

-20

-10

10

20

temperature (C)
K-I elastomer
ITS elastomer

K-I plastomer
ITS plastomer

30

-20

-10

10

20

30

temperature (C)
K-I elastomer
ITS elastomer

K-I plastomer
ITS plastomer

Figure 38. Apparent fracture toughness and indirect tensile strength as a


function of the temperature. Specimen: 150 x 50 mm. Displacement rate: 0.5
mm/s, except for figure top left: 0.05 mm/s. Top left (a): Dense asphalt concrete
0/16, unmodified. Top right (b): Dense asphalt concrete 0/16, polymer modified.
Bottom left (c): Porous asphalt 0/16, polymer modified. Bottom right (d): Stone
mastic asphalt 0/11, polymer modified.

Resistance to crack-growth and fracture

224

DAC 0/16, unmodified,


0.05 mm/s, -10C

DAC 0/16, unmodified,


0.05 mm/s, 0C
1.4

pertinent stress, o (MPa)

1.2

0.8

1.2

0.8

0.6

0.6
0

20

40

60

80

100 mm/-10C
220 mm/-10C

150 mm/-10C

20

40

60

80

specimen thickness (mm)

specimen thickness (mm)

100 mm/0C

150 mm/0C

220 mm/0C

DAC 0/16, unmodified, 0.05 mm/s


1.4

pertinent stress, o (MPa)

o (MPa)

1.4

pertinent stress,

Ch. 4

1.2

0.8

0.6
0

20

40

60

80

specimen thickness (mm)


100 mm/-10C
150 mm/0C

100 mm/0C
220 mm/-10C

150 mm/-10C
220 mm/0C

Figure 39. Pertinent stress, defined by equation (16b), as function of


the specimen thickness. Different markers are used to indicate
different combinations of specimen diameter and temperature. Top
left (a): -10C; top right (b): 0C; bottom (c): -10C and 0C.

Sec. 6

Results and analysis of fracture toughness tests

225

Table 16. Pertinent stress 0 , cf. equation (16b), of the notched specimen,
for different combinations of specimen diameter and temperature.
specimen
pertinent stress of notched specimen (MPa)
diameter/
Temperature
25 mm
50 mm
75 mm
100 mm/-10C
1.162
1.222
1.112
1.258
1.201
1.056
150 mm/-10C
1.214
1.204
1.064
1.088
1.131
1.058
220 mm/-10C
0.877
1.179
1.111
0.976
1.039
1.002
mean/-10C
1.10
1.16
1.07
stdev/-10C
0.15
0.09
0.04
100 mm/-0C
1.147
0.917
1.154
1.116
1.049
0.978
150 mm/-0C
0.863
0.908
0.988
0.955
1.060
1.028
220 mm/-0C
1.159
0.869
0.945
0.837
0.871
1.002
mean/0C
1.01
0.95
1.02
stdev/0C
0.15
0.09
0.07

Table 17. Pertinent stress, 0 , of the notched specimen, for different combinations
of specimen diameter and temperature.
specimen
pertinent stress of notched specimen (MPa)
thickness 100 mm, 100 mm, 150 mm, 150 mm, 220 mm, 220 mm,
(mm)
-10C
0C
-10C
0C
-10C
0C
1.159
0.877
0.863
1.214
1.147
1.162
25
0.837
0.976
0.955
1.088
1.116
1.258
25
0.869
1.179
0.908
1.204
0.917
1.222
50
0.871
1.039
1.060
1.131
1.049
1.201
50
0.945
1.111
0.988
1.064
1.154
1.112
75
1.002
1.002
1.028
1.058
0.978
1.056
75
mean
1.155
1.060
1.126
0.967
1.031
0.947
Stdev
0.075
0.097
0.069
0.074
0.106
0.120
67%
1.08-1.24 0.96-1.16 1.06-1.20 0.90-1.04 0.92-1.14 0.83-1.06

226

Resistance to crack-growth and fracture

Ch. 4

was plotted slightly deviating from the actual thickness (respectively 25


mm, 50 mm, and 75 mm). It is probable that what is shown in figure 39 is
the heteroscedasticity of 0 ; a parameter is called heteroscedastic, if its
standard deviation is not constant, in this case, as function of the specimen
thickness. It is probable that the heteroscedasticity is caused by the
material heterogeneity, meaning that the repeatability decreases as the
specimen thickness decreases. If true, the pertinent stress is independent of
the specimen thickness.
In table 16, the 0 -values from figure 39 are organised according to
specimen thickness, specimen diameter, and temperature. The mean value
and the standard deviation are given per specimen thickness and
temperature. The standard deviations are quite good, considering that the
mixture is a 0/16 mixture; therefore, it is not probable that better (i.e.
smaller) standard deviations will be obtained if the investigation is
repeated.
The averages in table 16 do not show 0 to increase with decreasing
specimen thickness; for example the averages for the combinations [25
mm, 0C] and [50 mm, 0C] are lower than the average for the
combination [75 mm, 0C]. Considering the average standard deviation,
0.10, the average pertinent stresses for the combinations [75 mm, -10C]
and [75 mm, 0C] are not statistically lower than the values for the
combinations [25 mm, -10C], [50 mm, -10C], [25 mm, 0C] and [50
mm, 0C]. The data in table 16 may be rearranged to give table 17. The
67% probability interval is indicated in the last row of table 17. The data
outside this interval are shown bold-faced. Relatively few data fall outside
the 67% probability interval. The mean pertinent stress and corresponding
67% probability intervals are shown in figure 40 as function of the
specimen diameter. The mean pertinent stress seems to decrease as
function of the diameter. There is no explanation for such a decrease. It is
also possible that the decrease is not real, since the 67% probability
intervals strongly overlap (per temperature). Therefore, the means per
temperature for the three diameters are not statistically different.
Substituting the mean values of the pertinent stress from table 16 into
equation (16a), and using the values of YI in table 3, yields the estimated
average values and standard deviations of the stress intensity factor listed
in table 18. The average values in table 18 give no indication that K Ic
increases with decreasing specimen thickness. Taking account of the
standard deviation, it seems that the average values in table 18 can be
considered to be independent of the specimen thickness.
Substituting the mean values of the pertinent stress in table 17 into
equation (16a), and using the values of YI in table 3, yields the estimated
average values and standard deviations of the stress intensity factor listed
in table 19.

Sec. 6

Results and analysis of fracture toughness tests

227

1.4

pertinent stress, o (MPa)

-10C
0C
1.2

0.8

0.6
50

100

150

200

250

specimen diameter (mm)

Figure 40. Mean pertinent stress and 67% probability interval as function
of specimen diameter.
Table 18. Calculated K Ic , using the mean values and standard
deviations of the pertinent stress in table 16, and YI in table 3.
K Ic (MPamm)
temperature
25 mm
50 mm
75 mm
-10C
28.2 3.8
33.3 2.0
32.2 1.2
0C
26.0 3.8
27.1 2.5
30.7 2.2
Table 19. Calculated K Ic , using the mean values and standard deviations of
the pertinent stress in table 17, and YI in table 3.
K Ic at -10C
K Ic at 0C
specimen diameter
(MPa.mm)
(MPa.mm)
(mm)
27.3 2.5
29.7 1.9
100
27.7 2.1
32.3 2.0
150
28.6 3.6
31.1 3.2
220
all diameters
31.0 2.5
27.9 2.7

The average K Ic seems to increase as function of the diameter. However,


considering the standard deviation, the averages per diameter are not
statistically different from the average for all diameters. The average K Ic
is constant within the standard deviation, at constant temperature. The
values of K Ic in table 19 are in agreement with the values in figure 36 for
the corresponding temperatures.
Summarising, the mean pertinent stress, 0 , at -10C and 0C shows a

228

Resistance to crack-growth and fracture

Ch. 4

weak dependence on the specimen diameter (figure 40). This dependence


is statistically not significant. The K Ic at -10C and 0C do not show an
increase as function of decreasing specimen thickness; in other words give
no indication of developing plane stress with decreasing specimen
thickness. The K Ic at -10C and 0C do not show a statistically significant
dependence on the specimen diameter (table 19).

6.4.2 Dependence of the apparent fracture toughness on the specimen size


at 15C
If the data from figure 37 are plotted as function of the diameter, then
figure 41 is obtained. Figure 41 clearly shows a dependence of K Ic on the
specimen diameter at 15C.
A possible (and probable) explanation is the stress reversal along the
specimen vertical, and the existence of a large zone of compressive stress
in the specimen, cf. 8. It is made plausible in 8 that the maximum applied
force is greater if the tensile stress, necessary for the specimen to fracture,
develops in the presence of a compressive stress. The smaller the
specimen, the greater is the relative importance of the zone of the
compressive stress. It is not clear, how this causes a dependence of K Ic on
the diameter at 15C, while such dependence is not found at lower
temperatures (0C and lower). A possible explanation might be that the
fracture process zone is smaller at low temperature, so that it is not
influenced by the presence of the compressive zone.
6.4.3 Validity of the fracture toughness
In order for the fracture toughness to be valid, it must fulfil the conditions
according to equation (17). In the following, first condition (17d) is
investigated, and then conditions (17a)-(17c). The apparent fracture
toughness was calculated using equation (16). Table 20 summarises the
score of specimens satisfying condition (17d).
The first number indicates the number of observations satisfying the
condition, and the number between brackets indicates the total number of
observations available for the given testing conditions (specimen size,
displacement rate and temperature). The results in table 20 indicate that
there is too much plasticity at 15C, causing Pmax to become large relative
to P0 . Therefore, only the results obtained for -10C and 0C are used in
the following analysis.
In the following, conditions (17a)-(17c) are investigated. Let us assume,
the stress strain behaviour is linearly elastic, and that the yield strength is
equal to the tensile strength. This condition is fulfilled if there is no strain
hardening. For asphalt mixture it is fulfilled at temperatures of 0C or
lower. Let us consider equation (7),

Sec. 6

Results and analysis of fracture toughness tests

35

30

30

K-IQ (MPamm)

K-IQ (MPamm)

35

229

25

25

20

20

15

15

10

10

0
0

100

200

300

diameter (mm)
v=0.005

v=0.05

100

200

300

diameter (mm)
v=0.5

v=0.005

v=0.05

v=0.5

35

K-IQ (MPamm)

30
25
20
15
10
5
0
0

100

200

300

diameter (mm)
v=0.005

v=0.05

v=0.5

Figure 41. Apparent fracture toughness, K IQ as a function of the specimen


diameter, and the displacement rate, for unmodified dense asphalt concrete
0/16. Temperature: 15C. Top left (a): specimen thickness: 25 mm. Top right
(b): specimen thickness: 50 mm. Bottom (c): specimen thickness: 75 mm.
Displacement rates, v, are in mm/s.

K I = max a f (a / W )
(64a)
max = k Pmax / D B
(64b)
where k is a constant, k = 4.263. At low temperatures, 10C - 0C, the
force-displacement diagram is of the type of that in figure 12a, hence in
equation (16b)
P0 = Pmax
(64c)
Applying equation (13) to the notched specimen, using equation (64c)
yields
max = BTS n = k 0
(64d)
where BTS n is the bending tensile strength of the notched specimen. It

230

Resistance to crack-growth and fracture

Ch. 4

Table 20. K IQ -validity score for condition (17d). The number before the brackets
indicates the number of tests satisfying condition (17d). The number between
brackets indicates the total number of tests performed for the particular testing
conditions (specimen size, temperature, and displacement rate).
v = 0.05 mm/s
100 x 25
150 x 25
220 x 25
100 x 50
150 x 50
220 x 50
100 x 75
150 x 75
220 x 75

-10C
2(2)
2(2)
2(2)
2(2)
2(2)
2(2)
2(2)
2(2)
2(2)

0C
2(2)
2(2)
1(2)
2(2)
2(2)
2(2)
2(2)
2(2)
2(2)

15C
0(2)
0(2)
0(2)
0(2)
1(2)
0(2)
1(2)
0(2)
0(2)

25C
2(2)
0(2)
0(2)
1(2)
0(2)
0(2)
2(2)
0(2)
0(2)

can be assumed that the yield strength is equal to the bending tensile
strength, since the force-displacement diagram is of the type of that in
figure 12a; hence

ys = k 0
(65)
where ys is the yield strength of the notched specimen. Note, that ys is
lower than the yield strength of the material, ys . Therefore, substitution
of equation (65) in equations (17a)-(17c) yields a conservative estimate of
the critical specimen dimensions. Substitution of equation (64d) and
equation (65) into equations (17a)(17c) yields:
a 2.5 a f ( a / W ) 2
(66a)

(
)
B 2.5 ( a f ( a / W ) )
W 5 ( a f (a / W ) )

(66b)

(66c)
Unfortunately, f ( a / W ) according to equation (7c), is given only for the
case where s / r in table 3 is equal to 0.8; i.e. for the 100 mm diameter
specimen. If, in spite of that, equation (7c) is also applied to the 150 mm
diameter specimen and the 220 mm diameter specimen, then the f ( a / W )
and the critical specimen dimensions given in table 21 are obtained.
Table 21. Critical specimen dimensions according to the ASTMstandard E399-78A (ASTM 1979), required for plane strain.
Specimen
f (a / W )
B
W
diameter (mm)
(mm)
(mm)
236
118
1.230
100
210
105
1.156
150
136
68
0.932
220
158
79
1
220

Sec. 6

Results and analysis of fracture toughness tests

231

According to table 21, the specimen thickness, B , must be greater than


118 mm for the 100 mm diameter specimen, must be greater than 105 mm
for the 150 mm diameter specimen, and must be greater than 68 mm for
the 220 mm diameter specimen in order to fulfil the ASTM size
requirements. Similarly, the critical width (radius) is shown in the right
column. Thus, table 21 shows, that the ASTM-requirements for plane
strain are not fulfilled.
If f ( a / W ) is assumed to be equal to 1 for the 220 mm specimen diameter
(since f ( a / W ) is a correction for the finiteness of the specimen, which is
normally greater than 1), then the required minimum thickness is 79 mm,
and the required width 158 mm.
Lim et al. (1994), based on extensive referenced research, have indicated
that the ASTM-standard provides a very conservative estimate of the
minimum size requirements, and that a fairly small specimen may be
used. The reason is that the fracture process zone in the SCB-specimen is
rather small.
Let us assume that Lims remark regarding the ASTM-requirement is
correct, and that the stress state of 220/75 mm specimen is plane strain.
We may then compare the results of the other thicknesses, to see if the
apparent fracture toughness increases owing to developing plane stress as
the thickness decreases. The experimental data in figure 36 (tables 16 and
17) give no indication that the apparent fracture toughness increases as the
specimen thickness decreases, cf. 6.4.1. Thus, the data for -10C and 0C
support Lims remark regarding the ASTM-requirements, that a relatively
small specimen can be used. The (provisional) conclusion is that the
apparent fracture toughnesses for -10C and 0C in figure 36 can be
considered valid fracture toughnesses irrespective of the specimen
diameter and the specimen thickness.

6.4.4 Dependence of the apparent fracture toughness on the displacement


rate
The apparent fracture toughness depends on the displacement rate. This is
shown by the experimental data in figure 37. It is also shown in figure 42a.
This was obtained using the data in figure 37.
To obtain figure 42a, first, the K IQ in figure 37 (15C) were averaged per
displacement rate and per specimen diameter over the specimen thickness.
The results are represented by the solid curves. The K IQ from figure 36
(0.05 mm/s) were averaged per specimen diameter and per temperature
over the specimen thickness. This yielded two additional points in figure
42a; these are indicated by a marker in a dotted curve, labelled as 0C and
25C.
Similarly, figure 42b was constructed using the K IQ in figure 36. To
obtain figure 42b, the K IQ from figure 36 (0.05 mm/s) were averaged per

232

Resistance to crack-growth and fracture

Ch. 4

40
0C

apparent K Ic (MPa.mm1/2)

35
30

25

15C

20
15
10
5
25C

0
-2.5

-2

-1.5

-1

-0.5

log v (mm/s)
100 mm/15C
150 mm/0C

150 mm/15C
150 mm/25C

220 mm/15C

35

0.5 mm/s
apparent K Ic (MPa.mm1/2)

30

0.05 mm/s
25
20
15
10

0.005 mm/s
5
0
-15

-10

-5

100 mm/0.05 mm/s


220 mm/0.05 mm/s
150 mm/0.5 mm/s

5
10
temperature (C)

15

20

25

30

150 mm/0.05 mm/s


150 mm/0.005 mm/s

Figure 42. Time temperature equivalence and specimen size dependence of


apparent fracture toughness in data in figures 36 - 37. Markers represent
average values. Top (a): K IQ as function of log v. Bottom (b): K IQ as function of
the temperature.

Sec. 7

Results and analysis of tensile strength

233

specimen diameter and per temperature over the specimen thickness. The
results are represented by a solid curve and two dashed curves for 0.05
mm/s. The K IQ from figure 37, i.e. those for 15C/0.005 mm/s and
15C/0.5 mm/s, were averaged per displacement rate and per specimen
diameter over the specimen thickness. This yielded two additional points
in figure 42b; these are indicated by a marker in a dotted curve, labelled as
0.005 mm/s and 0.5 mm/s.

7 Results and analysis of tensile strength using the uniaxial


tensile (UT) specimen, the semi-circular bending (SCB) specimen,
and the indirect tensile (IT) specimen
The conditions of linear elasticity and plane strain state of stress which
have to be fulfilled in order for the fracture toughness to be valid in
respect of the ASTM requirements, are very strict. At temperatures above
approximately 10C, these conditions are not fulfilled. Then, it is not
possible to obtain a valid fracture toughness. As an alternative it may be
useful to determine the tensile strength instead of the fracture toughness.
At temperatures above 10C the stress strain behaviour of asphalt mixtures
will be elasto-viscoplastic. Under those conditions, the tensile strength is
probably influenced by the heterogeneity of the material, its nonlinear
stress strain behaviour, and the three-dimensional stress distribution. As a
result, different values of the tensile strength may be obtained when using
different specimen geometries. This was investigated in an experimental
study using the uniaxial tensile test specimen, the semi-circular bending
test specimen, and the indirect tensile test specimen. The results are
described in the following.
7.1 Case 1: Fine porous asphalt - Uniaxial tensile strength and bending
tensile strength Influence of temperature and displacement rate
Figure 43, for fine porous asphalt, PA 4/8, compares the uniaxial tensile
strength, UTS, the bending tensile strength of the un-notched specimen,
BTS, and the bending tensile strength of the notched specimen, BTS-N,
for different displacement rates, as function of the temperature. Each data
point represents an average of 24 tests. It can be seen in figure 43, that the
uniaxial tensile strength is somewhat lower than the bending tensile
strength of the un-notched specimen, and is practically equal to the
bending tensile strength of the notched specimen.
In figure 43, the different tensile strengths are compared under comparable
conditions (material, temperature, and displacement rate), using different
specimens. However, the stress states of the different specimens differ. It
is expected that the tensile strength obtained with the un-notched SCBspecimen is greater than that obtained with the notched SCB-specimen,
since the notch weakens the specimen.

Resistance to crack-growth and fracture

234

Bending tensile strength


un-notched specimen

5
BTS (MPa)

UTS (MPa)

Uniaxial tensile strength


3.5
3
2.5
2
1.5
1
0.5
0

Ch. 4

4
3
2
1
0

-20

-10

0
10
temperature (C)

0.01 mm/s
0.10 mm/s

20

30

-20

0.05 mm/s

-10

0
10
20
temperature (C)

0.01 mm/s
0.10 mm/s

30

0.05 mm/s

BTS-N (MPa)

Bending tensile strength


notched specimen
3,5
3
2,5
2
1,5
1
0,5
0
-20

-10

0
10
20
30
temperature (C)
0.01 mm/s
0.05 mm/s
0.10 mm/s

Figure 43. Fine porous asphalt 4/8. Tensile strength as function of the temperature
for different displacement rates. Top left (a): Uniaxial tensile strength, 100 x 50
mm height x diameter. Top right (b): Semi-circular bending tensile strength unnotched specimen, 150 x 50 mm diameter x thickness. Bottom (c): Semi-circular
bending tensile strength, notched specimen, 150 x 50 mm diameter x thickness.

Figure 44 shows variation coefficients for the various conditions of


temperature and displacement rate. In this investigation, the variation
coefficient of the UTS varied between 2.1% and 6.2%, the variation
coefficient of the BTS (un-notched SCB-specimen) varied between 4%
and 12%, and the variation coefficient of the BTS-N (notched SCBspecimen) varied between 4.5% and 14.5%. These are representative
variation coefficients for asphalt mixture. In general, a variation
coefficient lower than 10% may be considered good, and a variation
coefficient of up to 15% fair. (With coarse graded asphalt mixtures, 0/11
and coarser, in particular when produced in a practical context, variation
coefficients greater than 20% are not uncommon, cf. figures 47-50).
7.2 Case 2: Porous asphalt - Uniaxial tensile strength and bending tensile
strength Influence of mixture grading, bitumen content, and type of
bitumen
In figure 45, the uniaxial tensile strength is given for twelve porous
asphalt mixtures, as function of the displacement rate, cf. table 6, section
C. The specimen used was cylindrical, 50 x 100 mm diameter x height.

Sec. 7

Results and analysis of tensile strength

235

16

variation coefficient (%)

14
12
10
8

UTT
BTS
BTS-N

6
4
2
0

Figure 44. Variation coefficients for the tests and testing conditions of
temperature and displacement rate of figure 43. UTS = uniaxial tensile strength,
BTS = bending tensile strength (un-notched SCB-specimen), BTS-N = bending
tensile strength of notched SCB-specimen.

The mixtures are arranged in four groups, indicated by the letters A, B, C,


and D, which indicate the mixture grading and the bitumen content: A:
porous asphalt, PA 0/16, 4.0% bitumen, B: PA 0/16, 5.0% bitumen, C: PA
0/11, 4.0% bitumen, and D: PA 0/11, 5.0% bitumen. The type of polymer
modification is indicated in the legend by the letters F, S, and P. The letter
F indicates a plastomer modified binder, the letter S indicates an elastomer
modified binder, the letter P indicates a pure 70/100 bitumen. Twelve sets
of three bars are shown. The left bar of each set is for displacement rate
0.01 mm/s, the middle one for 0.05 mm/s, and the right one for 0.10 mm/s.
Figure 45 shows that a 1% m/m increase in binder content has a positive
effect on the tensile strength, but only at displacement rates greater than or
equal to 0.05 mm/s. This can be seen by comparing group B to group A,
group D to group C, and by comparing in group B and group D the bars
obtained for the different displacement rates. In addition, figure 45 shows
that the displacement rate can have a significant effect on the tensile
strength. For example, for PA 0/16 with 5,0% 70/100 bitumen (group B,
bitumen P), the tensile strength increases from 1.7 MPa to 3.1 MPa.
However, not always is the effect significant; for example, the
displacement rate in group A does not seem to have a significant effect on
the tensile strength.

236

Resistance to crack-growth and fracture

Ch. 4

uniaxial tensile test, 1C


3.5
F 0.01 mm/s
F 0.05 mm/s
F 0.10 mm/s

tensile strength (MPa)

3
2.5

S 0.01 mm/s
S 0.05 mm/s
S 0.10 mm/s

2
1.5

P 0.01 mm/s
P 0.05 mm/s
P 0.10 mm/s

1
0.5
0
A

porous asphalt mixture

Figure 45. Uniaxial tensile strength of four porous asphalt (PA) mixtures, A, B,
C, D, for three displacement rates: 0.01 mm/s, 0.05 mm/s, and 0.10 mm/s.
Specimen size: 50 x 100 mm diameter x height. Mixture A: PA 0/16, 4.0%
bitumen, mixture B: PA 0/16, 5% bitumen, mixture C: PA 0/11, 4% bitumen,
mixture D: PA 0/11, 5% bitumen. F: plastomer modified bitumen, S: elastomer
modified bitumen, P: pure (unmodified) bitumen. Each bar represents an
average of three measurements.

In figure 46, the bending tensile strengths and corresponding variation


coefficients are given for the same porous asphalt mixtures as in figure 45.
The specimen was semi-circular, 150 x 50 mm diameter x thickness. The
applied displacement rate was 0.085 mm/s. Note, that the trend in figures
46a and 46c is similar to that in figure 45, showing a slightly higher tensile
strength for the higher bitumen content. Note further, that the bending
tensile strength of the notched specimen, figure 46c, is lower than the
bending tensile strength of the un-notched specimen. This was expected,
because the notch weakens the specimen.
7.3 Case 3: Various asphalt concrete and porous asphalt mixtures Uniaxial tensile strength, bending tensile strength, and indirect tensile
strength - Influence of temperature and displacement rate
In figure 47, the uniaxial tensile strength and the corresponding variation
coefficient are given for five projects (road trials), A..E, for each of which
five crushed gravel asphalt concrete 0/22 mixtures, indicated as MD, MD-,
MD+, PLA, and PAV, were manufactured under different conditions, cf.
table 6, section D. Each bar in figure 47 represents an average of four
tests. The specimen was cylindrical, 200 x 50 mm height x diameter. The
applied displacement rate was 0.5 mm/s.

Sec. 7

Results and analysis of tensile strength

semi-circular bending test


unnotched specimen

4
F
S
P

3
2
1

variation coefficient (%)

tensile strength (MPa)

0
A
B
C
D
porous asphalt mixture

A
B
C
D
porous asphalt mixture

semi-circular bending test


notched specimen

semi-circular bending test


notched specimen
20
18
16
14
12
10
8
6
4
2
0

4
F
S
P

3
2
1
0
A
B
C
D
porous asphalt mixture

variation coefficient (%)

5
tensile strength (MPa)

semi-circular bending test


unnotched specimen
20
18
16
14
12
10
8
6
4
2
0

237

F
S
P

F
S
P

A
B
C
D
porous asphalt mixture

Figure 46. Bending tensile strength and variation coefficients of twelve porous
asphalt mixtures. Specimen size: 150 x 50 mm diameter x thickness.
Displacement rate: 0.085 mm/s. Mixtures A: porous asphalt 0/16 with 4.0%
bitumen F, S, or P. Mixtures B: porous asphalt 0/16 with 5.0% bitumen F, S, or
P. Mixtures C: porous asphalt 0/11 with 4.0% bitumen F, S, or P. Mixtures D:
porous asphalt 0/11 with 5.0% bitumen F, S, or P. F: plastomer modified
bitumen, S: elastomer modified bitumen, P: pure (unmodified) 70/100 bitumen.
Each tensile strength bar represents an average of three measurements.

Figures 47b/d/f show quite large variation coefficients. Values greater than
25% are found in laboratory manufactured mixtures (MD, MD-, and
MD+) and in mixtures processed under practical circumstances (PLA and
PAV). The fact that variation coefficients greater than 25% occur is
disappointing. There is a variety of possible causes (which were not
investigated in detail as a part of this study): Manufacturing conditions
(the mixtures are coarse grained, 0/22, and contain large quantities of
reclaimed asphalt, cf. 4.3.3; poor homogenisation and/or varying dosage of

238

Resistance to crack-growth and fracture

uniaxial tensile test, 1C


C

uniaxial tensile test, 11C


C
variation coefficient (%)

tensile strength
(MPa)

5
4
3
2
1
0
A

60
50
40
30
20
10
0
avg

project
MD

MD -

MD +

PLA

PAV

MD

4
3
2
1
0

MD -

MD +

PLA

0,3
0,2
0,1
0

MD -

project

MD +

PLA

MD +

PLA

PAV

20
15
10
5
0

MD

variation coefficient (%)

uniaxial tensile strength


(MPa)

0,4

MD

MD -

project

MD +

PLA

E
PAV

uniaxial tensile test, 40C


40 C

0,5

25

uniaxial tensile test, 40C


40 C

MD -

avg

PAV

0,6

30

project
MD

uniaxial tensile test, 15


C
15C
variation coefficient (%)

uniaxial tensile strength


(MPa)

project

uniaxial tensile test, 15


C
15C

Ch. 4

25
20
15
10
5
0
avg

E
PAV

project
MD

MD -

MD +

PLA

PAV

Figure 47. Uniaxial tensile strength of crushed gravel asphalt concrete 0/22. Specimen size: 50 x 200 mm diameter x height. Displacement rate: 0.5 mm/s. Per project,
A..E, five asphalt mixtures were investigated: MD: mixture according to Marshall
mixture design method, MD-: MD mixture minus 0.5% bitumen, MD+: MD mixture
plus 0.5% bitumen, PLA: the plant-mixed mixture, slab-compacted at the site, PAV:
the mixture from the road pavement. Top (a,b): Average values and variation
coefficients at 1C. Middle (c,d): As (a,b), for 15C. Bottom (e,f): As (a,b), for 40C.
Each bar represents an average of three measurements. The left most bar in figs. b,
d, and f represents the average of the 25 bars on the right of it.

Sec. 7

SCB, un-notched specimen, 1C


10
8
6
4
2
0

SCB, un-notched specimen, 1C

20
15
10
5
0

A
MD

25

variation coefficient (%)

bending tensile strength


(MPa)

12

MD -

project

MD +

PLA

avg

PAV

MD

SCB, notched specimen, 1C

25

bending tensile strength


(MPa)

variation coefficient (%)

6
5
4
3
2
1

MD -

project
MD +

PLA

E
PAV

SCB, notched specimen, 1C

20
15
10
5
0

0
A
MD

239

Results and analysis of tensile strength

MD -

project
MD +

PLA

avg

E
PAV

MD

MD -

project

MD +

PLA

E
PAV

Figure 48. Bending tensile strength of crushed gravel asphalt concrete 0/22.
Specimen size: 150 x 50 mm diameter x thickness. Displacement rate: 0.085 mm/s.

Per project (road trial), A..E, five asphalt mixtures were investigated: MD:
mixture according to the Marshall mixture design method, MD-: MD mixture minus
0.5% bitumen, MD+: MD mixture plus 0.5% bitumen, PLA: the plant-mixed
mixture, slab-compacted at the site, PAV: the mixture from the road pavement. Top
(a,b): Average values and variation coefficients at 1C, un-notched specimen.
Bottom (c,d): As figure (a,b), for the notched specimen. Each bar represents an
average of three measurements. The left most bar in figs. b and c represents the
average of the 25 bars on the right of it.

reclaimed material), specimen preparation and handling (incidentally bad


specimens), and testing conditions (gluing and centring of the specimen).
Because of the many possible causes, the variation coefficients were not
investigated in detail as a part of the present study.
Figures 4849 show the bending tensile strengths and corresponding
variation coefficients for the same mixtures for which the uniaxial tensile
strength is shown in figure 47. The specimen was semi-circular, 150 x 50
mm diameter x thickness. The applied displacement rate was 0.085 mm/s.
Each bar in figures 4849 represents an average of four tests. Figure 48

240

Resistance to crack-growth and fracture

5
4
3
2
1
0
B

MD -

4.5

project
MD +

25
20
15
10
5

PLA

avg

PAV

MD

SCB, notched specimen, 15C

30

variation coefficient (%)

MD

4
3.5
3
2.5
2
1.5
1
0.5
0

MD -

project

MD +

PLA

E
PAV

SCB, notched specimen, 15C

25
20
15
10
5
0

A
MD

30

0
A

bending tensile strength


(MPa)

SCB, un-notched specimen 15C

variation coefficient (%)

bending tensile strength


(MPa)

SCB, un-notched specimen, 15C


6

Ch. 4

MD -

project

MD +

PLA

avg

E
PAV

MD

MD -

project

MD +

PLA

E
PAV

Figure 49. As figure 45, 15C. Top left (a): As figure 45a 15C. Top right (b): As
figure 45b, 15C. Bottom left (c): As figure 45c, 15C. Bottom right (d): As
figure 45d, 15C.

shows the bending tensile strength of the un-notched specimen and the
notched specimen at 1C. Figure 49 shows the bending tensile strength of
the un-notched specimen and the notched specimen at 15C.
Figure 50 shows the indirect tensile strength and the corresponding
variation coefficient for the same mixtures as in figure 47. The specimen
was cylindrical, 150 x 50 mm diameter x thickness. The applied
displacement rate: 0.85 mm/s). Each bar in figure 50 represents an average
of four tests.
7.4 Case 4: Asphalt concrete - Indirect tensile strength Influence of
specimen size, loading strip, temperature, and mixture composition
In figure 51, the indirect tensile strength is given for gravel asphalt
concrete, GAC, and dense asphalt concrete, DAC, as a function of the
temperature, cf. table 6, section E. For the GAC specimens, the specimen
diameter was 148 mm. For the DAC specimens, the specimen diameter

Sec. 7

3
2.5
2
1.5
1
0.5

30
25
20
15
10
5
0

0
A
MD

indirect tensile test, 15C

indirect tensile test, 15C


variation coefficient (%)

3.5

indirect tensile strength


(MPa)

241

Results and analysis of tensile strength

MD -

project

MD +

PLA

avg

E
PAV

MD

MD -

project
MD +

PLA

E
PAV

Figure 50. Indirect tensile strength at 15C. Specimen size: 150 x 50 mm diameter x
thickness. Displacement rate: 0.85 mm/s. Per project (road trial), A..E, the mean
value of three measurements and the corresponding variation coefficients are
given. MD: the mixture according to the Marshall mixture design method, MD-:
the MD mixture minus 0.5% bitumen, MD+: the MD mixture plus 0.5% bitumen,
PLA: the plant-mixed mixture slab-compacted at the site, PAV: the mixture from
the road pavement. Left (a): Mean values. Right (b): Corresponding variation
coefficients. The left most bar represents the average of the 20 bars right to it.

was 97 mm. Figure 51 shows the results for different specimen


thicknesses, 25 mm, 50 mm and 75 mm, and two different loading strip
widths, 12.7 mm (0.5 inch), and 25.4 mm (1 inch). In all cases the
displacement rate was 0.85 mm/s. The corresponding standard deviations
and variation coefficients are listed in tables 22 and 23.
Figure 52 shows the indirect tensile strength of the laboratory
manufactured mixtures in table 6, section B. The specimen was 150 x 50
mm diameter x thickness, and the displacement rate was 0.5 mm/s. The
mixtures are:
. unmodified dense graded asphalt concrete, DAC 0/16;
. the same DAC 0/16 mixture, containing a polymer modified binder, i.e.
a) elastomer modified binder,
b) plastomer modified binder;
. porous asphalt, containing a polymer modified binder, i.e.
a) elastomer modified binder,
b) plastomer modified binder;
. stone mastic asphalt mixture, containing a polymer modified binder, i.e.
a) elastomer modified binder,
b) plastomer modified binder.
DAC 0/16 unmod is the reference mixture, an unmodified standard
dense graded asphalt concrete. Figure 52a shows that a polymer
modification increases the indirect tensile strength. In the present study,
the plastomer modified bitumen yields a greater increase than the

242

Resistance to crack-growth and fracture

Ch. 4

GAC 0/32

ind. tensile strength (MPa)

6
5
4
3
2
1
0
-10

-5

25 mm/0.5"
25 mm/1"

10

15

temperature (C)
50 mm/0.5"
50 mm/1"

20

75 mm/0.5"
75 mm/1"

DAC 0/16

ind. tensile strength (MPa)

7
6
5
4
3
2
1
0
-10

-5

10

temperature (C)

15

20

25 mm/0.5"

50 mm/0.5"

75 mm/0.5"

25 mm/1"

50 mm/1"

75 mm/1"

25

Figure 51. Indirect tensile strength as a function of the temperature, for different
specimen thicknesses: 25 mm, 50 mm, and 75 mm, and different widths of the
loading strip: 12.7 mm (0.5 inch) and 25.4 mm (1 inch). Each point represents
the mean of four measurements. Top (a): Gravel asphalt concrete 0/32.
Specimen diameter: 148 mm, displacement rate: 0.85 mm/s. Bottom (b): Dense
graded asphalt concrete 0/16. Specimen diameter: 97 mm, displacement rate:
0.85 mm/s.

Sec. 7

Results and analysis of tensile strength

243

Table 22. Indirect tensile strength in MPa of gravel asphalt concrete, GAC 0/32
(figure 51a). Mean (m) of four measurements, standard deviation (s), and variation
coefficient (v), for different specimen thicknesses: 25 mm, 50 mm, and 75 mm, and
test temperatures.
GAC 0/32 specimen thickness:
specimen thickness:
specimen thickness:
25 mm
50 mm
75 mm
temp.
m
s
v
m
s
v
m
S
V
(C)
0.5 loading strip
0.44 0.11 4.43 0.76 0.17
3.55 0.57 0.16 4.15
-5
0.61 0.15 3.98 0.30 0.08
3.66 0.47 0.13 4.04
0
0.20 0.08 2.77 0.15 0.05
2.29 0.15 0.07 2.60
10
0.10 0.08 1.33 0.03 0.02
1.11 0.07 0.06 1.33
20
1 loading strip
0.15 0.03 5.21 0.27 0.05
4.63 0.53 0.11 5.22
-5
0.42 0.09 4.64 0.34 0.07
4.40 0.24 0.05 4.71
0
0.21 0.07 3.42 0.16 0.05
2.64 0.42 0.16 2.99
10
0.08 0.06 1.47 0.07 0.05
1.36 0.10 0.07 1.45
20

Table 23. Indirect tensile strength in MPa of dense asphalt concrete, DAC 0/16
(figure 51b). Mean (m) of four measurements, standard deviation (s), and variation
coefficient (v), for different specimen thicknesses: 25 mm, 50 mm, and 75 mm, and
test temperatures.
DAC 0/16 specimen thickness:
specimen thickness:
specimen thickness:
25 mm
50 mm
75 mm
temp.
m
s
v
m
s
V
m
S
V
(C)
0.5 loading strip
4.93 0.83 0.17 4.84 0.42 0.09 3.71 0.62 0.17
-5
4.16 0.61 0.15 4.39 0.31 0.07 3.73 0.39 0.10
0
2.26 0.25 0.11 2.25 0.23 0.10 2.30 0.25 0.11
10
1.04 0.06 0.06 0.97 0.12 0.12 0.91 0.06 0.07
20
1 loading strip
6.46 0.59 0.09 6.30 0.75 0.12 6.18 0.41 0.07
-5
5.93 0.29 0.05 5.59 0.30 0.05 5.12 0.53 0.10
0
3.02 0.19 0.06 2.90 0.15 0.05 2.99 0.16 0.05
10
1.39 0.11 0.08 1.25 0.21 0.17 1.17 0.09 0.08
20

244

Resistance to crack-growth and fracture

Ch. 4

elastomer modified bitumen. This result depends on the specific binders


used. It is noted, that apart from the polymer, also the base bitumen
contributes to the tensile strength of the asphalt mixture, and that the base
bitumen of an elastomer modified normally differs from that of a
plastomer modified bitumen.
Note, that the indirect tensile strength of porous asphalt is lower than the
indirect tensile strength of dense asphalt concrete and stone mastic asphalt.
This is attributed to the open graded structure of porous asphalt.
Note further, that, in the present study, the indirect tensile strength of the
polymer modified SMA 0/11 mixtures is lower than the indirect tensile
strength of the polymer modified DAC 0/16 mixtures. This is merely an
experimental observation, which should not be generalised.
7.5 Analysis of the results
The purpose of the following analysis is to investigate the influence of the
specimen geometry. Data were gathered for three different types of
specimen, the uniaxial tensile (UT) test specimen, the semi-circular
bending (SCB) test specimen, and the indirect tensile (IT) test specimen.
Unfortunately, the influence of the specimen geometry is obscured,
because the data were obtained for various materials and testing
conditions. It is still possible to investigate the influence of the specimen
size, but not without simultaneous consideration of the influences of the
material composition and the testing conditions. As a result, the data have
lost the power of proof of the influence of the specimen geometry.
Moreover, the analysis becomes relatively complicated. For the
convenience of the reader the results of the analysis are summarised in the
framed parts of each of the following subsections.
The followed approach was divided in two ways:
1 in specimen geometry; a comparison of the results for the UT- specimen
and the SCB-specimen, 7.5.1 and 7.5.2, and a separate of analysis of the
IT test results, 7.5.3 and 7.5.4;
2 in temperature; an analysis of the results at low temperature, i.e. 0C and
lower, 7.5.1 respectively 7.5.3, and high temperature, i.e. 15C and higher,
7.5.2 respectively 7.5.4.
In the previous sections, it was shown that the tensile strength as
determined in different tests seems to depend on the type of test. This
indicates that the tensile strength value obtained might not always be the
true tensile strength, which implies that it is a specimen property rather
than a material property.
In the following, the argument is used that the tensile strength is greater if
the tensile stress develops in the presence of compressive stress. This
argument is based on a finite element study of the SCB-specimen, which is
described in 8. It is probable that the same argument is applicable to the

Sec. 7

245

Results and analysis of tensile strength

ind. tensile strength (MPa)

6
5
4
3
2
1
0
-20

-10

10

20

30

temperature (C)
DAC 0/16 unmod
PA 0/16 elast
SMA 0/11 plast

DAC 0/16 elast


PA 0/16 plast

DAC 0/16 plast


SMA 0/11 elast

variation coefficient (%)

25

20

15

10

0
-10

15

25

temperature (C)
DAC 0/16 unmod
PA 0/16 elast
SMA 0/11 plast

DAC 0/16 elast


PA 0/16 plast

DAC 0/16 plast


SMA 0/11 elast

Figure 52. Average indirect tensile strength and corresponding variation


coefficients of different asphalt mixtures, as function of the temperature.
Specimen: 150 x 50 mm diameter x thickness, displacement rate: 0.85 mm/s,
loading strip width: 25.4 mm. DAC 0/16: Dense graded asphalt concrete 0/16;
unmod: with unmodified 80/100 bitumen. PA 0/16: Porous asphalt 0/16. SMA
0/11: Stone mastic asphalt 0/11; elast: with elastomer modified bitumen; plast:
with plastomer modified bitumen. Top (a): Average indirect tensile strength.
Each data point represents the average of two measurements. Bottom (b):
Corresponding variation coefficients.

246

Resistance to crack-growth and fracture

Ch. 4

IT-specimen, since there is also a compressive stress in the IT-specimen.


However, this was not investigated in detail. In opposition to the increase
of the tensile strength in the presence of a compressive stress, there is the
limitation of the tensile strength, caused by the presence of the stress/strain
concentration near the loading strip, which limits the maximum applied
force. Depending on which of these two influences rules the stress strain
behaviour, the indirect tensile strength, ITS, may be greater or smaller
than the uniaxial tensile strength, UTS, under similar conditions of
temperature and displacement rate.
7.5.1 Uniaxial tensile strength and bending tensile strength at low
temperature (0C and lower)
It is made plausible in 7.5.1 that, for a given material, the BTS at low
temperature, 0C and lower, is consistently greater than the UTS
under similar conditions of temperature and displacement rate. This is
explained by the presence of a compressive stress in the SCBspecimen.
At 0C and lower, no strain hardening was observed. This means that the
stress strain behaviour is approximately linearly elastic. Let us therefore
first investigate the data obtained at 0C and lower.
It can be seen in figure 43, that the tensile strength increases with
increasing displacement rate. This is well known as a phenomenon, not
only for asphalt mixtures. It is in agreement also with expectation based on
the time temperature equivalence, which tells that the material responds
stiffer as the loading-time is shorter, i.e. as the displacement rate is faster.
However, also deviations from this expectation may occur, for example in
figure 45. In figure 45, four groups, A, B, C, and D of porous asphalt (PA)
mixtures are shown; in group A, PA 0/16 with 4.0% bitumen, in group B,
PA 0/16 with 5% bitumen, in group C, PA 0/11 with 4.0% bitumen, in
group D, PA 0/11 with 5% bitumen. Twelve sets of three bars are shown
from left to right. In each set of three bars, the left one is for displacement
rate 0.01 mm/s, the middle one for 0.05 mm/s, and the right one for 0.10
mm/s. In five sets, the tensile strength increases as function of increasing
displacement rate, in agreement with expectation. However, in seven sets,
the result deviates from the expectation. Two explanations can be given
for this:
1 the asphalt mixtures are relatively coarse, i.e. with 0/16 grading;
2 the number of observations (test repetitions) per result is only four.
The coarser the material, the greater is the probability that the material
heterogeneity causes deviations from the average behaviour. The smaller
the number of observations, the greater is the probability that observed
behaviour deviates from the average behaviour.

Sec. 7

Results and analysis of tensile strength

247

If the displacement rate increases from 0.01 mm/s to 0.05 mm/s, then the
average increase in tensile strength in figure 45 is9:
4.0% bitumen (group A and group C):
0.45 MPa
(67a)
5.0% bitumen (group B and group D):
0.71 MPa
(67b)
If the displacement rate increases from 0.05 mm/s to 0.10 mm/s, then the
average increase in tensile strength is1:
4.0% bitumen (group A and group C):
0.25 MPa
(67c)
5.0% bitumen (group B and group D):
0.23 MPa
(67d)
Thus, the influence of the displacement rate on the tensile strength seems
to be smaller at higher displacement rate.
Apart from the displacement rate, it can be expected that the tensile
strength depends on the material composition, and the material structure.
As for the material composition, it can be seen, in figure 45, that the
tensile strengths are higher in group B and group D than in group A and
group C. The difference in tensile strength between groups B and D on the
one hand, an groups A and C on the other, is attributed to the bitumen
content, which is 1% higher for the mixtures in groups B and D. In figure
45, the tensile strength increases by roughly 1 to 1.5 MPa if the bitumen
content is increased by 1%, irrespective of the mixture grading. A possible
explanation might be that the cohesive strength increases with increasing
bitumen content, if this is lower than the optimum bitumen content.
As for the material structure, it was expected that the mixture grading
would have an effect on the tensile strength, based on the assumption that
the cohesive strength would increase as function of the number of stonestone bonds per unit volume specimen. Figure 45 does not support this
expectation, since the tensile strengths in group A (0/16 grading) are not
significantly different from those in group C (0/11 grading).
The uniaxial tensile strength, UTS, at 1C in figure 43a, may be compared
to the UTS in figure 45, which is also at 1C, the only difference being the
material. The average UTS at 1C in figure 43a range between 1.6 MPa
and 2.8 MPa. The average UTS at 1C in figure 45 range between 1.6
MPa and 3.2 MPa. The UTS in figure 43a for displacement rate 0.01 mm/s
(average: 1.6 MPa) corresponds to the UTS in figure 45 for displacement
rate for 0.01 mm/s (which range between 1.6 MPa and 2.1 MPa,
irrespective of the mixture composition). Thus, the finer 4/8 grading does
not seem to have much of an effect on the UTS at 0.01 mm/s.
In conclusion at this point, the UTS was reproduced in the two
investigations of which the results are shown respectively in figure 43a
and figure 45.

Averaged only over those combinations which show an increase of more than
0.1 MPa.

248

Resistance to crack-growth and fracture

Ch. 4

Let us now consider figure 47a. This shows the UTS for crushed gravel
asphalt concrete at 1C for a displacement rate of 0.5 mm/s. It can be seen
that the UTS on average, group B excluded, is slightly greater than 3 MPa,
which is greater than the average UTS for 0.10 mm/s in figure 43a. The
greater UTS in figure 47a can be attributed to two causes:
1 the crushed gravel asphalt concrete mixture is a denser mixture, i.e.
contains less air voids, than the porous asphalt mixture, and therefore
shows a higher cohesive strength, hence a higher tensile strength;
2 at a displacement rate of 0.5 mm/s (figure 47a) a higher tensile strength
may be expected than at 0.10 mm/s (figure 43a).
From these two observations, it seems, that the influence of the
displacement rate on the UTS diminishes above 0.10 mm/s , since part of
the increase can be attributed to the greater density of the crushed gravel
asphalt concrete mixtures. This is consistent with the observation made
previously, that the influence of the displacement rate decreases with
increasing displacement rate, cf. figure 43, and (67).
Let us now consider figure 43b. This shows the bending tensile strength,
BTS. The average BTS-values for -10C and 1C and the different
displacement rates in figure 43b are all greater than the corresponding
UTS-values for -10C and 1C and the respective displacement rates in
figure 43a. It can be seen also in figure 43b, that the influence of the
displacement rate diminishes between 0.05 mm/s and 0.10 mm/s.
Therefore, the BTS obtained at a displacement rate of 0.085 mm/s is not
expected to differ greatly from that obtained at 0.10 mm/s. This is
supported by the BTS of the un-notched specimen in figure 46a, which
were obtained for a displacement rate of 0.085 mm/s. The BTS-values in
figure 46a range between 3.4 MPa and 4.8 MPa. The overall average BTS
from figure 46a is slightly smaller than 4 MPa, which corresponds to the
value for 1C and 0.085 mm/s expected from figure 43b.
Let us now consider figure 48a. This shows the BTS for crushed gravel
asphalt concrete at 1C for a displacement rate of 0.085 mm/s. In figure
48a, the BTS is on average approximately equal to 6 MPa. This is
considerably greater than the average BTS at 1C and 0.10 mm/s in figure
43b. The greater BTS in figure 48a can be attributed to the greater density
(i.e. the lesser air voids content) of the crushed gravel asphalt concrete
mixture, which is therefore expected to show a higher cohesive strength,
hence a higher tensile strength.
Thus, assuming the available data are representative, the BTS seems
consistently greater than the UTS under similar conditions of temperature
and displacement rate. This can be explained, cf. 8, by the presence of a
compressive stress in the SCB-specimen.

Sec. 7

Results and analysis of tensile strength

249

7.5.2 Uniaxial tensile strength and bending tensile strength at high


temperature (15C and higher)
It is made plausible in 7.5.2 that, for a given material, the BTS at high
temperature, 15C and higher, is consistently greater than the UTS
under similar conditions of temperature and displacement rate. This is
explained by the presence of a compressive stress in the SCBspecimen.
As mentioned previously, the tensile strength increases with increasing
displacement rate. This is best seen in figure 43. If the displacement rate
increases from 0.01 mm/s to 0.05 mm/s, then the average increase in
tensile strength is:
figure 43a:
0.20 MPa
(68a)
If the displacement rate increases from 0.05 mm/s to 0.10 mm/s, then the
average increase in tensile strength is:
figure 43a:
0.12 MPa
(68b)
By comparison of (68) to (67), it can be seen that the effect of the
displacement rate on the tensile strength is relatively small at high
temperature. In addition, between 0.05 mm/s and 0.10 mm/s, the influence
of the displacement rate is relatively small. Based on this it might be
assumed that also the influence of the displacement rate on the result in
figure 47c is small. This means that the result in figure 47c can be
compared to the result in figure 43a. The higher uniaxial tensile strength
of the crushed gravel asphalt concrete mixtures (figure 47c) in comparison
to that of the fine porous asphalt mixture (figure 43a) is not explained by
the displacement rate. It is probable that the higher UTS of the crushed
gravel asphalt concrete mixture is to be attributed to the dense graded
structure of the mixture.
Similarly, the result in figure 49a can be compared to the result in figure
43b. The higher bending tensile strength of the crushed gravel asphalt
concrete mixtures (figure 49a) in comparison to that of the fine porous
asphalt mixture (figure 43b) is not explained by the displacement rate. It is
probable that the higher BTS of the crushed gravel asphalt concrete
mixture is to be attributed to the dense graded structure of the mixture.
In figure 47c, the average UTS is 3.36 MPa, with minimum UTS and
maximum UTS respectively 2.02 MPa and 4.76 MPa. In figure 49a, the
average BTS of the un-notched specimen is 4.58 MPa, with minimum
BTS and maximum BTS respectively 3.84 MPa and 5.68 MPa. Thus,
assuming the available data are representative, the BTS seems consistently
greater than the UTS. This can be explained, cf. 8, by the presence of a
compressive stress in the SCB-specimen.

250

Resistance to crack-growth and fracture

Ch. 4

7.5.3 Indirect tensile strength at low temperature (0C and lower)


It is made plausible in 7.5.3 that, for a given material, the ITS at low
temperature, 0C and lower, is approximately equal or slightly greater
than the UTS under similar conditions of temperature and
displacement rate. The ITS is determined by two major influences.
On the one hand, the (calculated) ITS may be limited, owing to
localised plastic deformation near the loading strip (which limits the
maximum applied force).
On the other hand, the (calculated) ITS may be relatively high, if
localised plastic deformation is negligible, as it is usually at low
temperature. This can be explained by the presence of a compressive
stress in the IT-specimen.
The indirect tensile strength data available in this study are shown in
figure 51 (gravel asphalt concrete and dense graded asphalt concrete, -5C
and 0C, 0.85 mm/s, specimen diameter: 97 mm (DAC 0/16) and 148 mm
(GAC 0/32), specimen thickness: 25 mm, 50 mm, 75 mm, loading strip
width: 0.5 and 1.0), and in figure 52 (various types of asphalt mixture,
various temperatures, 0.85 mm/s, 150 x 50 mm specimen).
Effect of the loading strip width in figure 51
In figure 51, the indirect tensile strength, ITS, is shown as obtained for
gravel asphalt concrete, GAC, 0/32 and dense graded asphalt concrete,
DAC, 0/16. The ITS was determined at a displacement rate of 0.85 mm/s.
In figure 51, the effect of the width of the loading strip is shown. It can be
seen in figure 51, that the ITS values obtained with a 0.5 loading strip are
on average consistently lower than those obtained with a 1 loading strip.
This can be attributed to the concentration of stress and strain, which is
greater in the case of the smaller loading strip (0.5). In the case of the
0.5 loading strip, the stress/strain concentration takes a smaller volume of
specimen than in the case of the 1 loading strip. The probable effect is
that the maximum applied force is limited by plastic flow that is mainly
localised in the stress concentration. That explains that the calculated ITS
is lower with the 0.5 loading strip. In the following, the effect of the
loading strip is not further considered.
Comparison of the ITS values of GAC 0/32 and DAC 0/16 in figure 51
In the investigation shown in figure 51a and 51b, the ITS values of DAC
0/16 seem to be somewhat higher than those of GAC 0/32. This might be
attributable to the bitumen content, which is higher in the DAC mixture. A
possible explanation might be that the cohesive strength increases with
increasing bitumen content, if this is lower than the optimum bitumen
content.

Sec. 7

Results and analysis of tensile strength

251

Comparison of the ITS values of various materials in figure 52


Figure 52 compares the ITS values of different types of asphalt mixture.
The unmodified DAC 0/16 mixture was used as the reference mixture. It is
a dense graded asphalt concrete mixture that fulfils the RAW requirements
for standardised asphalt mixtures. A binder containing an elastomer, and a
binder containing a plastomer replaced the binder of this mixture. Thus,
two polymer modified asphalt mixtures were obtained, indicated in figure
52 respectively as DAC 0/16 elast, and DAC 0/16 plast. The same
elastomer modified binder and plastomer modified binder were used in the
porous asphalt mixture and stone mastic asphalt mixture, of which the
results are also shown in figure 52. Thus, also two polymer modified
porous asphalt mixtures were obtained indicated respectively as PA
0/16 elast and PA 0/16 plast and two polymer modified stone mastic
asphalt mixtures were obtained indicated respectively as SMA 0/11
elast and SMA 0/11 plast.
Figure 52 shows that polymer modification increases the indirect tensile
strength. The plastomer modified bitumen yields a consistently greater
increase in tensile strength in the dense graded asphalt concrete mixture,
the porous asphalt mixture, and the stone mastic asphalt mixture, than the
elastomer modified bitumen. This result should not be generalised, since it
probably depends on the specific binders used; it can be noted, that apart
from the polymer, also the base bitumen contributes to the tensile strength
of the asphalt mixture, and that the base bitumen of an elastomer modified
normally differs from that of a plastomer modified bitumen.
The ITS values of the porous asphalt mixtures are lower than those of the
DAC mixtures and lower than that of the SMA mixtures. This was
expected, and can be attributed to the open graded structure of porous
asphalt.
The ITS values of the SMA 0/11 elast mixture are lower than those of
the DAC 0/16 elast mixture, and the ITS values of the SMA 0/11 plast
mixture are lower than those of the DAC 0/16 plast mixture. This result
was not quite expected. At this point, it may not be useful to propose a
possible explanation, since the result may be accidental, depending on the
quality of the particular mixtures. A possible explanation may be that the
SMAs binder content (read: mastic = sand, filler and bitumen) is
relatively high, as compared to the DAC binder content (read: mortar =
filler and bitumen)10, and the fact that the base bitumen of the elastomer
modified binder is relatively soft (normally 200 pen). Yet, at this point it
may be preferred to consider it merely an experimental observation.
10

In stone mastic asphalt, the sand particles are to be considered part of the
binder, rather than the grain skeleton; in dense graded asphalt concrete, the sand
particles are to be considered part of the grain skeleton.

252

Resistance to crack-growth and fracture

Ch. 4

Comparison of the ITS values of DAC 0/16 in figures 51 and 52


In figure 52, the ITS values of the DAC 0/16 unmod mixture at 0C are
approximately 3 MPa. This is clearly lower than the ITS values of the
DAC 0/16 mixture of figure 51. An explanation for this, possibly
deviating, result is not available. The results might be attributable to the
mixture composition, e.g. the binder content, and the type and content of
filler of the asphalt mixture.
Comparison of the ITS values of GAC 0/32 in figure 51 and the UTS
values of CGAC 0/22 in figure 47a
Although obtained at a higher displacement rate (0.85 mm/s), the ITSvalues of GAC 0/32 for 0C in figure 51 might be compared to the UTS
values obtained for the crushed gravel asphalt concrete, CGAC, 0/22
mixtures in figure 47a. The latter were obtained at a displacement rate of
0.50 mm/s. The values in figure 51a range between 3.5 MPa and 4.8 MPa,
while the average is approximately 4.1 MPa. Those in figure 47a range
between 1.9 MPa and 5.2, while the average is approximately 3.2 MPa.
Thus, the range of ITS-values of GAC 0/32 for 0C in figure 51 is covered
completely by the range of UTS-values of CGAC 0/22 for 1C in figure
47a; however, the average ITS-value in figure 51a is greater than the
average UTS-value in figure 47a. Apart from the possibility that these
results are accidental, depending on the quality of the particular mixtures,
two probable causes can be given for the higher ITS values in figure 51:
. the higher displacement rate;
. the presence of a compressive stress in the IT-specimen.
Thus, the available data are somewhat ambiguous, and not pointing quite
consistently in the same direction. It seems that, in general, the ITS at low
temperature is approximately equal to the UTS. Occasionally, the ITS is
observed to be higher than the UTS under similar conditions of
temperature and displacement rate.
It seems, that the calculated ITS is mainly determined by the following
two influences. Firstly, the localisation of stress and strain near the loading
strip. This influence is usually absent at low temperature. If not absent, the
calculated ITS is limited, because the plastic flow in the stress/strain
concentration limits the maximum applied force.
Secondly, the tensile strength is greater if the tensile stress develops in the
presence of compressive stress. This was investigated for the SCBspecimen, cf. 8. It seems probable that this is also true for the ITspecimen, since there is a compressive stress in that specimen. However,
this was not investigated in detail.

Sec. 7

Results and analysis of tensile strength

253

7.5.4 Indirect tensile strength at high temperature (15C and higher)


It is made plausible in 7.5.4 that, for a given material, the ITS at high
temperature, 15C and higher, is consistently lower than the UTS
under similar conditions of temperature and displacement rate. This is
explained by the presence of a stress/strain concentration near the
loading strip, which limits the maximum applied force.
The indirect tensile strength, ITS, of crushed gravel asphalt concrete,
CGAC, 0/22, in figure 50a is not greatly different from that of the gravel
asphalt concrete, GAC 0/32, at 15C, interpolated in figure 51a. In figure
50a, the values range between 1.6 MPa and 3.2 MPa, while the average is
approximately 2.5 MPa. In figure 51a, the interpolated values range
between 1.6 MPa and 2.5 MPa, while the average is approximately 2 MPa.
This slightly lower average ITS of GAC in figure 51a, as compared to that
of CGAC in figure 50a, might be attributable to the lesser internal friction
in the grain skeleton of GAC, owing to use of round, i.e. uncrushed, and
coarser aggregate in this type of mixture.
The ITS-values in figure 50a and figure 51a were determined at a
displacement rate of 0.85 mm/s, which is somewhat higher than that (0.5
mm/s) at which the UTS in figure 47c was determined. In spite of the
higher displacement rate, the ITS in figure 50a is lower than the UTS in
figure 47c. Apart from the possibility that these results may be accidental,
depending on the quality of the particular mixtures, this might be
explained by localisation of plastic deformation near the loading strips in
the IT specimen. In that case, the force applied to the specimen is limited
by plastic flow of the material in the specimen, as a result of which the
calculated ITS is relatively low. Note, that, in that case, the calculated ITS
is not the true ITS of the material.
7.5.5 Summary
Summarising, an investigation was performed using three types of
specimen in order to compare the uniaxial tensile strength, UTS, the
bending tensile strength, BTS, and the indirect tensile strength, ITS. The
results were obtained for a variety of asphalt mixtures, for different
temperatures, and displacement rates, and, in case of the ITS, using
different loading strips.
The investigation provides evidence which shows that the BTS is
consistently higher than the UTS in the temperature range between -10C
and +25C. A probable explanation is that the tensile strength is greater if
the tensile stress develops in the presence of compressive stress. The
experimental evidence also suggests that the indirect tensile strength is
approximately equal, or slightly greater than the uniaxial tensile strength
in the temperature range between -10C and +10C. The cause of the

254

Resistance to crack-growth and fracture

Ch. 4

relatively low ITS at temperatures above 15C was not investigated in


detail. However, based on experience, i.e. the observation of localised
plastic deformation near the loading strips, it is plausible that the lower
ITS above 15C is explained by the limitation of the maximum applied
force caused by plastic flow in the material.
7.6 Relationship between fracture toughness and bending tensile strength
in the SCB-test
The fracture toughness and the bending tensile strength differ by a
constant factor, if there is no strain hardening. This is easily seen as
follows. If there is no strain hardening, then the yield strength is equal to
the tensile strength; i.e. the force displacement diagram is of the type of
figure 12a. This condition is fulfilled at temperatures of 0C and lower.
Under those conditions,
P0 = Pmax
(69a)

P0 = Pmax
(69b)
where P0 is the pertinent force on the notched specimen, P0 is the
pertinent force on the un-notched specimen, Pmax is the maximum applied
is the maximum applied force on
force on the notched specimen, and Pmax
the un-notched specimen. According to equation (16), the fracture
toughness differs by a constant factor, ac YI , from the pertinent stress
of the notched specimen, 0 , where ac is the notch-length. According to
equation (12), the bending tensile strength differs by a constant factor,
k = 4.263 , from the pertinent stress of the un-notched specimen, 0 . It is
probable that the pertinent stress of the notched specimen, and that of the
un-notched specimen differ by a constant factor; hence the fracture
toughness and the tensile strength differ by a constant factor.
It can be shown more formally as follows. According to Van de Ven et al.
(1997), cf. equation (13), the bending tensile strength is given by
/DB
BTS = 4.263 Pmax
s / r = 0.8
(70a)
where BTS is the bending tensile strength. Rewriting,
BTS = k 0
where k depends on the specimen geometry. Recall equation (16),
K I = 0 a YI
P
0 = 0
DB
Combining equations (70b) and (71),
a c YI
K Ic
= 0
BTS
k 0

(70b)
(71a)
(71b)

(72)

255

Results and analysis of tensile strength

apparent fracture toughness (MPamm)

Sec. 7

45
40
35
30

DAC-unmod
DAC-E
DAC-P
PA-E
PA-P
SMA-E
SMA-P

25
20
15
10
5
0
0

indirect tensile strength (MPa)

Figure 53. Correlation between the apparent fracture toughness and the indirect
tensile strength for the data between -10C and +5C of figure 38. Displacement
rate: 0.5 mm/s, except for K IQ of DAC-unmod: 0.05 mm/s. DAC: dense graded
asphalt concrete, unmod: unmodified (i.e. not polymer modified), PA: porous
asphalt, SMA: stone mastic asphalt, E: containing elastomer modified bitumen,
P: containing plastomer modified bitumen.

Equation (72) shows that the fracture toughness and the (bending) tensile
strength differ by a constant factor, provided there is no strain hardening.
Unfortunately, in the investigation presented in figures 36-38, the bending
tensile strength of the un-notched SCB-specimen was not determined.
However, the indirect tensile strength was determined. In figure 53, the
average apparent fracture toughness is shown as function of the average
indirect tensile strength; only the data from figure 38 for -10C, 0C, and
5C were used. Figure 53 clearly shows a correlation between the average
apparent fracture toughness and the average indirect tensile strength.

256

Resistance to crack-growth and fracture

Ch. 4

8 Results and interpretation of finite element computations on the


SCB-specimen geometry
A finite element model of the SCB-specimen was developed to investigate
the stress state of the specimen. This was necessary because the stress state
of the SCB-specimen is three-dimensional, and the stress strain behaviour
of the material is nonlinear. Under those conditions, the property
determined from the test is expected to depend on the stress state, which,
in turn, depends on the specimen geometry. Therefore it is probable that
the property determined from the test is not a true material property.
The material model, ACRe, was developed by Scarpas et al. (1997), and is
implemented in the finite element system INSAP. The model consists of a
combination of the theory of dynamic viscoplasticity and the classical
technique of modelling smeared crack-growth. This permits a realistic
description of the stress strain behaviour in compression and in tension
(e.g. de Borst et al. 1998). For more details about the material model, the
reader is referred to appendix 5.
Loading

Y
X
Z

Figure 54. Schematic representation of the finite


element mesh.

8.1 Specimen geometry


The specimen is half diametrical with a diameter of 100 mm and a
thickness of 25 mm. For symmetry reasons, the damage in only a quarter
of the specimen volume needs to be computed. The finite element mesh
was made up of 1326 20-nodded elements, and is shown in figure 54. The
width of the loading strip at the top of the specimen was put equal to 5
mm. The width of the supporting strip, was put equal to 20 mm. The

Sec. 8

Results and interpretation of finite element computations


on the SCB-specimen geometry

257

specimen was subjected to a constant deformation speed of the loading


strip, vertically downward, of 0.5 mm/s. In the analysis, 1000 time steps
correspond to a displacement of 0.5 mm. The finite element code for the
semi-circular bending test specimen was created and run by Liu (2001).
8.2 Definition of damage
In the literature, different definitions of damage can be found. In this
investigation, the damage is defined as the magnitude of the total
permanent (irreversible) strain of the material. Mathematically, this
permanent (plastic or viscoplastic) strain can be written as:

= d p = d ijp d ijp

(72)

The damage, , can be considered as the sum of a volumetric and a


deviatoric component. Volumetric damage by compressive stress is
associated to inelastic compression of the material. Volumetric damage by
tensile stress is associated to cracking of the material. Deviatoric damage
is associated to shear of the material and crack formation as a result of
that.
Table 24. Values of material parameters.
E (MPa)
1500

0.2
f t
temperature and strain dependent

fc
>> f t
T (C)
15
0
0.17
lim
0.03
0
250
1 , 2
10.8
3
0.02
n
2.5

8.3 Material
The modelled material represents a dense graded asphalt concrete mixture
with 6% 80/100 bitumen. The material model requires the following
properties, cf. table 24: the stiffness modulus, E , Poissons constant, ,
the tensile strength as a function of the temperature and the deformation
rate, f t , the compressive strength as a function of the temperature and the
deformation rate, f c , the temperature, T , three parameters for the
simulation of the strain hardening process, 0 , lim , and 0 , three
parameters for the simulation of the response degradation process, 1 ,
2 and 3 , and a parameter which determines the shape of the Desai flow
surface, n .

258

Resistance to crack-growth and fracture

a).

b).

Ch. 4

Figure 55. Damage development in time step 1500. Left (a): tensile damage.
Right (b): compressive damage.

The values of the model parameters are based on results of monotonic


uniaxial tensile tests and monotonic uniaxial compression tests (Scarpas et
al. 1997). They can be considered representative for asphalt concrete and
enable a realistic simulation of the behaviour in the SCB-test.

8.4 Results of computations


Figure 55 shows the tensile damage (figure 55a) and the compressive
damage (figure 55b) at an early moment in the deformation process, i.e. in
time step 1500. Figure 55a shows that the damage initiates at the specimen
base where the bending moment is at maximum. Figure 55b shows that
the compressive damage initiates at the edge of the loading strip, at the

a).

b).

Figure 56. Damage progression in time step 2100. Left (a): tensile damage.
Right (b): compressive damage.

Sec. 8

Results and interpretation of finite element computations


on the SCB-specimen geometry

259

Figure 57. Deviatoric damage. Left (a): in time step 2100. Right (b): in time step
2400.

right-hand edge of the support strip, and in a volume section ahead of the
tensile zone. Notice the contour scales: In this phase, the maximum tensile
damage is approximately one order of magnitude greater than the
maximum compressive damage.
Figure 56 shows the damage in time step 2100. Figure 56a shows that the
tensile damage is concentrating along the vertical. Note the sharp damage
gradient. Figure 56b shows that the compressive damage has spread from
under the loading strip over a large portion of the specimen, and is also
extending from the edge of the support strip. However, the maximum
compressive damage increases slower than the maximum tensile damage,
and is approximately 20 times smaller than the maximum tensile damage.
Figure 57 shows the deviatoric damage in time step 2100 and time step 2400.

Figure 58. Damage in step 10500. Left (a): tensile damage. Right (b):
compressive damage.

260

Resistance to crack-growth and fracture

Ch. 4

The deviatoric damage is associated to shear. The greatest concentration is


located as expected near the edge of the loading strip (figure 57a). Notice
the contour scales. The maximum deviatoric damage in figure 57a is 20 x
smaller than the maximum tensile damage in figure 56a. In figure 57b the
deviatoric damage near the loading strip is smaller than in figure 57a. From
this it follows that the specimen has already collapsed: There is a crack at the
middle of the specimen base. The specimen halves will now depart from each
other. During that process hardly new deviatoric damage in the specimen is
formed. In the mesh the shear concentrates along the vertical.
Figure 58 shows the damage in time step 10500. Figure 58a shows that the
tensile damage is strongly concentrated along the vertical. Figure 58b
shows that the compressive damage is negligible almost everywhere in the
specimen except at the middle of the specimen base where the crack
started. Figure 58 shows that the compressive damage and the deviatoric
damage have increased along the vertical but not in the remainder of the
specimen (i.e. near the edges of the loading strip and the support strips),
and thus have in fact decreased relatively, i.e. compared to the maximum
compressive damage in the location where the crack started.

(a). Non-linear simulation

(b). Linear simulation

Figure 59. Deformation of the finite element mesh after 10800 time steps. Left
(a): nonlinear case. Right (b): linear case.

Figure 59 shows the difference in deformation of the mesh in the linear


and the nonlinear case (exaggerated). In the nonlinear case (figure 59a) the
specimen halves deviate from each other. The damage is localised in the
elements along the lower half of the vertical. At a distance from the
vertical, there is hardly shear at all. This is in agreement with what is
observed experimentally. In the linear case (figuur 59b) this damage
localisation does not take place. The (elastic) deformation is concentrated

Sec. 8

Results and interpretation of finite element computations


on the SCB-specimen geometry

261

12500

linear case

force (N)

10000

7500

nonlinear case
5000

2500

0
0

10

15

displacement (mm)

Figure 60. Comparison of the force displacement curves in the linear and
the nonlinear case.

in the bending of the specimen base. The corresponding computed force


displacement curves are shown in figure 60. In the nonlinear case, the
maximum force (and the tensile strength) is lower than in the linear case.
The reason is that the applied force is limited by the plastic flow of the
specimen, which is absent in the linear case.

8.5 Discussion and conclusions of computational results


The computational results show that the stress state of the SCB-specimen
is three-dimensional. The results show that the tensile damage starts to
develop at the middle of the specimen base, and develops from there along
the vertical upwards. During the test, the tensile damage concentrates
narrowly along the vertical.
Apart from tensile damage, there is compressive damage and deviatoric
damage. The compressive damage starts to develop on the outer edges of
the loading strip and the support strips. During the test, the compressive
damage extends over a large portion of the specimen volume.
The deviatoric damage starts to develop from under the edge of the
loading strip and spreads from there over almost the entire specimen
volume. The results show further that at the point in time where the crack
starts to develop at the specimen base, the maximum tensile damage is
approximately a factor of 20 greater than the maximum compressive
damage, and approximately a factor of 20 greater than the maximum
deviatoric damage. After that moment in time, the compressive damage
and the deviatoric damage increase further along the vertical, but decrease
relatively in the remainder of the specimen volume.

262

Resistance to crack-growth and fracture

Ch. 4

Thus, it is shown that the stress state of the specimen is three-dimensional.


However, the tensile damage greatly overshadows the compressive and
deviatoric damage. In fact, the compressive damage and the deviatoric
damage become significant only in the location where the crack started
and along the vertical. Only in those locations, the compressive damage
and deviatoric damage reach maximum values of approximately 5% of the
tensile damage. In the remainder of the specimen volume, the compressive
damage and the deviatoric damage remain negligible. Therefore, a
conclusion is, that the SCB-test is suitable for the determination of a
tensile property.
The results might be further interpreted as follows. The results show the
development of a large compressive zone in the specimen up until the
moment in time where the crack starts to develop. Thus, the tensile stress
which causes the crack, develops in the presence of developing
compressive stress, which is spread over a large portion of specimen
volume. The compressive stress causes the stiffness of the specimen, and
hence the applied force to generate the required tensile stress to initiate the
crack to be greater, than when this stress were absent. Consequently, the
calculated tensile strength is greater than when the compressive stress
were absent.

9 General discussion
It was found that the dynamic crack-growth behaviour of asphalt mixture
can be characterised by means of Paris Law. The dynamic crack-growth
properties are the constants A and n of the Paris equation. Using the
CCT-specimen and the 4PB-specimen, the constants A and n were
obtained independent of the specimen thickness. However, the
constants A and n were not found to be reproducible using the two
specimen geometries.
Also the critical stress intensity factor, K c , was determined. It was
calculated from the stress and the crack-length at the moment where the
residual fracture occurred. The K c obtained from the 4PB-specimen was
approximately 50% greater than that obtained from the CCT-specimen.
The greater K c in the 4PB-specimen was attributed to creep. It was found
that creep in the 4PB crack-growth test is unavoidable. This will likely
cause an error on the calculated K c .
The experimental results give no indication of an influence of the
specimen thickness on K c . It was assumed that K c is independent of the
specimen thickness.
The stress intensity factor describes the stress field at the crack-tip. The
critical stress intensity factor is the stress at which fracture occurs. It is
known from the literature for metals, that the K c depends on the specimen
thickness, and that for sufficient thickness, the critical stress intensity factor

Sec. 9

General discussion

263

is independent of the specimen thickness. It is then called the plane strain


critical stress intensity factor, also called fracture toughness. It is common
to consider the fracture toughness a material property, which characterises
the resistance to fracture of the material.
Provided the residual fracture is linearly elastic, and the specimen fulfils
the ASTM size requirements for the stress state of the specimen to be
plane strain, the critical stress intensity factor from a dynamic crackgrowth test can be expected to correspond to the fracture toughness
obtained from a fracture toughness test, K Ic .
It is important that, in principle, the resistance to crack-growth,
characterised by A and n , and the resistance to fracture, characterised by
K c , are related parameters, and that the relatedness can be shown in a
dynamic crack-growth test.
An advantage of the fracture toughness, K Ic , as a parameter to
characterise the resistance to fracture, is that it can be obtained from a
relatively simple fracture toughness test. An example is the SCB-test,
which, compared to a dynamic crack-growth test, is more suitable for
routine purposes, such as asphalt mixture design. The fracture toughness
from the SCB-test was found to be valid for temperatures of 0C and
lower. At -10C and 0C, it was found practically independent of the
specimen size. The numerical value of a dense graded asphalt 0/16
mixture was found to correspond approximately to that of the sand asphalt
mixture tested using the CCT-specimen.
The fracture toughness is an important engineering property. For example
in aerospace applications, it is used to avoid plane strain or low energy
fracture (Ewalds and Wanhill 1984).
A disadvantage of the K Ic is that the conditions of linear elasticity and a
plane strain state of stress are severe. One might consider to use the tensile
strength instead of the fracture toughness if these conditions cannot be
fulfilled. In the absence of strain hardening, the fracture toughness and the
tensile strength differ by a constant factor. Thus, in principle, the fracture
toughness and the tensile strength are related properties; the relatedness
can be shown in a fracture toughness test using a notched specimen, and
the same test as a tensile test using the same specimen un-notched. In this
manner, the SCB-test was used in the present study.
Tensile tests using different specimens were performed to investigate
whether the tensile strength could be reproduced independent of the
specimen geometry. It was found that this was not the case. Experimental
evidence shows that the bending tensile strength is consistently greater
than the uniaxial tensile strength in the temperature range between -10C
and +25C. Experimental evidence also suggests that the indirect tensile
strength is approximately equal tending to be greater than the uniaxial
tensile strength in the temperature range between -10C and +10C. At

264

Resistance to crack-growth and fracture

Ch. 4

temperatures above +15C, the calculated indirect tensile strength was


found lower than the uniaxial tensile strength. This is ascribed to the
limitation of the maximum applied stress by localised plastic deformation
near the loading strips.
An explanation for the consistently higher bending tensile strength was
found based on finite element computations on the SCB-geometry. Those
computations have shown that a large compressive zone in the specimen
develops. This may explain that the specimen responds stiffer to the
applied load as compared to the same specimen in which a compressive
stress is absent, hence that a higher applied force is needed for fracture.
Thus can be explained that the bending tensile strength in the SCB-test is
consistently higher than the uniaxial tensile strength in the uniaxial tensile
test at similar temperature and displacement rate. The experimental
evidence confirms this. It is probable that the same argument applies to
the indirect tensile test for temperatures of +10C and lower.
Generalising this result, it can be concluded that the tensile strength will
be found higher if the tensile stress develops in the presence of
compressive stress. More generally, mechanical asphalt mixture properties
are normally not obtained as true material properties, owing to nonlinear
stress strain behaviour.
The finite element computations of the SCB-test have shown further for
given input parameters, that the compressive damage is approximately a
factor of 20 smaller than the tensile damage. Therefore, it seems that the
SCB-test is suitable as a tensile test, i.e. for the determination of the
(bending) tensile strength.
Finally, it can be concluded, that it is not sufficient for a constitutive
model parameter or variable to be a true material property if it can be
reproduced independent of the specimen size; it must be reproducible
independent of specimen size and shape.

10 Conclusions
1 Under strict conditions, the crack-growth behaviour of an asphalt
mixture obeys Paris Law.
2 It is plausible that the stress controlled crack-growth of an asphalt
mixture is to be considered as creep crack-growth, and not as fatigue
crack-growth.
3 The applicability of the concept of stress intensity factor implies that a
more simple and efficient method of characterising the resistance to crackgrowth is possible, based on the fracture toughness.

Sec. 10

Conclusions

265

4 The minimum size requirements normally used with metals, which have
to be fulfilled in order to guarantee a plane strain state of stress, cannot be
satisfied completely in the case of asphalt mixture. Despite that, it is
plausible that the fracture toughness of asphalt mixture at approximately
1C or lower, can, from a practical standpoint, be considered a material
property.
5 The SCB-test is a suitable test for the determination of the fracture
toughness at a temperature of 1C or lower. The fracture toughness has
been obtained practically independent of the specimen size (i.e. within the
standard deviation).
6 Finite element computations have shown that the compressive damage
at 15C is approximately a factor of 20 lower than the tensile damage. It is
probable that the compressive damage is relatively smaller, i.e. negligible,
at 1C and lower. Therefore, the SCB-test is a suitable tensile test. It is
probable that the SCB-test is suitable for the determination of the bending
tensile strength of asphalt mixture at 1C and lower.
7 It is probable that the bending tensile strength is somewhat higher than
the uniaxial tensile strength, owing to the mixed compression/tensionstress state of the SCB-specimen.
8 The tensile strength is fairly reproducible in the uniaxial tensile test, the
bending tensile test and the indirect tensile test.

266

Resistance to crack-growth and fracture

Ch. 4

5
Performance judgement of
asphalt mixtures

1 Introduction
In chapter 2, the current pavement design methodology was analysed, and
it was shown that the requirements for asphalt mixtures are compositionrelated. In this chapter, the feasibility of a performance related pavement
design methodology is investigated. In 2, it is shown that the current
composition-related requirements for asphalt mixture are an impediment
for innovation in the field of paving materials. In 3, it is made plausible
that property-related requirements for asphalt mixture are the key to
enable innovation in the field of paving materials, and that it is possible to
control the pavements performance based on properties of the applied
materials determined in the laboratory. In 4, it is explained that the elastoviscoplastic nature of asphalt mixture causes the mixtures properties to be
difficult to reproduce independent of the specimen geometry. That is a
serious difficulty, because it means that it is difficult to obtain the mixture
properties as true material properties predictive for the behaviour of the
material in the pavement. In 5, the constitutive model is described. It is
shown that not only the material properties as such, but also their
interrelatedness is important in the characterisation of the material
behaviour. In 6, the importance of the interrelatedness for the control of
the pavements performance is discussed. This chapter is concluded in 7.

268

Performance judgement of asphalt mixtures

Ch. 5

2 Composition-relatedness of requirements for asphalt mixture:


An impediment to innovation in road building
The current requirements for asphalt mixture are composition-related. The
current system of requirements was developed based on practical
experience. A major disadvantage of such a development is that it is slow
and inefficient, because the system cannot be adapted easily to the needs
arising from socio-economic changes of society. There is considerable
pressure from society to develop more durable and environmentally
friendly paving materials. This pressure is a logic consequence of the
attention of the public to the nations civil infrastructure, that has
increased strongly owing to difficulties associated to its serviceability,
quality, cost, maintenance, space, and influences on our natural
environment and personal well being. Apart from that, up scaling of
economic processes in general, and civil infrastructure projects in
particular, demand more efficient working methods, new types of contract,
and changed relationships between contract partners. In line with these
more general socio-economic developments, also a greater efficiency is
required in the development of processes, techniques and materials in road
building.
Let us consider a few developments from the recent past, to get a feeling
of the difficulties that arise when the current system of requirements for
asphalt mixtures needs adaptation. Let us consider the following four
examples from the past up into the present.
Example 1. Asphalt recycling. The development of asphalt recycling in the
Netherlands began shortly after 1974. Early practical experience by 1980
made it possible to manufacture gravel asphalt concrete satisfying the
requirements using up to 25% reclaimed asphalt. In 1985, a maximum
percentage reclaimed asphalt up to 50% was made possible, but use of
reclaimed asphalt was allowed only in lower pavement layers. Still later,
1989, up to 50% reclaimed asphalt was permitted in all types of asphalt
concrete in all pavement layers.
If the current requirements for asphalt concrete are compared to those of
1978, it can be seen that, after all, minor changes of the mixture
formulations were needed. However, the point is that the risks involved
with those changes were at first unknown. The necessary changes and the
risks involved had to be determined based entirely on practical experience.
Example 2. Porous asphalt. The first road applications of porous asphalt
in the Netherlands date back to the nineteen seventies. At that time, road
authorities, confronted with over 3000 dead in traffic a year, were forced
to improve traffic safety. They expected to improve traffic safety by
means of porous asphalt pavements. In 1989, the government decided to
start the so-called porous asphalt programme aiming to have the entire
main road network equipped with porous asphalt pavement by the year

Sec. 2

Composition-relatedness of requirements
An impediment to innovation in road building

269

2010. It turned out that the diminished toll, approximately 1300 dead a
year, is to be attributed to other measures that improved traffic safety. The
greatest benefit of the porous asphalt pavement, not expected in the early
stage of its development, is its capacity to reduce the traffic noise.
Porous asphalt is a mixture with a minimum of 20% design voids content.
To achieve this voids content, the mixture is relatively single size grained
and low in sand. The risk of its application was its durability because of
the mixtures open structure. The risk consisted of two components:
1 the pavements functional service-life, i.e. the duration of its noise
reducing capacity;
2 the pavements durability, i.e. its material service-life.
Early experience indicated a relatively short functional service-life owing
to contamination of the pavement (clogging), and indicated a reduced
durability of on average 3 years, or 25%. However, at present, the average
durability of standard porous asphalt is 11 years, based on the current
maintenance experience, i.e. only one year less than the nominal servicelife of a standard dense graded asphalt pavement.
The current requirements for porous asphalt are composition-related. The
Marshall mixture design method is not used, since this is not suitable for
the design of stone skeleton mixtures.
Example 3. Artificial mineral aggregate. It seemed possible to fabricate an
artificial coarse aggregate by the sintering of dredge sludge from rivers
and harbours. It was attempted to use this as a replacement for natural
coarse mineral aggregate in asphalt mixtures. It was found that the
traditional Marshall mixture design method was not suitable. One reason
was the shape of the artificial coarse aggregate particles which differed
from that of natural coarse aggregate particles. Another was the porosity
and chemical activity of the artificial aggregate, which caused a selective
absorption of the mineral oils from the binder. This implied that the
method of determination of the mixtures optimum bitumen content,
which is a part of the Marshall mixture design method, had to be reinvented.
Example 4. Polymer modified asphalt mixtures. The addition of polymer
modified binder to asphalt concrete has a strong influence on the Marshall
properties. It is known from experience (not discussed here) that the
influence depends on the type of polymer and the polymer content of the
binder. This shows once again that Marshall properties are not predictive
for the behaviour of the material in the pavement. New requirements for
the Marshall properties of polymer modified asphalt (concrete) mixtures
can be developed only based on practical experience.
The above examples may illustrate that new requirements for new paving
materials have been developed repeatedly in history, based on practical

270

Performance judgement of asphalt mixtures

Ch. 5

experience with the limitation of using wholly empirical methods1. If the


process of origination and subsequent development of requirements for
asphalt mixture is critically reconsidered, then the following observations
might be made:
1 The composition-relatedness of the requirements, and the empirical
nature of the Marshall mixture design method and the Marshall properties,
are the cause that current requirements for asphalt mixture lose their
applicability when the material is modified.
2 Composition-related requirements for a given paving material cannot be
used to judge the materials cost-effectiveness or risk of failure.
3 Information about a materials cost-effectiveness and risk of failure,
needed to justify its standardisation, is gathered on the basis of practical
experience prior to its standardisation.
4 The process of origination and subsequent development of the current
requirements for asphalt mixture is a process in which practical experience
with cost-effectiveness and risk of failure is continually evaluated to
optimise mixture formulations and composition-related requirements.
5 The process of origination and development of new requirements has to
be re-initiated each time when it is recognised that current requirements
are not applicable to a newly developed paving material.
6 The process of origination and development of requirements based on
practical experience is a slow process it takes time to build up the
practical experience, first in road trials, later in accommodating
experiences in up scaling.
7 The research effort invested in the currently practised method of
developing new requirements for one asphalt mixture yields little or no
added value for improvements of other asphalt mixtures.
Summarising, the current requirements for asphalt mixture are
composition-related, and based on practical experience. History has shown
that new composition-related requirements for new paving materials have
been developed repeatedly based on practical experience with the
limitation of using wholly empirical methods, and that compositionrelated requirements for asphalt mixture lose applicability when the
material is modified. For different reasons mentioned (1, 3, 5, 6, 7), the
currently practised method of developing new composition-related
requirements is inefficient, and a drawback for innovation in the field of

Empirical knowledge relies on practical experience rather than theories.


Composition-related requirements for asphalt mixtures represent empirical
knowledge. The term empirical property is sometimes used to indicate a
material property that does not fit in any theory of physics. Examples are the
Marshall properties, cf. chapter 2, 2.3.

Sec. 3

Property-related requirements for asphalt mixture


The key to innovation in the field of paving materials

271

paving materials. Thus, the current requirements themselves are an


impediment to innovation in the field of paving materials.

3 Property-related requirements for asphalt mixture The key


to enable innovation in the field of paving materials
The natural alternative to a composition-related requirement is a propertyrelated requirement. The present study is an investigation into the
feasibility of property-related requirements for asphalt mixtures.
The main question is: Is it possible to overcome the disadvantage of the
current system of composition-related requirements for asphalt mixtures,
i.e. its empirical character and methods that represent a drawback for
innovation in the field of paving materials? Could a system of propertyrelated requirements be the answer to the problem, which is that the
current requirements lose applicability when the material is modified. Or:
Is it possible to create property-related requirements for asphalt mixtures
that are not equally empirical as the current composition-related
requirements?
3.1 Definition of performance relatedness
Based on historical experience, it can be asked whether a break-through
from the present situation, not beneficial to innovation, to a new situation,
beneficial to innovation, might be possible if the following conditions are
fulfilled:
1 requirements remain applicable when the material is modified;
2 material properties referred to in the requirements have a predictive
value for the behaviour of the material in the pavement, making costeffectiveness and risk of failure predictions possible.
If these conditions are fulfilled, then
. no new requirements have to be developed each time a newly developed
paving material is brought to the market; and
. the required information to justify the cost-effectiveness and the risk of
failure of the application can be made available by means of tests in the
laboratory; the required information is available much faster than when
gathered based on practical experience.
Requirements which fulfil the above two conditions could be considered
performance related requirements. Therefore the authors definition of the
performance relatedness of a requirement for asphalt mixture is:

272

Performance judgement of asphalt mixtures

Ch. 5

The authors definition of performance relatedness of a requirement


A performance related requirement for an asphalt mixture is a
requirement, which
i remains applicable when the material is modified;
ii provides the possibility to judge the cost-effectiveness and the risk of
failure of the material in an application;
iii acts on a material property to show that the requirement is fulfilled.
Could the fulfilment of the two conditions mentioned at the beginning of
this section be the key to enable innovation in the field of paving
materials? If a property-related requirement is based on an arbitrarily
defined property, then it is likely to be equally empirical as a compositionrelated requirement. It would lose its applicability just as easily upon
modification of the material, and it would have no predictive value for the
behaviour of the material in the pavement.
What is an arbitrarily defined property? The authors definition is: A
property that does not fit in any theory of physics. Examples are the
Marshall properties. If arbitrary properties are excluded, then which
properties can be used?
3.2 Physical meaningfulness and predictive value of constitutive model
variables and parameters
Why is it important to consider and define physical meaningfulness?
Basically, to avoid that asphalt mixture properties are selected which are
arbitrary, and therefore just as empirical as, for example, Marshall
properties. That is, to avoid that so-called performance related
requirements are developed which lose applicability upon modification of
the material.
Let us assume the following definition: A mechanical property is
physically meaningful, if it fulfils the following two conditions:
1 the property represents a constitutive model variable or parameter;
2 the property can be obtained as a true material property.
A physical model, i.e. a constitutive model, is needed as a context in
which material properties are defined. The physical meaningfulness of the
material properties exists by virtue of the model.
Let us assume that, in order for a mechanical asphalt mixture property to
be predictive for the behaviour of the material in the pavement, it suffices
to fulfil the above two conditions. Thus, according to this assumption, an
asphalt mixture property is in principle predictive for the behaviour of the
material in the pavement, if it is physically meaningful; that is, apart from
the loading conditions.

Sec. 3

Property-related requirements for asphalt mixture


The key to innovation in the field of paving materials

273

In this study, it is assumed, that a mechanical material property is


predictive for the mechanical behaviour in the pavement, if it is predictive
for a geometry that differs from the geometry in which it was determined.
A mechanical material property that satisfies this condition is called a true
material property. Conversely, a true material property is considered to be
predictive for the mechanical behaviour in a geometry that differs from
that in which it was determined. Alternatively, a material property can be
considered as such, if it is reproducible in different specimen geometries.
This alone is not sufficient for it to be physically meaningful. In order to
be physically meaningful, a true material property must also be a
constitutive model variable or parameter.
It is important that a physical meaning can be attributed to the constitutive
models and their variables or parameters, which are used to characterise
the mechanical behaviour of asphalt mixtures. This is important because
the characterisation of the mechanical behaviour based on tests in the
laboratory is made to be predictive for the mechanical behaviour of the
material in the pavement, i.e. is made to be able to judge the costeffectiveness and risk of failure based on tests in the laboratory rather than
on practical experience.
3.3 Controlling pavement performance
The driving force of innovation is the same as that of improving
productivity in the economy, namely: competition to achieve improved
efficiency. Therefore, it can be assumed there is a natural driving force to
innovation in the economy. As a part of that, the road owner will try to
realise improved cost-effectiveness by reducing damage to the road. As he
tries that, and on his way finds new possibly suitable candidate materials,
he will also want to know the risk of failure of the application. In the case
the risk of failure increases, he will find it irresponsible to run the
application. Thus, it seems that two criteria determine the performance of
a road: cost-effectiveness and risk of failure. This assumption is
elaborated in the following.
3.3.1 Cost-effectiveness and risk of failure
In the following, considerations will be limited to the pavement and the
applied paving materials. Let us assume that the pavements costeffectiveness can be defined as
total cost of build and maintenance
(1)
cost-effectiveness =
reference total cost of build and maintenance
The cost-effectiveness of the reference pavement is equal to 1, or 100%.

274

Performance judgement of asphalt mixtures

Ch. 5

In a simplified formulation, the risk of failure could be defined as2

risk of failure =

cause

effect-damagei

(2)

where causei represents any of the causes of failure listed in Ch. 1, table
1, and effect-damagei is the effect-damage of a particular cause of failure.
Note, that the risks in table 1 of Ch. 1 include risks of structural failure
and risks of failure of surface functionality.
Ideally, the pavements cost-effectiveness and risk of failure are calculable
based on behaviour predicted for a given pavement design. However,
many influence factors are unpredictable. Factors the influence of which
on pavement performance is difficult to predict or model are
. material selection;
. production and paving conditions;
. service conditions.
Material selection. Normally, the asphalt mixture constituents are in
accordance with the asphalt mixture design. Practical experience has
shown that not always the constituents are used, which are specified in the
mixture design. When old asphalt granulate is recycled, it is not possible
to know the properties of the constituents of the granulate to the extent of
the properties of the virgin materials.
Production conditions are the conditions under which the asphalt mixture
is produced. Apart from the production plant, the most important
conditions are temperature, mixing time, and storage time.
Paving conditions are paving equipment, weather conditions, transport
distance (temperature), and the asphalt mixtures workability
(compactability). The pavements surface characteristics are in part
realised during paving (evenness and slant), are in part realised by the
control of the asphalt mixtures bulk properties (noise emission, hydraulic
conductivity, resistance to permanent deformation, fatigue-cracking, and
ravelling), and depend in part on properties of the asphalt mixtures
constituents (skid resistance and light reflectivity). The paving conditions
cause differences between the realised properties of the mixture in the
pavement and the corresponding properties of the designed mixture.
Service conditions are, e.g. the condition of the soil, weather, traffic load
(axle load spectrum), attack by chemical agents (motor oil-spill, de-icers,
etc.), calamities, etc.
In addition, the pavements performance depends on the pavements
geometry and structure. The realised pavement performance differs from
that expected on the basis of design properties of the applied asphalt
2

Equation (2) is a simplified model, because the formulation implicitly assumes


that the individual terms i are independent. In reality, the effect-damage of a
particular cause i may depend on interactions with other causes.

Sec. 3

Property-related requirements for asphalt mixture


The key to innovation in the field of paving materials

275

mixtures, even that expected on the basis of realised initial properties of


the asphalt mixtures. The total cost of maintenance, hence the costeffectiveness, is difficult to evaluate based on equation (1). Let us
therefore redefine the cost-effectiveness,
service - life
cost-effectiveness =
(3)
costs
An asphalt mixtures cost-effectiveness and risk of failure realised in the
pavement depend on the mixtures properties and on the pavements
geometry, structure, and loading conditions. The mixtures realised costeffectiveness and risk of failure in the pavement are also influenced by
interactions between failure (damage) mechanisms. Altogether, this causes
the maintenance-free service-life and the total service-life, i.e. the time to
material replacement to be stochastical variables that can be represented
by statistical distributions. In the following, the maintenance-free servicelife is denoted as {t1}, and the time to material replacement as {t2}.
Note, that the influences mentioned, material selection, production
conditions, service conditions, pavement geometry, structure, loading
conditions, as well as the influence of interacting failure (damage)
mechanisms, disappear if cost-effectiveness and risk of failure are
determined in terms of the stochastical variables {t1} and {t2}. This is in
fact how cost-effectiveness and risk of failure of current standardised
asphalt mixtures are known. The same method could be adopted for a
newly developed asphalt mixture. However, a difference between a
standardised and a newly developed mixture is then that {t1} and {t2} are
known for the standardised mixture, but not for the newly developed
mixture.
Let us assume that any cause of failure is associated to a maintenance
criterion, which calls for maintenance when the criterion is met. The end
of the pavements (maintenance-free) service-life could be defined in the
form of a Boolean expression of the applicable maintenance criteria,
ESL = MC1 or MC2 or MCi
(4)
where ESL denotes a Boolean parameter, which can be true or false, and
MCi denotes the i-th maintenance criterion, which can be formulated as a
Boolean expression. An example of a maintenance criterion is that the rutdepth should not exceed 18 mm. ESL is true (1) if any of the maintenance
criteria is true, else is false (0). The service-life (time to failure), could be
defined as
SL = tESL t0
(5)
where tESL denotes the time at which ESL changes from false (0) to true
(1), and t0 denotes the time of the start of the utilisation of the pavement.
Let us assume the service-life of the pavements running surface for a
given asphalt mixture type is a stochastical variable, which can be

Performance judgement of asphalt mixtures

asphalt mixture
type 2

relative count (%)

asphalt mixture
type 1

SL

time
maintenance-free service-life of
running surface of pavement
applying maintenance criteria, MCi

cumulative count (%)

276

Ch. 5

asphalt mixture
type 1

asphalt mixture
type 2
time
maintenance-free service-life of
running surface of pavement
applying maintenance criteria, MCi

Figure 1. Left (a): Statistical distribution of the maintenance-free service-lives of


the running surface of the pavement, for two types of asphalt mixture, type 1, and
type 2. For convenience a normal distribution is assumed. Right (b): Cumulative
distributions of distributions shown in figure (a).

represented by a statistical distribution, e.g. the normal (Gaussian)


distribution, cf. figure 1a, then the probability that the running surfaces
service-life for that given asphalt mixture type will be greater than a given
SL is equal to the shaded area under the curve. The corresponding
cumulative distribution is shown in figure 1b.
In general, a maintenance criterion, MCi, has a margin. That means that
when the criterion is met, the damage is still limited to below the critical
damage where the material would have to be replaced. The reason is to
avoid loss of capital; more severe and costly measures of rehabilitation
would be required as more damage is allowed to meet a given
maintenance criterion. An economic optimum arises from practical
experience.
Example. Ravelling of porous asphalt. Let us assume the following two
maintenance criteria, denoted as MCrav1 and MCrav2:
MCrav1 = false:
maintenance-free service-life;
MCrav1 = true:
end of maintenance-free service-life;
MCrav2 = false:
prolonged service-life after maintenance;
MCrav2 = true:
end of service-life; time to material replacement.
Up to the damage defined by the maintenance criterion, MCrav1, the
running surface is treatable by means of different maintenance techniques,
so as to prolong its service-life. Possibly, a second similar treatment may
be repeated later, applying the same maintenance criterion, MCrav1.
However, there will be a time when another similar treatment is no longer
cost-effective, meaning that the running surface will have to be replaced.
In that situation, the maintenance criterion MCrav2 is applicable, which

relative count (%)

Sec. 3

Property-related requirements for asphalt mixture


The key to innovation in the field of paving materials

277

porous asphalt
mixture type X
{t2} for
MCrav2

{t1} for
MCrav1

{t1}
maintenance-free service-life
(time to failure) of
porous asphalt pavement
maintenance criterion, MCrav1

time

{t2}
total service-life, or
time to material replacement of
porous asphalt pavement
maintenance criterion, MCrav2

Figure 2. Two maintenance criteria for ravelling of a porous asphalt pavement,


denoted as MCrav1 and MCrav2 , and the associated statistical service-life
distributions {t1} and {t2}. The maintenance-free service-life ends when MCrav1
changes (Boolean parameter); i.e. the time to failure applying MCrav1 (left). The
total service-life, i.e. the time to material replacement, is determined similarly by
MCrav2 (right). The population averages of {t1} and {t2} are denoted as,
respectively, {t1} and {t2}.

differs from MCrav1 in that the allowable surface deterioration can be


greater. The pavements maintenance-free service-life can be represented
by a statistical distribution; let us for convenience assume a Gaussian
distribution, cf. figure 2. Note that, the pavements maintenance-free
service-life is equal to the time of failure of the running surface (a
structural element of the pavement), which, in turn, is equal to the time of
failure of the applied material, and is determined by MCrav1. MCrav2
indicates the running surfaces end of service-life, which is also the
materials end of service-life, i.e. the time to material replacement. In this
example, the running surfaces end of service-life is not the pavements
end of service-life, since the running surface can be removed and
replaced.
If the maintenance-free service-life, {t1}, and the time to material
replacement, {t2}, are stochastical variables, then the parameter t1/t2 is also
a stochastical variable that can be represented by a statistical distribution,
0 < t1/t2 < 1. The cumulative distribution is shown schematically in figure
3. The maintenance-free service-life corresponds to a given maximum
acceptable risk of failure, since maintenance is called for if the maximum
acceptable risk of failure exceeds a critical value. Thus, in the definition of
cost-effectiveness, equation (3), the service-life is associated to a given
maximum acceptable risk of failure. If the cost-effectiveness increases,

cumulative count (%)

278

Performance judgement of asphalt mixtures

Ch. 5

max. acceptable
risk of failure

{t1}/{t2}

t1/t2
Figure 3. Cumulative risk of failure of four types of asphalt mixture. The dashed
horizontal line corresponds to the maximum acceptable risk of failure, {t1}/{t2}
(cf. figure 2). (a) Reference mixture type with known standard deviation of the
time to failure. (b) Reference mixture type a improved by a decrease of the
standard deviation of the time to failure. (c) Modified mixture type with
improved average service-life, compared to a. (d) Modified mixture type c
improved by a decrease of the standard deviation of the time to failure.

then the service-life increases, which means that the maximum acceptable
risk of failure is reached after a longer time, cf. figure 3. Figure 3
illustrates two possible methods to improve the cost-effectiveness:
1 to reduce the standard deviation of applied materials time to failure;
2 to apply material with improved average durability.
In figure 3, the maintenance-free service-life represents a maximum
acceptable risk of failure associated to a given cause of failure. In figure 3,
the maximum acceptable risk of failure corresponds to MCrav1. Let us
denote the population averages of {t1} and {t2} as, respectively, {t1} and
{t2}, cf. figure 2, then the ratio of the parameter {t1}/{t2} could be
considered to represent the maximum acceptable risk of failure. The
maximum acceptable risk of failure changes if the ratio of {t1} and {t2}
changes.
The RHEIs pavement design method uses design relationships for the
pavement service-life as function the thickness and stiffness of the various
asphalt layers, and as a function of the traffic load. The method assumes
the application of standardised asphalt mixtures. The design-relationships
and the maintenance criteria were developed on the basis of practical
experience for application of the standardised road materials. The design
service-life is implicitly related to the regular maintenance criteria and
their associated maximum acceptable risks of failure.

Example 1. Criterion depending not only on the applied material itself.


Let us assume the following two maintenance criteria for rutting, MCrut1

Sec. 3

Property-related requirements for asphalt mixture


The key to innovation in the field of paving materials

279

and MCrut2:
MCrut1 = false: maintenance-free service-life;
MCrut1 = true: end of maintenance-free service-life; rut-depth is 18 mm;
MCrut2 = false: prolonged service-life after maintenance;
MCrut2 = true: end of service-life; time to material replacement.
The 18 mm maximum rut-depth is an example of a maintenance criterion
that depends not only on the applied material itself, but also on the
comfort and safety of the road user. Apart from comfort, this criterion
contributes to the prevention of skidding caused by standing water in ruts.
Let us assume a non-standardised asphalt mixture, e.g. a polymer
modified mixture, which exhibits improved resistance to permanent
deformation (rutting in the pavement). Let us denote the following
parameters:
. MCrut1: 18 mm maximum rut-depth criterion;
. MCrut2: material replacement criterion;
. {t1}: statistical population of maintenance-free service-life;
. {t2}: statistical population of time to material replacement;
. {t1}: population average of {t1};
. {t2}: population average of {t2}.
Let us assume as an example, that the maintenance-free service-life is
improved from {t1} = 6 years for the standardised mixture to {t1} = 7
years for the modified mixture, and that the time to material replacement
is improved from {t2} = 12 years for the standardised mixture to {t2} = 16
years for the modified mixture. The increase in {t2} is relatively greater
than the increase in {t1}. Let this be an intrinsic property of the material,
since the rate of permanent deformation (rutting in the pavement)
decreases as a function of increasing permanent deformation. Note, that
the numbers used here are completely fictitious.
For the standardised mixture, MCrut1 corresponds to the maximum
acceptable risk of failure represented by {t1}/{t2} = 6/12 = 0.5, cf. figure
3. If the same maintenance criterion, MCrut1, is applied to the modified
mixture, then MCrut1 corresponds to the maximum acceptable risk of
failure, {t1}/{t2} = 7/16 = 0.44. If the 18 mm maximum rut-depth were not
required for the safety of the road user, the same maximum acceptable risk
of failure applied to the non-standardised would allow a maintenance-free
service-life of 8 years.
The maximum acceptable risk of failure, {t1}/{t2}, decreases, because the
increase in {t2} is relatively greater than the increase in {t1}. This is
shown schematically by the arrows A in figure 4. Figure 4 illustrates that
application of a regular maintenance criterion to a newly developed
material with improved properties causes the maximum acceptable risk of
failure to reduce.

Performance judgement of asphalt mixtures

280

Ch. 5

cumulative count (%)

100%
maximum time to
material replacement in {t2}

max. maintenance-free
service-life in {t1}
max. acceptable
risk of failure

{t1}/{t2}

MC1

MC2

time
MC1: maintenance-free
service-life

MC2: time to
material replacement

Figure 4. Cumulative risk of failure of a type of asphalt mixture. The maximum


acceptable risk of failure, {t1}/{t2} decreases, indicated by arrows A, if the
increase in time to material replacement, represented by {t2} (i.e. time until MC2
= true), is relatively greater than the increase in maintenance-free service-life,
represented by {t1} (i.e. time until MC1 = true).

Example 2. Criterion depending only on the applied material itself. Let us


assume the following two maintenance criteria for surface cracking, MCsc1
and MCsc2:
MCsc1 = false: maintenance-free service-life;
MCsc1 = true:
end of maintenance-free service-life;
MCsc2 = false: prolonged service-life after maintenance;
MCsc2 = true:
end of service-life; time to material replacement.
MCsc1 depends on the crack-depth, i.e. the probability that penetrating
water damages the pavement structure, and/or the number of cracks per
unit area of running surface. MCsc2 depends on economic reasons to
replace (a part of) the pavement, and otherwise on the crack-depth and the
number of cracks per unit area of running surface i.e. the probability that
the pavements bearing capacity falls below a critical value.
Apart from economic reasons to replace (a part of) the pavement, this
surface cracking maintenance criterion depends wholly on the material, its
integrity, and not on environmental requirements. If, owing to improved
material properties, applying the regular maintenance criteria, MCsc1 and
MCsc2, the increase in {t2} of {t2} is relatively greater than the increase in
{t1} of {t1}, then that causes a reduction of the maximum acceptable risk
of failure.
In example 1, the pavement designer has no alternative but to apply the
regular maintenance criterion, since the safety of the road user depends on

Sec. 3

Property-related requirements for asphalt mixture


The key to innovation in the field of paving materials

281

it. In example 2, a reduced maximum acceptable risk of failure is not


required for road safety or other environmental reasons. Then the
pavement designer might be willing to accept a greater (the normal) risk
of failure; in other words, he might wish to adjust the maintenance criteria,
MCsc1 and MCsc2, in case the modified material is applied.

3.3.2 Pavement performance


If only two criteria determine the performance of a road, costeffectiveness and risk of failure, then pavement performance might be
considered as a summation of cost-effectiveness and risk of failure.
For standardised asphalt mixture types the maximum acceptable risk of
failure is not expected to vary, since neither {t1} and {t2} nor the
maintenance criteria change. The cost-effectiveness varies with the
service-life, as explained by equation (5). This is shown schematically in
figure 5.
In contrast, for non-standardised asphalt mixtures the maximum
acceptable risk of failure is likely to be reduced if {t1} and {t2} increase
applying the regular maintenance criteria. In general, it is undesirable that
the maximum acceptable risk of failure decreases, since that causes an
unnecessary reduction of the maintenance-free service-life and the time to
material replacement. If a reduced maximum acceptable risk of failure is
not required for road safety reasons or other environmental reasons, the
pavement designer might be willing to accept a greater maximum
acceptable risk of failure. For a pavement design, the total maximum
acceptable risk of failure is equal to the statistical summation of the all the
risks associated to all the maintenance criteria. In general, i.e. when {t1}
and {t2} change, or when maintenance criteria are adjusted, the maximum
acceptable risk of failure is expected to vary. This is shown schematically
in figure 6.
Thus has been made plausible that two criteria determine the performance
of a pavement: the pavements cost-effectiveness and the pavements
maximum acceptable risk of failure.
The maximum acceptable risk of failure might conveniently be
abbreviated as the risk of failure.
3.3.3 Performance related asphalt mixture properties
Let us assume as in 3.3.1, that to each performance related property
determined in the laboratory (in type testing) is associated a maintenance
criterion, MCi, for a given damage phenomenon, and that to each
maintenance criterion is associated a maintenance-free service-life {t1},
and a time to material replacement, {t2}, cf. figure 2, then, based on the
correlations3 between the performance related property considered and
3

Those correlations are as yet not available.

Performance judgement of asphalt mixtures

282

Ch. 5

varying cost-effectiveness as a
result of varying service-life

max. acceptable
risk of failure
(more or less constant for
standardised asphalt mixtures)

cost-effectiveness

Figure 5. Schematic representation of an asphalt mixtures performance as a


summation, similar to a vector summation, of the mixtures cost-effectiveness
and the mixtures maximum acceptable risk of failure. For standardised asphalt
mixtures, applying regular maintenance criteria, the maximum acceptable risk of
failure is more or less constant. The cost-effectiveness of the standardised
asphalt mixtures varies as a result of varying service-life.

greater max.
acceptable risk
of failure

risk of failure of
pavement design

performance of
pavement design

reference pavement
design

reference risk
of failure
smaller max.
acceptable risk
of failure

cost-effectiveness of
pavement design

reference costeffectiveness

1
cost-effectiveness

Figure 6. Schematic representation of the pavement performance as a


summation, similar to a vector summation, of the pavements cost-effectiveness
and the pavements maximum acceptable risk of failure. The cost-effectiveness of
the pavement design is compared to the reference cost-effectiveness of a
reference or standard pavement design, and the maximum acceptable risk of
failure of the pavement design is compared to the reference or standard risk of
failure.

Sec. 3

Property-related requirements for asphalt mixture


The key to innovation in the field of paving materials

283

{t1} and {t2}, an improved performance related property indicates an


improved cost-effectiveness and a reduced risk of failure, cf. figure 4.
If performance related asphalt mixture properties are indeed performance
related, then they should show correlations with {t1} and {t2}. However,
evidence supporting this is as yet scarce4. Let us assume that a given
asphalt mixture property and the associated {t1} and {t2} are positively
correlated; i.e. an improved property correlates to an improved servicelife. Then it is improbable that upon application of a newly developed
asphalt mixture the pavements cost-effectiveness is reduced and the risk
of failure is increased if the characterising material properties, determined
in the laboratory, indicate improved behaviour (other factors than material
behaviour influencing pavement performance excluded from
consideration), since those characterising asphalt mixture properties are
predictive for the behaviour of that mixture in the pavement.
Thus is made plausible that a rational approach is possible which permits
to obtain information required to judge a mixtures cost-effectiveness and
risk of failure on the basis of laboratory testing and if necessary some
accelerated pavement testing.
One might consider the possibility that the pavements cost-effectiveness
and risk of failure can be controlled based on material properties
determined in the laboratory instead of information gathered from
practical experience. That creates the possibility to speed up the
acceptance and standardisation of newly developed paving materials, to
enhance innovation in the field of pavement design and paving materials.

3.4 Summary
To develop property-related requirements for asphalt mixtures, the
properties referred to in the requirements cannot be arbitrary properties.
Those properties must be physically meaningful properties; otherwise
property-related requirements will be equally empirical as current
composition-related requirements for asphalt mixtures. The predictive
value of a material property of a given paving material might be
considered in terms of the possibility to judge the pavements costeffectiveness and risk of failure upon application of the material.
A physical model, i.e. constitutive model, serves two main purposes:
1 it is a context in which material properties and physical meaningfulness
of material properties is defined, which is necessary in order for a material
property to be predictive;
4

In a vast research effort around the world correlations between pavement


service-life and performance related properties of applied materials are searched
for. The benefit of that effort is limited for the reason that a great deal of it is
empirical research, concerned with empirical properties (which lose their
applicability if the material is modified); i.e. research with little added value.

284

Performance judgement of asphalt mixtures

Ch. 5

2 it is a model, i.e. a set of mathematical equations, which, in principle,


permits to predict the behaviour of a material in a geometry which differs
from that in which the properties of the material were determined.

4 Asphalt mixture: the complicating factor in the development of


performance related requirements
4.1. Dependence of asphalt mixture properties on specimen size and shape
From the studies in chapter 3 and chapter 4, it seems that some asphalt
mixture properties can be obtained almost independent of the specimen
size. It seems that this is true for the following properties:
. the complex modulus;
. the viscoelastic creep susceptibility, i.e. the slope of the log-log complex
modulus mastercurve;
. the coefficient A and the exponent n of the Paris equation;
. the fracture toughness.
However, it must be noted, that only under very strict conditions, those
asphalt mixture properties can be considered approximately as true
material properties. Those conditions are:
. the stress strain behaviour is (approximately) linearly (visco)elastic;
. the creep, although not entirely absent, is practically negligible.
Those conditions can be realised:
. in strain controlled dynamic bending;
. in stress controlled dynamic crack-growth, if the cumulative creep is
negligible.
From the study in chapter 4, it seems that not all asphalt mixture
properties can be obtained independent of the specimen geometry, i.e. size
and shape. The tensile strength was determined using different specimen
geometries, as the uniaxial tensile strength, the bending tensile strength,
and the indirect tensile strength, cf. Ch. 4, figure 14. Considering that
different values are obtained using different specimen geometries, it seems
that the tensile strength cannot be obtained independent of the specimen
geometry. The observation that the bending tensile strength is found
consistently greater than the uniaxial tensile strength was interpreted as
follows: The tensile strength is greater when the tensile stress develops in
the presence of a compressive stress than when the tensile stress develops
in the absence of a compressive stress (other conditions being equal). It
means, that, if the specimen geometry and the load application cause a
compressive stress in the specimen, then the observed tensile strength will
be greater. It means also that the tensile strength is found to depend on the
specimen shape, if the materials stress strain behaviour is nonlinear.
Thus, it seems, that asphalt mixture properties are in general not obtained
as true material properties, because of the combination of three properties
(causes):

Sec. 4

Asphalt mixture: the complicationg factor in the


development of performance related requirements

285

i the asphalt mixtures material heterogeneity;


ii the intrinsic nonlinear response to loading of both the mineral aggregate
skeleton and the binder of the asphalt mixture;
iii the stress state of the specimen, which, in general, is three-dimensional,
i.e. non-uniform.
From a theoretical viewpoint one might consider to redefine a true
material property as independent of the specimen size, and relate it to the
pure uniaxial stress state, compressive or tensile, in order to avoid a threedimensional stress state. However, this is hard to realise in a practical
context, e.g. that of asphalt mixture design and quality control. This is
discussed further in the following.
4.2 The stress state of the specimen
In the following, two examples are considered.
Example 1. Originally, the creep test was used in the absence of
confinement pressure (Bolk 1980). It was found that the resistance to
creep of porous asphalt (permanent deformation in the pavement) is less
than that of dense graded asphalt concrete (RHEI 1976). However, from
practical experience it was known that the resistance to permanent
deformation of porous asphalt in the pavement is much better than that of
dense graded asphalt concrete. The different creep behaviours of these
mixture types can be explained by the influence of confinement. Upon
loading of an asphalt mixture in the pavement, a natural confinement is
generated, which can be attributed to the internal friction (a reaction force)
under compressive stress in the material. Porous asphalt shows a greater
resistance to permanent deformation than dense graded asphalt, in the
presence of this confinement, which can be attributed to the rigid stone
skeleton. This experience was the main reason to abandon the creep test in
the absence of confinement, in favour of the triaxial creep test, i.e. the
creep test in the presence of confinement. The application of a
confinement pressure causes a three-dimensional stress state in the
specimen. Examples of analytically calculated stress states, reproduced
from Balla (1961), are shown in figure 7. Note, that the stress states in
figure 7 are two-dimensional, because Balla assumed the material was
homogeneous, and the specimen geometry is cylindrical. In a
heterogeneous material in a cylindrical geometry, the stress state is threedimensional.
Example 2. Originally, the indirect tensile test was used to determine the
tensile strength of asphalt mixture. At a first glance, it may seem peculiar
to use a compression test to determine the tensile strength. However, the
tensile stress is distributed practically uniform over most of the vertical,
cf. figure 8. Therefore, in principle, the indirect tensile test is suitable to

286

Performance judgement of asphalt mixtures

Ch. 5

Figure 7. Examples of non-uniform stress-states of the cylindrical triaxial test


specimen, calculated for maximum friction between the specimen and the
loading plate. 1 represents the applied axial stress, 3 the applied radial
stress. Contour lines represent the so-called intensity of the tangential stress, 0 ,
defined by 0 = 16 [( x r ) 2 + ( r ) 2 + ( x ) 2 ] + 2 , where x is the
axial stress, r is the radial stress, and is the tangential stress. Top left (a):
1 = 1, 3 ; uniaxial compression (no confinement). Top middle (b): 1 = 0,
3 = 1; only a uniformly distributed radial stress acts upon the specimen. Top
right (c): 1 = 1, 3 = 1; hydrostatic stress; in this case the intensity of the
tangential stress is rather small; the stress state is practically uniform. Bottom
(d), (e), (f): The intensity of the tangential stress increases as departing from (c),
the applied axial stress is increased, respectively to 1.5 (d), 4.0 (e), and 6.0 (f).
[From Balla 1961].

determine the tensile strength of the material, provided this is (a lot)


smaller than the compressive strength, so that the specimen fractures
owing to the tensile stress. There are different disadvantages to the
indirect tensile test:

Sec. 4

Asphalt mixture: the complicationg factor in the


development of performance related requirements

287

Figure 8. Schematic representation of the indirect tensile test. Two conditions


must be fulfilled to determine the tensile strength: (i) the tensile stress
distribution is uniform (shown); (ii) the maximum tensile stress is sufficiently
lower than the maximum compressive stress, so that the fracture is caused by the
tensile stress (not shown).

. the test is not applicable to all types of asphalt mixture;


. at temperatures above 15C, localised plastic deformation disturbs the
determination of the tensile strength;
. at temperatures above 15C, the difference between the tensile strength
and the compression strength may be insufficient to allow the
determination of a valid tensile strength (Hagemann 1980).
Therefore, two alternative tensile tests were investigated:
1 uniaxial tensile test;
2 semi-circular bending test (SCB).
The primary advantage of the uniaxial tensile test is the stress state of the
specimen; this is probably the best feasible approximation of a uniaxial
state of stress of a specimen of asphalt mixture. Unfortunately, the
uniaxial test is not always possible on samples from the pavement5.
An advantage of the indirect tensile test and the semi-circular bending test
is that these tests can be used on material in pavement layers thinner than
50 mm. The tests were used on specimens as thin as 25 mm, cf. Ch. 4,
figures 3637. An advantage of the semi-circular bending test specimen in
comparison to the indirect tensile test is the relatively low maximum

Normally, the standard-test is performed using a cylindrical specimen, 50


mm in diameter. This cannot always be made available, e.g. if the pavement
layer is thinner than 50 mm. From layers thinner than 50 mm a prismatic
specimen could be prepared. That has not yet been practised.

288

Performance judgement of asphalt mixtures

Ch. 5

applied force. This was found to be approximately nine times lower


(Krans et al. 1996). Therefore, in principle, the chance of plastic strain
localisation near the loading strip and support strips has decreased.
Accordingly, the quality of the bending tensile strength as a physically
meaningful material property was expected to improve in comparison to
the indirect tensile strength. The scatter on the data was expected to be
reduced; therefore the bending tensile test was expected to be more
discriminatory.
The above two examples may illustrate that the uniaxial stress state of the
specimen, although interesting from a research viewpoint, might have
limited possibilities in the practical context of performance related
specifications for asphalt mixtures.
4.3 Relative (qualitative) predictive value
Asphalt mixture properties determined in the laboratory do not have
absolute (quantitative) predictive value for the behaviour of the material in
the pavement. There are two main reasons for this6:
1 Asphalt mixture properties are time dependent, i.e. they depend on the
shape of the waveform of the applied load; cf. appendix 1;
2 The shape of the waveform induced in the pavement by a passing wheelload differs from the waveforms used in tests in the laboratory.
Therefore, at best, any mechanical asphalt mixture property, even a true
material property, is qualitatively (relatively) predictive for the behaviour
of the material in the pavement. There is no need to have all asphalt
mixture properties as a true material property. This is discussed in 6.
4.4 Summary
An asphalt mixtures heterogeneity, the materials intrinsic nonlinear
stress strain behaviour, and the chosen tests, in which a three-dimensional
stress state developes, represent a complication to the development of
property-related requirements. The physical meaningfulness and
predictive value of a constitutive model variable or parameter are in part
lost if the variable or parameter cannot be obtained as a true material
property. Ideally, from a characterisation viewpoint, the three-dimensional
stress state is preferably avoided. One might consider redefining a
material property as related to the uniaxial stress state, and independent of
specimen size. However, that is hard to realise in the practical context of
quality control in road building, because of the limited relevance of testing
paving materials uniaxially.
6

Other reasons are unpredictable conditions related to climate, traffic, and


settlement of the soil during the pavements service-life, and production and
processing conditions during paving. These are not considered.

Sec. 5

Interrelatedness of constitutive model parameters

289

For the reasons that asphalt mixture properties are time dependent, hence
depend on the shape of the waveform, and that the shape of the waveform
induced in the pavement by a passing wheel-load differs from the waveforms used in tests in the laboratory, a mechanical asphalt mixture
property has at best relative (qualitative) predictive value. There is no
need to have all asphalt mixture properties available as a true material
property.

5 Interrelatedness of constitutive model variables and parameters


In the following, the constitutive model of asphalt mixture is reviewed, in
the form in which it was the subject of the study in chapters 3 and 4. The
model characterises the asphalt mixtures elasto-viscoplastic behaviour in
terms of the following properties:
. complex modulus,
. creep compliance,
. dynamic crack-growth properties,
. fracture toughness,
. tensile strength.
The above-mentioned properties are interrelated. For two main reasons the
interrelationships are important:
1 As a compensation for the partial loss of predictive value caused by the
fact that, in general, asphalt mixture properties are not obtained as a true
material property. The compensation is the possibility to create fixed
defined relationships between at least one asphalt mixture property that is
a true material property, and all other properties relevant to pavement
performance. The result is that not all the predictive value of the model is
lost. This is discussed further in 6.
2 The interrelationships can be useful to verify that the model
characterises the behaviour of a newly developed paving material.
It is important that the same model characterises a newly developed
asphalt mixture. That means, that the newly developed mixture must not
exhibit new and unexpected failure mechanisms (types of damage). That
is a basic assumption which has to be fulfilled in order to control the
pavements cost-effectiveness and risk of failure based on the constitutive
model parameters as obtained from laboratory tests.
The interrelationships, which emerged from the present study, are
reviewed in the following.
5.1 Complex modulus
Figure 9 shows an example of a complex modulus master curve of an
asphalt concrete mixture. It was obtained using the four point bending test
as frequency sweep test. Let us consider the physical meaning of the
phenomenon complex modulus master curve. It means two things. Firstly,

Performance judgement of asphalt mixtures

290

complex modulus (MPa)

35000

Ch. 5

0C
10C
20C
30C

30000
25000
20000
15000
10000
5000
0
1

10

frequency (Hz)

100

-5

10

log (reduced frequency)

Figure 9. Example showing the complex modulus of an asphalt concrete mixture.


Left (a): Original frequency sweep data for 0C, 10C, 20C and 30C. Right (b):
Complex modulus master-curve, S * ( f r ) , where f r is the reduced frequency.

it means, that the time temperature superposition principle is valid. This


can be expressed mathematically in the form of the following equations7:

(t ) = J (t )d ( )

(6a)

(t ) = S (t )d ( )

(6b)

where (t ) is the creep strain, J (t ) is the creep compliance, (t ) is the


stress, S (t ) is the stiffness modulus, and is the reference time. The time
temperature superposition principle is the basic principle of linear
viscoelasticity. Secondly, it means that a necessary condition for its
validity is satisfied, which is that the stress strain behaviour is linearly
viscoelastic. In the case of an asphalt mixture, this is in a way peculiar,
because the stress strain behaviour of asphalt mixtures is, in general, not
linearly viscoelastic. The fact that the time temperature superposition
principle is applicable implies that the nonlinearity of the stress strain
behaviour of asphalt mixture can be negligible, in which case it is
approximately linear, cf. Intermezzo 1. The approximate linearity implies
that the stress strain behaviour is approximately independent of the stress;
7

The constitutive equations (6) are forms of the so-called principle of


superposition. They were introduced as purely empirical laws in the theory of the
viscoelastic after-effect (Boltzmann 1874). The principle of superposition
appears also in the laws of dielectrics (Hopkinson 1876) and in the laws of linear
electrical networks (Carson 1926). The integrals were first introduced in pure
mathematics where they are called Duhamels integrals (Duhamel 1833).

Sec. 5

Interrelatedness of constitutive model parameters

291

that is, of the stress distribution in the specimen. That explains why more
or less the same complex modulus is found using different specimen
geometries in the two-point bending test, the three-point bending test, and
the four-point bending test, as was shown in an international
interlaboratory investigation (Partl and Francken 1997). Thus, it seems
that the complex modulus can be obtained independent of the specimen
geometry. It means, that the complex modulus can be considered a true
material property, at least for practical purposes.
5.2 Linear viscoelastic creep
Let us consider linear viscoelastic creep. The Burgers model is a linear
rheological model, cf. Ch. 3, figure 62. For a constant applied stress, ,
the response strain is described by
t
(t ) 1
=
+
+ C (1 e t / 2 )
(7)
E1 1

where E1 is the series spring of the Burgers model, 1 is the series


dashpot of the Burgers model, C is a constant, and 2 is the relaxation
time of the Voigt-Kelvin element of the Burgers model. Letting t ,
the third term in equation (7) vanishes, reducing it to the following form
(t )
= J 0 + J 1t m
J (t ) =
(8)

where J 0 is the instantaneous elastic compliance (negligible), J 1 is the


creep compliance at 1 s, and m is the linear viscoelastic creep
susceptibility. Note that m = 1 , when a linear rheological model
consisting of springs and dashpots is used. This arises as a consequence of
the use of Newtons viscosity law to define the behaviour of the dashpot.
For real viscoelastic materials, m is smaller than unity. Equation (8)
represents linear viscoelastic creep. That is: creep which is linear as
function of the stress, and not necessarily also as function of the time.
Mathematically, the following equation can be derived from equation (8),
(ln S * )
=m
(9)
(ln )
where S * is the complex modulus, is the angular frequency, and m is
the slope of the master curve on log-log scale. Thus, if the complex
modulus is a true material property, independent of the specimen
geometry, then m is a true material property.
In table 1, the m -values of 25 asphalt concrete mixtures are recalled from
Ch. 3, table 5. The m -values are given for conditions of temperature and
frequency normally used in the triaxial creep test. Table 1 shows that the
m -value for the conditions in the triaxial test is practically equal to the
maximum limiting value in the low frequencies. The minimum value
obtained was approximately 0.26, the maximum value was approximately

Performance judgement of asphalt mixtures

292

Ch. 5

Table 1. m -values of 25 crushed gravel asphalt concrete mixtures,


CGAC 0/22.
CGAC 0/22

Mean
0.365
0.359
0.357
0.362
0.362

m0

m40o C , 0.5 Hz
m 40o C ,1 Hz

m50o C , 0.5 Hz
m 50o C ,1 Hz

stdev
0.047
0.048
0.049
0.048
0.048

min
0.276
0.266
0.264
0.271
0.270

Max
0.482
0.479
0.478
0.481
0.481

0.48, and a representative average value is 0.36. The parameter m is the


same in equation (8) and equation (9). In equation (9), m can be regarded
as the linear viscoelastic creep susceptibility. In equation (9), m
represents the slope of the complex modulus master curve. Hence, in the
linear case, dynamic stress strain behaviour and creep are directly related.
Theoretical proof for this is available from the literature, and is
reproduced here for the interested reader, cf. Intermezzos 1 and 2. In
Intermezzo 1, it is shown that equation (6) represents a mathematical
formulation of the time temperature superposition principle. The validity
of this principle implies the complex modulus master curve, which can be
represented by equation (9). In Intermezzo 2, it is shown that equation (9)
can be derived from equation (8). Thus, it is shown that linear viscoelastic
creep and linear dynamic stress strain behaviour are related, and that both
types of behaviour are governed by the time temperature superposition
principle, constitutive equation (6).
___________________________________________________________
Intermezzo 1
The simplest, realistic viscoelastic constitutive equation for which response under
constant temperatures can be used to predict response under transient temperatures is
that for a so-called thermorheologically simple material (Morland and Lee 1960).
By definition, the one-dimensional equation in terms of compliance is

x (t ) = J ( )
0

dx
d
d

(ia)

or, equivalently,
t

x (t ) = J ( )
0

dx
d t
d t

(ib)

where x is the axial strain due to stress, and


t

= (t )

d / aT ,

(t ) =

d / a

(ic)

It has been assumed that x = x = 0 when t < 0, which enables later use of
Laplace transform theory; no actual generality is lost since the time origin always
can be selected so as to satisfy this condition. The parameter is called reduced
time and

Sec. 5

Interrelatedness of constitutive model parameters

293

aT = aT (T ( ) )
(ii)
is the so-called temperature shift-factor. The only temperature-dependent material
property is this shift-factor; it reflects the influence of temperature on the internal
viscosity. The inverse of equation (i), in which stress is expressed as a functional of
strain, can be derived from equation (i) by first introducing the Laplace transform
with respect to reduced time,

f e s f ( ) d

(iii)

where f is a function of reduced time, and f is its Laplace transform. Next,


equation (ia) is transformed using the familiar rule for the Laplace transform of
convolution integrals (e.g. Flgge 1967), and then solved for x ; thus
x = s S x
(iv)
where, by definition,

s S x = ( s J ) 1

(v)

Laplace inversion of equation (iv) by means of the convolution rule yields

x (t ) = S ( )
0

d x
d
d

(via)

or, equivalently,
t

x (t ) = S ( )
0

d x
d t
d t

(vib)

Conversely, it can be shown in a similar manner, that if equation (vi) is known to be


valid for a given material, then equation (i) also will be true.
The experimental bases for equations (i) and (vi) under constant temperature
conditions can be demonstrated by considering isothermal creep and relaxation tests,
respectively. For the creep test, with
x = x H (t ) and x = constant,
(vii)
where
0 t < t
(viii)
H (t t )
1 t > t
represents the Heavyside unit step function, equation (ib) yields
x = J ( ) x
(ix)
and for the relaxation test, with
x = x H (t ) and x = constant,
equation (vib) yields
x = S ( ) x
For both types of isothermal tests,
= t / aT or log = log t log aT

(x)
(xi)
(xii)

Equation (xii) shows that the effect of temperature on compliance or modulus for a
thermodynamically simple material produces only horizontal translations when the
property is plotted against log t . Conversely, if it is found that constant temperature
curves can be superposed so as to form a single curve, which is called the

Performance judgement of asphalt mixtures

294

Ch. 5

mastercurve, by means of only rigid, horizontal translations, then the relaxation


modulus depends on time and temperature through the one parameter . Thus it is
shown that the validity of the time temperature superposition principle implies the
modulus or compliance mastercurve.

___________________________________________________________
Intermezzo 2
Let the Carson transform of f be defined by
~
f sf

(xiii)

where

f = e s t f (t ) d t

(xiv)

is the Laplace transform of f . By definition, cf. Intermezzo 1, equation (v),


~ ~
S = J 1
The response to a steady-state vibrational stress of frequency is defined by the
complex compliance or modulus, respectively
J * = (t ) / (t )
S * = (t ) / (t )
(xv)
where
(t ) = exp (i t + )
(t ) = exp (i t )
i = 1
(xvi)
Thus,
S * = J *1
(xvii)
It is also known that steady-state vibration functions and Carson transforms are
simply related (e.g. Pipkin 1972),
~
~
S * ( ) = S | s i
J * ( ) = J | s i
(xviii)
By using the power law creep compliance, equation (8), we find
~
J = J 0 + J 1 (1 + n) s m
(xix)
where (1 + n) is the Gamma function with argument n + 1 . From (xviii),
J * = J i J
(xxa)
where, the real part, J , and the imaginary part, J , are
J = J 0 + J 1 (1 + m) cos (m / 2) m
(xxb)
J = J 1 (1 + m) sin (m / 2) m

From which it follows that


(ln J * )
=m
(ln )

(xxc)

(xxi)

Thus it is shown that linear dynamic viscoelastic stress strain behaviour and linear
viscoelastic creep are directly related.

___________________________________________________________

5.3 Creep of asphalt mixture


Let us consider the creep of asphalt mixture. Two results from chapter 3
are recalled here. The first result, in figure 10, is for a 60 x 100 mm height
x diameter specimen, cf. Ch. 3, figures 5152. The second result, figure
11, is for a 200 x 100 mm height x diameter specimen, cf. Ch. 3, fig. 48.

Sec. 5

Interrelatedness of constitutive model parameters

295

For both results the material was dense graded asphalt concrete. The testtemperature was 50C, and the applied load was a block-wave with a
loading time of 0.2 s, and a rest-time of 0.8 s.
As for the 60 x 100 mm specimen, figure 10, the creep curves are
described by the logarithmic creep model according to the equation
J (t ) = J 1 + ln t z
(10)
where J (t ) is the creep compliance as function of the time, J 1 is the creep
compliance at 1 s, and z is the viscoplastic creep susceptibility. Figure 10
shows the exponent, z , of equation (10) as function of q / p . Figure 10
shows that z ( q / p ) obeys a power law relationship.
As for the 200 x 100 mm specimen, figure 11 and table 2, different creep
models, depending on the testing conditions, describe the creep curves.
There are a couple of interesting aspects of the result shown in figure 11.
Firstly, a transition can be observed from the logarithmic model, equation
(10), to the power law model, as a function of increasing stress
~ ~
J (t ) = J 0 + J 1 t z
(11)
~
where J 0 is the instantaneous elastic compliance at 0 s (negligible), J 1 is
the creep compliance at 1 s, and ~z is the viscoplastic creep susceptibility.
This transition takes place as the ratio of the deviatoric stress and the
volumetric stress increases, somewhere between q / p = 0.43 and 0.95.
Figure 11 shows that ~z ( q / p ) obeys a linear relationship beyond q / p =
0.95. It is noticed, that this relationship differs entirely from that for the 60
x 100 mm specimen of figure 10. So, it seems, the creep properties depend
strongly on the specimen geometry. If that is indeed the case, then creep
properties cannot be regarded predictive for the behaviour in a geometry
which differs from the one in which they were measured, e.g. the
pavement. In other words, the creep properties of asphalt mixture cannot
be considered true material properties.
Another interesting aspect of figure 11 is that the exponent of the power
law model, ~z , starts at a value lower than the m -value obtained from the
master curve, then increases as function of increasing q / p , and becomes
greater than the m -value. The minimum value of ~z is approximately
0.30, and the maximum value approximately 0.43. Alternatively, one
could say the exponent is equal to m within the scatter band between the
minimum and maximum values of m . So, it seems that m can be used as
an estimator for the creep susceptibility. That is interesting, because m is
obtained as a true material property, independent of the specimen
geometry.
It is interesting for another reason, which is the mathematical
correspondence of the elasto-viscoplastic power law model and the linear
viscoelastic power law model. It was shown in chapter 3, that, based on

Performance judgement of asphalt mixtures

296

150

DAC 0/16, block-wave, 1 Hz

1000

exponent, z10^4

exponent, z10^4

200

y = 29.312x1.6396
R2 = 0.976

100
50
0
0.00

1.00

2.00
q/p

3.00

p=0.37

p=0.17
p=0.4

p=0.23
p=0.43

p=0.3
p=0.47

p=0.33
p=0.5

DAC 0/16, block-wave, 1 Hz

100

10

y = 29.312x1.6396
R2 = 0.976

1
0.10

4.00

Ch. 5

p=0.37

1.00
q/p
p=0.17
p=0.4

p=0.23
p=0.43

10.00
p=0.3
p=0.47

p=0.33
p=0.5

Figure 10a. Exponent z of equation (10) as function of the quotient of the


deviatoric stress, q , and the volumetric stress, p . The legend shows the
volumetric stress in MPa. DAC 0/16, block-wave form of applied stress.
Hydraulic load control. Specimen: 60 x 100 mm height x diameter. (a): On
linear scale. Right (b): On log-log scale. In the regression equations, y(x), y
represents z104, and x represents q/p.

exponent, z10^4

100
80

DAC 0/16, sinus, 1 Hz


y = 18.374x1.514
R2 = 0.9157

60
40
20
0
0.00

p=0.37

1.00

2.00
3.00
4.00
q/p
p=0.17
p=0.23
p=0.3
p=0.43

DAC 0/16, sinus, 1 Hz

100

exponent, z10^4

120

10
y = 18.374x1.514
R2 = 0.9157
1
0.10
p=0.3

1.00
q/p
p=0.17
p=0.37

10.00
p=0.23
p=0.43

Figure 10b. Exponent z of eq. (10) as function of the quotient of the deviatoric
stress, q , and the volumetric stress, p . The legend shows the volumetric stress
in MPa. DAC 0/16. Sinusoidal applied stress. Hydraulic load control.
Specimen: 60 x 100 mm height x diameter. Left (a): On linear scale. Right (b):
On log log scale. In the regression equations, y(x), y represents z104, and x
represents q/p.

this correspondence, a physical meaning can be attributed to the exponent


of the elasto-viscoplastic power law model, ~z , even though ~z depends on
the specimen geometry. Because a physical meaning can be attributed to
~
z , also a physical meaning can be attributed to the exponents of all other
creep laws. For more details, the reader is referred to Ch. 3, 13.
Why are creep and dynamic stress strain behaviour related in the linear
case? Are these types of behaviour also related in the nonlinear case?
In the opinion of the author, the answer to the first question is that creep

297

Interrelatedness of constitutive model parameters

DAC 0/16, block-wave, 1 Hz

1.000

0.100

0.010

0.001
1

10

100

1000 10000

DAC 0/16, block-wave, 1 Hz

0,5

exponent z of eq. (11)

compliance J(t) (MPa^-1)

Sec. 5

0,45

z = 0.0713 q/p + 0.2353

0,4

0,35

time (s)
0,9/0,3 MPa
0,3/0,2 MPa
0,3/0,03 MPa
0,6/0,05 MPa

0,6/0.25 MPa
0,6/0,1 MPa
0,75/0,15 MPa
0,6/0,15 MPa

0,3
0

0,5

1,5

2,5

q/p

Figure 11. Dynamic creep of dense graded asphalt concrete, DAC 0/16. Left (a):
Each curve represents the average of three tests. The waveform of the applied
load was a block-wave with a loading time of 0,2 s and a rest-time, of 0,8 s.
Load control: pneumatic. Temperature: 50C. Specimen: 200 x 100 mm height x
diameter. In the legend, combinations of axial stress and radial stress in MPa
are given; see also table 2. Right (b): Exponent ~z of eq. (11) as function of the
quotient of the deviatoric stress, q, and the volumetric stress, p.
Table 2. Creep models belonging to the creep curves in figure 11.
1 / 3
creep model
q/ p
R2
99.9
0.6/0.05
J (t ) = 0.0032 t 0.4255
2.36
0.3673
99.7
0.3/0.03
2.25
J (t ) = 0.0040 t
99.4
0.6/0.1
1.88
J (t ) = 0.0029 t 0.3648
0.3750
99.7
0.75/0.15
1.71
J (t ) = 0.0020 t
0.3363
99.6
0.6/0.15
1.5
J (t ) = 0.0020 t
0.3192
99.1
0.9/0.3
1.2
J (t ) = 0.0010 t
0.3040
99.4
0.6/0.25
0.95
J (t ) = 0.0010 t
99.1
0.3/0.2
0.43
J (t ) = 0.0001 + 0.0004 ln t

and dynamic stress strain behaviour are described by the same constitutive
model, equation (6), but only if the following conditions are satisfied:
. the material is a homogeneous continuum;
. the stress strain behaviour is linear;
. the stress is uniform.
The answer to the second question is that creep and dynamic stress strain
behaviour are in general not related in the nonlinear case. A probable
explanation for this lies in the interaction between the nonlinearity as a
function of the stress, and the fact that the stress distribution in the
specimen is not uniform. The nonlinearity comprises two aspects:

298

Performance judgement of asphalt mixtures

Ch. 5

. the intrinsic nonlinearity of the binder, since this is pseudoplastic and


thixotropic;
. the nonlinearity of the grain skeleton, since this responds differently to
compressive stress and tensile stress.
The non-uniformity of the stress distribution is caused by two factors:
. the specimen geometry;
. the heterogeneous material structure.
5.4 Crack-growth in asphalt mixture
The following model was given to the author in a private communication
with R.L. Lytton of Texas A&M University. The fundamental relation of
fracture mechanics as stated by Schapery (1984) is
2 = E R J (t ) J v
(12)
where is the surface energy density, i.e. the energy required per unit
area of crack surface to separate a material from itself (in N/m), E R is the
reference modulus used, if necessary, in representing a nonlinear
viscoelastic material as an equivalent nonlinear elastic material (in N/m2),
J (t ) is the tensile creep compliance corresponding to the time, t , that is
required for a crack to move over the distance , which is the length of
the fracture process zone ahead of the crack tip (in m2/N), and J v is the
strain energy or pseudo-strain energy J -integral. This is the rate of
change of released energy per unit crack growth area from one tensile load
to the next (in N/m).
Equation (12) is an energy balance; the energy given up on the right-hand
side of the equation is taken up by the newly created crack surfaces on the
left-hand side of the equation.
The reference modulus is defined by the equation
1 t

0 =
E (t )
d
(13)

ER 0

where 0 is the pseudo-strain, E (t ) is the relaxation modulus, t is the


time, and is a reference time. The reference modulus makes the
equation dimensionally correct and may be set at any constant value
which produces a physically meaningful value of the pseudo-strain. One
may define the time required for the fracture to traverse the length of the
fracture process zone as
k
t =
(14)
a&
where k is a constant, and a& is the crack speed, da / dt . Substitution of
equation (8) into equation (12), with t = t yields
m

k
2 = E R J 0 + J 1
(15)
Jv
a&

Sec. 5

Interrelatedness of constitutive model parameters

299

or
1/ m

J1 E R J v

(16)
a& = k
2
J
E
J

0
R
v

If it is assumed that the J -integral energy is released according to the


normalised wave form with time, w(t ) ,
J v = J v 0 w(t )
(17)
where J v 0 is the maximum value of J v during the time interval, t , then

( J 1 E R )1 / m w(t )1 / m
da t
1/ m
dt
= k
( J v0 )
1/ m
[ 2 J 0 E R J v 0 w(t )]
dN 0

This is of the form


da
= A ( J v0 ) n
dN
where
t
( J 1 E R )1 / m w(t )1 / m
dt
A = k
[ 2 J 0 E R J v 0 w(t )]1 / m
0

(19)

n0 = 1 / m

(21)

(18)

(20)

If J 0 is much smaller than J 1 , as is commonly the case, then equation


(20) simplifies to
1/ m

k m J 1 E R t w(t )1 / m
A=
dt
(22)

1 / m
2

0
According to the theory, the exponent n0 is a function which depends on
the characteristics of the failure zone. In different articles, Schapery
discussed different possible functions for n0 as function of m . For
example, if the materials fracture energy, , and failure stress within the
fracture process zone are constants, then n0 = 1 + 1 / m . If the fracture
process zone size and are constants, then n0 = 1 / m . For the linear
elastic case, the J -integral can be written as
~
(23)
J = K I2 / E
~
where K I is the stress intensity factor, and E is the modulus of elasticity.
~
E=E
for plane stress
(24a)
~
E = E /(1 2 )
for plane strain
(24b)
where is Poissons constant. Substitution into equation (19) yields
da
= A (K I ) n
(25a)
dN
n = 2/ m
(25b)
n = 2 (1 + 1 / m)
(25c)
where n depends on the properties of the fracture process zone. At this
point it is recalled from Ch. 4, 3.4 that the properties which characterise

300

Performance judgement of asphalt mixtures

Ch. 5

the resistance to crack growth, A and n , depend on the complex modulus


(the slope of the master curve, m ), the creep compliance (the coefficient
of the creep law, J 1 ), the tensile strength, and the fracture energy. An
indirect characterisation of the resistance to crack growth based on these
properties is relatively simple and permits the use of more practical tests.
5.5 Fracture toughness and tensile strength
From the previous section, the dynamic crack-growth properties are the
coefficient A and the exponent n of the Paris equation, equation (25).
From a dynamic crack-growth test, the critical stress intensity factor, K c
may be determined. If the critical stress intensity factor is independent of
the specimen thickness, as seems to be indicated by the results in Ch. 4,
table 8, then it is expected to be the plane strain critical stress intensity
factor, i.e. the fracture toughness (provided the ligament is linear elastic,
and unweakened by fatigue). To investigate this, the fracture toughness
was determined using the SCB-specimen. From the analysis in Ch. 4, 6, it
seems that the value is indeed reproducible using a different specimen. If
true (it needs to be reconfirmed), it shows that the dynamic crack-growth
properties and the fracture toughness are related.
The conditions that have to be fulfilled, linear elasticity, and a plane strain
state of stress, to determine the fracture toughness are very strict. These
conditions cannot be fulfilled for temperatures above approximately 10C.
It can be shown that the fracture toughness and the tensile strength differ
by a constant factor in the absence of strain hardening, cf. Ch. 4, 7.6. If
the conditions for the determination of the fracture toughness cannot be
fulfilled, then one might consider to determine the tensile strength, since
this can be related to the fracture toughness, and hence indirectly to the
dynamic crack-growth properties.
5.6 Summary
The asphalt mixture properties considered in this study are:
. complex modulus,
. creep compliance,
. dynamic crack-growth properties,
. fracture toughness,
. tensile strength.
These properties are interrelated. Not only the properties as such, also the
interrelationships, characterise the materials behaviour. Upon application
of the constitutive model (i.e. the entire set of equations) to a newly
developed asphalt mixture, the interrelationships between the model
parameters can be useful to verify that the model characterises the
behaviour of a newly developed paving material, and also to exclude that
a newly developed paving material exhibits new and unexpected failure
mechanisms (types of damage).

Sec. 6

Discussion

301

6 Discussion: Importance of the interrelatedness of constitutive


model variables and parameters
Any mechanical asphalt mixture property, even a geometry-independent
(true) material property is at best relatively (qualitatively) predictive for
the behaviour of the material in the pavement, cf. 4.3.
Some asphalt mixture properties can be obtained approximately
independent of the specimen geometry. Some asphalt mixture properties
can be obtained approximately independent of the specimen size.
Examples:
i The complex modulus is reasonably reproducible using different
specimens with different geometries (Partl and Francken 1997), if bending
is the dominant phenomenon. Also properties directly related to the
complex modulus, e.g. the linear viscoelastic creep susceptibility, m in
equation (8), can be considered approximately geometry-independent.
ii The crack-growth properties, i.e. the constants of the Paris equation,
within the scatter on the data, were found not to depend on the specimen
size.
A mechanical asphalt mixture property, if geometry-dependent, need not
lose all predictive value for the behaviour of the material in the pavement,
if it is reproducible. A geometry-dependent mechanical property is
reproducible, if its defined relationship to at least one geometryindependent (true) material property is fixed. For example, the tensile
strength of an asphalt mixture is a geometry-dependent mechanical
property. Values obtained using different specimens are reproducible, if
specimens and testing conditions are standardised. Standardisation of the
specimen and the testing conditions will result in that the defined
relationship of the tensile strength to the mixtures true material
properties, e.g. the complex modulus, is fixed this study shows that the
tensile strength and the complex modulus of an asphalt mixture are related
(see also Heukelom 1966).
Thus, a geometry-dependent material property is reproducible if its
relationship to at least one geometry-independent (true) material property
is fixed:
. by means of a constitutive model;
. by a standardised test method (test specimen and testing conditions).
A constitutive model provides a theoretical definition of all mechanical
properties that are relevant to the mechanical behaviour of the material in
the pavement. The uniqueness of the definition of a mechanical property is
completed if the test method to determine it is standardised. Geometryindependent (true) properties serve as fixed references for the predictive
value of the geometry-dependent material properties. Therefore it seems
unnecessary to have all mechanical properties relevant to the behaviour in
the pavement independent of the specimens geometry (size and shape). It

302

Performance judgement of asphalt mixtures

Ch. 5

suffices to have at least one property as a true material property, provided


the relatedness of that property to other properties required to judge
pavement performance is fixed.
The behaviour of a non-standardised paving material can be characterised
and judged in comparison to a standardised paving material for a similar
application, for which the behaviour is known.

7 Conclusions
1 A method is proposed to control a pavements cost-effectiveness and
risk of failure, based on property-related requirements for the applied
paving materials (asphalt mixtures), and material properties determined in
the laboratory.
2 Since the properties of a given asphalt mixture, determined in the
laboratory, are predictive for the behaviour of that material in the
pavement, it is improbable that upon application of that material the
pavements cost-effectiveness is reduced or the pavements risk of failure
is increased if the material properties which are relevant to those
pavement properties indicate improved behaviour (other factors than
material behaviour influencing pavement performance excluded from
consideration).
3 In order for a property-related requirement for asphalt mixture not to be
equally empirical as current composition-related requirements, properties
shall be used that are not defined arbitrarily. A physical model is
necessary to define physically meaningful properties that are predictive
for the behaviour of the material in the pavement.
4 For different reasons, it suffices that asphalt mixture properties, defined
in the form of constitutive model parameters or variables, have a relative
predictive value. This means, that the values of the properties must be
considered in comparison to the corresponding properties of a comparable
standardised asphalt mixture and in comparison to the performance, i.e.
cost-effectiveness and risk of failure, of that standardised mixture in the
pavement.
5 It was made plausible that property-related requirements for asphalt
mixture can help enhance innovation in the field of paving materials,
because property-related requirements do not lose their applicability upon
modification of the material, and because the properties referred to in the
requirements are predictive for the behaviour of the material in the
pavement.

6
General discussion
1 Introduction
The present study was undertaken with the following general objective
and goals, cf. Ch. 1, 3:
1 General objective: to make a characterisation of the mechanical
behaviour of an asphalt mixture possible in as much as that is relevant to
the pavements main functions, bearing capacity, surface characteristics,
and long-term performance. This characterisation should be based on
sound engineering principles rather than on practical experience. The
ultimate aim is to facilitate the acceptance of new and non-standardised
paving materials that are needed to enhance the durability of our heavily
trafficked main road network.
2 Practical goal: to make possible a characterisation of an asphalt
mixtures mechanical behaviour relevant to its behaviour in an asphalt
pavement allowing the use of tests that are suitable for the practical
purposes of material selection in pavement design, asphalt mixture design
(type testing), and production quality control.
3 Research goal: to develop a method for the validation of simple tests for
practical purposes.

2 General objective
The ultimate aim is to facilitate the acceptance of new and nonstandardised paving materials that are needed to enhance the durability of
our heavily trafficked main road network.

304

General discussion

Ch. 6

The mechanical behaviour of an asphalt mixture in an asphalt pavement


comprises properties that are relevant to the pavements main functions,
bearing capacity, surface characteristics, and long-term performance, cf.
Ch. 1, table 1.
Important for the pavements long-term performance is also the
characterisation of the asphalt mixtures physiochemistry, i.e. the adhesion
and disbonding of mineral aggregate and bitumen, and the ageing of
bitumen. However, the physiochemical aspects are beyond the scope of
this study.
Traditionally, acceptance for application of new products has been a
tedious process that relies heavily on practical experience. This implies
that potentially promising products need long time to be implemented in
practice. Moreover, practical experience can tell whether a new product is
successful but cannot tell why. It is believed that modelling and
characterisation can help shortening the implementation period of newly
developed materials. It is also believed that knowing the behaviour of the
materials is the basis for risk of failure and cost-effectiveness analyses.
In general, the stress strain behaviour of an asphalt mixture is stress
dependent (nonlinear), and the stress, i.e. the stress distribution in the
material, is not uniform but three-dimensional. Two factors contribute to
the nonuniformity of the stress:
1 macro-geometrically: specimen geometry, load application, size effects;
2 meso-geometrically: the granular structure, and the different mechanical
properties of the constituent phases, mineral aggregate and binder.
This combination of factors often leads to geometry-dependence of
mechanical properties. A geometry dependent mechanical property has no
predictive value for the mechanical behaviour in a different geometry.
Therefore, measured mechanical asphalt mixture properties must often be
considered as specimen properties, and not as material properties.
This conclusion results immediately in the need to develop material
models to overcome the abovementioned problems. Ideally, models
describing the material behaviour of the asphalt mixtures on the mesolevel, i.e. on the level of the granular structure, would be needed. Such
models are not yet feasible for everyday use. Therefore, it seems that
models capable of predicting the material behaviour on the macro-level,
i.e. on the level of a structure (pavement or specimen) under well known
three-dimensional states of stress, are adequate for the time being. This
study aimed to contribute to the development of such models.

3 Practical goal
The practical goal of this study was to make possible a characterisation of
an asphalt mixtures mechanical behaviour relevant to its behaviour in an
asphalt pavement allowing the use of tests that are suitable for the

Sec. 3

Practical goal

305

practical purposes of material selection in pavement design, asphalt


mixture design (type testing), and production quality control. Therefore,
the following tests were investigated:
. a frequency sweep test, to characterise the linear dynamic viscoelastic
stress strain behaviour;
. a dynamic triaxial creep test, to characterise the nonlinear dynamic
elasto-viscoplastic stress strain behaviour;
. a tensile test, to characterise the resistance to crack-growth;
. a fracture toughness test, to characterise the resistance to fracture.
The investigations yielded the following practical results.
Four point bending test as frequency sweep test
An international interlaboratory investigation (Partl and Francken 1997)
showed that the complex modulus is reproducible within a normal band of
scatter using different specimens, in the two-point bending test, the three
point bending test, and the four point bending test. Based on that
investigation it can be concluded that the complex modulus is fairly
reproducible using different specimen geometries. Therefore, from a
practical standpoint, the complex modulus, based on the tests mentioned,
can be considered a true material property, i.e. when bending is the
dominating phenomenon.
Triaxial creep test
Creep properties depend on the specimen height. The experimental
evidence in the present study supports this, cf. Ch. 3, 8. A possible
explanation emerged from the finite element model of the creep specimen,
cf. Ch. 3, 9. It should be mentioned, however, that some part of the
observed dependence is to be attributed probably to insufficient friction
reduction in the interface between the specimen and the loading platens.
This is supported by the data in App. 2, figure 1.
In spite of the geometry dependence, a relative predictive value can be
retained, if the reproducibility is secured. As mentioned, cf. Ch. 5, 6, this
requires that the defined physical relatedness to at least one true material
property is fixed, which requires that the specimen and the testing
conditions are standardised.
Tensile test or fracture toughness test
Three types of tensile test were investigated: the uniaxial tensile (UT) test,
the semi-circular bending (SCB) test, and the indirect tensile (IT) test. The
three tests yield different values of the tensile strength, all relevant testing
conditions being equal apart from the specimen geometry (and the
geometry of the load application).

306

General discussion

Ch. 6

It is believed that confusion has in part to do with the fact that road
engineers are somewhat loose in their terminology. The SCB-test is in fact
a bending test, whereas the IT-test is a biaxial test. From principles of
material behaviour it is known that such tests do not yield similar values
of the stress at rupture of the specimen. The finite element analysis
described in Ch. 4, 8 also showed that the material behaviour in the SCBspecimen is highly nonlinear, implying that one must be careful in
choosing testing conditions to allow application of the linear elastic
theory, if that is indeed valid at all.
The fracture toughness may be determined as an alternative for the tensile
strength1. The fracture toughness is commonly considered to be the
residual tensile strength of a cracked (i.e. notched) specimen (however,
expressed in MPam instead of MPa). A valid fracture toughness is
independent of the specimen thickness, and is considered a true material
property. In this study, valid fracture toughnesses of asphalt mixture were
obtained at a temperature of 1C or lower.

4 Research goal
The research goal was to develop a method for the validation of simple
tests for practical purposes. This was done by making a contribution to the
modelling and characterisation of the mechanical behaviour of asphalt
mixtures. It is believed that the following important results were obtained.
In Ch. 4, 5 was described how fundamental crack-growth properties, i.e.
the constants of the Paris equation, can be determined in dynamic crackgrowth tests. The test using the centre-cracked tensile (CCT) specimen is
less suitable for use on a routine basis in the day-to-day practice. Such
tests are called here advanced tests2. It was made plausible that the
value of the critical stress intensity factor, K c , in the CCT-specimen is
reproducible in the form of the fracture toughness, K Ic , in the semicircular bending (SCB) test, cf. Ch. 4, 6. It was also made plausible that
the fracture toughness and the tensile strength differ by a constant factor in
the absence of strain hardening, Ch. 4, 7.6. Thus was made plausible that
the dynamic crack-growth properties and the tensile strength are linked, as
expected based on theory, cf. Ch. 4, equation (8). The dynamic crack1

And advantage of the fracture toughness is its reproducibility as a true material


property, since it is independent of the specimen geometry, cf. Ch. 4, 6.4.1.
Another advantage, only with fine graded asphalt mixtures (0/8 and finer), is that
the standard deviation is approximately 50% of that of the tensile strength
obtained using the un-notched specimen, which is beneficial to the
discriminatory ability of the test. This was not discussed.
2
A so-called advanced crack-growth test is here a crack-growth test that is less
suitable for use on a routine basis in the day-to-day practice.

Sec. 4

Research goal

307

growth properties were obtained independent of the specimen size, cf. Ch.
4, figure 23. Also the fracture toughness was obtained independent of the
specimen size, cf. Ch. 4, 6.4.1 and independent of the specimen
geometry if different thickness to diameter ratios and support spans are
considered. Note, that the dynamic crack-growth properties and the
fracture toughness were determined at respectively 0C and 1C.

308

General discussion

Ch. 6

7
General conclusions
1 Introduction
In chapters 3 and 4, the experimental evidence of this study was presented.
In chapter 5, the feasibility of property-related requirements to replace
composition-related requirements was investigated; that is, the possibility
to control the pavements performance by controlling the cost-effectiveness and maximum acceptable risk of failure of the pavement as a whole
and of the applied paving materials. A section containing the conclusions
of the investigation concludes each of those chapters.
In this chapter, the general objective and goals as mentioned in the
introduction on page 15 are addressed; the general objective in 2, the
practical goal in 3, and the research goal in 4. The ultimate aim to
enhance innovation in the field of pavement design and paving materials
to be achieved on the basis of property-related requirements, is addressed
in 5.

2 General objective
The general objective of this study was to make a characterisation of the
mechanical behaviour of an asphalt mixture possible in as much as that is
relevant to the pavements main functions, bearing capacity, surface
characteristics, and long-term performance.
The ultimate aim was to know the material behaviour of a paving material,
to make possible that the acceptance for application of non-standardised
paving materials can be facilitated as a means to enhance innovation in the
field of pavement design and paving materials.

310

General conclusions

Ch. 7

The following is concluded regarding the general objective:


i Available theories permit a physical characterisation of the behaviour of
asphalt mixtures relevant to the pavements main functions, bearing
capacity, surface characteristics, and long-term performance.
ii An advantage of a physical characterisation as compared to currently
used empirical characterisations of asphalt mixture is that a physical
characterisation has a more general predictive value.
iii To allow a physical characterisation based on an analytical approach,
such as the approach to viscoelastic and viscoplastic stress strain
behaviour in chapter 3, and the approach to crack-growth and fracture in
chapter 4, a simplification of the mechanical material behaviour was
unavoidable.
iv The following three mechanical asphalt mixture properties contribute to
the pavements bearing capacity, surface characteristics, and long-term
performance:
. the resistance to deformation;
. the resistance to crack-growth;
. the resistance to fracture.
The materials resistance to deformation can be characterised by its
dynamic elasto-viscoplastic stress strain behaviour. The materials
resistance to crack-growth can be characterised by its dynamic crackgrowth behaviour. The materials resistance to fracture can be
characterised by the stress intensity at the tip of a crack in the material.
The characterisation of the resistance to fracture can be considered a less
demanding alternative of the characterisation of the resistance to crackgrowth.

3 Practical goal
The practical goal of this study was to make possible a characterisation of
an asphalt mixtures mechanical behaviour relevant to the pavements
main functions, bearing capacity, surface characteristics, and long-term
performance, allowing the use of tests that are suitable for the practical
purposes of material selection in pavement design, asphalt mixture design
(type testing), and production quality control.
The following is concluded regarding the practical goal:
v The following four tests were investigated and were found to be suitable
for a physical characterisation of the mechanical behaviour of asphalt
mixture for the practical purposes of material selection in pavement

Sec. 3

Practical goal

311

design, asphalt mixture design (type testing) and production quality


control:
. a frequency sweep test, to characterise the linear dynamic viscoelastic
stress strain behaviour;
. a dynamic triaxial creep test, to characterise the nonlinear dynamic
elasto-viscoplastic stress strain behaviour;
. a tensile test, to characterise the resistance to crack-growth;
. a fracture toughness test, to characterise the resistance to fracture.
vi A frequency sweep test1 is designed to determine the complex modulus.
The complex modulus is a form of stiffness modulus. It characterises an
asphalt mixtures dynamic linearly viscoelastic stress strain behaviour.
vii A dynamic triaxial creep test is designed to determine the creep
compliance. The creep compliance is the inverse of the stiffness modulus.
It characterises an asphalt mixtures elasto-viscoplastic stress strain
behaviour.
viii A tensile test is designed to determine the tensile strength. The tensile
strength is one of the properties that characterises an asphalt mixtures
resistance to crack-growth indirectly.
ix A fracture toughness test is designed to determine the fracture
toughness. The fracture toughness characterises an asphalt mixtures
resistance to fracture2. Valid fracture toughnesses were obtained at
temperatures of 1C and lower.
x The complex modulus is used as an input parameter in the pavement
design method to determine the pavement thickness, and is therefore
relevant to the pavements bearing capacity.
xi The creep compliance characterises an asphalt mixtures resistance to
permanent deformation, or rutting in the pavement, and is therefore
relevant to the pavements surface characteristics.

There are several tests possible, but in this study only the four point bending
test was used.
2
The criteria for a valid fracture toughness in the literature were developed for
metals. These can be generalised for asphalt mixture. A valid fracture toughness
for asphalt mixture is a fracture toughness that is independent of the specimen
thickness, and that is obtained if the stress strain behaviour is to a good
approximation linearly elastic.

312

General conclusions

Ch. 7

xii The tensile strength characterises an asphalt mixtures resistance to


crack-growth indirectly. The resistance to crack-growth is relevant to the
pavements bearing capacity, if the crack-growth occurs near the bottom
of the pavement, and is relevant to the pavements surface characteristics,
if the crack-growth occurs at the pavements surface3. Depending on the
crack-depth and the number of cracks per unit area, surface cracking
affects the pavements bearing capacity on the long term.
xiii For the practical purposes of material selection in pavement design,
asphalt mixture design (type testing) and production quality control, and
the control of an asphalt mixtures cost-effectiveness and risk of failure, it
suffices that asphalt mixture properties have relative (i.e. qualitative)
predictive value, since the influences of different external influence
factors are unpredictable, e.g. production conditions and service
conditions.
xiv In order to achieve relative predictive value, it suffices that asphalt
mixture properties are performance related, i.e. can be related to failure
(damage) phenomena observed in the pavement, and are reproducible.
xv To achieve reproducibility of mechanical asphalt mixture properties
that are often not obtained as true material properties, it is necessary,
. that at least one of those properties required to judge the mixtures
performance in the pavement, can be considered a true material property,
. that all properties required to judge the mixtures performance in the
pavement are uniquely defined and physically related, i.e. in the form of a
physical model, to at least one property that can be considered a true
material property, and
. that specimens and testing conditions to determine required properties
are standardised.
xvi Physically, the complex modulus of asphalt mixture is not a true
material property, but, based on an international interlaboratory
investigation, which showed its reproducibility using different specimen
geometries, can be considered a true material property from a practical
standpoint.

It is assumed, and there is also limited experimental evidence however not


conclusive, that the tensile strength and/or the fracture toughness characterise the
resistance of an asphalt mixture to ravelling. Ravelling of asphalt mixture is a
complex damage phenomenon, in which also physiochemical properties play a
role. This was not discussed in detail.

Sec. 3

Practical goal

313

xvii The physical model used to characterise an asphalt mixtures


mechanical behaviour shows that the complex modulus relates to many of
the mixtures properties including properties that are relevant to the
pavements surface characteristics and long-term behaviour. For example,
the model shows that the complex modulus is related to the crack-growth
properties and the tensile strength.

4 Research goal
The research goal of this study was to develop a method for the validation
of simple tests for practical purposes. The ultimate aim was to limit the
empirical character of the methods used for pavement design and asphalt
mixture design (type testing), and to find methods for validation of
methods other than methods using practical experience.
The following is concluded regarding the research goal:
xviii It was shown possible to validate the use of a simple test, namely the
semi-circular bending (SCB) test, based on a fundamental method, which
consisted of two elements, namely experimental work and a finite element
model of the specimen. The method is an example illustrating the
possibility to limit the empirical character of methods used for pavement
design, material selection, asphalt mixture design, and production quality
control.

5 Property-related requirements versus composition-related


requirements for asphalt mixture
This study was undertaken with the ultimate aim to know the material
behaviour of a paving material, to make possible that the acceptance for
application of non-standardised materials can be facilitated.
With regard to this ultimate aim the following comments are made in
addition to the previous conclusions.
xix It is believed that property-related requirements acting on material
properties that are relevant to the behaviour of the material in the
pavement remain applicable upon modification of the material.
In principle, once having performance related requirements for asphalt
mixtures, should make it possible to save time developing new
requirements for newly developed non-standardised paving materials.
Then it is possible to further rationalise the approach to pavement design,
material selection and asphalt mixture design.

314

General conclusions

Ch. 7

xx It was made plausible that property-related requirements for asphalt


mixture allow an asphalt mixtures expected relative4 performance in the
pavement to be judged.
xxi It is assumed that a rationalised approach to pavement design, material
selection, and asphalt mixture design (type testing), and facilitated
acceptation of newly developed standardised asphalt mixtures are
necessary to enhance innovation in the field of pavement design and
paving materials.

Relative means here in comparison to standardised asphalt mixtures for which


the performance is known.

315

Appendix 1: Influence of the shape of the waveform of the applied


stress on the response strain of the Burgers model
1 Introduction
With reference to Ch. 3, the aim of the present analysis is to show that the
response strain of a viscoelastic material depends on the shape of the
waveform of the applied stress. It is difficult to show that this is true also
for asphalt mixture, because the observed scatter on the mechanical
properties easily overshadows an influence of the shape of the waveform.
That the response strain depends on the shape of the waveform of the
applied stress, was made plausible experimentally using the accelerated
pavement tester of the Delft University and the RHEI by Groenendijk
(1998). Since the experimental evidence lacks the power of proof, owing
to the scatter on the data, the author felt it necessary to further make the
influence of the waveform plausible on the basis of theoretical predictions.
The dependence of the response strain on the shape of the waveform is
important since the shapes of waveform used in tests in the laboratory
differ from the shape of waveform of the strain induced in the pavement
by a passing wheel-load (Groenendijk 1998). This alone means that
constitutive model parameters as material properties cannot be
quantitatively (absolutely) predictive for the behaviour of the material in
the pavement! It means that constitutive model parameters as material
properties can at best be qualitatively (relatively) predictive for the
behaviour of the material in the pavement.
The author felt it also important to draw attention to the influence of the
shape of the waveform on the material response, because it is his
experience during this investigation that this fact and its consequences are
not always correctly understood. A theoretical model, even when
simplified, explains the dependence of the response strain on the shape of
the waveform of the applied stress.
Viscoelasticity is a phenomenon in which the time is of particular
importance. The properties of the viscoelastic continuum vary as a
function of the time during which a load is present1. This implies that the
shape of the waveform of the applied stress influences the test result. The
author decided to solve the Burgers equation analytically for different
shapes of waveform of applied stress, not having been successful to trace a
similar set of solutions in the literature.

A load can be an applied stress or an applied strain. In a stress controlled


experiment, the applied load is a controlled stress. All cases considered in the
following are stress controlled.

316

Influence of the shape of the waveform


in the Burgers model

App. 1

The Burgers model is shown in figure 1. It is noted, that analytical


solutions of the Burgers equation are possible only if the dashpots 1 and
2 are defined as Newtonian viscosities, i.e. constant as function of the
time.
E1

1
2
E2

Figure 1. The Burgers model.

This represents a limitation when applied to asphalt mixture, because the


binder used in an asphalt mixture, bitumen, is a pseudoplastic (i.e. shear
thinning) and thixotropic fluid.

2 Analysis
The differential equation for the Burgers model is written as
1
E

E1&& + 1 & = && + [ ] & +
2
1 2

(1a)

where

[ ] =

1 21
1 = 1 / E1
21 = 2 / E1
2 = 2 / E 2

(1b)

(1c)
(1d)
(1e)
Equation (1a) is a partial differential equation. Normally, a partial
differential equation is solved by means of the method of separation of
variables. However, in the following, it is solved by means of Laplace
transformation. The Laplace transformation method has a specific
advantage compared to the previous method: It is very efficient for more
complex shapes of waveform of the applied stress, (t ) , i.e. when the
applied stress has to be written in the form of a Fourier series. Solutions
(t ) are given for the following shapes of waveform of the applied stress:

Sec. 2

Analysis

317

1,5

2
1,5

0,5

0,5
0

0
0

1
2
1 + sin x x
constant

x (x pi)
in units

block-width = pi

(x )
pi)
xx(

1,5
1
0,5
0
0

halfsine x

x ( x)(x pi)

Figure 2. Different shapes of waveform of applied stress. (a) Left: constant


stress and haversine stress. (b) Right: unidirectional block-wave; the loading
time, or block-width is defined as 2 . (c) Right: half sine stress.

1 constant applied stress, cf. figure 2a;


2 sinusoidal applied stress (not shown in figure 2);
3 unidirectional sinusoidal applied stress, cf. figure 2a;
4 alternating block-wave of applied stress (not shown in figure 2);
5 unidirectional block-wave of applied stress, cf. figure 2b;
6 half sine of applied stress, cf. figure 2c.
The details of the analysis are given in 3.
2.1 Constant applied stress
For more details, the reader is referred to Case 1 in 3. For a constant
applied stress, cf. figure 2a,
(t )

t < 0 : = 0
(2)

(t )
t 0 :
=1

equation (1a) may be solved for ( t ) , to give

& (0)
t
1
(t ) ( 0 )
(1 e t /
=
+
+ 2

E11
E11


The initial conditions are, cf. Intermezzo 1,

(3)

318

Influence of the shape of the waveform


in the Burgers model

App. 1

Intermezzo 1
The Burgers model is a linear rheological model, consisting of springs and dashpots. The spring is defined by
= E
(i)
where E is the spring constant (elasticity modulus). The dashpot is defined by
= &
(ii)
where is the dashpot viscosity, and & is the strain rate. The linearity of the
Burgers model implies the following conditions:
= M + V = E 1 + 1 + 2

& = & M + &V = & E1 + &1 + & 2


= E 2 E 2 + 2 & 2

(iii)
(iv)
(v)

where M is the strain in the Maxwell element, [ E1 , 1 ], V is the strain in the


Voigt-Kelvin element [ E 2 , 2 ], E1 is the strain in the spring E1 , 1 is the strain
in the dashpot, 1 , E 2 is the strain in the spring E 2 , 2 is the strain in the dashpot, 2 , and &i are the time derivatives of i . From equation (v),
E 2 E 2
(vi)
& 2 =
2
Substitution in equation (iv) yields
1
1 E2 E 2 /
&

& =
+
+
(vii)
2
E1
1

Let us assume the stress, , is applied suddenly at t = 0, and is constant, cf. 2.1.
Then, from equation (i), it follows that the strain in the series spring at t = 0,
E1 (0) , is equal to the instantaneous elastic strain, / E1 . From equation (ii), it
follows that the strain in the series dashpot at t = 0, 1 (0) , is equal to zero, and
that the strain in the parallel dashpot at t = 0, 2 (0) , is equal to zero. Hence, the
total strain in the Burgers model at t = 0, (0) , is equal to the instantaneous elastic
strain, cf. equation (4a). Since the stress is constant, & = 0. On t = 0, E 2 (0) = 0 .
Substitution in equation (vii) yields equation (4b).
Let us assume the stress, , is sinusoidal, (t ) = sin (at ) , cf. 2.2. In this case,
the total strain at t = 0, (0) , is equal to zero, and E 2 (0) = 0 . Substitution in
equation (vii) yields
& (0) (0) a
+
=
(viii)
& (0) =
E1
E1
1

Sec. 2

Analysis

319

Intermezzo 2
Let us assume, that for t < 0 , = & = && = 0 . The integral of equation (1a) from 0
to t is
t
E1
1
( (t ) (0)) = & (t ) & (0) + [ ] ( (t ) (0)) +
E1 (&(t ) & (0) ) +
(t ) dt

1 2

(ix)
From equation (ix), the strain rate is obtained as
t

1
1
1

&
&
&
&
( (t ) (0)) + (t ) (0) + [ ] ( (t ) (0)) +
(t ) = (0)

(
t
)
dt

2
E1
1 2 0

(x)
Note, that the strain rate depends on the mathematical form of the applied stress,
(t ) , i.e. on the shape of the waveform of the applied stress.
Let us assume, that at some time, t , the applied stress is taken away, i.e. for
t , &(t ) , & (t ) , and () = 0 , then, from equation (ix),
t
E
1
(xi)
E1 (&(0) ) + 1 (t ) = & (0) [ ] (0) +
(t ) dt
2
1 2 0
The remaining, or permanent, strain is obtained as

(t ) = (0) + 2 & (0)

2
E1

& (0)

2
E1

[ ] ( 0) +

1
E11

(t ) dt

(xii)

Note, that the permanent strain differs from the creep strain. Let us define
permanent strain = creep strain delayed elastic strain

(xiii)

In equation (xii), the term in [ ] describes the delayed elastic term. The influence of
this term is limited, since it is a constant.

(0)
1
=
(4a)
$
E1
&(0)
1
1
(4b)
=
+
$
1 2
After substitution in equation (3), letting t ,
1
1
1 1
1
(t ) 1
t
t
=
=
+
+ 2 +

+
+
(5)

E1 E11
E11
1 2 E11 E1 E 2
In equation (5), ( t ) is linearly proportional to t . It is noted, that this
linear proportionality between (t ) and t for t is a property of the
Burgers model, which represents a limitation of the model when applied to
asphalt mixture, since for asphalt mixture the creep is proportional to t z ,
with z < 1. The strain rate is given by

&(t ) =
(6)
E11

320

Influence of the shape of the waveform


in the Burgers model

App. 1

In the present case of a constant applied stress, the strain rate is equal to
the creep strain rate, cf. 2.3. Hence,

&creep (t ) =
(7)
E11
Note, that this result is also obtained from equation (x) in Intermezzo 2, if
the begin-conditions, equation (4) and equation (6) are applied.
2.2 Sinusoidal applied stress
For more details, the reader is referred to Case 1 in 3. For the applied
stress,
t < 0 : (t ) = 0
(8)
t 0 : (t ) = k sin ( at )

equation (1a) can be solved for the response strain, ( t ) ,


k
+
(t ) = (0) +
E11 a

ka
k 22 a / 21 t /
(1 e t / ) +
+ 2 (0)
e
+
2 2
E
E

a
+
1
1
1

2
2

k 2 / 21
k 22 a / 21
1
+
+ 1 sin (at )
+
(9)
2 2
2 2
cos( at )
E1 2 a + 1 1 a
E1 2 a + 1

Equation (9) contains constant terms, transient terms, proportional to


e t / 2 , and sine and cosine terms. Those terms do not cause accumulation
of strain. Equation (9) does not contain a creep term, proportional to t .
Thus, if a sinusoidal stress is applied to the Burgers model, no creep is
predicted. Alternatively, this can be shown as follows. Since there was no
stress at t = 0, there is no strain at t = 0, hence
( 0) = 0
(10a)
From equation (vii) in Intermezzo 1, using (0) = 0 and & (0) = a ,
& (0) = a / E1
(10b)

Substitution in equation (xi) in Intermezzo 2 yields the permanent strain,


cos (t ) 1
(t ) =
(11)

+
E11
a
a
For any number of whole load repetitions, cos (t ) equals 1, hence
the permanent strain, (t ) , is equal to zero.
2.3 Unidirectional sinusoidal applied stress (haversine)
For more details, the reader is referred to Case 1 in 3. For the applied
stress, cf. figure 1a,
t < 0 : (t ) = 0
(12)
t 0 : (t ) = k + k sin ( at )
0
1

Sec. 2

Analysis

321

equation (1a) can be solved for the response strain, ( t ) ,


kt
k
(t ) = (0) + 0 + 1 +
E11 E11 a

k
k a
k 2 a /
+ 2 (0) 0 1 (1 e t / ) + 1 22 2 21 e t / +
E1 1
E1
E1 2 a +1

k1 2 / 21
k1 22 a / 21
1
+ 2 2
+ 1 sin (at )
+
2 2
cos ( at )
E1 2 a + 1
E1 2 a + 1 1 a

(13)

Subtracting equation (9) from equation (13),


kt
k
(14a)
equation (13) equation (9) = 0 0 2 (1 e t / )
E11 E11
Only the first term on the right hand side of equation (14a), which is
proportional to t , predicts accumulation of the strain with increasing t .
The second term on the right hand side tends to a constant value, since the
transient term proportional to e t / vanishes, if t . Hence, this term
does not cause accumulation of strain. Let us define the creep term as the
term which causes accumulation of strain,
creep (t )
k t
= 0
(14b)

E11
&creep (t )
k
= 0
(14c)

E11
where &creep (t ) is the creep strain rate. Generalising, the response, (t ) , to
an applied stress, (t ) , is found to be of the following general form:
(t ) = integration constant + creep term + transient terms + dynamic terms
(15)
2

The integration constant is equal to the instantaneous elastic deformation,


cf. equation (4a). The creep term is a term proportional to the time, t . The
transient terms are proportional to exp ( t / ) , where t is the time and
is the relaxation time. Those terms vanish if t . The dynamic terms
consist of sums of sines and cosines. Those are periodical terms. The
integration constant, the transient terms, and the dynamic terms do not
cause accumulation of strain. Only the creep term causes accumulation of
the strain. Hence, the creep strain may be defined as
creep (t ) = creep term, proportional to t
(16)
2.4 Alternating block-wave of applied stress
For more details, the reader is referred to Case 2 in 3. An alternating
block-wave can be written as
0< x < /2
1

/ 2 < x < 3 / 2
(17a)
f ( x ) = 1
1
3 / 2 < x < 2

322

Influence of the shape of the waveform


in the Burgers model

or, in the form of a Fourier series


4 cos x cos 3x cos 5x

f ( x) =

3
5
1
Accordingly, the applied stress can be written as

4 ( 1) i 1
(t ) =
cos (( 2i 1) 2 t / T ) )
i =1 2i 1

or

App. 1

(17b)

(17c)

(t ) = k 0 + k i cos (ai t )

(18a)

where
k0 = 0

(18b)

i =1

4 ( 1) i 1
(18c)
2i 1
a i = ( 2i 1) t
(18d)
For the applied stress according to equation (18), equation (1a) can be
solved for the response strain, ( t ) ,
n

[ ] k i

kt
k
i =1
+
(t ) = (0) + 0 + 2 (0) 0

E11
E11
E1

k
+ 2 (0) 0 e t / 2 +
E11

ki =

t / 2
22 ai2
k i
1 2

[
]
[
]
+

2
22 ai2 + 1
1 22 ai2 + 1
i =1 E1
n
1
1

k
sin (ai t ) cos (ai t ) +
+ i [ ] 2 + 2 22 2 ai2
2 ai +1 1
i =1 E1
2 ai

+ [ ] ( 2 ai sin ( ai t ) + cos ( ai t ) )
(19)

For the applied stress according to equation (18), k 0 is equal to zero.


Hence, the creep term vanishes. The remaining constant terms, transient
terms, and sine and cosine terms do not cause accumulation of ( t ) .
+

2.5 Unidirectional block-wave of applied stress


For more details, the reader is referred to Case 2 in 3. An arbitrary
unidirectional block-wave, see figure 2b, can be written as
0< x<
0

f ( x ) = 1
(20a)
< x< +
0
+ < x < 2

or, as a Fourier series

Sec. 2

f ( x) =

Analysis

2 sin cos x sin 2 cos 2 x sin 3 cos 3x


1
2
3

Accordingly, the applied stress can be written as


2
sin (i )

(t ) = + (1) i
cos (2i t / T )
i =1
i

or

323

(20b)

(20c)

(t ) = k 0 + k i cos (ai t )

(21a)

where
k0 = /

(21b)

i =1

2 ( 1)
sin (i )
(21c)
i
(21d)
a i = i t
For the applied stress according to equation (21), equation (1a) can be
solved for the response strain, ( t ) . The solution is the same as in the
previous case, because the mathematical form of the applied stress is
similar (compare equation (21a) an equation (18a)),
n

[ ] k i

kt
k
i =1
+
(t ) = (0) + 0 + 2 (0) 0

E11
E11
E1

k
+ 2 (0) 0 e t / 2 +
E11

ki =

t / 2
k i
22 ai2
1

[ ] 2 22
+ [ ] 2 2 2
+
e
2 ai + 1
1 2 ai + 1
i =1 E1
n
1
1

k
+ i [ ] 2 + 2 22 2 ai2
sin (ai t ) cos (ai t ) +
2 ai +1 1
i =1 E1
2 ai

+ [ ] ( 2 ai sin ( ai t ) + cos ( ai t ) )
(22)

For the applied stress according to equation (21), k 0 differs from zero.
The term in k 0 proportional to t is the only term causing ( t ) to
accumulate.
n

2.6 Half sine waveform of applied stress


For more details, the reader is referred to Case 3 in 3. The half sine
function, see figure 2c, can be written as
sin x 0 < x <
f ( x) =
(23a)
< x < 2
0
or, as a Fourier series

324

Influence of the shape of the waveform


in the Burgers model

App. 1

1
2 cos 2 x cos 4 x cos 6 x

sin x
+
+
+

35
5 7
2
1 3
Accordingly, the applied stress can be written as
1 1
2 cos(2i 2 t / T )
(t ) = + sin (2 t / T )
2
i =1 (2i 1) (2i +1)
or
f ( x) =

(23b)

(23c)

(t ) = k 0 + k10 sin (a1t ) + k i cos (ai t )

(24a)

where
k0 = 1/

(24b)

i =1

= 1/ 2
( 2)
ki =

0
1

a i = 2 i t

(24c)
1
(2i 1) ( 2i + 1)

(24d)

(24e)
Equation (24a) can be obtained as the summation of equation (12) and
equation (18a) minus k 0 of equation (18). Accordingly, the solution of
equation (1a) is the summation of the solutions according to equation (13)
and equation (19), taking care that the proper terms arising from k 0 are
included. Thus

kt
k
k1
k a
= ( 0) + 0 +
+ 2 (0) 0 1 (1 e t / 2 ) +
E11 E11 a
E11
E1

k1 2 / 21
k1 22 a / 21 t /
1
+
e
+
+

sin ( at ) +
2
2
E1 2 a + 1
E1 22 a 2 + 1

[ ]2 k i

k1 a / 21
1
i =2
+
+

cos ( at )
2
E1
E1 a + 1 1 a
n
t /
k i
22 ai2
1
[ ] 2 22
+
+ [ ] 2 2 2
e
2 ai + 1
1 2 ai + 1
i =1 E1
n
1
1

k
sin (ai t ) cos (ai t ) +
+ i [ ] 2 + 2 22 2 ai2
2 ai +1 1
i =1 E1
2 ai

+ [ ] ( 2 ai sin ( ai t ) + cos ( ai t ) )
(25)

For the applied stress according to equation (24), k 0 differs from zero.
The term in k 0 proportional to t is the only term causing ( t ) to
accumulate.

2
2
2
2

Sec. 2

Analysis

325

Table 1. Creep strain rate for different shapes of waveform of the applied stress
in the Burgers model.
shape of waveform
creep strain
rate
(t )
constant

=1
&creep =
(26a)

$
E1 1
(t )
haversine

= 1 + sin ( t )
(26b)
&creep =
$
E11
half sine
1
2 cos(2i 2 t / T )
(t ) 1 1
(26c)
= + sin(2 t / T )
&creep =
$

2
i = 1 (2i 1) (2i + 1)
E11
unidirectional (t ) 2
sin (i )


(26d)
&creep =
=
+
( 1) i
cos (2i t / T )
block-wave
$

i =1
i

E11

2.7 Influence of the shape of the waveform of the applied stress on the
creep strain rate
In the previous sections, the solutions are given for different waveforms of
applied stress. It was shown that only the creep term proportional to the
time, t , causes the strain to accumulate. This term can be written as
kt
(27a)
creep (t ) = 0
E11
The creep strain rate may be defined as
k
(27b)
&creep (t ) = 0
E11
If the constant applied stress in 2.1 is normalised, i.e. (t ) / = 1, then
k 0 is equal to . If the haversine form of applied stress in section 2.3 is
normalised, i.e. (t ) / = 1 + sin (at ) , then k 0 = k1 = k , and k is equal
to . If the half sine form of applied stress in section 2.6 is normalised,
i.e. (t ) / = 1, then k 0 is equal to / . If the unidirectional blockwave form of applied stress in section 2.5 is normalised, i.e. (t ) / = 1,
then k 0 is equal to / . The waveform shapes are shown in figure 2,
and the creep strain rates in table 1. E1 , 1 refer to the model parameters
of the Maxwell element of the Burgers model, 1 = 1 / E 1 . For the
block-wave shown in figure 2c, = / 2 . If this value of is
substituted in equation (26d), see table 1, then it is seen that the creep
strain rate is half as large that obtained for the constant applied stress, and
also half as large that obtained for the haversine waveform. If the blockwave is made up by a loading time of 0.2 s and a rest-time of 0.8 s, i.e. the
frequency is 1 Hz, then 2 / 2 = 0.2/1, i.e. = / 5 . Depending on
the loading time, the creep strain rate for the block-wave can be greater or
smaller than that for the half sine. For example, for the block-wave shown
in figure 1c, = / 2 , and the creep strain rate is greater than that for the
half sine, by a factor of / 2 . For a block-wave with a loading time of 0.2

326

Influence of the shape of the waveform


in the Burgers model

App. 1

s and a rest-time of 0.8 s, = / 5 , and the creep strain rate is smaller


than that for the half sine. Thus, table 1 shows that the creep strain rate
depends on the shape of the waveform of the applied stress.
Although the Burgers model is a simple rheological model, and, as such,
cannot predict the behaviour of a real asphalt mixture quantitatively
correctly, it is likely that also in the case of a real asphalt mixture, the
creep strain rate depends on the shape of the waveform of the applied
stress.

3 Detailed solutions of the Burgers equation


Let be given the differential equation,
1
E

E1&& + 1 & = && + [ ] & +
2
1 2

(28a)

where dotted symbols represent first order or second order time


derivatives, and
1
1
1
(28b)
[ ] =
+
+
1 21 2
1 = 1 / E1
(28c)
21 = 2 / E1
(28d)
2 = 2 / E2
(28e)
The Laplace transform of equation (28a) is
E
E 1 [ s 2 y sy (0) y ( 0)] + 1 [ sy y ( 0)] =

s 2 z sz ( 0) z ( 0) + [ ] ( sz z (0) ) +

1 2

(29a)

(29b)

or, after rearrangement,


y E1 s ( s + 1 / 2 ) y (0) E1 ( s + 1 / 2 ) y (0) E1 =
s 2 z sz ( 0) z ( 0) + [ ] ( sz z (0) ) +

1 2

where
y = L{ ( t )}
(30a)
z = L{ ( t )}
(30b)
and y( 0) and z( 0) represent the begin-values of the strain and the stress
respectively, i.e.
y( 0) = ( 0)
(31a)
z( 0) = ( 0)
(31b)
L{}
represents the Laplace transformation operator.

Sec. 3

Detailed solutions of the Burgers equation

327

Case 1
Let the applied stress be given by
( t ) = k 0 + k 1 sin (at )
(32)
then, by equation (31b),
z ( 0) = k 0
(33a)
z (0) = k 1 a
(33b)
The Laplace transform of the applied stress is given by equation (30b),
L{ ( t )} = L{k 0 + k 1 sin (at )}
(34)
k
ka
z = 0 + 2 1 2
(35)
s
s +a
Substitution of equation (33) into equation (29b) yields
y E1 s ( s + 1 / 2 ) y (0) E1 ( s + 1 / 2 ) y (0) E1 =

s2
s
k 0 s + k1a 2
k 0 s k 1a + [ ] k 0 + k 1a 2
k0 +
2
2
s +a
s +a

k 1
ka
1
+ 0
+ 1
(36a)
2
1 2 s 1 2 s + a 2
Rearranging,
ka
y ( 0)
y (0)
s
y=
+ 1
+
+
s
s ( s +1 / 2 )
E1 ( s +1 / 2 )( s 2 +a 2 )
ka
1
[ ] k1a
1
1
+
+
E1 s( s + 1 / 2 )
E1 ( s + 1 / 2 ) ( s2 + a 2 )
+

1
1
ka
+ 1
E112 s ( s + 1 / 2 ) E112 s( s + 1 / 2 ) ( s2 + a 2 )
k0

The last term on the right hand side can be transformed, using
1

2
= 2
s ( s + 1 / 2 )
s
s + 1 / 2
y=

(36b)

(37)

k a
y (0)
1
+ y (0) 1
+
s
E1 s ( s +1 / 2 )

[ ]k1a
1
k1a
s
+
2
2 +
E1 ( s + 1 / 2 ) ( s + a )
E1 ( s + 1 / 2 ) ( s2 + a 2 )
+

1
1
2
k a
+ 1 2
2 2
E112 s ( s + 1 / 2 ) E112 s
s + 1 / 2 ( s + a )
k0

(38)

The desired solution is given by the inverse Laplace transform of y ,


( t ) = y = L1 { y}
(39)
This is equal to the sum of the inverse Laplace transforms of the individual
terms of equation (38). To perform the transformation of these terms, the
following inverse transforms are needed:

328

Influence of the shape of the waveform


in the Burgers model

App. 1

1
L1{ } = 1
s
1
1
} = (1 e at )
L 1 {
s( s + a )
a
1
1 cos (at )
L1{ 2 2 } =
s( s + a )
a2
1
1
1
} = { t (1 e at )}
L 1 { 2
a
a
s (s + a)
1
1
1
b
}= 2
{e bt + sin ( at ) cos ( at )}

L 1 { 2
2
2
a
s + a s +b
a +b
1
s
b
a
}= 2
{e bt sin ( at ) cos (at )}

L 1 { 2
2
2
b
s + a s +b
a +b
Thus,

k a
y = y ( 0) + 2 y (0) 1 (1 e t / 2 ) +
E1

(40a)
(40b)
(40c)
(40d)
(40e)
(40f)

t /

[ ] k 1 a 22
1
+
sin (at ) cos(at ) +
e
2 2
E1 2 a + 1
2 a

k1 a 2
e t / 2 a sin (at ) cos( at )) +
(
2 2
E1 2 a + 1
k
k a 1 cos (at )
+ 0 t 2 (1 e t / ) + 1
+
E1 1
E11
a2

t /

k1a
22
1
+
sin (at ) cos(at )
e
2 2
E 1 1 2 a + 1
2 a

Rearranging,

k
kt
k a
k1
y = y ( 0) + 0 +
+ 2 y (0) 0 1 (1 e t / ) +
E11
E1
E11 E11 a

k /
k 2 a /
+ 1 22 2 21 e t / + 1 22 2 21 + 1 sin ( at ) +
E1 2 a + 1
E1 2 a + 1

(41a)

k1 22 a / 21
1
+
2 2
cos ( at )
E1 2 a +1 1 a
Letting k 0 = k 1 = k , then, from equation (32), $ = k ,
(t )
= 1 + sin ( at )
$
(t ) ( 0 )
1
t
=
+
+
+

E11 E11a

(41b)

(42)

Sec. 3

Detailed solutions of the Burgers equation

(0)
a
1
(1 e t /
+ 2

E1
E1 1

1 /
1
+ 2 2 2 21 + 1 sin (at )
E1 2 a + 1
E1

)+

329

1 22 a / 21 t /
e
+
E1 22 a 2 +1
2

22 a / 21
1
+
2 2
cos ( at )
2 a + 1 1 a
Letting k 0 = 0, and k1 = k , then from equation (32), $ = k ,
(t )
= sin ( at )
$
Equation (41b) reduces to

k
ka
(1 e t / ) +
y = y ( 0) +
+ 2 y (0)
E11a
E

(43)

(44)

k 2 / 21
k 22 a / 21 t /
+
e
+
+ 1 sin ( at ) +
2 2
2 2
E1 2 a + 1
E1 2 a + 1

2
1
k 2 a / 21

+
2 2
cos ( at )
E1 2 a +1 1 a
which yields
(t ) (0)
1
=
+
+

E11 a
2

(0)
a
1 22 a / 21 t /
(1 e t / ) +
+ 2

e
+
2 2

E
E

a
+
1
1
1

1 /
1 22 a / 21
1
+ 2 2 2 21 + 1 sin (at )
+
2 2
cos ( at )
E1 2 a + 1
E1 2 a + 1 1 a

Letting k 0 = k and k1 = 0, then from equation (32), $ = k ,


(t )
=1
$
Equation (41b) reduces to

kt
k
1 e t / 2
y = y ( 0) +
+ 2 y (0)
E11
E11

which yields
(0)
(t ) ( 0 )
1
t
1 e t / 2
=
+
+ 2

E11
E11

2

(45)

(46)

(47)

(48)

(49)

Case 2
Let the applied stress be given by
( t ) = k 0 + k 1 cos( pt ) + k 2 cos( qt )
(50)
then, by equation (31b),
z ( 0) = k0 + k1 + k2
(51a)
z ( 0) = 0
(51b)
The Laplace transform of the applied stress is given by equation (30b),

Influence of the shape of the waveform


in the Burgers model

330

L{ ( t )} = L{k 0 } + L{k 1 cos( pt )} + L{k 2 cos( qt )}


k
k s
k s
z = 0 + 2 1 2 + 2 2 2
s
s + p
s + q
Substitution of equation (53) into equation (29b) yields
y E1 s ( s + 1 / 2 ) y (0) E1 ( s + 1 / 2 ) y (0) E1 =
k 0 s + k1

App. 1

(52)
(53)

s3
s3
k
+
k 0 s k1 s k 2 s +
2
s2 + q 2
s2 + p2

s2
s2
+
[

]
k
+
2
s2 + p2
s2 + q 2
k2
k0 1
k1
s
s
[ ]k 0 [ ] k 1 [ ] k 2 +
+
+
1 2 s 1 2 s 2 + p 2
1 2 s 2 + q 2
(54a)
Rearranging,
[ ]( k1 + k 2 )
1
y (0) (k1 + k 2 ) / E1

+
+ y (0)
+
y=
s
s + 1 / 2
E1

s(s + 1 / 2 )
k0
1
+
+
2
E 1 1 2 s ( s + 1 / 2 )
+ [ ] k 0 + [ ] k 1

k1
k2
+
2
+
E 1 1 2 ( s + p 2 ) ( s + 1 / 2 )
( s 2 + q 2 ) ( s + 1 / 2 )

k1 s
k2 s
[ ]
+
2
+
E1 (s + p 2 ) (s + 1 / 2 )
(s 2 + q 2 ) (s + 1 / 2 )

k1 s 2
k2 s2
1
(54b)
+
+
2

E1 ( s + p 2 ) ( s + 1 / 2 )
(s 2 + q 2 ) (s + 1 / 2 )
The desired solution is given by the inverse Laplace transform of y ,
( t ) = y = L 1 { y }
(55)
This is equal to the sum of the inverse Laplace transforms of the individual
terms of equation (54b). To perform the transformation of these terms, the
following inverse transforms are needed:
1
L1{ } = 1
(56a)
s
1
L 1 {
} = e at
(56b)
s + a
1
1
L 1 {
} = (1 e at )
(56c)
s( s + a )
a
1
1
1
L 1 { 2
} = { t (1 e at )}
(56d)
a
a
s (s + a)

Sec. 3

Detailed solutions of the Burgers equation

s + a2 s
s

L 1 { 2
s + a2 s
L 1 {

1
+
1
+

331

b
1
{e bt + sin ( at ) cos ( at )}
(56e)
2
a
b
a +b
b
a
}= 2
{e bt sin ( at ) cos (at )} (56f)
2
b
b
a +b
}=

s2
1
a2
b
bt

}
=
e

{e bt + sin ( at ) cos( at )}
(56g)
2
2
2
2
s +a s+b
a +b
a
Thus,

k
k t k +k
[ ] (k1 + k 2 )
(1 e t / ) +
y = y (0) + 0 1 2 e t / + 2 y (0) 0
E11
E1
E11
E1

L1{

t /

1
+
sin ( pt ) cos ( pt ) +
e
2 p
E 1 1 2 p + 1

t /

1
+
sin
(
)

cos
(
)
e
qt
qt

+
2 q
E 1 1 2 22 q 2 + 1

22

k1

2
2

k2

22

k 1 [ ] 2
e t / 2 2 p sin ( pt ) cos ( pt ) ) +
(
2 2
E 1 2 p + 1
k [ ] 2
e t / 2 2 q sin ( qt ) cos ( qt )) +
2
(
2 2
E1 2 q + 1

k1 t / 2
22 p 2 t / 2
1
e
+
2 2
+
sin ( pt ) cos ( pt ) +
e
E1
2 p + 1
2 p

k
2 q 2 t / 2
1
e
+ 2 e t / 2 2 22
+
sin ( qt ) cos (qt )
E1
2 q + 1
2 q

Rearranging,

k
kt
[ ] ( k1 + k 2 )
+
y = y (0) + 0 + 2 y (0) 0

E
E
E11

1 1
1

k
2 y (0) 0 e t / 2 +
E11

t / 2
k
2 p 2
1
+ 1 [ ] 2 2 2 2 [ ] 2 22
+
e
E1
2 p +1
1 2 p + 1
t / 2
k 2
22 q 2
1

[ ] 2 22
+
+
[ ] 2 2 2
e
E1
2 q + 1
1 2 q + 1
1

k
2 1
+ 1 2 22
sin ( pt ) cos ( pt ) +
2 p
E1 2 p + 1 1
2 p

+ [ ] ( 2 p sin ( pt ) + cos ( pt ) )

(57a)

332

Influence of the shape of the waveform


in the Burgers model

2
k2
2 2
E1 2 q +1

App. 1

2 1
sin (qt ) cos (qt ) +
2 q
2 q

(57b)
+ [ ] ( 2 q sin ( qt ) + cos ( qt ) )

Equation (57b) shows that for each term in k i , i = 1, 2 , ..., in equation


(50), there is a transient term proportional to e t / 2 , and a dynamic term
which is a sum of sine and cosine terms. Generalisation yields, for the
applied stress
n

(t ) = k 0 + k i cos ( ai t )

(58)

i =1

and for the response strain


n

[ ] k i

kt
k
i =1
(t ) = (0) + 0 + 2 (0) 0

E11
E11
E1

k
+ 2 (0) 0 e t / 2 +
E11

t / 2
k i
22 ai2
1

[ ] 2 22
[
]
+

2
2 2
2 ai + 1
1 2 ai + 1
i =1 E1
n
1
1

k
sin (ai t ) cos (ai t ) +
+ i [ ] 2 + 2 22 2 ai2
2 ai +1 1
i =1 E1
2 ai

+ [ ] ( 2 ai sin ( ai t ) + cos ( ai t ) )

(59)

Case 3
Let the applied stress be given by
(t ) = k 0 + k1 sin (at ) + k 2 cos ( pt ) + k 3 cos (qt )
(60)
then, by equation (31b),
z ( 0) = k0 + k2 + k3
(61a)
z (0) = k1 a
(61b)
The Laplace transform of the applied stress is given by equation (30b),
L{ (t )} = L{k 0 } + L{k1 sin ( at )} + L{k 2 cos ( pt )} + L{k 3 cos ( qt )}
(62)
k
k s
ka
k s
z = 0 + 2 1 2 + 2 2 2 + 23 2
(63)
s
s +a
s +p
s +q
Substitution of equation (63) into equation (29b) yields
y E1 s ( s + 1 / 2 ) y (0) E1 ( s + 1 / 2 ) y (0) E1 =
k 0 s + k1 a

s2
s3
s3
+
k
+
k
k 0 s k 2 s k 3 s k1 a +
2 2
3 2
s2 + a2
s + p2
s + q2

Sec. 3

Detailed solutions of the Burgers equation

333

s
s2
s2
+
[

]
k
+
[

]
k
+
2 2
3
s2 + a2
s + p2
s2 + q2
[ ]k 0 [ ] k 2 [ ] k 3 +
k 1
k
k
k
s
s
1
+ 0
+ 1 2
+ 2 2
+ 3 2
(64)
2
12 s 12 s + 1 12 s + p
12 s + q 2
Equation (64) can be obtained as the summation of equation (38) and
equation (54b) minus the terms k 0 in equation (54b) which arise from k 0
in equation (50). Accordingly, the solution of equation (64) is the
summation of the solutions of equation (41b) and equation (57b), minus
the terms in k 0 in equation (57b) arising from k 0 in equation (50). Thus

kt
k
k1
k a
= ( 0) + 0 +
+ 2 (0) 0 1 (1 e t / 2 ) +
E11 E11 a
E11
E1

+ [ ]k 0 + [ ]k1 a

k /
k1 22 a / 21 t / 2
e
+ 1 22 2 21 + 1 sin ( at ) +
2 2
E1 2 a + 1
E1 2 a + 1

[ ]2 k i

k1 22 a / 21
1
i =2

+
+
2 2
cos ( at )
E1
E1 2 a + 1 1 a
n
t / 2
k i
22 ai2
1

[ ] 2 22
+
+ [ ] 2 2 2
e
2 ai + 1
1 2 ai + 1
i =1 E1
n
1
1

k
sin (ai t ) cos (ai t ) +
+ i [ ] 2 + 2 22 2 ai2
2 ai +1 1
i =1 E1
2 ai

+ [ ] ( 2 ai sin ( ai t ) + cos ( ai t ) )
(65)

4 Conclusion
For a Burgers material, the response strain, ( t ) , to an applied stress,
( t ) , depends on the shape of the waveform of the applied stress. If the
creep term of ( t ) is defined as the term which causes accumulated creep
strain as function of the time, creep (t ) , then the creep strain rate, &creep ,
also depends on the shape of the waveform of the applied stress.

334

Influence of the shape of the waveform


in the Burgers model

App. 1

335

Appendix 2: Influence of the friction reduction system


1 Introduction
The tests described in chapter 3 were performed in 1994 using the normal
friction reduction method. This method was developed for the static creep
test in the early nineteen seventies when the creep test was developed for
asphalt mixture (EMPA 1977, Jongeneel et al. 1985), and has been in use
in the Netherlands and other European countries since then. Bolk (1980)
described different friction reduction systems and other details concerning
the use of the creep test on asphalt mixture in the Netherlands. The
method had proven satisfactory for testing asphalt concrete mixtures.
There was some barrelling in the static creep test, cf. Ch. 3, figure 47, but
this was considered acceptable. During the investigation of 1994 it was
found that barrelling was more serious in the dynamic creep test in the
absence of confinement, cf. Ch. 3, figure 47. Molenaar et al. (1995)
suggested applying a confinement pressure, expecting that would improve
the state of stress of the specimen and suppress barrelling. This was
confirmed by the investigation on porous asphalt described in Ch.3, 8.2
(Molenaar and Molenaar 2000), which showed no significant barrelling
and no significant specimen volume increase, cf. Ch. 3, figure 56.
Therefore, it was assumed that the friction reduction method was
adequate.
According to an investigation conducted by the United States National
Council of Highway Research (NCHRP project 9-19), the greatest friction
reduction is achieved using a system consisting a layer of silicone grease
between two membranes of the membrane material used to water proof
seal the specimen for the triaxial test. Also Erkens investigated the friction
reduction (2002). Erkens recommended using a thin plastic foil, Luflexen,
thickness 40 to 75 m, wetted on two sides with liquid soap (green
soap).

2 RHEI investigation into friction reduction


Based on these experiences a small investigation was performed at the
RHEI to compare the different friction reduction systems. The following
friction reduction systems were used:
. system 1: talcum powder and glycerine, strewed with graphite powder;
. system 2: three layer system: green soap, Luflexen membrane, 40-75
m thickness, green soap;
. system 3: three layer system: latex rubber membrane silicon grease
latex rubber membrane.

336

Influence of the friction reduction system

App. 2

70000

creep strain (m/m)

60000
50000
40000
30000
no friction reduction
20000

system 1
system 2

10000

system 3
0
0

2000

4000

6000

8000

10000

number of load repetitions (N)

Figure 1. Creep using different friction reduction systems. Specimen:


50 x 100 mm height x diameter. Axial applied stress: 0.45 MPa, radial
applied stress: 0.05 MPa. Temperature: 50C. Shape of waveform of
applied stress: block-wave; loading time: 0.2 s, rest-time: 0.8 s. Load
control: hydraulic. Material: crushed gravel asphalt concrete 0/22.
Each curve represents one test (one specimen).

The test material was a crushed gravel asphalt concrete 0/22 mixture. The
specimen used was 50 x 100 mm height x diameter. The applied axial
stress was 0.45 MPa, the radial stress was 0.05 MPa. The test temperature
was 50C. The shape of the waveform was a block-wave with a loadingtime of 0.2 s and a rest-time of 0.8 s. The results are shown in figure 1.
Each curve represents a single test (one specimen).
The curve with the square markers was obtained with no friction
reduction; the curve with the lozenge-shaped markers was obtained with
the traditional friction reduction system, system 1; the curve with the
triangular markers was obtained with the green soap system, system 2; the
curve with the cross-shaped markers was obtained with the latex rubber
membrane system, system 3.
Figure 2 shows the results of a subsequent series of tests on DAC 0/16
with 6.0% 80/100 bitumen, slab compacted, using friction reduction
system 3. The tests were performed under the same conditions as those of
figure 1. Each curve represents a single test (one specimen). The results in
figure 2 confirm those obtained with system 3 in figure 1.

3 Discussion
The results in figure 1 and figure 2 indicate that the traditional creep
reduction system, system 1, is hardly effective, whereas the recently
developed systems, system 2 and system 3, are effective. The results show
that the creep depends strongly on the friction reduction.

Discussion

creep compliance J(t) (MPa^-1)

Sec. 3

337

0.35
0.3
0.25
0.2

test #1
test #2
test #3
test #4

0.15
0.1
0.05
0
0

100

200

300

400

500

creep compliance J(t) (MPa^-1)

time (s)

0,1

test #1
test #2
test #3
test #4
0,01
1

10

100

1000

time (s)

Figure 2. Creep of DAC 0/16 with 6.0 % bitumen, using friction


reduction system #3. Specimen: 50 x 100 mm height x diameter. Axial
applied stress: 0.45 MPa, radial applied stress: 0.05 MPa. Temperature:
50C. Shape of waveform of applied stress: block-wave; loading time:
0.2 s, rest-time: 0.8 s. Load control: hydraulic. Material: crushed
gravel asphalt concrete 0/22. Each curve represents one test (one
specimen). Top (a): creep compliance J (t ) versus time t on linear
scales. Bottom (b): creep compliance J (t ) versus time t on log-log scale.

Interpreting the results in figures 1 and 2, the following two assumptions


could be made:
1 it seems that the friction between the specimen and the loading plate
holds the specimen from deforming in radial directions at the loading
plate;
2 it seems that the friction between the specimen and the loading plate
does not have the capacity to hold the specimen from deforming in
radial directions at a distance from the loading plate.

338

Influence of the friction reduction system

App. 2

Further interpreting the first assumption, the friction between the


specimen and the loading plate seems to be equivalent to a confinement
stress, however applied locally, i.e. at the circumference of the specimen
at the loading plate.
Further interpreting the second assumption, it seems that barrelling can be
explained as the result of an unevenly distributed confinement.
Continuing this reasoning, it is expected that the friction between the
specimen and the loading plate will no longer influence the creep if the
confinement pressure is (sufficiently) greater than the friction per unit
friction area. Then, in principle, quantitatively similar results as in the
present study, i.e. Ch. 3, are expected regardless of the friction reduction
system.
In other words, the new friction reduction systems, systems 2 and 3, are
expected to be effective especially at low confinement. This is in
agreement with the experiments, since the confinement, 0.05 MPa, was
rather low.
The purpose of the friction reduction system is to obtain the best feasible
approximation of a homogeneous stress state of the specimen. It is
important to have a homogeneous stress state, as the stress strain
behaviour of an asphalt mixture is stress dependent, and the stress state of
the material may be inhomogeneous, cf. Ch. 5, figure 7, in order to obtain
properties that characterise the material behaviour.
However, creep properties of asphalt mixture are not expected to be
independent of the specimen geometry or, if the shape is constant, of the
specimen size since they will depend on the specimen height, cf. Ch. 3,
9, even when the stress state of the specimen is homogeneous.
Since it appears thus that the dependence of the creep properties on the
specimen geometry (shape and size) is itself an intrinsic property of the
material, it can be eliminated only by standardisation of the test, i.e. the
specimen geometry, and the testing conditions.

339

Appendix 3: List of creep models


Table 1. Dense asphalt concrete 0/16. Static creep, cf. Ch. 3, 6.1.
J (t ) = J 1 + z ln t , J (t ) in MPa-1.
applied creep model
Mixture
temp.
stress
J (t ) = J 1 + z ln t
(MPa)
(C)
J1
z
DAC 0/16
40
0.2
nine tests (specimens)
0.0061
0.0405
0.0075
0.0190
0.0059
0.0246
0.0050
0.0273
0.0062
0.0140
0.0058
0.0099
0.0037
0.0188
0.0051
0.0075
DAC 0/16
40
0.2
average J 1 : 0.0202
stdev. J 1 : 0.0106
average z : 0.0057
stedev. z : 0.0011
DAC 0/16
40
0.2
J (t ) = 0.0063 ln t + 0.0161
large x : y = 0.0033 x 0.0843

R2

0.986
0.984
0.984
0.987
0.986
0.991
0.985
0.993

0.979

Table 2. DAC 0/16, polymer modified. Static creep, cf. Ch. 3, 6.2.
Different temperatures and applied stresses.
J (t ) = J 1 + z ln t , J (t ) in MPa-1.
J1
z
DAC 0/16, elastomer modified
40C/0.1 MPa
0.0213
0.0079
50C/0.1 MPa
0.0382
0.0060
40C/0.2 MPa
0.0209
0.0042
50C/0.2 MPa
0.0311
0.0040
DAC 0/16, plastomer modified
40C/0.1 MPa
0.0009
0.0164
50C/0.1 MPa
0.0398
0.0137
40C/0.2 MPa
0.0191
0.0093
50C/0.2 MPa
0.0293
0.0076

340

List of creep models

Table 3 (Part I). DAC 0/16. Dynamic creep, cf. Ch. 3, 7.1.1.
Block-wave load with constant rest-time, 1.8 s.
Different temperatures and applied stresses

J (t ) = J 10 (log t ) z , J (t ) in MPa-1.
J 10
z
Loading-time/rest-time
40C/0.1 MPa
2.2579
0.0029
600 s/1.8 s
2.7115
0.0027
10 s/1.8 s
2.9543
0.0028
1 s/1.8 s
2.3369
0.0115
0.2 s/1.8 s
3.2100
0.0024
0.05 s/1.8 s
40C/0.2 MPa
600 s/1.8 s
0.00002
5.0293
10 s/1.8 s
0.0013
3.2306
1 s/1.8 s
0.0054
2.5592
0.2 s/1.8 s
0.0066
2.2940
0.05 s/1.8 s
0.0032
2.6239
Table 3 (Part II). DAC 0/16. Dynamic creep, cf. Ch. 3, 7.1.1.
Block-wave load with constant rest-time, 1.8 s.
Different temperatures and applied stresses

J (t ) = J 10 (log t ) z , J (t ) in MPa-1.
J 10
z
Loading-time/rest-time
50C/0.1 MPa
600 s/1.8 s
0.0002
4.0136
10 s/1.8 s
0.0050
2.9456
0.0149
2.3447
1 s/1.8 s
0.0226
2.0893
0.2 s/1.8 s
0.05 s/1.8 s
0.0047
3.0735
50C/0.2 MPa
600 s/1.8 s
0.00004
5.3661
10 s/1.8 s
0.0017
3.4764
1 s/1.8 s
0.0109
2.0719
0.2 s/1.8 s
0.0126
2.3001
0.05 s/1.8 s
0.0032
2.5520
Table 4. DAC 0/16. Dynamic creep, cf.. Ch. 3, 7.1.2.
Block-wave load with constant loading-time, 0.2 s.

J (t ) = J 10 (log t ) z , J (t ) in MPa-1.
J 10
z
Loading-time/rest-time
40C/0.2 MPa
0.2 s/0.2 s
0.0262
1.3171
0.2 s/0.5 s
0.0191
1.4775
0.2 s/1.0 s
0.0173
1.7211
0.2 s/1.8 s
0.0063
2.3530
0.2 s/5.0 s
0.0039
2.6339

App. 3

App. 3

List of creep models

Table 5 (Part I). DAC 0/16, elastomer modified, cf. Ch. 3, 7.2.1.
Dynamic creep, block-wave loading with constant rest-time, 1.8 s.
Different temperatures and applied stresses. J (t ) = J 10 (log t ) z ,
J (t ) in MPa-1.
J 10
z
Loading-time/rest-time
40C/0,1 MPa
0.2 s/1.8 s
0.0158
1.2988
1.0 s/1.8 s
0.0390
0.9715
40C/0,2 MPa
0.2 s/1.8 s
0.0152
1.1163
1.0 s/1.8 s
0.0290
0.8432
50C/0,1 MPa
0.2 s/1.8 s
0.0300
0.9621
1.0 s/1.8 s
421
0.7498
50C/0,2 MPa
0.2 s/1.8 s
0.0286
0.8750
1.0 s/1.8 s
0.0323
0.7725
Table 5 (Part II). DAC 0/16, plastomer modified, cf. Ch. 3, 7.2.1.
Dynamic creep, block-wave loading with constant rest-time, 1.8 s.
Different temperatures and applied stresses. J (t ) = J 10 (log t ) z ,
J (t ) in MPa-1.
J 10
z
Loading-time/rest-time
40C/0,1 MPa
0.2 s/1.8 s
0.0065
2.3385
1.0 s/1.8 s
0.0191
1.6306
40C/0,2 MPa
0.2 s/1.8 s
0.0102
1.5998
1.0 s/1.8 s
0.0297
1.1104
50C/0,1 MPa
0.2 s/1.8 s
0.0186
1.7122
1.0 s/1.8 s
0.0341
1.3382
50C/0,2 MPa
0.2 s/1.8 s
0.0163
1.4653
1.0 s/1.8 s
0.0480
0.9015

341

342

List of creep models

Table 6. DAC 0/16. Dynamic creep, cf. Ch. 3, 7.3.1.


Sinusoidal load. Different temperatures, frequencies, and
applied stresses. J (t ) = J 10 (log t ) z , J (t ) in MPa-1.
J 10
z
frequency (Hz)
40C/0.1 MPa
0.1
0.0513
1.1837
1.0
0.0187
1.4184
10
0.0072
1.7473
40C/0.2 MPa
0.1
0.0276
1.3008
1.0
0.0128
1.6693
10
0.0063
1.6383
50C/0.1 MPa
0.1
0.0293
1.6238
1.0
0.0384
1.2841
10
0.0168
1.6168
50C/0.2 MPa
0.1
0.0345
1.6121
1.0
0.0222
1.4044
10
0.0122
1.5314

App. 3

App. 3

List of creep models

343

Table 7 (Part I). DAC 0/16, elastomer modified, cf. Ch. 3, 7.4.1.
Dynamic creep, sinusoidal load. Different temperatures, frequencies,
and applied stresses. J (t ) = J 10 (log t ) z , J (t ) in MPa-1.
J 10
z
frequency (Hz)
40C/0.1 MPa
0.1
0.0501
0.5564
1.0
0.0335
0.8645
10
0.0108
1.4259
40C/0.2 MPa
0.1
0.0404
0.5066
1.0
0.0274
0.6744
10
0.0107
1.3009
50C/0.1 MPa
0.1
0.0599
0.5111
1.0
0.0558
0.656
10
0.0266
1.0586
50C/0.2 MPa
0.1
0.0545
0.4542
1.0
0.0437
0.5672
10
0.0236
0.8089
Table 7 (Part II). DAC 0/16, plastomer modified, cf. Ch. 3, 7.4.1.
Dynamic creep, sinusoidal load. Different temperatures, frequencies,
and applied stresses. J (t ) = J 10 (log t ) z , J (t ) in MPa-1.
J 10
z
frequency (Hz)
40C/0.1 MPa
0.1
0.0487
0.8017
1.0
0.0201
1.5771
10
0.0074
1.9969
40C/0.2 MPa
0.1
0.0575
1.6493
1.0
0.0196
1.0840
10
0.0081
1.7933
50C/0.1 MPa
0.1
0.0884
1.7697
1.0
0.0429
1.1247
10
0.0126
1.6083
50C/0.2 MPa
0.1
1.0
0.0266
1.1206
10
0.0153
1.5123

344

List of creep models

App. 3

Table 8. DAC 0/16. Dynamic creep in the presence of confinement,


cf. Ch. 3, 8.1. Block-wave loading, loading-time: 0.2 s, rest-time: 0.8 s.
Temperature: 50C. Specimen: 200 x 100 mm height x diameter.
Pneumatic load control.
1 / 3
creep model
q/ p
R2
J (t ) = 0.0032 t 0.4255
0.6/0.05
2.36
99.9
0.3/0.03

J (t ) = 0.0040 t 0.3673

2.25

99.7

0.6/0.1

J (t ) = 0.0029 t

0.3648

1.88

99.4

0.75/0.15

J (t ) = 0.0020 t 0.3750

1.71

99.7

0.6/0.15

J (t ) = 0.0020 t

0.3363

1.5

99.6

0.9/0.3

J (t ) = 0.0010 t 0.3192

1.2

99.1

0.6/0.25

J (t ) = 0.0010 t 0.3040
J (t ) = 0.0001 + 0.0004 ln t

0.95

99.4

0.43

99.1

0.3/0.2

App. 3

List of creep models

Table 9. DAC 0/16. Dynamic creep in the presence of confinement,


cf. Ch. 3, 8.2. Block-wave loading, loading-time: 0.2 s, rest-time: 0.8 s.
Temperature: 50C. Specimen: 60 x 100 mm height x diameter.
Hydraulic load control.
1 / 3
creep model
q/ p
R2
0.10/0
J (t ) = 0.0841t 0.2171
3
0.978
0.30/0.1
J (t ) = 0.0041+ 0.0054 ln t
1.2
0.993
0.50/0.1
1.72
J (t ) = 0.0077 + 0.0072 ln t
0.997
J (t ) = 0.0155 + 0.0085 ln t
0.70/0.1
2
0.999
J (t ) = 0.0098 + 0.0097 ln t
0.90/0.1
2.18
0.998
0.30/0.2
J (t ) = 0.0006 + 0.0006 ln t
0.43
0.990
J (t ) = 0.0040 + 0.0032 ln t
0.50/0.2
1
0.979
J (t ) = 0.0049 + 0.0045 ln t
0.70/0.2
1.36
0.999
J (t ) = 0.0055 + 0.0063ln t
0.90/0.2
1.63
1.000
0.40/0.3
J (t ) = 0.0057 + 0.0004 ln t
0.3
0.996
0.50/0.3
0.54
J (t ) = 0.0038 + 0.0014 ln t
0.973
J (t ) = 0.0064 + 0.0015 ln t
0.60/0.3
0.75
0.960
J (t ) = 0.0051+ 0.0025 ln t
0.70/0.3
0.93
0.981
J (t ) = 0.0040 + 0.0037 ln t
0.80/0.3
1.07
0.978
J (t ) = 0.0011+ 0.0036 ln t
0.90/0.3
1.2
0.995
Table 10. DAC 0/16. Dynamic creep in the presence of confinement,
cf. Ch. 3, 8.2. Sinusoidal loading, 1 Hz. Temperature: 50C. Specimen:
60 x 100 mm height x diameter. Hydraulic load control.
1 / 3
creep model
q/ p
R2
0.10/0
J (t ) = 0.1024 t 0.1939
3
0.989
0.30/0.1
J (t ) = 0.0186 + 0.0041ln t
1.2
0.997
J (t ) = 0.0186 + 0.0041ln t
0.50/0.1
1.72
0.997
J (t ) = 0.0184 + 0.0056 ln t
0.70/0.1
2
1.000
J (t ) = 0.0259 + 0.0063ln t
0.90/0.1
2.18
0.999
0.30/0.2
J (t ) = 0.0111+ 0.0005 ln t
0.43
0.941
0.50/0.2
1
J (t ) = 0.0111+ 0.0015 ln t
0.998
J (t ) = 0.0203 + 0.0028 ln t
0.70/0.2
1.36
0.998
J (t ) = 0.0155 + 0.0032 ln t
0.90/0.2
1.63
0.995
0.40/0.3
0.3
J (t ) = 0.0080 + 0.0009 ln t
0.50/0.3
0.54
0.966
J (t ) = 0.0117 + 0.0011ln t
0.70/0.3
0.93
0.993
0.90/0.3
1.2
-

345

346

List of creep models

App. 3

Table 11. PA 0/16. Dynamic creep in the presence of confinement,


cf. Ch. 3, 8.2. Block-wave loading, loading-time: 0.2 s, rest-time: 0.8 s.
Temperature: 50C. Specimen: 60 x 100 mm height x diameter.
Hydraulic load control.
1 / 3
creep model
q/ p
R2
0.10/0
J (t ) = 0.0448 x 0.6611
0.985
3
0.30/0.1
0.50/0.1
0.70/0.1
0.90/0.1

J (t ) = 0.0030 + 0.0037 ln t
J (t ) = 0.0154 + 0.0048 ln t
J (t ) = 0.0153 + 0.0064 ln t
J (t ) = 0.0453 + 0.0070 ln t

1.2
1.72
2
2.18

0.949
0.999
0.995
0.979

0.30/0.2
0.50/0.2
0.70/0.2
0.90/0.2

J (t ) = 0.0111+ 0.0010 ln t
J (t ) = 0.0026 + 0.0018 ln t
J (t ) = 0.0082 + 0.0027 ln t
J (t ) = 0.0208 + 0.0026 ln t

0.43
1
1.36
1.63

0.980
0.999
0.995
0.998

0.40/0.3
0.50/0.3
0.70/03
0.90/0.3

J (t ) = 0.0072 + 0.0001ln t
J (t ) = 0.0103 + 0.0006 ln t
J (t ) = 0.0140 + 0.0014 ln t
J (t ) = 0.0096 + 0.0015 ln t

0.3
0.54
0.93
1.2

0.812
0.988
0.999
0.965

Table 12. PA 0/16. Dynamic creep in the presence of confinement,


cf. Ch. 3, 8.2. Sinusoidal loading, 1 Hz. Temperature: 50C. Specimen:
60 x 100 mm height x diameter. Hydraulic load control.
1 / 3
creep model
q/ p
R2
0.10/0
J (t ) = 0.0903t 0.3755
0.997
3
0.30/0.1
0.50/0.1
0.70/0.1
0.90/0.1

J (t ) = 0.0157 + 0.0021ln t
J (t ) = 0.0246 + 0.0027 ln t
J (t ) = 0.0233 + 0.0033ln t
J (t ) = 0.0148 + 0.0065 ln t

1.2
1.72
2
2.18

0.964
1.000
0.998
0.917

0.30/0.2
0.50/0.2
0.70/0.2
0.90/0.2

J (t ) = 0.0074 + 0.0004 ln t
J (t ) = 0.0122 + 0.0011ln t
J (t ) = 0.0134 + 0.0017 ln t
J (t ) = 0.0141+ 0.0023ln t

0.43
1
1.36
1.63

0.990
0.998
0.998
0.999

0.40/0.3
0.50/0.3
0.70/0.3
0.90/0.3

J (t ) = 0.0072 + 0.0004 ln t
J (t ) = 0.0097 + 0.0005 ln t
J (t ) = 0.0123 + 0.0011ln t
J (t ) = 0.0095 + 0.0013ln t

0.3
0.54
0.93
1.2

0.855
0.980
0.998
0.999

347

Appendix 4: The material model of the FEMMASSE finite


element model of a heterogeneous viscoelastic-viscoplastic
creep specimen
1 FEMMASSE finite element code
Finite element computations of the creep of a heterogeneous specimen
were conducted, using the FEMMASSE finite element code owned by
Intron (Verburg et al. 1995).

2 The aggregate grains


The material was assumed to be a two-dimensional heterogeneous
granular material containing air voids. It is shown in figure 1. This
structure was created artificially, by hand. The air voids were created by
elimination of elements from the finite element mesh. The mineral
aggregates were assumed to be infinitely rigid, which means that they can
only translate or rotate. Note, that this artificial structure differs from a
real dense asphalt concrete structure. This can be explained as follows. An
asphalt mixture can be thought to constitute a granular phase embedded in
a matrix phase, called mastic. The granular phase can be a mineral
aggregate skeleton, or can be coarse mineral aggregate grains floating in
the matrix. The matrix also contains small mineral aggregate particles, of
2 mm and smaller, i.e. sand and filler. In the case of a skeleton asphalt
mixture, these particles may also be a part of the skeleton, but are
considered to be a part of the binder (mastic). In creating the artificial
structure of figure 1, it was a question how many of these small particles
had to be introduced in the structure as part of the granular phase. Thus, in
principle, the model described in the following can be extended to three
dimensions, but a difficulty is to define a realistic asphalt mixture
structure.
In a preliminary version of the model, it was assumed that the aggregate
grains were perfectly linearly elastic. From preliminary analyses, it was
concluded that the stiffness of the coarse aggregate grains is very large
with respect to the stiffness of the bituminous matrix. This large
difference caused numerical instabilities in the solution of the boundary
value problem. To solve this problem, the aggregate grains were assumed
to be infinitely rigid, and allowed only to translate or rotate. The
numerical modelling of this assumption was realised by imposing
kinematic relations between the degrees of freedom of the nodes lying on
an aggregate grain contour and the three basic degrees of freedom of the
particular grain. The imposed kinematic relations are:

348

Finite element material model of creep specimen

Figure 1. Two dimensional artificial


granular structure representing an
asphalt mixture.

App. 4

Figure 2. Schematic representation of


the viscoelastoplastic model.

ui uk k xi
(1)
=

vi v k + k yi
where ui and vi are the degrees of freedom of node i , and uk , v k , and
k are the degrees of freedom of grain k . This is illustrated in figure 1.

3 The bituminous matrix


The bituminous matrix was modelled by means of the generalised
viscoelasto-viscoplastic rheological model, shown in figure 2. The total
strain is divided in a viscoplastic part and a viscoelastic part. The plastic
sliders can be activated and deactivated by means of a yield function, f ,
which is related to the stress state. The following stress strain relationships
were used:
t + t = [ I + t A P t + t ] 1 [ A (1 ) t A P t t + ~ t ]
(2a)
= ve + vp

(2b)

vp = t P t + t t + t + (1 ) t P t t
where

(2c)

A=

( D

i =1

~ t =

+ i t N i )

(D
i =1

(2d)

+ i t N i )

1 ei

1
1
Di =
Ei

(D

(1 i ) t N i ) it

ei

ei
2(1 + ei )
0

1
0
0

(2e)

(2f)

Sec. 3

1 vi

1
1
Ni =
i

vi

vi
2(1 + vi )
0

1
1
1
+

6 I2
3 I2
1

+
I 2 + I 1 k1
3 I2

P=

I 1 = xx + yy + zz

349

The bituminous matrix

0
0

(2g)

1
+
3 I2

1
6 I2
1

6 I2

0
0
1
I2

(2h)

(2i)

1 2
(2j)
+ yy2 + zz2 xx yy yy zz zz xx + xy2
3 xx
where represents the stress, t is the time, t is the time increment, I
is the unit matrix, is a time integration constant, which determines the
mode of time integration1, is the total strain, ve is the total
viscoelastic strain, vp is the total viscoplastic strain, i is the i th unit of
the rheological model, n is the number of units in the rheological model,
E i is the elasticity modulus of unit i , ei is Poissons elastic ratio for unit
i , i is the viscosity of unit i , vi is Poissons viscous ratio for unit i , I 1
is the first stress invariant, I 2 is the second stress invariant, and where k1 ,
k 2 , and are material constants. The yield function, f , is defined by
f g
& vp =
(3)
I2 =


f = I 2 + I 1 k1

(4a)

I 2 + I1 k 2

(4b)

g=

It was assumed that the nonlinear stress strain behaviour can be explained
by the displacement in mainly horizontal directions of bituminous matrix
in more or less horizontal layers between aggregate particles. To model
this, large strains were permitted in the model, by use of the midpoint
strain approach (Roelfstra 1989).

For i = 0 , the time integration is explicit, and for i = 1 , the time integration
is implicit. In the present model, i was assumed constant, and equal to 1/2.

350

Finite element material model of creep specimen

App. 4

351

Appendix 5 The Asphalt Concrete Response (ACRe) Model


1 The material model
The material model, ACRe, was developed by Scarpas et al. (1997) and
Scarpas and Blaauwendraad (1998), and is implemented in the finite
element system called INSAP. The model consists of a combination of the
theory of dynamic viscoplasticity and the classical technique of modelling
smeared crack-growth. This permits a realistic description of the stress
strain behaviour in compression and in tension (e.g. de Borst et al. 1998).
compressive
volumetric
deformation

(a)
tensile
volumetric
deformation

deviatoric
deformation

(b)

Figure 1. Schematic representation of different


types of damage.

1.1 Definition of damage


In the literature, different definitions of damage can be found. In this
investigation, the damage is defined as the magnitude of the total
permanent (irreversible) volumetric strain of the material. Mathematically,
this permanent (plastic or viscoplastic) strain can be written as:

= d p = d ijp d ijp

(1)

The damage, , can be considered as the sum of a volumetric and a


deviatoric component, figure 1. Volumetric damage by compressive stress
is associated to inelastic compression of the material. Volumetric damage
by tensile stress is associated to cracking of the material. Deviatoric
damage is associated to shear of the material and crack formation as a
result of that.

1.2 Plastic flow criterion


In analogy with the theory of plasticity, the viscoplastic strain rate can be
defined by the following flow rule:
f
& p =
(2)

The Asphalt Concrete Resonponse (ACRe) Model

352

Ch. 4

Figure 2. Representation of the Desai flow surface.

J2

I1

Figure 3. The parameter deter- Figure 4. The parameter determines the size of the Desai flow mines the slope of the Desai flow
surface during the hardening phase.
surface during the degradation phase.

In equation (2) is a constant, which is positive for a material element in


a viscoplastic state ( f = 0 en f& = 0 ), and zero for a material element in an
elastic state ( f < 0 ). The function f is designated as the flow rule (flow
criterion). The plastic flow is described by the criterion of plastic flow,
f = 0 . In INSAP the flow criterion originally developed by Desai is
implemented (Desai et al. 1991):
J
f = 22 Fa Fb = 0
(3a)
pa
where
n

I + R
I + R
Fa = 1
+ 1

pa
pa

Fb = (1 cos 3 )
cos 3 =

3 3 J3

2 J 23/ 2

1/ 2

(3b)

(3c)
(3d)

Sec. 1.3

Simulation of the hardening process

353

response
degradation
phase
hardening
phase

Figure 5. Schematic representation of the hardening


phase and the degradation phase.

In equation (3), I 1 , J 2 and J 3 are stress invariants, pa is the atmospheric


pressure, and R is the triaxial tensile strength. Equation (3a) can be
represented as a closed surface in the ( I 1 , J 2 , )-space, cf. figure 2. The
flow surface defines the stress states where the material begins to deform
plastically. Fa defines the shape and size of the flow surface in the
I 1 , J 2 -plane. The size of the flow surface is determined by the
parameter, , cf. figure 3. The flow surface increases as decreases, and
attains its maximum size for = 0. The slope of the flow surface in the
I 1 , J 2 -plane is determined by the parameter , figure 4. The slope of
the flow surface increases as increases. The cross-section of the surface
in the octahedral plane is determined by the parameter . The crosssection is circular if = 0 . The cross-section becomes increasingly
triangular as increases. The parameter n determines the position of the
apex of the flow surface in the I 1 , J 2 -plane. Two processes in
compressive loading have to be modelled, cf. figure 5: the hardening
process, and the degradation or softening process.

1.3 Simulation of the hardening process


It is assumed that is a function of the deformation rate, & , the
temperature T , and the damage, :
= & ,T ,
(4)
The relationship for a specific asphalt mixture is determined
experimentally in the laboratory, by means of mechanical tests. In the
present study, is given by:
( )
(5)
= 0 lim
1 0
In equation (5), 0 is a constant, which determines the initialisation of the
material plasticity, lim is the effective plastic strain at the maximum

The Asphalt Concrete Resonponse (ACRe) Model

354

Ch. 4

0.09
0.08

0.07
0.06

0.05
0.04
0.03

0.02
0.01
0
0

Figure 6.

0.05

0.1

0.15

p,c

Figure 7. versus

versus .

response, and 0 is a constant. In figure 6, is given as a function of .

1.4 Simulation of the degradation process


It is assumed that is a function of the deformation rate, & , the
temperature T , and the equivalent post fracture plastic strain, consisting
of increments of compressive principal plastic strain components only
(similar to equation (1)), p, c :
= (& ,T , p ,c )
(6)
where
p ,c = d p , c
(7)
where

d p ,c = ( d I d I )

1/ 2

d I < 0

(8)

The definitions (6) (8) define an isotropic softening which simulates an


overall material response degradation that is observed as a result of
compressive loading. The degradation process is modelled by the
following decay function of :
= f + (1 ) r
(9a)

= 1 + 1 p ,c e
f = max
r = 3 f

2 p , c

(9b)
(9c)
(9d)

In equation (9), 1 , 2 , and 3 are material parameters, and max is the


value of at the start of the degradation. During the degradation the slope
of the flow surface, , decreases from the maximum value, f , to the
value which corresponds to the residual strength of the material, r .
According to equation (9a), see figure 7, the variation is controlled by

Sec. 1.5

Simulation of crack-growth

355

Figure 8. Stresses on the crack plane.

Figure 9. Schematic representation of


the Hoffman crack-growth criterion on
the crack plane.

the parameter . The material has maximum strength if = 1, and has


residual strength if = 0.
1.5 Simulation of crack-growth
The tensile hardening in tensile tests could not be satisfactorily described
by the hardening parameter obtained from compression tests (Erkens
2002). A separate tensile hardening law had to be introduced. The nature
of tensile failure differs from that of compressive failure, the former
occurring on a localised plane and the latter more distributed. Erkens
found that the compressive strength is not reduced by tensile damage,
however, that the tensile strength is reduced by compressive damage. As a
result, two types of response degradation were observed, isotropic
softening related to compressive loading and a second degradation
mechanism localised on crack planes owing to tensile loading (Erkens et
al. 2000). On the crack plane, a Hoffman type crack-growth criterion was
assumed (Scarpas and Blauwendraad 1998):
(10)
2 + q ( 2 + 2 ) = f 2 &, T ,
s

where is the normal stress on the crack plane, s and t are shear stress
components, f R is the actual tensile strength after crack initiation, and
is a constant, cf. figures 8 and 9. If the tensile stress direction is taken as
the reference direction, then the coefficient of the Hoffman surface, q , is
given by
f2
q = R2
(11)

where f R is the uniaxial tensile strength, and u is the shear strength. In


the present investigation the following decay function was assumed:
f R = f tu e
(12)
p
cr

356

The Asphalt Concrete Resonponse (ACRe) Model

Ch. 4

where f tu is a constant, is a material parameter, and crp is the cracking


strain on the crack plane.
Table 1. Values of material parameters.
E (MPa)
1500

0.2
temperature and strain dependent
f t
>> f t
f c
T (C)
15
0
0.17
lim
0.03
0
250
1 , 2
10.8
3
0.02
n
2.5

2 Determination of material parameters


The used values of the model parameters are summarised in table 1. These
values are based on results of monotonic uniaxial tensile tests and
monotonic uniaxial compression tests (Scarpas et al. 1997). The material
uses was a dense asphalt concrete 0/16 mixture with 6% 80/100 bitumen.
Parameter . For a given combination of temperature and deformation
rate, & , the compressive strength and the tensile strength are plotted in the
I 1 - J 2 -plane. The value of then follows from the line connecting the
points thus obtained. It is defined by:
2
1 / 3 f c(T , &)
&

(T , ) =
(13a)
f c( T , &) + 3 f t (T , &)
where f c( T , &) is the monotonic uniaxial compressive strength, and
f t ( T , &) is the monotonic uniaxial tensile strength. The mixture specific
relationships for the compressive strength respectively the tensile strength
as function of the temperature and the deformation rate are:
& exp (3.25 0.045 T )
(13b)
f c( T , &) = &
. exp ( 0.035 T ) )
+ 0.85 (12
e1.3

&
f t ( T , ) = 1.3 510 T &
1.75 +
e
5

for T 10 o C

3.3

for T > 10 o C

(13c)

Sec. 2

Determination of material parameters

e1.3

f t ( T , &) = 1.3 510 T &


1.75 +
e
5

357

for T 10 o C

3.3

for T > 10 C

(13c)

Parameter . The parameter was put equal to zero. The reason is that
at the time of the investigation only experimental data of uniaxial
compression tests and uniaxial tensile tests were available, and that
experimental data for two or three-dimensional stress states are required to
be able to model the influence of on the response.
Parameter n . The parameter n was arbitrarily put equal to 2.5.

358

The Asphalt Concrete Resonponse (ACRe) Model

Ch. 4

Appendix 6: Computer programme to compute the


da / dN K -relationship
//Computer programme to compute the crack-growth rate in a dynamic
//crack-growth test based on the crack-growth rate in a static crack-growth
//test (creep crack-growth test)
//The crack-growth law is: da/dt = A*K^exp
INPUT specimen width [m] ; W
INPUT specimen thickness [m] ; B
INPUT minimum force [N] ; Pmin
INPUT maximum force [N] ; Pmax
INPUT frequency [Hz] ; f
INPUT integration step size [s] ; dt
INPUT coefficient of creep law [m/cycle] ; A
INPUT exponent of creep law ; exp
INPUT initial crack-length [m] ; a0
INPUT final crack-length [m] ; af
OUTTEXT W = , W, B = , B, Pmin = , Pmin, Pmax = , Pmax, f =
, f, dt = , dt, A = , A, exp = , exp, a0 = , a0, af = , af
OUTTEXT N [cycles] a [m] Kavg [MPam] delta-K [MPam]
dadN [m/cycle]
pi:=3.14159
time:=0
a1:=a0
WHILE a<af DO
BEGIN
a:=0
FOR t:=1 TO 100 DO
BEGIN
Kavg:=(0.5*(Pmin+Pmax)/(W*B))*((pi*a)^0.5)*(1/(cos(pi*a/W)))^0.5
Kt:=(0.5*(PmaxPmin)/(W*B)*sin(w*t/100))*((pi*a)^0.5)*
(1/(cos(pi*a/W)))^0.5
K:=(Kavg + Kt)/1000000 //in MPam
IF K > 0.55 THEN STOP
IF K < 0 THEN da:=da ELSE da:=da+A*(K^exp)/100
END
//continue on next page

360

dK:=(PmaxPmin)/(W*B)/1000000*((pi*a)^0.5)*(1/(cos(pi*a/W)))^0.5
a:=a+da
time:=time+1
N:=time*f
dadN:=da/f*1000000 //in m/cycle
IF aa1>=0.001 THEN
BEGIN
WRITE N, a, Kavg/1000000, dK, dadN
a1:=a
END
END

361

Bibliography
ASTM (1979): ASTM Standard E 399-78A, Standard Test Method for
Plane-Strain Fracture Toughness of Metallic Materials, 1979 Annual Book
of ASTM Standards, Vol. 3.01, American Society for Testing Materials.
ASTM (1993): Standard test method for measurements of fatigue crackgrowth rates, ASTM Standard E 647-93.
Balla A. (1961): Stress conditions in triaxial compression, Transactions
of the American Society of Civil Engineers, 1074-1101.
Bolk H.J.N.A. (1980): The creep test, Publication M, Study Centre Road
Building (Stichting Studie Centrum Wegenbouw), Delft; today: National
Bureau of Standardisation and Research for Civil Works, C.R.O.W., Ede
(in Dutch).
Boltzmann (1874): Zur Theorie der elastischen Nachwirkung,
Sitzungsbericht der Kaiserlichen Akademie der Wissenschaften, Wien,
Mathematische Naturwissenschaft Kl. 70(2), 275-306.
Borst R. de, Bicanic N., Mang H., Meschke G. (1998): Computational
modelling of concrete structures, Proceedings of the Euro-C 1998
Conference on Computational Modelling of Concrete Structures, March
31-April 3, Badgastein, Austria, 193-202.
Broek D. (1986): Elementary engineering fracture mechanics, Fourth
revised ed., Kluwer Academic Publishers, Boston, ISBN 90 247 2656 5.
Carson J.R. (1926): Electrical circuit theory and operational calculus,
McGraw-Hill, New York.
Clard B. (1977): Esso Road Design Technology, Proceedings of the
Fourth International Conference on the Structural Design of Asphalt
Pavements, University of Michigan, Ann Arbor, 249-268.
Claessen A.I.M., Edwards J.M., Sommer P., Ug P. (1977): Asphalt
pavement design, Proceedings of the Fourth International Conference on
the Structural Design of Asphalt Pavements, University of Michigan, Ann
Arbor, 39-74.
CROW (1996): Modified bitumen, Publication Nr. 104, C.R.O.W.
(National Bureau of Standardisation and Research for Civil Works), Ede
(in Dutch).

362

Bibliography

CROW (1998): Stone mastic asphalt, Publication Nr. 63, C.R.O.W., Ede
(in Dutch).
CROW (2000): Standard RAW Requirements, C.R.O.W., Ede (in Dutch).
Desai C.S., Sharma K.G., Wathugala G.W., Rigby D.B. (1991):
Implementation of hierarchical single surface 0 and 1 models in finite
element procedure, International Journal for Numerical and Analytical
Methods in Geomechanics, 15,649-680.
Duhamel J.M.C. (1833): Sur la mthode gnrale relative au mouvement
de la chaleur dans les corps solides plongs dans des milieux dont la
temprature varie avec le temps, Journal de lcole Polytechnique de
Paris, 14(22)20-77.
Eisenmann J., Hilmer A. (1987): Influence of wheel load and inflation
pressure on the rutting effect at asphalt pavements, Proceedings of the
sixth international conference Structural design of asphalt pavements,
University of Michigan, Ann Arbor, Michigan, July 13-17, 392-403.
Elber W. (1971): ASTM STP 486, American Society for Testing
Materials, Philadelphia, 230-242.
Elphingstone G.M. (1997): Adhesion and cohesion in asphalt-aggregate
systems, dissertation, Texas A&M University.
EMPA (1977). Colloquium 77: Plastische Verformbarkeit von
Asphaltmischungen, Mitteilung Nr. 37, Institut fr Strassen-, Eisenbahnund Felsbau an der Eidgenssischen Technische Hochschule Zrich,
Zrich, September 29-30.
Erkens S.M.J.G., Liu X., Scarpas A., Molenaar A.A.A. (2000): 3D Finite
element model for asphalt concrete response simulation, paper present at
the Second International Symposium on 3D Finite Element Modelling in
Pavement Engineering, Charleston, West-Virginia, U.S.A.
Erkens S.M.J.G. (2002): Asphalt Concrete Response (ACRe)
Determination, Modelling and Prediction, dissertation of the Delft
University of Technology, ISBN 90 407 2326 5.
Ewalds H.L., Van Doorn F.C., Sloof W.G., (1983): ASTM STP 801,
American Society for Testing Materials, Philadelphia, 115-134.
Ewalds H.L., Wanhill R.J.H. (1984): Fracture mechanics, First edition,
Delftse Uitgevers Maatschappij, Delft, ISBN 90 6562 024 9.

Bibliography

363

Findley W.N., Lai J.S., Onaran K. (1976): Creep and relaxation of


nonlinear viscoelastic materials, in H.A. Lauwerier, W.T. Koiter (eds.):
Applied mathematics and mechanics, Volume 18. ISBN 0 7204 2369 4.
North-Holland Publishing Company, Amsterdam
Flgge W. (1967): Viscoelasticity, Ginn (Blaisdell), Boston.
Groenendijk J. (1998): Accelerated testing and surface cracking of
asphaltic concrete pavements, Dissertation, Delft University of
Technology, ISBN 90-804590-1-1.
Hagemann R. (1980): Ein Verfahren zur Beurteiling flexibler
Fahrbahnbefestigungen unter Bercksichtigung von Festigkeitshypothesen
fr Asphalte. Mitteilungen aus dem Institut fr Baustoffkunde und
Materialprfung der Universitt Hannover, Heft 44.
Hagos E.T. (2002): Characterisation of polymer modified bitumen, MSc
Thesis, Delft University of Technology (TRE 123, IHE Delft).
Heukelom W. (1966): Observations on the rheology and fracture of
bitumens and asphalt mixes, Proceedings of the Association of Asphalt
Paving Technologists, 35, 358-399.
Hondros G. (1959): The evaluation of Poissons ratio and the modulus of
materials of a low tensile resistance by the Brazilian (indirect tensile) test
with particular reference to concrete, Australian Journal of Applied
Science 10(3), 243-268.
Hopkinson J. (1876): The presidual charge of the Leyden jar,
Philosophical Transactions of the Royal Society, London, 166, 489-494.
Irwin L.H. (1977): Use of fracture energy as a fatigue failure criterion,
Proceedings of the Association of Asphalt Paving Technologists, February
21-23, 41-63.
Jongeneel D.J., Haugh L.D., Gerritsen A.H. (1985): Creep testing.
Results of a European interlaboratory study of laboratory apparatuses and
test procedures, Proceedings of the Third Eurobitume Symposium, The
Hague, September 11-13, 294-300.
Jacobs M.M.J. (1995): Crack growth in asphaltic mixes, dissertation
Delft University of Technology, ISBN 90-9007965-3.

364

Bibliography

Kleemans C.P. (1994): Fatigue crack-growth in sand


MSc Thesis, Delft University of Technology (in Dutch).

asphalt,

Kleemans C.P., Zuidema J., Krans R.L., Molenaar J.M.M., Tolman F.


(1997): Fatigue and creep crack growth in fine sand asphalt materials,
Journal of Testing and Evaluation, Vol. 25, No. 4, 424-428.
Krans R.L., Van de Ven, M.F.C., Molenaar J.M.M., Kunst P.A.J.C.
(1993): Crack Growth experiments on asphalt concrete beams.
Proceedings of the International Conference on Safety in Europe and
Strategic Highway Research Program, The Hague, September 20-22,
1995, Vol. 3, 179-192.
Krans R.L. (1995): unpublished document.
Krans R.L., Tolman F., Van de Ven M.F.C. (1996): Semi-circular
bending test: a practical crack growth test using asphalt concrete cores in
L. Francken, E. Beuving, A.A.A. Molenaar (eds.) Reflective cracking in
pavements, Proceedings of the Third International Rilem Conference,
Maastricht, October 2-4, Published by E & FN Spon, London, 123-132.
Kuppens E.A.M., Sanches F., Nardelli L., Jongmans E.C. (1997):
Bitumen-ageing tests for predicting durability of porous asphalt,
Proceedings of the Fifth International Rilem Symposium, May 14-16,
Lyon (Mechanical Tests for Bituminous Materials, Editors: H. Di
Benedetto, L. Francken) 71-78.
Lee H-J., Kim Y.R. (1998): Viscoelastic constitutive model for asphalt
concrete under cyclic loading, Journal of Engineering Mechanics,
January, 32-40.
Lim I.L., Johnston I.W., Choi S.K. (1993): Stress intensity factors for
semi-circular specimens under three-point bending, Engineering Fracture
Mechanics, 44, 363-382.
Lim I.L., Johnston I.W., Choi S.K., Boland J.N. (1994): Fracture testing
of a soft rock with semi-circular specimens under three point bending. Part
1 - Mode I, Int. J. Rock Mechanical., Min. Sci. & Geomech. Abstr. 31,
185-197, Part 2 - Mixed-mode I, Int. J. Rock Mechanical. Min. Sci. &
Geomech. Abstr. 31, 199-212.
Liu X., Scarpas A. (2001): 3D Finite element investigation of localisation
issue in strain rate dependent material, Proceedings of the International
Conference on Composite Materials, Edited by Y. Zhang, Beijing.

Bibliography

365

Majidzadeh K., Kauffmann E.M., Saraf C.L. (1971): Analysis of fatigue of


paving mixtures from the fracture mechanics viewpoint. Fatigue of
compacted bituminous aggregate mixtures. ASTM report STP 508, 67-83.
Medani T.O., Molenaar A.A.A. (2000): A simplified practical procedure
for estimation of fatigue and crack growth characteristics of asphaltic
mixes, Road materials and pavement design Vol. 1, No. 4, 451-465
Mes B. (2003): Unravelling of ageing properties of bitumen in porous
asphalt (in Dutch), Hogeschool Rotterdam, Analytical Physical
Chemistry, also available at RHEI, report No. DWW 2003-075.
MH (1988): National Environmental Programme, Ministry of Housing,
Spatial Planning, and Environment, in Dutch only: Nationaal
Milieubeleidsplan, Ministerie van Volkshuisvesting, Ruimtelijke Ordening
en Milieu.
Molenaar A.A.A. (1983): Structural performance and design of flexible
road construction and asphalt concrete overlays, Dissertation, Delft
University of Technology.
Molenaar J.M.M. (1991): Polymer modified bitumens - Functional
properties and quality aspects, Proceedings of the International
Symposium Chemistry of Bitumens, Rome, June 5-8.
Molenaar J.M.M., Verburg H.A., Westera G.E. (1995): Characterisation
of permanent deformation behaviour of asphalt mixtures, Proceedings of
the International Conference on Safety in Europe and Strategic Highway
Research Program, Prague, September 20-22, Vol. 6, 28-53.
Molenaar J.M.M., Molenaar A.A.A. (2000): Susceptibility to permanent
strain of asphalt in the dynamic triaxial compression creep test,
Eurasphalt and Eurobitume Congress, Barcelona, September 20-22,
Molenaar J.M.M. Molenaar A.A.A. (2000): Fracture toughness of asphalt
in the semi-circular bend test, Eurasphalt and Eurobitume Congress,
Barcelona, September 20-22,
Monismith C.L., Secor K.E. (1962): Viscoelastic behavior of asphalt
concrete pavements, Proceedings of the International Conference on the
Structural Design of Asphalt Pavements, Ann Arbor, August 20-24, 476498.

366

Bibliography

Monismith C.L., Alexander R.L., Secor K.E. (1966): Rheologic behavior


of asphalt concrete, Proceedings of the Association of Asphalt Paving
Technologists, Vol. 35, 400-450.
Morland L.W., Lee E.H. (1960): Transactions of the Society of
Rheologists IV, 233.
MT (1988): Second Traffic and Transport Framework Programme,
Ministry of Transport, Public Works and Water Management (in Dutch:
Tweede Structuurschema Verkeer en Vervoer, Ministerie van Verkeer en
Waterstaat).
MT (1996a): Beating Congestion, Ministry of Transport, Public Works
and Water Management (in Dutch: Samen Werken aan Bereikbaarheid,
Ministerie van Verkeer en Waterstaat).
MT (1996b): Balancing Transport, Ministry of Transport, Public Works
and Water Management (in Dutch: Transport in Balans, Ministerie van
Verkeer en Waterstaat).
MT (2001): National Traffic and Transport Framework Programme,
Ministry of Transport, Public Works and Water Management (in Dutch:
Nationaal Verkeer en Vervoer Plan, Ministerie van Verkeer en
Waterstaat).
Murakami Y. (1986): Stress intensity factor handbook. Pergamon Press,
Oxford.
Partl M.N., Francken L. (1997): Rilem interlaboratory tests on stiffness
properties of bituminous mixtures, Proceedings of the Fifth International
Rilem Symposium on Mechanical Tests for Bituminous Materials, Lyon,
9-14.
Phillips M.C., Robertus C. (1995): Rheological characterisation of
bituminous binders in connection with permanent deformation in asphaltic
pavements - the zero shear viscosity concept, Proceedings of the First
European Workshop on the Rheology of Bituminous Binders, Eurobitume
Brussels, April 5-7, paper #50.
Phillips M.C., Robertus C. (1996): Binder rheology and asphaltic
pavement permanent deformation; the zero shear viscosity concept,
Proceedings of the Eurasphalt & Eurobitume Congress Strasbourg, May 79, paper #5.134.

Bibliography

367

Phillips M.C. (1997): Developments in specifications for bitumens and


polymer modified binders, mainly from a rheological point of view, in
J.G. Cabrera (ed.): Performance and Durability of Bituminous Materials,
Proceedings of the Second European Symposium, Leeds, April, ISBN 3931681 14 9, 3-18.
Phillips M.C. (1999): Module 2 - Plenary briefing, Stiffness and
permanent deformation, Proceedings Eurobitume Workshop 99
Performance Related Properties for Bituminous Binders, ISBN 2-93016005-5, 132.
Pipkin A.C. (1972): Lectures on Viscoelasticity Theory, SpringerVerlag, Berlin and New York.
RHEI (1976): Annual report, Verslag der werkzaamheden (in Dutch),
26-30.
RHEI (1996): Super singles and heavy traffic Costs and benefits
analysis on main lines (in Dutch), RHED report P-DWW-96-058.
RHEI (1998): Manual Road Design - Pavement Design Road and
Hydraulic Engineering Division, P-DWW-98-065, ISBN 90 369 3742 6,
ISBN 90 369 0051 4.
RHEI (2001): Investigation into the yearly maintenance costs of the main
road network caused by overloaded trucks in The Netherlands (in Dutch).
KOAC-report e01011, November 19.
Roelfstra P.E. (1989): A numerical approach to investigate the properties
of concrete, Numerical Concrete, dissertation no. 788 of the Swiss Federal
Institute of Technology, Lausanne, Switzerland.
Romer P.M. (1986): Increasing returns and long-run growth, Journal of
political economy 94(5)1002-1037.
Romer P.M. (1990): Endogeneous technological change, Journal of
political economy 98(5)71-102.
Scarpas A., Al-Khoury R., Gurp C.A.P.M. van, Erkens S.M.J.G. (1997):
Finite element simulation of damage development in asphalt concrete
pavements, Proceedings of the Eighth International Conference on
Asphalt Pavements, Seattle, Washington, August 10-14, 673-692.

368

Bibliography

Scarpas A., Blaauwendraad J. (1998): Experimental calibration of a


constitutive model for asphaltic concrete, paper presented at the Euro-C
conference on the computation modelling of concrete structures,
Badgastein, Austria, March 31-April 3, 1998.
Schapery R.A. (1973): Nonlinear fracture analysis of viscoelastic
composite materials based on a generalised J integral theory, Texas A&M
University, Report TX 77843.
Schapery R.A. (1974): Viscoelastic behaviour and analysis of composite
materials, in G.P. Sendeckyj (ed.): Mechanics of composite materials,
Vol. 2, Academic Press, New York, 85-168.
Schapery R.A. (1975): A theory of crack initiation and growth in
viscoelastic media, Part I. Theoretical development, International
Journal of Fracture, Vol. 11, pp. 141-159, Part II. Approximate methods
of analysis, Int. J. Fract., Vol. 11, pp. 369-388, Part III. Analysis of
continuous growth, Int. J. Fract., Vol. 11, 549-562.
Schapery R.A. (1981): On viscoelastic deformation and failure behaviour
of composite materials with distributed flaws in S.S. Wang, W.J. Renton
(eds): Advances in aerospace structures and materials, The American
Society of Mechanical Engineers, Publication AD-01, 5-20.
Schapery R.A. (1981b): Nonlinear fracture analysis of viscoelastic
composite materials based on a generalised J integral theory, in K.
Kawata, T. Akasaka (eds): Proceedings of the Japan-U.S. Conference
Composite materials, 171-180.
Schapery R.A. (1982): Models for damage growth and fracture in
nonlinear viscoelastic particulate composites, Proceedings of the 9th U.S.
National Congress on Applied Mechanics, The American Society of
Mechanical Engineers, 237-245.
Schapery R.A. (1984): Correspondence principles and a generalized J
integral for large deformation and fracture analysis of viscoelastic media,
Internationa Journal of Fracture Vol. 25, 195-223.
Schapery R.A.(1986): A micromechanical model for nonlinear
viscoelastic behavior of particle-reinforced rubber with distributed
damage, Engineering Fracture Mechanics Vol. 25, No. 5/6, 845-867.

Bibliography

369

Schijve (1977): Four Lectures on Fatigue Crack Growth, Delft


University of Technology, Faculty of Aerospace Engineering, Report LR254.
Schijve J. (1981): Engineering Fracture Mechanics Vol. 14, 467-475.
SVC (1982): Filler and the mechanical properties of dense graded asphalt
concrete (in Dutch), Stichting Vulstof Controle; see also: Bolk H.J.N.A.,
van der Heide J.P.J., Zandvliet M.C. (1982): Basic Research into the
effect of filler on the mechanical properties of dense graded asphalt
concrete, Proceedings of the Association of Asphalt Paving
Technologists, February 22-24, Vol. 51, 398-452.
Tangella R., Craus J., Deacon J.A., Monismith C.L. (1990): Summary
report on fatigue response of asphalt mixtures. SHRP Project A 003-A,
University of California.
Valkering C.P., Lanon D.J.L., Hilster E. de, Stoker D.A. (1990): Rutting
resistance of asphalt mixes containing non-conventional and polymer
modified binders, Journal of the Association of Asphalt Paving
Technologists, Vol. 59, 590-609.
Van de Ven M., de Fortier Smit A., Krans R. (1997): Possibilities of a
semi-circular bending test, Proceedings of the Eighth International
Conference on Asphalt Pavements, Seattle, Washington, August 10-14,
939-950.
Verburg H.A., Krans R.L., Salet A.T.M., Roelfstra P.E. (1995):
Numerical asphalt, Proceedings of the International Conference on
Safety in Europe and Strategic Highway Research Program, Prague,
September 20-22, Vol. 7, 41-52.
Voskuilen J.L.M., Molenaar J.M.M., Pietersen H.S. (1996): Adsorption
and desorption of bitumen/toluene mixtures on mineral aggregates,
Proceedings of the Eurobitume and Eurasphalt Congress, paper # 4.078.
Zuidema J. (1995): Square and slant fatigue crack-growth in Al 2024,
dissertation of the Delft University of Technology, ISBN 90 407 1191 7.

You might also like