You are on page 1of 8

Construction and Building Materials 165 (2018) 264–271

Contents lists available at ScienceDirect

Construction and Building Materials


journal homepage: www.elsevier.com/locate/conbuildmat

Durability assessment of asphalt binder


Ezequiel Angius, Haibo Ding, Simon A.M. Hesp ⇑
Department of Chemistry, Queen’s University, Kingston, Ontario K7L 3N6, Canada

h i g h l i g h t s

 AASHTO M 320 lacks sensitivity and passes many binders of low durability.
 DENT testing separates binders with a high degree of sensitivity in ductile failure.
 Grade losses due to extended PAV aging are correlated with losses in the extended BBR.
 Limiting EBBR grades are highly correlated with limiting phase angle temperatures.

a r t i c l e i n f o a b s t r a c t

Article history: Test results are presented for 24 asphalt binders with the aim to compare various measures of durability.
Received 1 October 2017 Surprisingly high correlations were found between disparate measures of durability but practical consid-
Received in revised form 3 January 2018 erations make certain tests preferable. A comparison is provided between AASHTO M 320 grades, critical
Accepted 5 January 2018
crack tip opening displacements, extended bending beam rheometer grades and grade losses, differences
in limiting temperatures (DTc), and simple limiting phase angle temperatures (T(d)), on aged residues.
The limiting phase angle temperature appears to be a convenient surrogate property that accurately pre-
Keywords:
dicts thermoreversible aging tendencies. Several binders performed poorly, likely because they contained
Double-edge-notched tension test (DENT)
Critical crack tip opening displacement
unwarranted reclaimed asphalt.
(CTOD) Ó 2018 Elsevier Ltd. All rights reserved.
Extended bending beam rheometer test
(EBBR)
Phase angle
Asphalt specification
Thermal and fatigue cracking

1. Introduction ( 34 °C to 36 °C) Superpave grades from six different suppliers.


In early 2004, the air temperature reached a record 48 °C, and
Shortly after the implementation of the SuperpaveÒ specifica- the pavement surface reached the low design of 34 °C twice. (In
tions in the late 1990s, the Ministry of Transportation of Ontario 2005, it dropped below 30 °C on six additional days.) In spring
(MTO) became concerned that roads were aging prematurely at 2005, the pavement had cracked to moderate degrees in four sec-
significant cost to taxpayers. In 2003, MTO commissioned a pave- tions. As of 2011, section 1 remained largely free of distress with
ment trial on Highway 655, north of Timmins, to validate the var- only few cracks (mostly through core holes) because it was made
ious test methods that were being proposed to improve the with a durable binder from Lloydminster [1]. The worst performer,
performance grading of asphalt [1]. The 3.5 km trial was built late section 4, was riddled with cracks and many potholes. Fig. 1 pro-
in the season with binders of nearly identical low-temperature vides representative crack maps and photographs for both sec-
tions. Performance as seen in section 1 is rare today while the
Abbreviations: BBR, bending beam rheometer; CTOD, critical crack tip opening state represented by section 4 is rather common. The other four
displacement; DCM, dichloromethylene; DENT, double-edge-notched tension; DSR, sections showed performance somewhere in between these two
dynamic shear rheometer; FHWA ALF, Federal Highway Administration accelerated extremes. In 2014, as measured by MTO’s Automated Road Ana-
loading facility; LS, laboratory standard; MTO, Ministry of Transportation of lyzer, section 1 had remained largely free of thermal cracking dis-
Ontario; PAV, pressure aging vessel; PPA or P31, polyphosphoric acid; PG,
tress, five sections had about 1000 m of cracks each, while the
performance grade; RAP, reclaimed asphalt pavement; REOB, recycled engine oil
bottoms; RTFO, rolling thin film oven; SHRP, Strategic Highway Research Program. worst performer, section 4, had 2160 m of cracks, many potholes,
⇑ Corresponding author. and will soon be in need of significant end-of-life care (MTO,
E-mail address: simon@chem.queensu.ca (S.A.M. Hesp).

https://doi.org/10.1016/j.conbuildmat.2018.01.037
0950-0618/Ó 2018 Elsevier Ltd. All rights reserved.
E. Angius et al. / Construction and Building Materials 165 (2018) 264–271 265

Fig. 1. Representative 2011 crack maps and photographs from 2003 Highway 655 pavement trial: (a) section 1 and (b) section 4. Note: Both crack maps are for randomly
selected 50 m stretches of pavement. Photographs were not necessarily taken in areas where cracks were mapped.

Unpublished Data, 2014). Hence, this trial has shown rather starkly The air blown (oxidized) residues lack colloidal stability since addi-
that the current AASHTO M 320 specification lacks accuracy and tional asphaltenes introduced through the air blowing process are
needs to be improved to provide a more correct measure of poorly accommodated and form gel-type structures that hinder
durability. stress relaxation during winter and spring. However, these binders
The term durability is defined by the Oxford Online dictionary do well in the grading process since the structure that forms the gel
as follows: ‘‘The quality of being durable. Capability of withstand- network takes time to consolidate when cold, and is therefore not
ing decay or wear.” [2]. In the context of asphalt science and tech- captured in most current specification tests that prescribe only
nology this means the ability to retain certain design properties minimal conditioning.
that allow a flexible pavement to remain without significant dis- Finally, the recent ‘‘greening” of the asphalt industry through
tress for at least the design life and preferably longer. It is well doc- the increased use of reclaimed asphalt pavement (RAP), recycled
umented in the literature that pavements of ostensibly the same engine oil bottoms (REOB), asphalt shingle waste, and other
design, subgrade, climate and traffic can show a wide range of unknowns has introduced performance challenges. Added to the
lifespans. This is often due to differences in durability of material problem is the increased use of wax-based warm asphalt technolo-
properties that are overlooked by current specification protocols gies that facilitate compaction. Early pavement failures have
[1]. An ideal durability test for asphalt binder can be defined as become quite common and the causes for these can nearly always
one that measures properties showing a high degree of correlation be traced back to deficient materials and processes. The incorrect
with performance in old age. It might be added that for practical and irresponsible use of certain recycled materials is costing user
reasons, the test should be easy to conduct on a small amount of agencies more for less performance than in years past [12].
material and be highly reproducible. An accurate test with a high Our long-term vision is that paved roads last the lifespan
degree of reproducibility lessens the risk for both users and pro- expected from quality Western Canadian asphalt instead of the sig-
ducers of asphalt binder. nificantly reduced lifespans currently obtained in Ontario and else-
Certain asphalts high in naphthene aromatics, and low in paraf- where [11–13]. Longer-lasting roads with less need for repair cause
fin and asphaltenes, can accommodate a significant amount of less congestion, delay and danger for users. To achieve this, we will
additional asphaltenes from oxidation and harden slightly over develop more comprehensive, theoretically-sophisticated tests to
time [3]. California Valley crude produces such binder since it is assess durability in asphalt. Our test methods will be available to
largely naphthenic and its residue is low in asphaltenes. Binders producers and user agencies in Ontario and beyond. Our short-
produced from aromatic crudes can be oxidized for long times term objective is to develop a practical test protocol that can be
without much hardening. Many Western Canadian naphthenic done in a reproducible manner, in a short amount of time, with lit-
crudes can produce material that is highly resistant to aging [4]. tle material. Placing a minimum durability requirement in asphalt
In contrast, inferior paraffinic crudes typically produce materials binder specifications will prevent premature failures in terms of
that are sensitive to both thermoreversible (wax crystallization, fatigue, low temperature cracking and moisture damage.
asphaltenes and resins precipitation, phase separation, collectively
known as paraffinic demixing) and irreversible (oxidative)
hardening [5–7]. Western Texas Intermediate and certain Argen- 2. Background
tinian crudes are examples that produce hardening-susceptible
binders [8]. The two types of aging that are most relevant for long-term
Air blown binders are another group of materials that lack dura- pavement performance are oxidative, or non-reversible aging,
bility. While they provide impressive grades in the laboratory, their and thermoreversible aging. In the literature there are a great
performance in service has consistently been disappointing [9–11]. number of studies related to oxidative aging. This is likely because
266 E. Angius et al. / Construction and Building Materials 165 (2018) 264–271

most asphalt is used in temperate and hot climates. A complete ductile state yet close to the ductile-to-brittle transition zone
review of the literature on oxidative aging is beyond the scope of [9,10,12,13,20]. The DENT test is modelled after the original ductil-
this paper. Instead, this paper will limit discussion to the way in ity test developed by Dow [21]. He observed with a keen eye that
which the Superpave specification protocol was meant to assess for binders that flow well, performance is generally satisfactory,
aging and how the AASHTO M 320 standard does it today, using while for those that fail abruptly the experience is rarely accept-
the rolling thin film oven (RTFO), the pressure aging vessel (PAV), able. The method was published by the MTO as Laboratory Stan-
the dynamic shear rheometer (DSR), and bending beam rheometer dard 299 [20], and it has now been implemented on all Ontario
(BBR). It has been recognized in several studies that those binders asphalt paving contracts ($300 million annual asphalt cost). The
that suffer from oxidative aging are typically also sensitive to ther- DENT was adopted by AASHTO as provisional standard TP 113-15
moreversible effects so the discussions that follow are applicable in 2015 [22], and is currently being evaluated by several research
to both [4,11]. groups around the world.

2.1. Adoption of RTFO, PAV and BBR in AASHTO M 320 2.4. Extended bending beam rheometer (MTO LS-308, AASHTO TP 122-16)

The researchers that worked under the Strategic Highway Along the same lines, a considerable amount of research was
Research Program (SHRP) to develop the Superpave asphalt binder done to show how variations in thermoreversible aging can help
specification included one direct measure of durability in terms of a to explain relative performance differences [1,9–13]. Materials
maximum limit on the weight loss during RTFO aging. This had from the 2003 MTO pavement trial near Timmins [1,9,11], along
been the norm and no significant changes were made. It is currently with materials from other trials and regular contracts [10,12,13],
set at one percent in order to avoid binders that potentially pose were able to show that extended conditioning separates the satis-
environmental, health and safety hazards during construction. factory from the unsatisfactory performing materials with a high
A considerable amount of research focused on the refinement of degree of success. The extended bending beam rheometer (EBBR)
the PAV [14,15]. The SHRP researchers looked at the effect of film protocol conditions samples for one, 24 and 72 h at 10 and 20
thickness, aging temperature, time and pressure on the perfor- degree Celsius above the lower pavement design temperature, Td
mance properties as measured with the penetration test and DSR + 10 and Td + 20, which is where most thermoreversible aging
[14,15]. The general finding confirmed earlier research that had occurs. The 24 h is often too short a time to obtain much significant
shown all these variables to be of importance, especially in the case hardening except for the very worst materials (round robin testing
of lesser quality binders [15]. However, for practical reasons, the has shown the reproducibility to be around ± 1 °C (MTO, Perfor-
PAV was standardized using only a single film thickness of 3.2 mance Graded AC Correlation Results, April 2015 [23]). Hence,
mm (50 g/pan), temperature of 100 °C, time of 20 h, and pressure 72 h was selected as a compromise between getting sufficient
of 2.08 MPa, and hence lacks a clear assessment of durability [4]. levels of hardening and keeping the test method practical. As such,
The original BBR specification contained provisions for creep the EBBR provides a discriminating assessment of low temperature
testing of the PAV residue after one and 24 h of cold conditioning durability. The method was published by the MTO as Laboratory
but the latter never appears to have caught on for reasons that Standard 308 [24], and it has now been implemented on all Ontario
are not well documented. Supposedly, the industry lobbied against asphalt paving contracts. The EBBR was adopted by AASHTO as
the use of BBR grading data after 24 h of conditioning at the test provisional standard TP 122-16 in 2016 [25].
temperature, claiming it to be impractical [16]. However, a rather
significant amount of work was done during the SHRP on the
2.5. Difference in limiting BBR temperatures (DTc)
effects that extended cold conditioning has on low temperature
BBR properties. It generally confirmed earlier findings that differ-
One drawback of the EBBR protocol is that it takes three days to
ent crude sources produce binders that show a wide range of ther-
obtain a grade and a grade loss. So, our group and other researchers
moreversible aging tendencies (age hardening, physical hardening,
have continued efforts to find alternative, less time consuming
physical aging, reversible aging, etc.) [17]. So, the BBR specification
ways to measure durability (e.g., [26–29], others). The difference
that ended up in AASHTO M 320 [18], as used in most jurisdictions
between limiting stiffness (S) and m-value temperatures, DTc,
today, also lacks a comprehensive assessment of durability.
being a measure of the degree of gelation in the binder, has been
investigated [26,27,29]. Binders that start as a gel-type will suffer
2.2. Modified pressure aging vessel (MTO LS-228)
most from the effects of additional conditioning in both the PAV
and BBR. However, the issue is not straightforward in that there
Our group took the original PAV research one step further and
are materials that actually improve upon additional PAV aging
used materials sampled during the construction of a 2003 Ministry
[30]. There are now suggestions to condition binders for longer
of Transportation of Ontario (MTO) pavement trial to show how
in the PAV and then measure the difference in the limiting BBR
important the actual PAV conditions can be in helping to explain
temperatures, T(S = 300 MPa) and T(m = 0.3), after one hour of con-
relative performance differences in the real world [4]. Aging in
ditioning. The DTc obtained can be limited as another approach to
thinner films or for longer times was shown to significantly
prevent the use of binders that lack durability.
enhance the performance grading. In spite of this, the modified
The objective of this study is to compare most of the current
PAV protocols published by the MTO as Laboratory Standard 228
tests for durability using a set of 24 binders obtained during and
[19] have yet to find traction in Ontario pavement design. A high
after the construction of a number of pavements.
degree of inertia makes it difficult to implement changes, no mat-
ter how much they will save the user agencies in the long term.
3. Experimental details
2.3. Double-edge-notched tension test (MTO LS-299, AASHTO TP 113-15)
3.1. Materials
The double-edge-notched tension (DENT) test was developed to
determine an approximate critical crack tip opening displacement Binders investigated for this study were obtained by extraction
(CTOD), which reflects a binder’s critical strain tolerance in the and recovery from hot-mix asphalt, sampled 1 and 2 years after
E. Angius et al. / Construction and Building Materials 165 (2018) 264–271 267

construction (core samples from two pavements) and during the 4. Results and discussion
construction of a number of new pavements (loose mix for the
remaining samples). All binders had been specified as PG 64–28 4.1. Standard AASHTO M 320 assessment of durability
in the contract documents, several according to regular AASHTO
M 320 specifications but most according to DENT and EBBR crite- The findings for the standard AASHTO M 320 [18] high (XX),
ria. The asphalt binders were obtained from three different com- intermediate (II) and low temperature (YY) grades, LS-299 CTOD
mercial suppliers and the paving work was performed by five [20], and LS-308 EBBR [24] grade and grade loss durability assess-
different contractors. Extracted and recovered binders were tested ments are in Table 1.
in order to deal with the use of RAP and other unknowns at the hot The two core samples C1 and C2 were extracted with both DCM
mix plant. and toluene, in order to compare the effect of the solvent-
extraction method on performance-based properties. It is evident
that the solvent does not have any significant effect on the find-
ings. The average grade temperature difference of 0.9 °C, and an
3.2. Binder extraction, recovery and aging
average CTOD difference of 1.55 mm between the samples are well
within the reproducibility levels for both test methods and agree
The mixes were left to soak overnight in reagent grade dichlor-
with results obtained for a significantly larger set of materials [33].
omethylene (DCM) or toluene and rinsed several times to extract
From the AASHTO M 320 XX and YY grading results it can be
nearly all the binder. The 4–6 L of asphalt solution was left to
concluded that all binders are expected to perform well since they
sediment overnight in order to remove the fines. The bulk of the
achieved the required 64 °C high temperature and 28 °C low tem-
solvent was evaporated at moderate temperatures and vacuum
perature grades set out in the contract documents.
(50–80 °C and 500 mbar) under a dry nitrogen gas atmosphere in
The intermediate environmental temperature for the grade
a rotary evaporator. Once no more solvent was visibly being dis-
specified in the contract documents is (64 + 28)/2–28 + 4 or
tilled, the temperature was increased in steps of 20 °C to a final
22 °C, which is above all the intermediate grades in Table 1. For
160 °C. After the temperature was stabilized, the flask was left to
the climatic zone it is (58 + 28)/2–28 + 4 or 19 °C, which is above
rotate at 10 rpm and a vacuum below 100 mbar for an additional
all but three of the grades in Table 1. For the three that failed, they
one hour.
do so by only marginal amounts (0.6 °C to 2.0 °C). Hence, the
The binders recovered from core samples were tested without
intermediate grading criteria also suggest these binders should
further aging while the materials from loose mix were PAV-aged
all perform well.
according to standard procedures (AASHTO R 28-09 [31]) prior to
performance grading.

4.2. Double-edge-notched tension (DENT) analysis

3.3. Rheological and failure testing In contrast with the AASHTO M 320 findings, the LS-299 DENT
protocol, which was adopted on selected contracts from which the
Recovered binders were tested in the DSR for high and interme- samples originated, failed a considerable number of samples. The
diate temperature performance properties according to standard specification required a minimum CTOD of 14 mm to pass in a
procedures (AASHTO R 29-08 [32]). Both complex modulus, G⁄ 28 °C zone. Out of the sample set tested, Table 1 shows that a
(Pa), and phase angle, d (°), were measured at 10 rad/s and various total of nine failed the minimum criterion. Three of the nine fail-
temperatures. ures were for samples extracted and recovered from loose mix
Aged binders were poured in silicone molds with aluminum used on a contract for which both LS-299 CTOD and LS-308 EBBR
inserts and tested in a ductilometer after 0.5–2 h of condition- acceptance criteria had been specified and thus should not have
ing at 15 °C prior to DENT testing. The displacement rate was failed. However, acceptance was based on binder supply tank sam-
kept constant at 50 mm/min. Specimens were tested with 5, ples, rather than material extracted and recovered from the hot-
10 and 15 mm ligaments at a thickness of 10 mm in 30 mm mix asphalt, so the contractor was not penalized. The remaining
width. The essential work of failure, we (J.m 2), was calculated six failures came from contracts which had asphalt cement speci-
as the specific total work (i.e., area under the force- fied according to conventional AASHTO M 320 criteria. These mixes
displacement curve) at zero ligament. The we was determined were likely contaminated with RAP where it was not sanctioned.
by extrapolation to zero from ligament lengths of 5, 10 and The acceptance is now based on LS-299 CTOD and LS-308 EBBR
15 mm. An approximate critical crack tip opening displacement properties of extracted and recovered asphalt cement.
was calculated as the essential work divided by the average net
section stress in the smallest ligament specimen, CTOD = we/
rnet section. The CTOD is used for specification grading of asphalt 4.3. Extended bending beam rheometer analysis
binders in the ductile state and provides a measure of durabil-
ity as it has been shown to correlate with field performance Table 1 shows that the LS-308 EBBR standard also managed to
[12,13]. pass only the 15/24 (62.5%) binders that did well in the LS-299
The PAV residues were finally poured into slender beams and CTOD test. There appears to be a moderate correlation between
tested according to Ontario’s EBBR method (LS-308 [24]). A total EBBR and DENT results, and both should therefore be good indica-
of 12 beams were placed in ethanol baths kept at 8°C and 18 °C tors of durability. The extended conditioning in the LS-308 EBBR
(six beams in each bath). Three samples each were tested at pass method penalizes certain binders more than others, which is
and fail temperatures after one, 24 and 72 h. The warmest of the reflected in its greater sensitivity compared to AASHTO M 320.
limiting temperatures where the creep stiffness, S(60 s), reached The grade range covered over 11.8 °C for the EBBR compared to
300 MPa or the m(60 s)-value reached 0.3 set the EBBR grade. 9.3 °C for the BBR, a difference of nearly 27%.
The EBBR grade was then subtracted from the one hour test result The LS-308 EBBR grade loss was positive for all binders, imply-
in order to obtain a 72 h grade loss, which is a measure of ing the extended BBR protocol produced limiting grades warmer
durability that has shown high correlation with old age pavement than those found using the AASHTO M 320 low temperature
performance [1,10,11]. standard by an average of 3.8 °C.
268 E. Angius et al. / Construction and Building Materials 165 (2018) 264–271

Table 1
Conventional Durability Measures.

Sample Code AASHTO M 320, °C LS-299, mm LS-308, °C


XX II YY CTOD Grade Grade Loss
C1T 77 17 29 10.3 25 3.7
C1D 78 17 28 12.4 25 3.1
C2T 79 17 31 8.9 26 4.8
C2D 80 19 30 7.9 24 6.2
L3T 69 14 33 14.0 31 2.7
L4T 77 14 34 18.0 31 2.8
L5T 74 16 33 15.1 28 5.0
L6T 67 10 36 36.5 34 2.0
L7T 71 20 28 8.8 22 5.2
L8T 71 21 28 9.7 24 3.6
L9T 66 12 35 15.9 32 3.0
L10D 70 7 37 22.3 34 2.7
L11D 68 13 32 15.1 29 2.7
L12D 69 13 33 15.7 29 3.9
L13D 72 11 35 18.3 32 3.0
L14D 72 13 33 16.7 30 3.1
L15D 77 21 29 8.7 23 6.2
L16D 75 19 30 9.7 25 5.5
L17D 72 19 31 12.1 26 5.3
L18D 72 15 34 16.1 31 2.6
L19D 71 13 33 21.4 30 3.4
L20D 67 12 34 23.4 32 1.8
L21D 71 13 33 22.6 30 3.5
L22D 69 15 33 18.3 28 4.1
Averages: 72 15 32 15.8 28 3.8
Ranges: 66 to 80 7 to 21 28 to 37 7.9 to 36.5 22 to 34 1.8 to 6.2
Pass Criteria: >64 >19 < 28 >14 < 28 <6

Note: C = core, L = loose mix, D = dichloromethylene (DCM), and T = toluene. XX = AASHTO M 320 high grade temperature based on G*/sin d, II = AASHTO M 320 intermediate
grade temperature based on G*sin d, and YY = AASHTO M 320 low grade temperature based on S(60 s) and m(60 s) (18). Grades are rounded to the nearest degree.

4.4. Modified pressure aging vessel analysis in Fig. 2(b) do not provide enough sensitivity for any meaningful
conclusions to be drawn.
All the binders were treated for an additional 20 h in the PAV for
a total of 40 h, according to the LS-228 Modified PAV protocol [19], 4.5. Differences in limiting temperatures (DTc)
to measure changes in rheological properties. Binders aged in this
way were once more tested according to both standard DSR and The difference in limiting BBR temperatures, TS = 300 MPa–Tm = 0.3,
BBR protocols at pass and fail temperatures for all specification commonly referred to as DTc, has gained a significant amount of
properties [18]. The results were compared with conventional attention as a potential specification property to assure a minimal
properties. level of durability (e.g., [4,27,29]). Binders that deviate from the
A comparison for the LS-308 EBBR grade after 20 h of PAV aging line of equality in a plot of TS versus Tm are more prone to form
and the limiting one hour BBR grade after 40 h of aging is provided gel-type structures with reduced relaxation ability, and have
in Fig. 2(a). The graph shows that there is a rather strong correla- shown to provide inferior performance [1,4,5,9–13,26,27]). The
tion between the limiting grade temperatures (R2 = 0.91), in agree- correlations between DTc values for each BBR/PAV protocol and
ment with earlier findings [4]. Additionally, the average difference the LS-308 EBBR grade loss are in Fig. 3(a)–(c). It is obvious that
in grade temperature between the two methods was found to be there is little to no correlation after one hour in the BBR, but some-
only 0.90 °C. As the LS-308 EBBR standard has been proven to accu- what of a trend after 72 h of cold conditioning prior to pass/fail
rately predict binder performance in the real world [1,5,10–13], it testing. This makes sense since most of the binders included for
can be concluded that extending the pressure aging treatment to this study start off close to a DTc of zero (Fig. 3(a)) and after 72 h
40 h will likely produce limiting grades which are less prone to some become m-controlled by moderate amounts (Fig. 3(c)).
error when predicting asphalt durability. This approach would Hence, DTc correlates with the grade loss albeit not strongly. Once
accomplish the grading in less time. However, caution is warranted again, the sample size used is limited in binder performance vari-
in that remarkable binders exist that appear to improve their BBR ation, therefore future DTc studies would benefit from a sample
performance upon additional PAV aging [30]. Hence, it appears the pool containing a greater range of materials with more varied
good correlation seen for this set of materials might not reflect all durability.
commercial binders that are currently available.
The grade losses for AASHTO M 320 at 40 h versus 20 h of PAV 4.6. Limiting phase angle and loss modulus analysis
aging are compared to the standard LS-308 grade losses from
Table 1 in Fig. 2(b). There again appears to be a general correlation The final comparison involves the AASHTO M 320 YY, EBBR
and the average difference between both grade losses was found to grades, limiting phase angle temperatures and intermediate M
be only 0.90 °C. 320 II grades. The latter is determined by the temperature at which
It should be noted that the grade loss correlation appears the binder loss modulus, G⁄sin d, reaches 5.0 MPa. The relevant
weaker as suggested by a lower R2 of 0.49. This can be attributed graphs are provided in Fig. 4.
to the small range of values seen in both the LS-308 EBBR grade The findings are most interesting in several respects. First, the
loss (1–72 h conditioning) and the AASHTO M 320 (40–20 h PAV) comparison between Fig. 4(a) and (b) shows how the EBBR test is
grade loss. Unlike the high correlation found in Fig. 2(a), the ranges more sensitive than the regular BBR test. The grades are spread
E. Angius et al. / Construction and Building Materials 165 (2018) 264–271 269

-20 4

ΔTc (20 h PAV, 1 h BBR), oC


y = 1.00x 2
M 320 (40 h PAV), °C

R² = 0.91
-25
0

-2
-30

-4 y = -0.33x + 1.40
R² = 0.14
-35 -6
-35 -30 -25 -20 (a) 0 3 6 9
(a) LS-308 EBBR (20 h PAV), °C LS-308 Loss, °C

9 4

y = 0.93x
y = -0.68x + 0.06

ΔTc (40 h PAV, 1 h BBR), oC


2
R² = 0.29
R² = 0.49
M 320 Loss (PAV), °C

0
6

-2

-4
3

-6

-8
0 (b) 0 3 6 9
0 3 6 9
LS-308 Loss, °C
(b) LS-308 EBBR Loss, °C
4
Fig. 2. (a) Correlation between AASHTO M 320 grade (40 h PAV and 1 h condition-
ing at the test temperature prior to BBR pass/fail testing) and LS-308 grade (20 h y = -0.87x + 1.37
2
ΔTc (20 h PAV, 72 h BBR), oC

PAV and 72 h conditioning prior to BBR pass/fail testing), and (b) Correlation
between PAV grade loss (BBR 40 h PAV with 1 h conditioning minus BBR 20 h PAV
R² = 0.51
with 1 h conditioning) and the EBBR grade loss (BBR 20 h PAV with 1 h conditioning
minus BBR 20 h PAV with 72 h conditioning). 0

out over a range of 11.8 °C for the EBBR tests versus 9.3 °C for the -2
regular BBR test (27% better for the EBBR). Second, more revealing
is that the limiting 30° phase angle temperature has even higher -4
sensitivity covering a range of 13.3 °C (43% better than the BBR).
This goes to show how the non-equilibrium state in which the bin-
der is tested in the BBR and EBBR creates performance issues due -6
to the under-design of the pavement for thermal cracking. A higher
degree of sensitivity with equal or better reproducibility lowers -8
risk and thus benefits both users and producers of asphalt cement.
(c) 0 3 6 9
Third, there appear to be very high correlations between the EBBR LS-308 Loss, °C
grade, measured after 72 h of cold conditioning, and the limiting
Fig. 3. Correlation between LS-308 grade loss (BBR 20 h PAV with 1 h conditioning
phase angle temperatures, measured after 15 min of conditioning minus BBR 20 h PAV with 72 h conditioning) and DTc property for regular and
at more moderate temperatures (R2 of 0.94–0.96). This suggests extended conditioning protocols.
that it is possible to get nearly the same information from a simple
phase angle measurement as what is obtained from a rather differ-
ent, and some would say cumbersome, extended BBR protocol. Our group and others have previously considered the interme-
Phase angles are determined in a DSR on very small amounts of diate and low temperature phase angles for the performance rank-
material after minimal conditioning, which would provide for a ing of asphalt binders [11,26,34–36]. Extracted and recovered
much improved and user-friendly test method. Finally, the materials were tested from a number of Eastern and Northeastern
AASHTO M 320 II grade appears to be less sensitive than the limit- Ontario trial sections and contracts and it was found that the phase
ing phase angle temperature. It suffers from reproducibility issues angle (tan d) provided a very accurate measure of pavement per-
related to the complex modulus measurement changing during formance after only minimal conditioning [11,26]. Binders with
equilibration [23,37], and it appears to pass nearly all binders sold phase angles >28° at 10 °C showed little cracking while those
in the Ontario market, making the acceptance criterion somewhat below 28° had mostly cracked prematurely and severely. This find-
questionable. ing agreed well with results previously and independently
270 E. Angius et al. / Construction and Building Materials 165 (2018) 264–271

-10 64-28) produced by three suppliers and used by five contactors.


A wider assortment of binders with different modifiers would need
R² = 0.85
to be investigated before the EBBR protocol can be replaced with a
simple DSR test. Further, the DENT test will likely continue to be
M 320 YY Grade, oC

R² = 0.87 needed in order to provide a second safeguard against premature


-20
cracking distress. It has been reported by Glover et al. [38] that
there is not always a clear relationship between low strain rheo-
R² = 0.86
logical properties and ductility. Hence, for the time being it is pru-
dent to use both rheological and failure tests to control cracking
-30 distress.

5. Summary and conclusions


-40
Given the results presented, the following summary and con-
-15 -5 5 15
(a) clusions are provided:
T(δ = 25o, 30o and 35o), oC
1) The regular AASHTO M 320 specification appears to lack sen-
sitivity and passes a significant number of asphalt binders
-10
deficient in durability.
R² = 0.95 2) The approximate critical crack tip opening displacement
(CTOD) as measured with the double-edge-notched tension
R² = 0.96
(DENT) test provides a significant degree of differentiation
LS-308 Grade, oC

-20 for binders in the high strain regime.


R² = 0.94 3) For binders investigated in this study, the extended pressure
aging vessel (PAV) conditioning for 40 h provides low tem-
perature grades that are very similar to those obtained by
-30 extended bending beam rheometer (EBBR) testing of the
20 h PAV residues.
4) The difference in limiting temperatures from the BBR mea-
surement (DTc parameter), as measured on 20 h PAV resi-
due, shows little to no correlation with the grade loss in
-40
the EBBR. It does show a modest correlation for 40 h PAV
-15 -5 5 15
residues.
(b) T(δ = 25o, 30o and 35o), oC 5) There appears to be a very high correlation between the
EBBR grade and limiting temperatures where the phase
25 angle reaches 25°, 30° and 35°. Hence, these parameters
are worthy of further investigation for future thermal and
y = 0.83x + 16 fatigue cracking specifications for asphalt binders.
R² = 0.87 6) For the time being it remains prudent to use both rheological
20
M 320 II Grade, oC

and failure tests to control thermal and fatigue cracking dis-


tress in old age.

15 Given the scale of thermal cracking problems in North American


pavements, the issue of which durability properties to use for the
control of this distress deserves continued attention.
10
Acknowledgements and Disclaimer

5 The authors thank the Natural Sciences and Engineering


-15 -10 -5 0 5 10 15
(c) Research Council of Canada and the user agency of the asphalt
T(δ = 30), oC for financial support. Haibo Ding thanks the China Scholarship
Council for the financial support that was provided to study at
Fig. 4. Correlation between limiting phase angle temperatures (T(d), 15 min
conditioning) and (a) AASHTO M 320 YY (1 h conditioning), (b) LS-308 EBBR (72 h
Queen’s University. The authors also thank user agency staff for
conditioning), and (c) AASHTO M 320 II grades (15 min conditioning). supplying the materials evaluated in this study.
Neither of the sponsors necessarily agree with, endorse or will
adopt the findings, conclusions, or recommendations either
inferred or expressly stated in subject data developed.
reported by Migliori et al. [35] who found a limit of 27° to be opti-
mal (at a slightly different oscillatory frequency).
Our findings suggest that the phase angle is worthy of further References
investigation. It may well be the preferred property for the control
of pavement performance in old age. However, caution is war- [1] A. Rigg, A. Duff, Y. Nie, M. Somuah, N. Tetteh, S.A.M. Hesp, Non-isothermal
ranted to make sure that a switch would result in improved pave- kinetic analysis of reversible aging in asphalt cement, Road Mater. Pavement
Des. 18 (4) (2017) 185–210.
ment performance rather than a continued struggle. The set of [2] Oxford Online Dictionary, www.oxforddictionaries.com, accessed on July 28,
materials used for this study was limited to a single grade (PG 2017.
E. Angius et al. / Construction and Building Materials 165 (2018) 264–271 271

[3] J.C. Petersen, A review of the fundamentals of asphalt oxidation, Double-Edge-Notched Tension (DENT) Test. TP 113-15, AASHTO, Washington,
Transportation Research Circular E-C140, Transportation Research Board, D.C., 2015.
Washington, D.C., 2009. [23] Bituminous Section, Materials Engineering and Research Office, Ministry of
[4] J. Erskine, S.A.M. Hesp, F. Kaveh, Another look at accelerated aging of asphalt Transportation of Ontario. Performance Graded AC Correlation Results for
cements in the pressure aging vessel, June 13-15, Proc., Fifth Eurasphalt & April 2015 – Round 1, Downsview, Ontario, July 15, 2015.
Eurobitume Congress, Istanbul, 2012. [24] Ministry of Transportation of Ontario. LS-308 Determination of Performance
[5] S.A.M. Hesp, S. Iliuta, J. Shirokoff, Reversible aging in asphalt, Energy Fuels 21 Grade of Physically Aged Asphalt Using EBBR Method. Revision 29 to the MTO
(2007) 1112–1121. Laboratory Testing Manual, 2015.
[6] G. Van Gooswilligen, F.T. DeBats, T. Harrison, Quality of paving grade bitumen– [25] AASHTO (American Association of State Highway and Transportation Officials).
A practical approach in terms of functional tests, in: Proc., Fourth Eurobitume Determination of Performance Grade of Physically Aged Asphalt Binder Using
Congress, 1989, pp. 290–297. the Extended Bending Beam Rheometer (BBR) Method. TP 122-16, AASHTO,
[7] A. Schmets, N. Kringos, T. Pauli, P. Redelius, T. Scarpas, On the existence of wax- Washington, D.C., 2016.
induced phase separation in bitumen, Int. J. Pavement Eng. 11 (6) (2010) 555– [26] A. Soleimani, S. Walsh, S.A.M. Hesp, Asphalt cement loss tangent as surrogate
563. performance indicator for control of thermal cracking, Transp. Res. Record
[8] I. Kodrat, D. Sohn, S.A.M. Hesp, Comparison of polyphosphoric acid modified 2162 (2009) 39–46.
asphalt binders with straight and polymer-modified materials, Transport. Res. [27] S.A.M. Hesp, S. Subramani, Another look at the bending beam rheometer for
Rec. 2007 (1998) 47–55. specification grading of asphalt cements July 8-10, Proc., Seventh International
[9] S.A.M. Hesp, S.N. Genin, D. Scafe, H.F. Shurvell, S. Subramani, Five year Conference on Maintenance and Rehabilitation of Pavements and
performance review of a northern ontario pavement trial, in: Proc., Canadian Technological Control, Mairepav6, Torino, Italy, 2009.
Technical Asphalt Association, 2009, pp. 99–126. [28] G. Rowe, Evaluation of the relationship between asphalt binder properties and
[10] M.O. Zhao, S.A.M. Hesp, Performance grading of the Lamont, Alberta C-SHRP non-load related cracking. prepared discussion, J. Assoc. Asphalt Paving
pavement trial binders, Int. J. Pavement Eng. 7 (3) (2006) 199–211. Technol. 80 (2011) 649–663.
[11] J.L. Freeston, G. Gillespie, S.A.M. Hesp, M. Paliukaite, R. Taylor, Physical [29] R.M. Anderson, G.N. King, D.I. Hanson, P.B. Blankenship, Evaluation of the
hardening in asphalt, in: Proc., Canadian Technical Asphalt Association, 2015, relationship between asphalt binder properties and non-load related cracking,
pp. 53–81. J. Assoc. Asphalt Paving Technol. 80 (2011) 615–663.
[12] S.A.M. Hesp, H.F. Shurvell, Waste engine oil residue in asphalt cement August [30] H. Xu, A. McIntyre, T. Adhikari, S.A.M. Hesp, P. Marks, S. Tabib, Quality and
28, Proc., Seventh International Conference on Maintenance and Rehabilitation durability of warm rubberized asphalt cement in Ontario, Transp. Res. Rec.
of Pavements and Technological Control, Mairepav7, Auckland, New Zealand, 2370 (2013) 26–32.
2012. [31] AASHTO (American Association of State Highway and Transportation Officials).
[13] S.A.M. Hesp, A. Soleimani, S. Subramani, T. Phillips, D. Smith, P. Marks, K.K. Standard Practice for Accelerated Aging of Asphalt Binder Using a Pressurized
Tam, Asphalt pavement cracking: analysis of extraordinary life cycle Aging Vessel (PAV), AASHTO R 28-09, AASHTO, Washington, D.C., 2009.
variability in Eastern and Northeastern Ontario, Int. J. Pavement Eng. 10 (3) [32] AASHTO (American Association of State Highway and Transportation Officials).
(2009) 209–227. Standard Practice for Grading or Verifying the Performance Grade (PG) of an
[14] J.W. Button, M. Jawle, Evaluation and development of a pressure aging vessel Asphalt Binder. R 29-08, AASHTO, Washington, D.C., 2008.
for asphalt cement, Transp. Res. Rec. 1391 (1993) 11–19. [33] H. Ding, J. Gyasi, S.A.M. Hesp, P. Marks, Y. Nie, M. Somuah, S. Tabib, N. Tetteh, I.
[15] H. Bahia, D.A. Anderson, The Pressure Aging Vessel (PAV): a test to simulate Ubaid, Performance grading of recovered asphalt binders from recent Ontario
rheological changes due to field aging physical properties of asphalt cement, paving contracts (Submitted for Publication, October 2017).
in: J.C. Hardin (Ed.), ASTM STP 1241, ASTM, Philadelphia, 1995. [34] F. Migliori, M. Pastor, G. Ramond, Etude statistique de quelques cas de
[16] D.A. Anderson, Discussion following the paper by Rigg et al., J. Assoc. Asphalt fissurations thermiques, in: Proc., Fifth Eurobitume Congress, Stockholm,
Paving Technol. 86 (2017). 1993, pp. 724–728.
[17] H.U. Bahia, D.A. Anderson, Glass transition behavior and physical hardening of [35] F. Migliori, G. Ramond, M. Ballie, B. Brule, C. Exmelin, B. Lombardi, J. Samanos,
asphalt binders, J. Assoc. Asphalt Paving Technol. 62 (1993) 93–129. A.F. Maia, C. Such, S. Watkins, Correlations between the thermal stress
[18] AASHTO (American Association of State Highway and Transportation Officials). cracking of bituminous mixes and their binders’ rheological characteristics
M 320-09, Standard Specification for Performance-Graded Asphalt Binder, Paper No. 046, Luxemburg, 3-6 May, in: Proc., Eurobitume Workshop on
AASHTO, Washington, D.C., 2009. Performance Related Properties for Bituminous Binders, 1999, pp. 1–5.
[19] Ministry of Transportation of Ontario. LS-228 Modified Pressure Aging Vessel [36] I. Widyatmoko, M.W. Heslop, R.C. Elliott, The viscous to elastic transition
Protocol. Revision 29 to MTO Laboratory Testing Manual, 2015. temperature and the in situ performance of bituminous materials, J. Inst.
[20] Ministry of Transportation of Ontario. LS-299 Method of Test for Asphalt Asphalt Technol. 14 (2005) 3–7.
Cement’s Resistance to Fatigue Fracture Using Double-Edge-Notched Tension [37] Kriz, P. DSR-PAV Test Improvement. Task Force Status Update, Asphalt
Test (DENT). Revision 23 to MTO Laboratory Testing Manual, 2007. Institute Technical Advisory Committee Meeting, December 7, 2016.
[21] A.W. Dow, Discussion on ductility test, Proc. Assoc. Asphalt Paving Technol. 5 [38] Glover, C.J., Davidson, R.R., Domke, C.H., Ruan, Y., Juristyarini, P., Knorr, D.B., Jung,
(1936) 139–141. S.H. Development of a New Method to Assess Asphalt Binder Durability with
[22] AASHTO (American Association of State Highway and Transportation Officials). Field Validation. Report No. FHWA-TX-05/1872-2. Texas Transportation Institute,
Determination of Asphalt Binder Resistance to Ductile Failure Using the Texas A&M University System, College Station, Texas 77843-3135, 2005.

You might also like