Sunteți pe pagina 1din 29

Road Materials and Pavement Design

ISSN: 1468-0629 (Print) 2164-7402 (Online) Journal homepage: http://www.tandfonline.com/loi/trmp20

Ageing and rejuvenators: evaluating their


impact on high RAP mixtures fatigue cracking
characteristics using advanced mechanistic
models and testing methods

Walaa S. Mogawer, Alexander Austerman, Reynaldo Roque, Shane


Underwood, Louay Mohammad & Jian Zou

To cite this article: Walaa S. Mogawer, Alexander Austerman, Reynaldo Roque, Shane
Underwood, Louay Mohammad & Jian Zou (2015): Ageing and rejuvenators: evaluating
their impact on high RAP mixtures fatigue cracking characteristics using advanced
mechanistic models and testing methods, Road Materials and Pavement Design, DOI:
10.1080/14680629.2015.1076996

To link to this article: http://dx.doi.org/10.1080/14680629.2015.1076996

Published online: 16 Sep 2015.

Submit your article to this journal

Article views: 47

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


http://www.tandfonline.com/action/journalInformation?journalCode=trmp20

Download by: [University of Western Ontario] Date: 13 December 2015, At: 13:45
Road Materials and Pavement Design, 2015
http://dx.doi.org/10.1080/14680629.2015.1076996

Ageing and rejuvenators: evaluating their impact on high RAP mixtures


fatigue cracking characteristics using advanced mechanistic models and
testing methods
Walaa S. Mogawera∗ , Alexander Austermana , Reynaldo Roqueb , Shane Underwoodc ,
Louay Mohammadd and Jian Zoub
Downloaded by [University of Western Ontario] at 13:45 13 December 2015

a Highway Sustainability Research Center (HSRC), University of Massachusetts, Fall River, MA, USA;
b University of Florida, Department of Civil and Coastal Engineering, Gainesville, FL, USA; c Arizona State
University, Civil, Environmental and Sustainable Engineering, Tempe, AZ, USA; d Department of Civil and
Environmental Engineering, Louisiana Transportation Research Center, Louisiana State University, Baton
Rouge, LA, USA

(Received 4 August 2014; accepted 26 October 2014 )

Fatigue cracking of asphalt mixtures is highly dependent on ageing. Using larger amounts
of reclaimed asphalt pavement (RAP) presents a concern that the resultant mixtures may be
prone to fatigue cracking because of the aged binder in the RAP. Several studies have indi-
cated that asphalt rejuvenators can allow more aged binder to be incorporated into asphalt
mixtures. The four-point flexural beam fatigue test, HMA (hot-mix asphalt) fracture mechan-
ics model, simplified viscoelastic continuum damage model, and the semi-circular bending
test were used to evaluate the effect of ageing on the fatigue characteristics of high RAP mix-
tures modified with rejuvenators. The results from these tests were compared to see if they
provided similar performance trends. The results indicated that the long-term ageing used in
this study did not have a significant effect on the fatigue characteristics of the high RAP mix-
ture with and without rejuvenators. Comparison of the fatigue tests did not show universal
agreement.
Keywords: ageing; reclaimed asphalt pavement; asphalt rejuvenator; fatigue cracking
performance

1. Introduction
Cracking is a major distress in asphalt mixtures that manifests in the field in the form of fatigue,
low temperature, longitudinal, block, reflection, and slippage cracking. For this study one of the
most typically encountered forms of cracking, fatigue cracking, was the focus of investigation.
Fatigue cracking is highly dependent on many factors including ageing. Asphalt mixtures
encounter both short- and long-term ageing. Ageing during mixing and construction is referred
to as short-term ageing. Ageing during the service life of the pavement is referred to as long-term
ageing. Short-term ageing is mainly a result of the production process (mixing and compaction)
including the elevated temperatures and presence of moisture. Long-term ageing is a result of
exposure to a combination of environmental conditioning factors after placement (ultraviolet
radiation, temperature, rainfall, and in-place service time) and also in-service loading (Abbas,
Chai, Masad, & Papagiannakis, 2002). It continues to occur throughout an asphalt mixture’s
service life. A prominent component of long-term ageing is oxidation which results from the

*Corresponding author. Email: wmogawer@umassd.edu

© 2015 Taylor & Francis


2 W.S. Mogawer et al.

interaction of asphalt binder components with both oxygen and ultraviolet radiation. During the
first Strategic Highway Research Program, laboratory procedures were developed to simulate
the field hardening of asphalt binders and mixtures including oxidative ageing (Bell, Wieder, &
Fellin, 1994). These procedures are American Association of State Highway and Transportation
Officials (AASHTO) R28 “Standard Practice for Accelerated Aging of Asphalt Binder Using a
Pressurized Aging Vessel (PAV)” and AASHTO R30 “Standard Practice for Mixture Condition-
ing of Hot-Mix Asphalt (HMA)” (American Association of State Highway and Transportation
Officials [AASHTO], 2011).
Mixture ageing is affected by both internal and external variables. Internal variables include
asphalt and aggregate properties, mixture asphalt content, asphalt film thickness, and air void
content (Kandhal & Chakraborty, 1996). External variables are the short- and long-term variables
noted previously. These variables interact throughout the service life of the pavement causing an
Downloaded by [University of Western Ontario] at 13:45 13 December 2015

asphalt mixture to undergo rheological (elastic, viscous, and cracking behaviours) and chemical
changes including an increase in the stiffness of the asphalt binder. The result of ageing is that
it increases an asphalt mixture’s susceptibility to cracking (Abbas et al., 2002; Glover et al.,
2005). Utilisation of larger amounts of previously aged materials like reclaimed asphalt pavement
(RAP) in new paving mixtures further complicates the efforts to quantify and understand the
effects of ageing and its impact on the fatigue cracking potential of mixtures.
Due to the increased cost of asphalt binder and Federal Highway Administration’s policy
to increase environmental stewardship, the industry has been looking for ways to increase the
amount of RAP in asphalt mixtures. RAP is comprised of an asphalt binder that has been hard-
ened due to long-term ageing along with re-usable aggregates. Using larger amounts of RAP in
new paving mixtures presents a concern that the resultant mixtures may be prone to more fatigue
cracking during the service life of the pavement. This is due to the asphalt binder in the RAP
being significantly aged. In new paving mixtures, this already aged binder will be exposed to
additional short- and long-term ageing. Studies have confirmed that increased ageing will reduce
the stress relaxation capacity of the binder which means it decreases the cracking resistance of
the mixture (Daniel, Gibson, Tarbox, Copeland, & Andriescu, 2013; Molenaar, Hagos, & Van de
Vev, 2010). To alleviate the effect of the hardened RAP binder on the cracking susceptibility of
new asphalt mixtures generally a softer binder is used. However, several studies have indicated
that asphalt rejuvenators can allow more aged binder to be incorporated into mixtures than a
softer binder alone. Even so, for recycling purposes, rejuvenators are not encouraged by some
state agencies because of potential rutting-related concerns.
For mixtures that incorporate materials with high RAP contents (with or without using a softer
binder and/or a rejuvenator) fatigue cracking models and tests that rely on the mechanistic and
volumetric properties of mixtures are needed. The classic method for evaluating asphalt mixtures
fatigue in the USA is the standardised flexural beam fatigue test outlined in AASHTO T321
“Standard Method of Test for Determining the Fatigue Life of Compacted Hot-Mix Asphalt
(HMA) Subjected to Repeated Flexural Bending” (AASHTO, 2011). The primary outcome of
this method, when carried out in a series of experiments with mixtures at different temperatures
and/or strain amplitudes, is a relationship between strain amplitude, modulus, and fatigue life.
This relationship can then be incorporated in some empirical or mechanistic-empirical algorithm
of varying complexity to extract the fatigue performance of asphalt pavements. There are certain
limitations in this approach with respect to the mathematical rigour of the pavement prediction
algorithm since it must rely on two separate analyses (one to extract the pavement response and
one to predict the damage). Likewise, the material model does not directly consider any specific
cracking mechanism and in the end it is not known what individual factors are responsible for the
empirically calibrated coefficients. For the methodology to yield reasonably accurate results the
calibration data set must be sufficiently large so as to properly cover the entire range of conditions
Road Materials and Pavement Design 3

expected even though this calibration set may in fact include some redundancy in characterising
the contribution from individual mechanisms.
Recently, researchers have used the theories of fracture mechanics (FM) and continuum dam-
age to develop models and tests that can evaluate the fatigue cracking initiation and cracking
propagation phenomenon of asphalt mixtures from a less empirical and more mechanism-centric
perspective (Roque, Birgisson, Drakos, & Dietrich, 2004; Transportation Research Board, 2012;
Underwood & Kim, 2010; Wu, Mohammed, Wang, & Mull, 2005). These models and tests pro-
vide another means to ascertain the effects of long-term ageing on the performance of high RAP
content mixtures. Also, since these models and tests were derived based on different theories, it
is important to determine if they provide similar performance in regard to each other and to the
flexural bending beam fatigue test.
Downloaded by [University of Western Ontario] at 13:45 13 December 2015

2. Objectives
The objectives of this study were as follows: (1) study the effect of long-term ageing on the
fatigue characteristics of high RAP mixtures modified with rejuvenators using both conventional
and recently developed mechanistic models and tests; and (2) to determine if these models/tests
provide similar performance trends in regard to each other and to the flexural bending beam
fatigue test.

3. Description of models and tests used in the study


The following outlines the tests and models utilised to evaluate the effects of long-term ageing
on the fatigue cracking characteristics of the mixtures included in this study. Four tests/models
were used:

• Four-point flexural beam fatigue test


• HMA-FM model
• Simplified viscoelastic continuum damage (SVECD) model
• Semi-circular bending (SCB) test

Each of these tests/models follows different testing protocols such as sample size, specimen
shape, testing temperatures, etc. Nevertheless, each of them has shown good correlation to field
performance. Accordingly, it is expected that, although these tests/models were developed based
on different theories, they would provide similar fatigue cracking performances for the mixtures
evaluated in this study.

3.1. Four-point flexural beam fatigue test


One of the most common and historically used laboratory test procedure chosen to evaluate
the fatigue cracking resistance of asphalt mixtures is the four-point flexural beam fatigue test.
The typical test protocols are AASHTO T321 “Determining Fatigue Failure of Compacted
Asphalt Concrete Subjected to Repeated Flexural Bending” (AASHTO, 2011) and ASTM D7460
“Determining Fatigue Failure of Compacted Asphalt Concrete Subjected to Repeated Flexural
Bending” (American Society for Testing and Materials [ASTM], 2009). In order to investigate
the effect of ageing and rejuvenators on the fatigue cracking characteristics of high RAP content
asphalt mixtures, beam fatigue testing was conducted in accordance with AASHTO 321. This
test was conducted in strain control mode at a load frequency of 10 Hz applied using a sinusoidal
4 W.S. Mogawer et al.

waveform. Specimens were tested at strain levels of 500 and 750 με. The number of cycles to
failure was determined by fitting an exponential function to the flexural stiffness versus number
of cycles and then evaluating the number of cycles it took to decrease the initial stiffness by 50%
of the initial stiffness measured at the 50th cycles.

3.2. HMA-FM model


The HMA-FM model was developed based on the fact that an asphalt mixture has a fundamen-
tal dissipated creep strain energy limit (DCSEf ) (also called the fracture threshold), which has
been determined to be independent of loading mode (controlled stress or controlled strain) and
history. Cracking will initiate or propagate in any region of an asphalt mixture where induced
damage exceeds the threshold. The following asphalt mixture properties are used in the model:
Downloaded by [University of Western Ontario] at 13:45 13 December 2015

(i) resilient modulus (M R ). (ii) creep rate (dependent on creep compliance power law parame-
ters D1 and m-value), (iii) tensile strength (S t ), (iv) fracture energy limit (FEf ), and (v) dissipated
creep strain energy limit (DCSEf ). These properties can be obtained using the Superpave indirect
tensile (IDT) test system. Detailed procedures regarding testing, data reduction, and interpreta-
tion are described elsewhere (Roque, Birgisson, Sangpetngam, & Zhang, 2002; Roque et al.,
1997; Zhang, Roque, Birgisson, & Sangpetngam, 2001).
Based on a detailed analysis and evaluation of 22-field test sections (in service for over 10
years) throughout the State of Florida, including conducting Superpave IDT tests on aged field
cores and comparing predicted cracking performance by the HMA-FM model with the observed
field performance, an energy-based fracture criterion was developed at the university of Florida,
consisting of an energy ratio (ER) and associated minimum values of ER required for adequate
cracking performance (Roque et al., 2004). The ER parameter, defined as the ratio of DCSEf
over the minimum dissipated creep strain energy required, is expressed in Equation (1):

a × DCSEf
ER = , (1)
m2.98 × D1
where
a = 0.0299σt−3.1 × (6.36 − St ) + 2.46 × 10−8 ,
where σ t (in psi) is the tensile stress in the asphalt layer; S t is tensile strength (in MPa),
DCSEf is dissipated creep strain energy limit (in kJ/m3 ), D1 (in 1/psi), and m-value are creep
compliance power law parameters. Previously established minimum values of ER (ERmin )
required are (i) 1.0 for no greater than 5 million equivalent single-axle loads (ESALs), (ii) 1.3
for no greater than 10 million ESALs, and (iii) 1.95 for no greater than 20 million ESALs,
which were determined by taking into account the change in pavement structure with varying
traffic levels.
The ER accounts for the effects of pavement structural characteristics (σ t ) and material prop-
erties (DCSEf , S t , m, D1 ) on cracking performance. The higher the value of the ER, the better the
expected cracking performance of the section. Therefore, ER can be used to integrate asphalt
mixture properties in the pavement design process as well as to predict the performance of
in-service pavement sections.

3.3. SVECD model


Researchers (Underwood & Kim, 2010) have shown that models developed based on continuum
damage theory (CDT) can be used to explain and predict the fatigue life of an asphalt mix-
ture. By theory, these methods are generally considered to be most applicable up to the point of
Road Materials and Pavement Design 5

macrocrack development, many times referred to as crack initiation. In the laboratory, this point
is clearly defined by a localisation of deformation at the macrocrack site. In the field, initiation is
less clearly defined since it is often taken as the time when a crack becomes visible on the pave-
ment surface. For top-down cracking such a definition may correspond to the localisation-based
definition. However, in the case of bottom-up cracking, this definition actually represents both
the initiation of cracking and the propagation of the crack through the pavement layers. While
their direct theoretical application is limited to initiation, CDT approaches may be applied to
model post-localisation behaviours of pavements (Kim et al., 2008; Park & Kim, 2013; Under-
wood, Kim, Savadatti, Thirunavukkarasu, & Guddati, 2009). Justification for this extension lies
in implicit correlations between the initiation and propagation mechanics of asphalt mixtures
(Roque et al., 2010). These correlations may be related to the observation that macrocrack prop-
agations in asphalt mixtures, at least at moderate to high temperatures where fatigue is a primary
Downloaded by [University of Western Ontario] at 13:45 13 December 2015

concern, are associated with a microcrack dominated process zone ahead of the macrocrack tip.
While CDT can be used for the prediction of highly developed pavement cracking, it is the more
commonly held belief that after the point of crack initiation, other techniques including FM
and approximate cohesive zone methods are needed for the further evolution of cracking. Even
without explicitly integrating these techniques with CDT, pavement predictions based on CDT
can be beneficial as the aim for pavement design is to produce pavements that do not exhibit
larger cracks.
CDT methods have been evolving for asphalt concrete (AC) since at least the late 1980s, and
their characterisation and analysis protocols involve more complex calculus than is used in tra-
ditional beam fatigue analysis. The method used in this paper is the SVECD model developed
by Underwood and Kim (2010). This technique allows the fatigue life of an asphalt mixture at
various strain–stress amplitudes under different temperatures to be predicted from its dynamic
modulus and cyclic fatigue data. For the purposes of this paper, cyclic fatigue tests were per-
formed at strain levels between approximately 250 and 550 με at 15°C. These conditions were
determined following AASHTO TP 107 “Determining the Damage Characteristic Curve of
Asphalt Concrete from Direct Tension Cyclic Fatigue Tests”. The experiments themselves do not
preclude any particular analysis and could be interpreted using the same protocols used in beam
fatigue testing. However, for modelling purposes the primary material function determined from
these tests is the relationship between the material integrity, denoted by the variable C, and the
amount of damage, S. This function is referred to as the damage characteristic function and has
been found to be a unique function irrespective of temperature and mode of loading (Underwood
& Kim, 2010).

3.4. Semi-circular bending


The SCB test has been favoured by many researchers due to the ease of sample preparation
including cores removed from the field and the quick and simple testing procedure (Adamson,
Dempsey, & Mulmule, 1996; Li & Marasteanu, 2010). It is based on FM principles (Ander-
son, 2005) that offers the potential of assessing the cracking resistance of asphalt mixtures in
the laboratory in the design phase as well as in QA (quality assurance) testing activities. Crack-
ing mechanisms in asphalt mixtures have been studied since the 1960s. Various test methods
were proposed to predict the fracture behaviour of asphalt mixtures, including the single-edged
notched beam test, the SCB test, the IDT test fracture parameter, and the disk-shaped tension test
test. Each of these test procedures has unique advantages and disadvantages (Li & Marasteanu,
2010). The SCB test is based on FM and the J-integral, which considers the elasto-plastic/visco-
plastic relationships of asphalt mixtures rather than the stress intensity factor (K) commonly used
in linear elastic FM (Mull, Othman, & Mohammad, 2006).
6 W.S. Mogawer et al.

Mull et al. (2006) evaluated the use of the semi-circular bend (SCB) configuration to char-
acterise the fatigue crack propagation of asphalt mixtures. Three mixtures were evaluated
for fatigue susceptibility using SCB fracture parameters and scanning electron microscopy to
identify fatigue damage species associated with each mixture. The research found that the SCB
specimen is suitable for both static and fatigue fracture characterisation (Mull et al., 2006).
Elseifi, Mohammad, Ying, and Cooper (2012) utilised the SCB test at 25°C in the evaluation
of a number of asphalt mixtures, including mixtures with a high content of RAP. Results of the
experimental programme were used to validate a three-dimensional finite element model, which
was used to interpret and to analyse the failure mechanisms in the SCB test. The experimental
programme showed that the SCB test results successfully predicted the fracture performance of
the mixtures and was able to differentiate between them in terms of cracking resistance (Elseifi
et al., 2012).
Downloaded by [University of Western Ontario] at 13:45 13 December 2015

4. Experimental plan
In order to address the objectives of this study, an experimental plan was developed as shown in
Figure 1.

Figure 1. Experimental plan.


Road Materials and Pavement Design 7

5. Materials
All materials for this study (binders, aggregates, and RAP) were obtained from sources in the
state of Massachusetts. Rejuvenators were obtained from multiple sources in the USA.

5.1. Rejuvenators
A total of four asphalt rejuvenators were selected for evaluation in this study. Rejuvenators were
selected so that different types were represented as well as those already commonly used and
those based on emerging green technologies. Details about each rejuvenator are outlined in
Table 1.
The proposed dose of each asphalt rejuvenator was based on a previous study (Mogawer,
Booshehrian, Vahidi, & Austerman, 2013) which suggested a dosage of 0.5% by weight of
Downloaded by [University of Western Ontario] at 13:45 13 December 2015

total RAP in the mixture. Based on the design binder content and per cent asphalt binder in
the RAP, this dosage equalled 9.0% rejuvenator by weight of recycled binder. This dosage was
utilised throughout this study. Each rejuvenator was added directly to the pool of heated binder
immediately prior to mixing for each specimen fabricated.

5.2. Asphalt binder


For this study, two different asphalt binders were utilised. The binders consisted of a PG64-28
which is typically specified in the Northeast and the typically available PG58-28 softer grade
binder utilised for RAP mixtures. These binders were used in an attempt to evaluate each tests
sensitivity to the binder grade used and for analytical comparisons of the different tests/models.
The PG64-28 binder was used only for an all virgin material mixture. Based on the viscos-
ity of the binder, the mixture mixing temperature range was 161–165°C (322–329°F) and the
compaction temperature was 153–157°C (308–315°F).
A PG58-28 binder was used in an attempt to offset the potential mixture stiffening due to the
use of high percentage of RAP in the 50% RAP mixtures. The PG58-28 was selected because it
was the softest grade available. Based on the viscosity of the binder, the mixture mixing temper-
ature was 150°C (300°F) and the compaction temperature was 138°C (280°F). For comparison
purposes the PG58-28 was also utilised in the control mixture without RAP.

5.3. Aggregates
The aggregates utilised were from a crushed stone source in Wrentham, Massachusetts. Two
aggregate stockpiles were obtained: 9.5 mm crushed stone and stone dust. Each aggregate stock-
pile was tested to determine their properties which are shown in Table 2. Sieve analysis was
completed in accordance with AASHTO test method T11 “Standard Method of Test for Materi-
als Finer Than 75-μm (No. 200) Sieve in Mineral Aggregates by Washing” and T27 “Standard
Method of Test for Sieve Analysis of Fine and Coarse Aggregates” (AASHTO, 2011).

Table 1. Rejuvenator descriptions and details.

Rejuvenator ID Details

Aromatic oil Commonly used in pavement preservation activities


Paraffinic oil Commonly used in pavement preservation activities
Organic blend #1 Green chemistry product
Organic blend #2 Organic oils based
8 W.S. Mogawer et al.

Table 2. Virgin aggregate, RAP, and recovered RAP binder properties.

Sieve size (mm) 9.5 mm Stone dust RAP aggregates post ignition

19.0 100 100 100


12.5 99.4 100 100
9.5 93.8 100 100
4.75 29.7 99.7 76.8
2.36 5.2 83.7 57.6
1.18 2.8 57.1 43.3
0.600 2.3 38.6 31.1
0.300 2.1 24.9 19.8
0.150 1.8 15.9 12.1
0.075 1.5 10.9 8.3
Downloaded by [University of Western Ontario] at 13:45 13 December 2015

RAP binder content, % 5.6


RAP binder continuous grade 82.0–21.8
RAP binder performance grade 82–16

5.4. Reclaimed asphalt pavement


RAP was obtained from the same contractor as the aggregates. The RAP stockpile was fractioned
in the laboratory in order to meet the gradation requirements for this study. As shown in Table 2,
the binder content of the RAP material was determined using the ignition oven in accordance
with AASHTO T308 “Determining the Asphalt Binder Content of Hot Mix Asphalt (HMA) by
the Ignition Method” (AASHTO, 2011). The aggregates in the RAP remaining after ignition
were tested to determine their properties which are also shown in Table 2. The RAP binder
was extracted and recovered in accordance with AASHTO T164 “Standard Method of Test for
Quantitative Extraction of Asphalt Binder from Hot Mix Asphalt (HMA)” and AASHTO T170
“Standard Method of Test for Recovery of Asphalt Binder from Solution by Abson Method”
(AASHTO, 2011). As shown in Table 2, the continuous and performance grade of the recov-
ered RAP binder was determined in accordance with AASHTO R29 “Grading or Verifying
the Performance Grade of an Asphalt Binder” and AASHTO M320 “Standard Specification for
Performance-Graded Asphalt Binder” (AASHTO, 2011).

6. Description of mixture ageing procedures


For this study, four different ageing schemes were utilised as described in the following. For
mixture fatigue cracking performance evaluations, testing and analysis was conducted using
the short-term, long-term, and long-term plus cyclic pore pressure conditioning (CPPC) ageing
schemes presented.
For long-term ageing, it should be noted that all Superpave gyratory compacted (SGC) spec-
imens were aged in the compacted state prior to any cutting/trimming for a specific test. Only
the long-term aged flexural beam fatigue specimens were cut prior to long-term ageing. This was
due to the large weight 25 kg (55 lb) and dimensions 150 mm wide, 150 mm tall, and 450 mm
long (6 in. wide, 6 in. tall, 17.5 in. long) of the slabs from which the beams were cut. Ageing of
the smaller cut beams was done in an attempt to ensure thorough and consistent ageing between
all beams tested in this study.

6.1. Mixture design and volumetric verification ageing


For mixture design and volumetric verifications, specimens of each mixture were batched, mixed,
and aged at the respective compaction temperature previously noted for 2 h ± 5 min in a
Road Materials and Pavement Design 9

loose state in accordance with AASHTO R30 “Standard Practice for Mixture Conditioning of
Hot Mix Asphalt (HMA)” Section 7.1 “Mixture Conditioning for Volumetric Mixture Design”
(AASHTO, 2011). Specimens were stirred after 60 ± 5 min to maintain uniform conditioning.
After ageing, specimens were immediately compacted.

6.2. Short-term oven ageing


Short-term oven ageing (STOA) was conducted in accordance with AASHTO R30 “Stan-
dard Practice for Mixture Conditioning of Hot Mix Asphalt (HMA)” Section 7.2 “Short-Term
Conditioning for Mixture Mechanical Property Testing” (AASHTO, 2011). For STOA ageing,
specimens of each mixture were batched, mixed, and aged at 135 ± 3°C (275 ± 5°F) for 4
h ± 5 min in a loose state. Specimens were stirred after 60 ± 5 min to maintain uniform condi-
Downloaded by [University of Western Ontario] at 13:45 13 December 2015

tioning. After ageing, specimens were returned to the compaction temperature and immediately
compacted.

6.3. Long-term oven ageing


Long-term oven ageing (LTOA) was conducted in accordance with AASHTO R30 “Stan-
dard Practice for Mixture Conditioning of Hot Mix Asphalt (HMA)” Section 7.3 “Long-
Term Conditioning for Mixture Mechanical Property Testing” (AASHTO, 2011). For LTOA
ageing, specimens of each mixture were first subjected to STOA ageing as described previ-
ously. Next, the compacted STOA specimens were cooled for 16 ± 1 h. In accordance with
AASHTO R30 Subsection 7.3.4 “Long-term Conditioning of Prepared Test Specimens” the
compacted specimens were aged in an 85 ± 3°C (185 ± 5°F) oven for 120 ± 0.5 h (5 days).
After ageing, the specimens were allowed to cool at room temperature which took approx-
imately 16 h. The specimens were not handled or disturbed until after completely cooled
to room temperature.

6.4. LTOA + CPPC


The CPPC system was developed at the University of Florida to induce damage in specimens
due to combined effects of moisture and load (Birgisson, Roque, Page, & Wang, 2007; Isola
et al., 2014). It was determined that CPPC can induce internal pressure (stress) in a tensile mode
within the air voids that is similar to the effect of repeated load induced by traffic on mixtures
in the field. Specifically, pore water under pressure in mixtures can cause premature failure of
the mixture, through loss of adhesion between asphalt binder and aggregates (i.e. stripping of the
asphalt film) or through loss of cohesion within the asphalt binder or mastic. In this study, the
CPPC system was used for additional conditioning of asphalt mixtures after LTOA to simulate
the effects of moisture and cyclic internal pressure on changes in mixture fracture properties after
oxidative ageing.
The structural core of the system is a triaxial cell modified for CPPC of asphalt spec-
imens, which consists of two round stainless steel plates separated by posts or struts and
encased with a Plexiglas cylinder. Once sealed, the entire package creates an enclosed cav-
ity capable of being pressurised. Figure 2 shows the tabletop triaxial chamber containing cut
specimens for Superpave IDT tests. As can be seen, spacers were used in between specimens
to facilitate water infiltration and to protect the gauge points from being damaged during the
conditioning process.
The conditioning of the asphalt specimens takes place by exerting cyclic pore pressure on all
water accessible voids of the specimens. This is accomplished by vacuum-saturating specimens
10 W.S. Mogawer et al.
Downloaded by [University of Western Ontario] at 13:45 13 December 2015

Figure 2. Tabletop triaxial chamber.

and then placing them into the airtight, water-filled triaxial chamber. Deairated water is forced
into the chamber to build up pressure. This pressure is transferred to every surface with which
the water is in contact. When this pressure is cycled in the chamber at a constant frequency in
the form of a sine wave, it can simulate the pumping action that occurs in the field when vehicles
travel over wet pavement.
Prior to insertion in the chamber, the specimens were first subjected to a two-cycle saturation
process. Each cycle included a 15-min vacuum saturation period at 85 kPa ( ± 7.0 kPa) followed
by a 20-min submergence period at atmospheric pressure. No specific saturation levels were
targeted since each mixture has a unique void structure that may enhance or reduce its satura-
tion capacity. The specimens were then placed into the tabletop triaxial chamber, carefully filled
with water, and subjected to a combination of pore pressure cycles and temperature determined
during previous research conducted at the University of Florida (Birgisson et al., 2007). Specif-
ically, water pressure in a sine waveform at a frequency of 0.33 Hz and an amplitude of 69 kPa,
ranging from 34.5 to 172.5 kPa, was applied for 5800 cycles at 25°C. Immediately after CPPC,
specimens were kept in a water bath for two days at 10°C, the temperature used for Superpave
IDT tests.

7. Mixture design
The target gradation for the mixtures utilised in this study is shown in Table 3. This target grada-
tion was developed to meet the requirements for a 9.5 mm Superpave mixture in accordance
with AASHTO M323 “Superpave Volumetric Mix Design” and AASHTO R35 “Superpave
Volumetric Design for Hot Mix Asphalt” (AASHTO, 2011).
Control mixtures utilising all virgin materials were designed utilising the control PG64-28
binder and the softer PG58-28 binder for comparison purposes. Then RAP was incorporated to
replace a portion of the virgin materials in the control mixtures. RAP was added to replace 50%
of the mixture aggregates with RAP aggregates. The aggregate gradation for the control and
50% RAP mixtures was identical. The 50% RAP mixture was then designed utilising only the
softer PG58-28 binder. Because the gradations were identical for the mixtures developed, mixture
design verifications were performed for the 50% RAP mixtures incorporating each rejuvenator
Road Materials and Pavement Design 11

Table 3. Target mixture gradation and specification.

Sieve size Sieve size (mm) Target gradation for all mixtures Superpave 9.5 mm specification

1/2 12.5 100 100 min


3/8 9.5 98 90–100
No. 4 4.75 85 90 max
No. 8 2.36 58 32–67
No. 16 1.18 42 –
No. 30 0.600 27 –
No. 50 0.300 15 –
No. 100 0.150 9 –
No. 200 0.075 6.0 2–10
Binder content 6.5% –
Downloaded by [University of Western Ontario] at 13:45 13 December 2015

at the design binder content determined for the control mixture. Verifications were completed
assuming 100% contribution of the RAP binder.
To incorporate the RAP into the mixtures, a procedure that was used in a prior study utilising
similar materials was followed (Mogawer et al., 2013). This procedure was utilised to eliminate
moisture in the RAP stockpile material and to optimise the blending between the aged and virgin
binders in the mixture. The procedure steps were:

(1) The RAP was air dried until a constant mass was achieved, which typically took three to
five days.
(2) The RAP was further dried for two days at 60°C (140°F).
(3) The RAP was added to heated aggregate 2 h prior to adding the binder during the mixing
process.

The design ESALs for this project was selected as 0.3 to < 3 million which is consistent with
surface course mixtures in New England. The design Superpave gyratory compactive effort for
this ESALs level was Ndesign = 75 gyrations. The results of the mixture design indicated that
all mixtures met the volumetric requirements within typical production tolerances.

8. Dynamic modulus results


The baseline for understanding fatigue performance was establishing the stiffness behaviour of
each mixture. The dynamic modulus, |E*|, was measured according to the method outlined in
AASHTO TP79 “Standard Method of Test for Determining the Dynamic Modulus and Flow
Number for Hot Mix Asphalt (HMA) Using the Asphalt Mixture Performance Tester (AMPT)”
with the exception that the test temperatures were limited to 4°C, 20°C and 40°C for the mul-
tiple loading frequencies. The results are summarised in the mastercurves shown in Figure 3(a)
and 3(b). The amount of stiffening that occurs due to ageing is summarised by temperature in
Figure 4. The ageing ratio shown on the y-axis of this plot is calculated by taking the ratio of
|E*|LTOA to |E*|STOA . From this figure it can be seen that the relative amount of stiffening caused
by LTOA varies by mixture and rejuvenator type. It is also seen in Figure 3(a) and 3(b) that
the modulus from the rejuvenated mixtures shows a larger spread in the STOA case relative to
the LTOA case. Since there was no reduction in dynamic modulus from the STOA to LTOA
condition for mixtures with rejuvenators, this suggested for these rejuvenators there is no contin-
ued softening of the mixtures. Therefore, increased rutting susceptibility of these mixtures with
rejuvenators may not be a concern.
12 W.S. Mogawer et al.

(a)
Downloaded by [University of Western Ontario] at 13:45 13 December 2015

(b)

Figure 3. (a) Results from |E*| characterisation STOA. (b) Results from |E*| characterisation LTOA.

9. Fatigue cracking tests and results


This section outlines the testing details and results of each fatigue cracking test. Unless otherwise
specified, these tests were conducted at a 15°C (59°F) test temperature which represents the
intermediate temperature for the Northeast.

9.1. Four-point flexural beam fatigue test


For this test, fatigue specimens were cut from slabs. Slabs with dimensions of 150 mm wide,
150 mm tall, and 450 mm long (6 in. wide, 6 in. tall, and 17.5 in. long) were fabricated for each
mixture using the IPC Global Pressbox slab compactor. From each slab, beams with dimensions
of 63 mm wide, 50 mm tall, and 380 mm long (2.5 in. wide, 2 in. tall, 15 in. long) were cut
such that the sides had smooth faces. The air voids of the final cut specimens were 7 ± 1%.
Beam specimens were conditioned at the test temperature of 15°C (59°F) for at least 2 h prior to
testing.
Road Materials and Pavement Design 13
Downloaded by [University of Western Ontario] at 13:45 13 December 2015

Figure 4. Results from |E*| characterisation: ageing ratio at 10 Hz.

Each beam fatigue test was conducted in strain control mode at a loading frequency of 10 Hz
applied using a sinusoidal waveform. Specimens were tested at strain levels of 500 and 750 με.
The 500 and 750 με were selected because all mixtures lost 50% of their initial stiffness after
at least 10,000 cycles. The number of cycles to failure was determined by fitting an exponential
function to the flexural stiffness versus number of cycles and then evaluating the number cycles
that it took to decrease the initial stiffness by 50%. The beam fatigue testing results are shown in
Figure 5(a) and 5(b).
Comparing the properties of the 50% RAP mixture with no rejuvenator and those of the control
PG 58-28 mixture it appeared that the introduction of high RAP content resulted in a reduction in
the number of cycles to failure. All mixtures with rejuvenators generally exhibited higher number
of cycles to failure. Based on the ageing condition only, the beam fatigue data showed opposite
trends. At 500 με the 50% RAP LTOA aged specimens had better fatigue performance than the
STOA specimens. This trend was reversed for the 750 με level.

9.2. HMA-FM model


9.2.1. Mixture properties determined by the Superpave IDT
Results obtained using the Superpave IDT tests are presented in Figures 6 and 7 showing the
effect of the ageing levels and the effects of the high RAP content as well as rejuvenators on
mixture properties.

9.2.1.1. Effect of ageing levels. FE limits (FEf ), representing the fracture tolerance of the mix-
tures, are shown in Figure 6(a). It is clear that the FE limits of all mixtures were reduced after
LTOA. The FE limits of the three mixtures with no rejuvenator (the control PG 58-28, the PG
64-28, and the 50% RAP mixtures) further decreased after CPPC, while the FE limits of all four
mixtures with rejuvenators increased. The results of creep rate in Figure 6(b) showed that the
damage-related property of all mixtures greatly decreased due to LTOA, indicating a significant
reduction in the rate of damage accumulation of the mixtures.
The creep rates, however, remained about the same after CPPC. The results of resilient modu-
lus representing stiffness of the mixtures are presented in Figure 7(a). As can be seen, the resilient
14 W.S. Mogawer et al.

(a)
Downloaded by [University of Western Ontario] at 13:45 13 December 2015

(b)

Figure 5. (a) Fatigue cracking results from the flexural beam fatigue – 500 με. (b) Fatigue cracking results
from the flexural beam fatigue – 750 με.

modulus of all mixtures increased due to LTOA. This property was affected by CPPC to a less
extent, but varied in both directions. Similarly, the tensile strength of all mixtures increased due
to LTOA, and in general, slightly increased after CPPC (Figure 7(b)). Overall, the conditioning
levels had greater effects on FE limit and creep rate than on the other two properties. It is impor-
tant to note that the results of a separate study on laboratory conditioning procedures (Roque
et al., 2010) showed that CPPC was able to induce changes in fracture properties consistent with
those observed in field cores, while oxidation alone (STOA or LTOA) was not able to induce
relevant changes. Therefore, it was determined that the predicted cracking performance based on
the properties obtained after CPPC should be used in the final ranking of all the mixtures.
Road Materials and Pavement Design 15

(a)
Downloaded by [University of Western Ontario] at 13:45 13 December 2015

(b)

Figure 6. (a) Mixture properties: FE limit. (b) Mixture properties: creep rate determined using Superpave
IDT at three ageing levels.

9.2.1.2. Effects of high RAP content and rejuvenators. Comparing the properties of the 50%
RAP mixture with no rejuvenator and those of the control PG 58-28 mixture (see Figures 6(a),
6(b), 7(a), and 7(b)) it appeared that the introduction of high RAP content resulted in fairly
large reduction in both FE limit and creep rate, but smaller increase in both stiffness and tensile
strength. In fact, after incorporating 50% RAP, the properties of the control PG58-28 mixture
became more similar to the PG 64-28 mixture than the other mixtures. All mixtures with rejuve-
nators generally exhibited higher FE limit, higher creep rate, lower stiffness, and lower strength
than the rest of the mixtures.

9.2.2. Design pavement structure and responses


In order to compare the performance of different mixtures, a typical three-layer pavement
structure including an AC layer, a crushed stone base, and a subgrade (see Table 4), and an
intermediate traffic level (10 million ESALs over a 20-year period) were selected.
16 W.S. Mogawer et al.

(a)
Downloaded by [University of Western Ontario] at 13:45 13 December 2015

(b)

Figure 7. (a) Mixture properties: resilient modulus. (b) Mixture properties: tensile strength determined
using Superpave IDT at three ageing levels.

Table 4. Design layer thickness for a typical three-layer pavement structure.

Layers Modulus, MPa (ksi) ai mi Design layer thickness, mm (in.)

AC MR 0.40 – 178 mm (7 in.)


Base 69 MPa (10 ksi) 0.14 1.2 267 mm (10.5 in)
Subgrade 276 MPa (40 ksi) – – –

The layer thickness values were then designed according to the AASHTO procedure (1993),
assuming a reliability of 95%, a standard deviation of 0.4, and a design serviceability loss of
2.0. As shown in Table 4, given the typical base and subgrade moduli, 276 and 69 MPa, respec-
tively (40 and 10 ksi, respectively), structural coefficient (ai ), and drainage coefficient (mi ), the
AC layer and base thickness values were determined to be 178 and 267 mm, respectively (7
and 10.5 in, respectively). The designed pavement structure was used along with measured mix-
ture properties to predict its relative cracking performance, determined as N f , the number of
load repetitions required to initiate and propagate a crack to a length of 50 mm (2 in.) (Roque
Road Materials and Pavement Design 17

et al., 2004). At a temperature of 10°C (a fairly critical condition for cracking) and for a standard
18-kip single-axle load, the calculated tensile stresses at the bottom of the AC layer are in the
range of 827–1034 kPa (120–150 psi) for AC modulus determined at the STOA condition, or in
the range of 897–1103 kPa (130–160 psi) for AC modulus obtained at both the LTOA and CPPC
conditions.

9.2.3. Results of HMA-FM model


Figure 8(a) presents the predicted evolution of crack length with increasing load repetitions for
the control PG58-28 mixture with properties determined at all three ageing levels. As can be
seen, the control mixture at LTOA had better performance than at the other two ageing levels,
in terms of larger numbers of loads to both crack initiation and crack propagation to 2 in. Also
Downloaded by [University of Western Ontario] at 13:45 13 December 2015

shown in Figure 8(a) are the calculated ER values using Equation (1), which support that the con-
trol mixture had the best performance at the LTOA ageing level. However, it should be noted that

(a)

(b)

Figure 8. (a) Model prediction: number of loads resulting in a 50 mm (2 in.) long crack (N f ) for the
PG58-28 control mixture at three ageing levels. (b) Model prediction: normalised N f with respect to that of
the control mixture at STOA.
18 W.S. Mogawer et al.

(a)
Downloaded by [University of Western Ontario] at 13:45 13 December 2015

(b)

Figure 9. (a) Model prediction: comparing normalised N f at LTOA. (b) Model prediction: comparing
normalised N f at LTOA + CPPC.

both the HMA-FM model and associated ER criterion are recommended for use in relative com-
parison of performance at a similar ageing level, because the healing potential is not considered
in the current form of the model, which can substantially affect the cracking performance (Zou,
Roque, & Byron, 2012). Therefore, the predicted performance of all mixtures was compared at
individual ageing levels subsequently. For simplicity, the predicted performance indicators for all
mixtures (both ER and N f ) were normalised with respect to the predicted values for the control
PG58-28 mixture. The final results are presented in Figures 8(b), 9(a), and 9(b) for the STOA,
LTOA, and CPPC ageing levels, respectively.
In general, the results showed that for all three ageing levels and in terms of both ER and N f
predictions, the PG64-28 mixture had the best performance, and the 50% RAP mixture with no
rejuvenator outperformed the control PG58-28 mixture. The performance (both ER and N f ) of
high RAP mixtures with rejuvenators were generally worse than the control PG 58-28 mixture
except for the CPPC ageing level, at which better performance was observed.
The ER and N f predictions generally provided consistent rankings among high RAP mixtures
with rejuvenators for both LTOA and CPPC ageing levels. At LTOA, the mixture with Paraffinic
Road Materials and Pavement Design 19

Oil and the one with Organic Blend #2 appeared to outperform the other two mixtures modified
with different rejuvenators. At CPPC, Organic Blend #2 appeared to be the best rejuvenator that
improved the cracking performance of the high RAP mixture. Surprisingly, the ER and N f ranked
differently the performance of mixtures with rejuvenators at the STOA ageing level. It should be
noted that the coefficients of the ER equation were calibrated based on measured properties and
observed performance of 22 field test sections (Roque et al., 2004). When mixture properties
beyond the range of properties used in the calibration are encountered, the ER equation may lead
to erroneous results. The N f results are directly predicted by the HMA-FM model, which are not
affected by the data sets used in the calibration of the ER equation. Therefore, the N f results are
more accurate than the ER results. In this specific case, it was found that the creep rates of all
mixtures modified with rejuvenators are far beyond the range used in the calibration, that is, 0.2–
6.2 ( × 10−3 1/Gpa/s), which appeared to desensitise the ER results to the creep rates. According
Downloaded by [University of Western Ontario] at 13:45 13 December 2015

to the N f results, the mixture with Organic Blend #2 outperformed the mixtures modified with
other rejuvenators at STOA.

9.3. SVECD model


The uniaxial tests were conducted in actuator controlled tension-only displacement mode
(pull–pull). In this test LVDTs are mounted to the sample surface to measure and record the
on-specimen strains, which are used in the model characterisation. The initial strain levels for
these tests were generally between 250 and 550 με (peak-to-peak). The exact strain levels used
for each test were chosen based on guidance provided in AASHTO TP109 so that each mixture
would have tests with failure at approximately 10,000, 50,000, and 100,000 cycles. These precise
targets were not always met; nevertheless, the strain levels chosen produced a range in the order
of magnitude of failure cycles for each mixture.
Once completed, the fatigue test results were coupled with the linear viscoelastic characterisa-
tion results to derive the damage characteristic curve for each mixture according to the procedure
presented by Underwood and Kim (2010). For the mixtures in this study, this characteristic func-
tion was described using the exponential analytical form shown in Equation (2), where a and b
are the fitting coefficients.
C = exp(aS b ). (2)
Failure was defined in the experiments as the cycle when the apparent phase angle showed a
rapid decrease. This point has been found to coincide with the onset of substantial load signal dis-
tortion and the appearance of surface cracks on the AC specimen. The model failure criteria was
calibrated to this experimentally determined N f according to the average released pseudo strain
criterion developed by Sabouri and Kim (2014). Subsequent to these characterisation procedures,
the SVECD model was used to predict the number of cycles to failure via Equation (3).

fred × 2α SNf
b −α
Nf = (−abS b−1 eaS ) (dS). (3)
[(ε0,pp )((β + 1)/2)(|E ∗ |)] K1

0

The predicted fatigue relationships for the study materials are shown in Figure 10 separated
by ageing, (a) STOA (b) LTOA. What is immediately apparent from this graph is the differences
in slope between the 50% RAP + Paraffinic oil sample and the other mixtures. The cause of
this behaviour can be traced back to a greater sensitivity of the failure criteria to the rate of
change in the average pseudo strain energy, GR , than the other mixtures. It should be noted that
the experimental data for this rejuvenator, although gathered over a more narrow range of strain
levels also showed a greater sensitivity to strain than the other cases.
20 W.S. Mogawer et al.

(a)
Downloaded by [University of Western Ontario] at 13:45 13 December 2015

(b)

Figure 10. (a) Predicted fatigue life from SVECD model STOA. (b) Predicted fatigue life from SVECD
model LTOA.

It is also interesting to observe that the failure curve for the 50% RAP with no rejuvenator
case is actually expected to outperform the control mixtures at lower strain amplitudes. These
behaviours are supported with the trends in the axial experiments; however, the predicted curves
do amplify the difference somewhat. One possible cause of this behaviour may be the use of only
two middle failure tests to define the failure criteria for this condition. Data at 500 and 750 με
have been extracted from these plots and summarised in Figure 11.
The data in Figure 11 indicated that generally the mixture performed better after STOA ageing
than LTOA ageing. This trend matched the beam fatigue results at the 750 με level, but not the
500 με level. Overall comparing the SVECD predicted results to those from the beam fatigue
test it is found that the rankings were similar. The largest exception to this case is the Control
PG64-28 case. These results are truly unexpected. The beam fatigue tests suggest relatively little
difference between the two control cases whereas the SVECD model suggests a reduction by
nearly 60–70% in the fatigue life. Reasons for the differences in behaviour are not immediately
Road Materials and Pavement Design 21

(a)
Downloaded by [University of Western Ontario] at 13:45 13 December 2015

(b)

Figure 11. (a) SVECD analysis results -500 με. (b) SVECD analysis results -750 με.

clear and require further investigation. It is noted, however, that this difference cannot be asso-
ciated with the SVECD model itself since the experimental data used in model characterisation
also demonstrated noticeably inferior performance from the PG64-28 asphalt mixture.

9.4. SCB test


Cracking potential was evaluated using the SCB test procedure, based on FM principles (Ander-
son, 2005) and suggested by Mohammad and co-workers (Wu et al., 2005), Figure 12. The
critical strain energy release rate, also called the critical value of J -integral (J c ) was used to
22 W.S. Mogawer et al.
Downloaded by [University of Western Ontario] at 13:45 13 December 2015

Figure 12. SCB test setup.

describe the mixture’s resistance to fracture:


 
U1 U2 1
Jc = − , (4)
b1 b2 a2 − a1

where J c is the critical strain energy release rate (kJ/mm2 ), b the sample thickness (mm), a the
notch depth (mm), U the strain energy to failure (N mm), and dU/da the change of strain energy
with notch depth.
To determine the critical value of J -integral (J c ) using Equation (4), semi-circular specimens
with at least two different notch depths should be tested to determine the change of strain energy
with notch depth (dU/da). In this study, three notch depths of 25.4, 31.8, and 38 mm were tested
to increase the accuracy of slope calculation (dU/da) by fitting a regression line to the change of
strain energy with notch depth.
The semi-circular specimen was loaded monotonically until fracture failure under a constant
cross-head deformation rate of 0.5 mm/min. in a three-point bending load configuration. The load

Figure 13. SCB test results.


Road Materials and Pavement Design 23

(a)
Downloaded by [University of Western Ontario] at 13:45 13 December 2015

(b)

Figure 14. (a) Comparative analysis of normalised test results- STOA ageing 500 με. (b) Comparative
analysis of normalised test results- STOA ageing 750 με.

and deformation were continuously recorded. This test was performed at a temperature of 15°C.
Specimens were compacted in the SGC to an air void level of 7 ± 0.5%. Triplicate specimens
were utilised for this test. In general, the coefficient of variation was within 15% for the samples
tested. High J c values are desirable for fracture-resistant mixtures.
Figure 13 presents the calculated critical fracture resistance (J c ) values for the asphalt mix-
tures evaluated in the study. Control mixture containing PG 64-28 binder exhibited better J c
values than the control mixture containing PG 58-28 binder at STOA condition; however, at
LTOA condition, both mixtures possessed similar J c values. This indicates that the LTOA con-
dition affected control mixture PG 64-28 more than the one with PG 58-28. Mixtures containing
50% RAP and no rejuvenator; 50% RAP and aromatic oil; and 50% RAP and organic blend 1
showed similar fracture resistance performances to control mixture containing PG 64-28. On the
other hand, mixtures containing 50% RAP and paraffinic oil, and 50% RAP and organic blend
2 presented similar fracture resistance performances to control mixture containing PG58-28. It
24 W.S. Mogawer et al.

(a)
Downloaded by [University of Western Ontario] at 13:45 13 December 2015

(b)

Figure 15. (a) Comparative analysis of normalised test results- LTOA ageing 500 με. (b) Comparative
analysis of normalised test results- LTOA ageing 750 με.

is worth noting that the addition of aromatic oil rejuvenator did improve the fracture resistance
at LTOA condition when compared to control mixture PG64-28. On the other hand, paraffinic
oil and organic blend 2 rejuvenators did not improve the mixture resistance to fracture at both
STOA and LTOA conditions.

10. Comparative analysis of the different fatigue tests


An objective of this study was to evaluate whether different fatigue tests show the same trends.
In order to accomplish this objective, the results from each test for the PG64-28 control mixture
were normalised. A performance indicator ratio was calculated for each test result by dividing
it by the value for the control PG64-28 mixture. Therefore, if a specific test indicated for a
particular mixture the same performance as the PG64-28 mixture, the performance indicator ratio
was a value of 1.0. The control PG64-28 mixture was selected because it utilised the typically
specified binder in the state of Massachusetts. The asphalt binders, aggregates, and RAP for the
Road Materials and Pavement Design 25

mixtures were obtained from a location in Massachusetts. The aim was to develop a 50% RAP
mixture that performed similar to a typical mixture, which in this study was the PG64-28 control
mixture.
The performance indicator ratios are shown in Figures 14 and 15 for STOA and LTOA, respec-
tively. Figures 14(a) and 15(a) show specific performance indicator ratios for beam fatigue and
SVECD at a 500 με while all other indicator ratios remain constant based on the STOA age-
ing condition. Similarly, Figures 14(b) and 15(b) show performance indicator ratios for the beam
fatigue and SVECD at 750 με while all other indicator ratios remain constant based on the LTOA
ageing condition.
If two tests were comparable their performance indicator ratios would have the same trends
for the same mixtures tested, but not necessarily the same magnitude. Overall, the performance
indicator ratios did not agree among the various methodologies. Furthermore, the performance
Downloaded by [University of Western Ontario] at 13:45 13 December 2015

indicator ratios did not show universal agreement in the trends of the mixtures, meaning whether
the ratio went up or down compared to a ratio of 1.0 for the PG64-28 control mixture. Also,
the repeated load cyclic tests (beam fatigue and SVECD) exhibited a wide range of performance
compared to the constant rate fatigue tests and analysis (HMA-FM models and SCB). Further-
more, as shown by the performance indicator ratios, the two repeated cyclic load tests (beam
fatigue and SVECD) often provided large increases in the performance of the 50% RAP mix-
ture when a rejuvenator was added. The ratios for these mixtures were very high compared to
the ratios for PG58-28 and PG64-28 mixtures. It is hypothesised that this means blending in
the binders was incomplete, causing these mixtures to have areas of binder with relatively low
stiffness. These two tests are sensitive to mixture stiffness. However, the very high performance
indicator ratios provided by some of the SVECD tests strongly suggest that this test might be
overly sensitive to stiffness. If this hypothesis concerning blending is correct, in general the con-
stant rate fatigue tests and analyses were not sensitive to these areas of low stiffness. Repeated
loads must be used to measure this effect. The reason why some of the constant rate fatigue
tests and analyses provided no change, or even a decrease, in the performance of the 50% RAP
mixture when a rejuvenator was added is not known.
Field trials are needed to determine the actual ranking trend that should be obtained in the
laboratory tests. This may be the only way to separate and evaluate the quality of the results
obtained by these methods.

11. Conclusions
Based on the analysis of the data collected, the following conclusions were made:

• The ageing ratio calculated as |E*|LTOA to |E*|STOA indicated that the relative amount
of stiffening caused by LTOA varies by mixture and rejuvenator type. It was seen that
the moduli from the rejuvenated mixtures showed larger variations relative to the 50%
RAP mixture with no rejuvenators in the STOA case relative to the LTOA case. Larger
variations would have been indicative of continuous softening after LTOA. This finding
may suggest that these rejuvenators, after exposure to long-term ageing, will not continue
to rejuvenate or soften the mixtures.
• The flexural beam fatigue tests indicated that mixtures with rejuvenators generally
exhibited higher number of cycles to failure. LTOA aged specimens had better fatigue
performance than the STOA specimens at a 500 με test level. This trend was reversed
for the 750 με test level. This led to the conclusion that a strain level should be used
that is representative of the actual loading conditions where these mixtures will be
placed.
26 W.S. Mogawer et al.

• In general, the HMA-FM results showed that for all three levels of ageing in terms of both
ER and N f predictions that the high RAP mixtures with rejuvenators generally performed
worse than the control PG58-28 mixture (except for the CPPC conditioning level for which
better performance was observed). This was opposite to the trend observed in the flexural
beam fatigue.
• Generally, the SVECD analysis illustrated that LTOA caused a decrease in number of
cycles to failure at the different strain levels relative to the STOA. The decrease is function
of the mixture and rejuvenator type. This agreed with the flexural beam fatigue results at
the 750 με level but not at 500 με.
• The data from the FM-based SCB test indicated the J c is function of rejuvenator type and
ageing. The J c value dropped for all mixtures except the control mixture with PG58-28
after the long-term ageing. Mixtures containing 50%RAP and no rejuvenator; 50% RAP
Downloaded by [University of Western Ontario] at 13:45 13 December 2015

and aromatic oil; and 50% RAP and organic blend 1 showed similar fracture resistance
performances to control mixture containing PG 64-28 after STOA and LTOA. On the
other hand, mixtures containing 50% RAP and paraffinic oil, and 50% RAP and organic
blend 2 presented similar fracture resistance performances to control mixture containing
PG58-28. It is worth noting that the addition of aromatic oil rejuvenator did improve the
fracture resistance at LTOA condition when compared to control mixture PG 64-28. On
the other hand, paraffinic oil and organic blend 2 rejuvenators did not improve the mixture
resistance to fracture at both STOA and LTOA conditions.
• Both FM-based J c values and HMA-FM results showed that the PG64-28 mixture had the
best performance, and the 50% RAP mixture with no rejuvenator outperformed the control
PG58-28 mixture after STOA.
• Comparison of the performance indicator ratios showed that the fatigue tests did not
provide universal agreement. Also, the two repeated load cyclic tests (beam fatigue and
SVECD) exhibited a wide range of performance compared to the constant rate fatigue
tests and analysis. The reason for these variations is not known at this time and requires
further study. The repeated load cyclic tests indicated the mixtures with RAP and reju-
venators had the best fatigue performance, except when using aromatic oil, whereas, the
constant rate fatigue tests and analysis generally provided no constant trend in fatigue per-
formance among the mixtures. The reason for this trend is not known and requires further
study.
• Field trials are needed to determine the actual ranking trend that should be obtained in the
laboratory tests. In addition to field trials, chemical and rheological analyses of the binders
and rejuvenators are recommended to provide an insight into the mixtures’ performance.
This may help to separate and evaluate the quality of the results obtained by these methods.
• Future work is needed to establish the causes of the differences in trends among the fatigue
tests. Because of these differences no particular test can be recommended at this time.

Disclosure statement
No potential conflict of interest was reported by the authors.

References
Abbas, A., Chai, B. C., Masad, E., & Papagiannakis, T. (2002). The influence of laboratory aging method
on the rheological properties of asphalt binders. American Society for Testing and Materials (ASTM)
Journal of Testing and Evaluation, 30(2), 171–176.
Adamson, R. M., Dempsey, J. P., & Mulmule, S. V. (1996). Fracture analysis of semi-circular and semi-
circular-bend geometries. International Journal of Fracture, 77, 213–222.
Road Materials and Pavement Design 27

American Association of State Highway and Transportation Officials. (1993). AASHTO guide for design of
pavement structures. Washington, DC: Author.
American Association of State Highway and Transportation Officials. (2011). Standard specifications
for transportation materials and methods of sampling and testing (31st ed.). Washington, DC:
Author.
American Society for Testing and Materials. (2009). Annual book of ASTM standards. ASTM International,
Section 4, 04.03.
Anderson, T. J. (2005). Fracture mechanics – fundamentals and application (3rd ed.). Boca Raton, FL:
CRC Press, Taylor and Francis Group.
Bell, C. A., Wieder, A. J., & Fellin, M. J. (1994). Laboratory aging of asphalt-aggregate mixtures:
Field validation (Report SHRP-A-390). Washington, DC: Strategic Highway Research Program,
Transportation Research Board, National Research Council.
Birgisson, B., Roque, R., Page, G. C., & Wang, J. (2007). Development of new moisture conditioning pro-
cedure for hot-mix asphalt. Transportation Research Record: Journal of the Transportation Research
Downloaded by [University of Western Ontario] at 13:45 13 December 2015

Board, (2001), 46–55.


Daniel, J. D., Gibson, N., Tarbox, S., Copeland, A., & Andriescu, A. (2013). Effect of long term aging on
rap mixtures: Laboratory evaluation of plant produced mixtures. Journal of the Association of Asphalt
Paving Technologists, 82, 327–365.
Elseifi, M., Mohammad, L. N., Ying, H., & Cooper, S. (2012). Modeling and evaluation of the cracking
resistance of asphalt mixtures using the semi-circular bending test at intermediate temperature. Journal
of Road Materials and Pavement Design, 13(S1), 124–139.
Glover, C. J., Davison, R. R., Domke, C. H., Ruan, Y., Juristyarini, P., Knorr, D. B., & Jung, S. H.
(2005). Development of new method for assessing asphalt binder durability with field validation (Report
FHWA/TX-05/1872-2). Austin: Texas Department of Transportation.
Isola, M., Chun, S., Roque, R., Zou, J., Koh, C., & Lopp, G. (2014). Development and evaluation of labora-
tory conditioning procedures to properly simulate mixture property changes in the field. Transportation
Research Record: Journal of Transportation Research Board. doi:10.3141/2447-08
Kandhal, P. S., & Chakraborty, S. (1996). Effect of asphalt film thickness on short-and long-term aging of
asphalt paving mixtures (National Center for Asphalt Technology (NCAT) Report 96-1).
Kim, Y. R., Baek, C., Underwood, B. S., Subramanian, V., Guddati, M. N., & Lee, K. (2008). Application of
viscoelastic continuum damage model based finite element analysis to predict the fatigue performance
of asphalt pavements. KSCE Journal of Civil Engineering, 12(2), 109–120.
Li, X. J., & Marasteanu, M. O. (2010). Using semi circular bending test to evaluate low temperature fracture
resistance for asphalt concrete. Journal of Experimental Mechanics, 50, 867–876.
Mogawer, W. S., Booshehrian, A., Vahidi, S., & Austerman, A. J. (2013). Evaluating the effect of rejuve-
nators on the degree of blending and performance of high RAP, RAS, RAP/RAS mixtures. Journal of
the Association of Asphalt Paving Technologist, 82, 403–436.
Molenaar, A., Hagos, E., & Van de Vev, M. (2010). Effects of aging on the mechanical characteristics of
bituminous binders in PAC. Journal of Materials in Civil Engineering, 22(8), 779–787.
Mull, M. A., Othman, A., & Mohammad, L. (2006). Fatigue crack growth analysis of HMA employing
the semi-circular notched bend specimen. Journal of the Transportation Research Board, 85th Annual
meeting compendium of papers CD-ROM, Washington, DC.
Park, H. J., & Kim, Y. R. (2013). Investigation into top-down cracking of asphalt pavements in North
Carolina. Transportation Research Record: Journal of the Transportation Research Board, (2368),
45–55.
Roque, R., Birgisson, B., Drakos, B., & Dietrich, B. (2004). Development and field evaluation of energy-
based criteria for top-down cracking performance of hot mix asphalts. Journal of the Association of
Asphalt Paving Technologists, 73, 229–260.
Roque, R., Birgisson, B., Sangpetngam, B., & Zhang, Z. (2002). Hot mix asphalt fracture mechanics:
A fundamental crack growth law for asphalt mixtures. Journal of the Association of Asphalt Paving
Technologist, 71, 816–827.
Roque, R., Buttlar, W. G., Ruth, B. E., Tia, M., Dickson, S. W., & Reid, B. (1997). Evaluation of SHRP indi-
rect tension tester to mitigate cracking in asphalt pavements and overlays. Final report of the Florida
Department of Transportation, University of Florida, Gainesville, FL.
Roque, R., Zou, J., Kim, Y. R., Baek, C. M., Thirunavukkarasu, S., Underwood, B. S., & Guddati,
M. N. (2010). NCHRP 1-42A: Top-down cracking of hot mix asphalt layers: Models for initiation
and propagation (Final report). Washington, DC: National Cooperative Highway Research Program,
Transportation Research Board, National Research Council.
28 W.S. Mogawer et al.

Sabouri, M. R. & Kim, Y. R. (2014). Development of a failure criterion for asphalt mixtures under dif-
ferent modes of fatigue loading. Washington, DC: Transportation Research Record Journal of the
Transportation Research Board, Transportation Research Board of the National Academies.
Transportation Research Board. (2012, January). Applications of advanced models to understand behavior
and performance of asphalt mixtures (Transportation Research Circular E-C 161). Washington, DC:
Transportation Research Board, National Research Council.
Underwood, B. S., & Kim, Y. R. (2010). Improved calculation method of damage parameter in viscoelastic
continuum damage model. International Journal of Pavement Engineering, 11(6), 459–476.
Underwood, B. S., Kim, Y. R., Savadatti, S., Thirunavukkarasu, S., & Guddati, M. N. (2009). Simpli-
fied fatigue performance modeling of ALF pavements using VECD + 3-D finite element modeling. 7th
International RILEM symposium on advanced testing and characterization of bituminous materials,
Rhodes, Greece.
Wu, Z., Mohammed, L. N., Wang, L. B., & Mull, M. A. (2005). Fracture resistance characterization of
superpave mixtures using the semi-circular bending test. Journal of ASTM International, 2(3), 127–141.
Downloaded by [University of Western Ontario] at 13:45 13 December 2015

Zhang, Z., Roque, R., Birgisson, B., & Sangpetngam, B. (2001). Identification and verification of a suitable
crack growth law. Journal of the Association of Asphalt Paving Technologist, 70, 206–241.
Zou, J., Roque, R., & Byron, T. (2012). Effect of HMA aging and potential healing on top-down cracking
using HVS. International Journal of Road Materials and Pavement Design, 13(3), 518–533.

S-ar putea să vă placă și