You are on page 1of 23

1 Developing Simple Binder Indices for Cracking Resistance of Asphalt Binders at

2 Intermediate and Low Temperatures


3
4
5
6
7
8 Raquel Moraes, Ph.D. (Corresponding Author)
9 Asphalt Research Engineer
10 Modified Asphalt Research Center
11 University of Wisconsin-Madison
12 1415 Engineering Drive, 3352 Engineering Hall, Madison, WI, 53706
13 Email: moraes@uwalumni.com
14
15
16
17
18
19 Hussain Bahia, Ph.D.
20 Professor
21 Department of Civil and Environmental Engineering
22 Modified Asphalt Research Center
23 University of Wisconsin-Madison
24 1415 Engineering Drive, 3350 Engineering Hall, Madison, WI, 53706
25 Email: bahia@engr.wisc.edu
26
27
28
29
30
31
32
33
34
35
36
37
38 Submission date: August 1st, 2017
39 Word count 5750 plus 3 Tables and 4 Figures
40 Total number of words: 7500
41
42
43 Submitted for publication and presentation at the
44 Transportation Research Board Annual Meeting
45 January 7–11, 2018
46 Washington, D.C.
Moraes and Bahia 2

1 ABSTRACT
2
3 Oxidative aging causes hardening of asphalt binders and, consequently, contributes to
4 deterioration of asphalt pavements. Non-load related cracking of asphalt pavements (i.e.
5 transverse and block cracks) are related to original properties and hardening of the asphalt
6 binder. In recent years, researchers have proposed new indices derived from Superpave PG
7 testing to identify changes in asphalt cracking susceptibility with aging. These indices include
8 the parameter G′/(η′/G′) and the difference between continuous low temperature binder grade
9 measured via Superpave creep stiffness and m-value (ΔTc). This study aims to develop
10 interrelationships between tests allowing choice selection for the determination of simpler
11 parameters that could be used for specification tests of asphalt binders. Two simpler asphalt
12 binder indices for cracking resistance are investigated by means of Dynamic Shear Rheometer
13 (DSR), Bending Beam Rheometer (BBR) and Single Edge Notched Bending (SENB) tests
14 results. At intermediate temperatures, the slope of the |G*|-frequency curve obtained from a
15 simple frequency sweep test is proposed as an alternative approach to directly calculate the
16 durability parameter G′/(η′/G′). At low temperatures, results indicated a direct correlation
17 between failure energy at fracture obtained from SENB and ΔTc. To confim the validity of these
18 indices and the changes at molecular level, Gel Permeation Chromatograph (GPC) results are
19 presented to indicate that asphalts with higher content of large molecular size (LMS) molecules
20 are likely to crack. Limits for specifcations for the slope of |G*|-frequency curve and DTc can be
21 derived based on testing a wide range of binders and field experience.
22
23
24
25
26 Keywords: Oxidative aging, cracking, durability, failure energy, rheology, ΔTc
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
Moraes and Bahia 3

1 INTRODUCTION
2
3 In 1987, the Strategic Highway Research Program (SHRP) was established to improve the
4 performance and durability of roads in the United States, and the Superpave (Superior
5 Performing Asphalt Pavements) asphalt binder grading system was a product of this research
6 effort. The Superpave Performance Graded (PG) asphalt binder specification was developed to
7 relate binder properties to field performance, and allow the selection of an appropriate binder for
8 a specific climatic condition.
9 The long-term pavement performance is a function of traffic load and volume, material
10 properties, construction practices, and environmental factors. Pavement deterioration manifests
11 itself in several common types of distress including rutting, fatigue cracking, low temperature
12 cracking, reflective cracking, aging, raveling, and stripping. The conventional materials used in
13 the asphalt mixture may perform satisfactorily relative to one distress type, but fail prematurely
14 relative to the others. Asphalt production, however, is not one of the main profit-generating
15 processes in the refining industry, since most refineries in the United States deal with asphalt as a
16 by-product of crude fractionation (1). Therefore, the production of better-performing asphalts is
17 not considered one of the common strategies in planning refining practices.
18 As the asphalt binder is responsible for the visco-elastic behaviour characteristic of
19 asphalt pavements, it plays a large part in determining many aspects of road performance,
20 particularly resistance to permanent deformation and cracking (2). In general, the proportion of
21 any induced strain in asphalt that is attributable to viscous flow, i.e. non-recoverable, increases
22 with both loading time and temperature. Non-load related cracking of asphalt pavements (i.e.
23 transverse and block cracks) are related to oxidation and hardening of the asphalt binder.
24 Transverse cracks are usually formed because of thermal movement. It may occur because of
25 shrinkage of the binder surface due to low temperatures or asphalt binder hardening. Block
26 craking is typically caused by an inability of asphalt binder to expand and contract with
27 temperature cycles because of aging, as the binder starts to lose its durability.
28 Researchers have spent significant efforts to classify cracking resistance of asphalt
29 binders using an index parameter. Kandhal et al. in 1977 (3) noted that the decrease in
30 low-temperature ductility is an important factor as asphalt binder ages. Following this work and
31 by using a mechanical-empirical relationship with observed cracking, Glover et al. (4) suggested
32 a parameter relating storage modulus (G’) and dynamic viscosity (η′) to ductility at 15 °C and
33 0.005 rad/s.
34 Anderson et al. in 2011 (5) investigated the relation between ductility and binder
35 properties to non-load associated cracking for airport pavements. The findings of the study
36 identified the Glover parameter G′/(η′/G′), the fatigue parameter B, and the difference between
37 the continuous low temperature binder grade measured via the Superpave creep stiffness
38 and m-value (ΔTc) as parameters to identify changes in cracking susceptibility with aging.
39 Rowe in 2011 (6) proposed modifications of the Glover parameter by introducing the
40 Glover-Rowe (G-R) parameter G*(cosδ)2/sinδ, which focuses attention on complex modulus
41 (|G*|) and phase angle (δ) of asphalt binders at 15 °C and frequency of 0.005 rad/s. On his
42 proposal, Rowe ignored the frequency term expressing the G-R parameter purely in terms of |G*|
43 and δ, allowing users to plot the ductility-based failure planes in Black Space diagrams.
44 In addition to these efforts, the AASHTO TP 101 “Estimating Damage Tolerance of
45 Asphalt Binders Using the Linear Amplitude Sweep” was proposed in 2012 (7, 8) and used in
46 various studies to estimate fatigue cracking resistance.
Moraes and Bahia 4

1 In the present study, a simpler asphalt binder index for cracking resistance at intermediate
2 temperature is investigated. At low temperatures, failure energy at fracture obtained from the
3 Single Edge Notched Bending (SENB) test, relaxation properties obtained from DTc, and
4 changes at molecular level by using Gel Permeation Chromatograph (GPC) are investigated. The
5 study aims to either confirm the validity of the cracking indices proposed by previous
6 researchers, or suggest a replacement of these values with simpler parameters that could be used
7 for specification tests of asphalt binders.
8
9
10 BACKGROUND
11
12 Asphalt Chemistry
13 The chemical composition of asphalts cannot be defined exactly, because it is extremely complex
14 and variable, depending on the source of the crude oil from which the asphalt originates and on
15 modification induced by treatments in the refinery or during the “in service” life. Based on
16 differences in solubility and polarity, the asphalt can be separated into four main chemical
17 classes, designated as saturates (S), aromatics (A), resins (R), and asphaltenes (A).
18 Asphalt is often described as a colloidal dispersion of asphaltenes into an oily matrix
19 constituted by saturates, aromatics, and resins. It has been established that asphaltenes are
20 stabilized in crude oils by natural resins which are surfactant-like agents. The action of resins on
21 the asphaltenes aggregation and precipitation processes is thought to be important. Due to their
22 molecular constitution, asphaltenes and resins have a mutual intrinsic effect on the stability of
23 molecular self-assembling, either in the form of asphaltene-resin association (which promotes
24 re-dispersion) or asphaltene-asphaltene association (which promotes precipitation) (9). It is
25 important to mention that both asphaltenes and resins are responsible for the visco-elastic
26 properties of the asphalt at ambient temperatures (2).
27
28 Aging Process in Asphalts
29 The aging process in asphalt pavements may be detrimental when excessive hardening or
30 stiffening is observed, subjecting the pavement to cracking, either under traffic load or thermal
31 stress. There are two types of aging in asphalt: (i) physical aging caused by temperature, and (ii)
32 oxidative aging caused by air or oxygen. Physical aging is a reversible phenomenon that takes
33 place as a consequence of cooling or quenching of amorphous materials from melt temperatures
34 to below the glass-transition temperature. Oxidative aging is irreversible hardening of asphalts.
35 During the oxidative aging process of asphalt binders, the concentration of polar functional
36 groups increases, resulting in an immobilization of molecules through intermolecular association
37 (10). Hence, the molecules or molecular agglomerates lose sufficient mobility to flow past one
38 another under thermal or mechanical stress. This results in embrittlement of the asphalt, making
39 it more susceptible to fracturing or cracking and resistant to healing (11).
40 According to the literature (12, 13), thermo-oxidative aging increases the content of
41 large asphalt molecules and decreases the content of small molecules, leading to an increase in
42 the molecular weight of the asphalt binders. Because the various asphalt fractions have different
43 reactivities toward oxidation, during oxidative aging a reduction of aromatics content is observed
44 along with an increase in the content of resins and asphaltenes (13). Oxidation causes the oils
45 (i.e., maltene fraction) to convert to resins, and the resins to convert to asphaltenes (14). Changes
46 in the content of saturates are almost negligible due to their inert nature. Because of this change
Moraes and Bahia 5

1 in asphaltenes-maltene ratio that occurs during oxidation, the stiffness property of the asphalt is
2 affected.
3
4 Asphalt Modification to Improve Its Performance
5 For several decades, blending different types of asphalt was the applicable solution to obtain
6 improvements in the asphalt binder properties. In recent years, the modification of the asphalt
7 binder to enhance its performance at both high and low temperatures and under traffic loading
8 has served as one of the cost-effective engineering solutions to allow the construction of superior
9 pavements (15, 16, 17).
10
11 Polymer Modified Asphalts
12 Over the years, polymer modification of asphalt binders for pavements is finding increasing
13 usage, since demands made upon roads increase year by year. This type of modification provides
14 increased viscosity at high pavement temperatures, resulting in a reduced tendency for
15 permanent deformation (i.e., rutting). At the same time, polymers can provide increased ductility
16 at low pavement temperatures and thus a reduced tendency for thermal cracking (18). Other
17 advantages of polymer modification are improved resistance to abrasion as well as to oxidation
18 and aging, and the possibility of applying a thinner pavement in comparison to the one where
19 conventional binder is used.
20 Blends of asphalt with polymers form multiphase systems. Such systems contain a phase
21 rich in polymer, a phase rich in asphaltenes not adsorbed by the polymer, and a phase formed by
22 maltenes (19). As a result, the properties of the modified asphalt binders depend on the
23 concentration and the type of polymer used for modification. The main point in polymer
24 modification is, therefore, the formation of a thermodynamically unstable but kinetically stable
25 mixture, in which the polymer is partially swollen by the lighter asphalt components (20).
26 Although great advances have been achieved in the field of asphalt polymer modification,
27 there are still various drawbacks which are limiting its future developments, such as higher costs,
28 polymer low aging resistance and poor storage stability.
29
30 Rejuvetanors
31 With increased demand and limited aggregate and binder supply, hot mix asphalt (HMA)
32 producers have began using reclaimed asphalt pavement (RAP) as a valuable component in
33 HMA. Recycled roofing shingles (RAS) are also identified as a viable replacement or
34 supplement to RAP due to their relatively high asphalt binder content and the presence of hard
35 mineral aggregates and fibrous filler materials (18).
36 RAP and RAS are useful alternatives to virgin materials because they reduce the use of
37 virgin aggregate and the amount of virgin asphalt binder required in the production of HMA.
38 However, the hard, oxidized nature of reclaimed asphalt binder is a major concern when
39 incorporating recycled materials into hot mix asphalt mixtures, since the stiff binder that is
40 present in RAP/RAS would cause premature fatigue and low temperature cracking failures (20,
41 21).
42 In general, there are two types of chemical additives that can be added to pavements
43 containing RAP/RAS: rejuvenating agents and softening agents. The primary difference between
44 a rejuvenating agent and a softening agent is that a rejuvenating agent will restore the chemical
45 structure of aged asphalt via the restoration of the original binder asphaltenes to maltenes ratio
46 (22), while a softening agent will reduce the overall viscosity of the binder.
Moraes and Bahia 6

1 For optimal rejuvenation of the aged binder, consideration should be given not only to
2 the viscosity reducing capacity of the rejuvenator, but also to its chemical composition.
3 Furthermore, the degree of diffusion of the rejuvenator into the aged binder is of the utmost
4 importance, since it will allow chemical changes to take place, affecting the properties of the
5 recycled asphalt mixture.
6 Researchers have used a wide range of oils to enhance binder performance, specifically
7 at low service temperature (23, 24). Concerns exist, however, with respect to oil modification as
8 it relates to high temperature performance and aging susceptibility. Softening effects of oil
9 modification can detrimentally affect high temperature performance and increase the rate of
10 oxidative aging.
11
12 Hybrid Modification
13 Hybrid modification is a term that is applied when two different types of modifiers are combined
14 to produce an improved modification. The main interest in hybrid modification of binders is to
15 capture the properties of the different modifiers.
16
17
18 MATERIALS AND TESTING PROCEDURE
19
20 Materials
21 Four unmodified asphalt binders and six modified asphalt binders were selected for investigation
22 in this study. The modification type, content and the final performance grades for each selected
23 binder are shown in Table 1. The target factor for the selection of both oil and polymer content
24 was the required climatic needs found in the location of the project. It should be noted that oil
25 type A is a bio-oil while the type B is a refined waste oil.
26 To validate lab findings, heavily aged binders from the Pooled Fund TPF-5 (302)
27 project (25) were also included in the experimental matrix of this study. Two asphalt binders
28 commonly used in the Mid-West region of the United States were selected to produce ‘artificial’
29 reclaimed asphalt pavement (RAP) materials: an unmodified (neat) PG 64-22 binder (A-RAP1),
30 and a polymer modified asphalt (PMA) PG 70-28 binder (A-RAP2) collected from the member
31 state Wisconsin. The artificial RAP materials were produced by exposing the virgin binders to
32 extended cycles of aging in the Pressure Aging Vessel (PAV) (40 and 60 hours for A-RAP1 &
33 A-RAP2, respectively). The A-RAP materials (A-RAP1 & A-RAP2) were blended with polymer
34 modified asphalts sampled from partner states. The two PMAs are WI 70-28 and KS 64-34,
35 which were nominated PMA1 and PMA2, respectively. The combinations of PMA/A-RAP
36 prepared for this study are: 100%/0%, 80%/20%, 60%/40%, and 0%/100%. The combinations of
37 PMA/A-RAP were prepared using a low shear mixer for one 1 h and 30 min at a temperature of
38 150 °C.
39 The asphalt binder modified with 5% of oil-A (sample M-A) was prepared using a high
40 shear mixer for 30 min at temperature of 150 °C. The asphalt binders modified with oil and
41 elastomer (samples M-C and M-G) were prepared using a high shear mixer for 60 min at
42 temperature of 180 °C. After 1 hr of mixing, the cross-linker (Sulfur) was added, and the shear
43 mixing was kept for another 3 hrs. The samples were then cured overnight at 145-150 °C prior
44 further testing. The asphalt binders modified with oil and plastomer (samples M-B, M-D, and
45 M-F) were prepared using a low shear mixer for 60 min at temperature of 150 °C.
46
47
Moraes and Bahia 7

1 Methods
2 A summary of the test methods selected for testing materials in this study is provided in Table 2.
3
4 Linear Amplitude Sweep (LAS) Test
5 The Linear Amplitude Sweep test is a method proposed to evaluate asphalt binder fatigue at
6 intermediate temperature using the Dynamic Shear Rheometer (DSR) (8). The LAS test
7 procedure was performed using 8 mm parallel plate geometry. Prior to testing, all samples were
8 aged in RTFO followed by PAV aging for 20 h at 100 °C. The test consists of two steps: a
9 frequency sweep and an amplitude sweep. In the first part of the LAS procedure, an initial 100
10 cycles is applied at small strain (0.1 %) to determine undamaged linear viscoelastic properties.
11 The second part of the procedure consists of ramping strain amplitude, beginning at 0.1 % and
12 ending at 30 % applied strain, over 3100 cycles of loading at 10 Hz. Once the strain sweep is
13 applied to the sample, damage accumulation can be then determined through Viscoelastic
14 Continuum Damage (VECD) analysis, resulting in the fatigue power law damage model
15 (Equation 1), and the corresponding coefficients, A and B.
16
17 Nf = A (γmax)-B (1)
18
19 Nf is the traffic volume failure criteria and defines the number of cycles to fatigue failure
20 at a user-defined damage level. γmax is the maximum tensile strain expected in the binder phase
21 under traffic loading, which will be a function of pavement structure. A is the LAS power-law
22 parameter representing the intercept at 1 % strain. B is the LAS power-law parameter
23 representing the slope of the Nf-strain curve. The logarithmic slope of the storage modulus
24 (G’(ω)) as a function of angular frequency is used to calculate the damage accumulation and the
25 parameter B.
26
27 Single Edge Notched Bending (SENB) Test
28 The SENB system was developed for measuring asphalt binder, mastic, and mortar cracking
29 resistance at low service temperatures (31, 32). The SENB test uses a modified Bending Beam
30 Rheometer (BBR) with the addition of a loading motor that controls the displacement rate during
31 testing, a load cell with a higher capacity than the regular BBR, and a modified beam placement
32 fixture. In the SENB analysis, the failure energy (Gf) is defined as the total area under the entire
33 load-deflection (P-u) curve, divided by the area of the ligament, as shown in Equation 2. The
34 displacement at failure, uf, is also reported from this test procedure as a characterization
35 parameter.
36
𝑾𝒇 ∫ 𝑷𝒅𝒖
37 𝑮𝒇 = 𝑨 = 𝑨𝒍𝒊𝒈
(2)
𝒍𝒊𝒈

38
39 Wf is work of failure, Gf is failure energy, P and u are the load and displacement
40 measured by the SENB, and Alig is the area of the ligament. In this study, SENB tests were
41 performed at a constant displacement rate of 0.01 mm/sec, at low temperature PG grade of each
42 asphalt binders. Prior to testing, all samples were aged in RTFO followed by PAV aging for 20 h
43 at 100 °C.
44
45 Gel Permeation Chromatograph (GPC)
46 In this study, the Gel Permeation Chromatograph was used to measure the molecular size
47 distribution of unmodified and modified asphalt binders. In this method, the asphalt binder is
Moraes and Bahia 8

1 dissolved in a solvent and is injected into the GPC. The injected sample travels through a series
2 of columns which separates the sample based on molecular size. The larger molecular size
3 particles exit the columns first and are detected by the GPC system. The smaller molecular size
4 particles travel into the pores of the columns and, therefore, have longer retention times. A
5 molecular size distribution (analogous to a sieve analysis on a smaller scale) is then obtained. For
6 the GPC analysis of the asphalt binders selected in study, the detector used was a differential
7 refractometer, which is a concentration sensitive detector that measures the difference in
8 refractive index (dRI) between the eluent in the reference side, and the sample + eluent in the
9 sample side. The GPC columns used were PolyPore 5 µm mixed, 300 x 7.5 mm. The asphalt
10 samples were prepared in concentrations of 5 mg/2 mL. The flow rate was 1 mL/min and the
11 testing temperature was 40 °C. The tetrahydrofuran (THF) solvent was selected for the GPC
12 analysis since previous researchers had successfully selected it for GPC applications in asphalt
13 (13). To better perform the analysis of the asphalt binders, each GPC chromatogram was divided
14 into 13 equal elution time areas, and the molecules eluted during the first third of the elution
15 period were classified as large molecular size (LMS), those eluted during the second third as
16 medium molecular size (MMS), and those eluted in the last third as having small molecular size
17 (SMS) (13).
18
19
20 DETERMINATION OF LIMITS AT INTERMEDIATE AND LOW TEMPERATURE TO
21 COVER DURABILITY OF ASPHALTS
22
23 Correlation Between G’/(η’/G’) and Fatigue Parameter B
24 Research conducted by Glover et al. (4) indicated that asphalt ductility results of unmodified
25 binders obtained at 15 °C and 1 cm/min correlated well with G’/(η’/G’) for different binders
26 aged at different conditions. Upon further investigation, the same authors concluded that, if
27 ductility at 15 °C and 1 cm/min can be used as an indication of durability, then so can G’/(η’/G’)
28 at 15 °C, 0.005 rad/s. Considering that this low frequency is not accessible to most instruments
29 except through data taken at several temperatures, and over a range of several orders of
30 magnitude in frequency, the same researchers followed the time-temperature superposition
31 principle and shifted this correlation to measurements at 44.7 °C and 10 rad/s.
32 In this present study, an alternative approach is considered to directly calculate the
33 durability parameter G′/(η′/G′) developed by Glover et al. (4). DSR test results collected from the
34 LAS frequency sweep procedure were used to obtain the storage modulus (G’) and loss modulus
35 (G’’) of each asphalt binder at fixed frequency of 10 rad/s. In order to calculate G′/(η′/G′) at this
36 frequency value, an interpolation of the testing results for frequency versus G’ and G’’ was
37 perfomed for each binder. From the determined values of storage modulus (G’) and loss modulus
38 (G’’), the durability parameter was then calculated for each binder at its intermediate
39 performance grade temperature and 10 rad/s.
40 Since,
41
42 η′ = G’’/ ω (4)
43
44 Therefore,
45
46 G′/(η′/G′) = (G’)2/(G’’).ω (5)
47
Moraes and Bahia 9

1 The proposed approach does not require manipulations using time-temperature


2 superposition. Therefore, measured values can be plotted among each other without the use of
3 shift factor and sigmoidal fit functions.
4 Figure 1(a) shows the absolute value of the LAS B parameter for the unmodified and
5 modified binders selected in this study. Figure 1(b) shows the absolute value of the LAS B
6 parameter for the artificialy recycled binder, prepared in laboratory under heavy aging in PAV
7 (40 and 60 hours for A-RAP1 & A-RAP2, respectively). Please note that the fatigue parameter B
8 (slope parameter) is derived from the frequency sweep procedure, that is conducted prior to the
9 strain sweep procedure in the LAS test.
10 Anderson et al. (5) showed a correlation between the absolute value of the fatigue
11 parameter B and the durability parameter G′/(η′/G′), calculated at 15 °C and 0.005 rad/s from a
12 temperature-frenquency sweep produced mastercurve. Figure 1(c) shows the correlation between
13 the absolute value of the fatigue parameter B and the durability parameter G′/(η′/G′) calculated at
14 10 rad/s and at the intermediate performance grade temperature of each binder. Because the
15 fatigue parameter B simply represents the slope of the |G*|-frequency curve at a given
16 temperature, it is expected that it will relate to G′/(η′/G′), indicating a materials’ response to
17 oxidative aging.
18 As it can be seen in Figure 1(c), given the fairly good correlation indicated by R2 = 0.76,
19 it does appear that the absolute value of the LAS B parameter (measured at the binders’
20 intermediate PG temperature) is related to G′/(η′/G′) parameter (calculated from LAS frequency
21 sweep results collected at intermediate PG temperature and 10 rad/s). Therefore, the frequency
22 sweep step of LAS test could be used as a substitute for the “temperature-frequency sweep test
23 with mastercurve determination” for calculation of the durability parameter G′/(η′/G′). From the
24 results it can be see that this new approach works for unmodified, modified and artificial
25 recovered asphalt binders, which is very positive since the correlation found by Glover et al. (4)
26 only included unmodified binders.
27 According to Glover et al. (4), a value of G′/(η′/G′) larger than 0.003 MPa/s at 15 °C,
28 0.005 rad/s (i.e., ductility of about 3 cm at 15 °C, 1 cm/min) should be taken as a critical value
29 for pavement binder failure, while a value of 0.0009 MPa/s (ductility of about 5 cm) should be
30 taken as an indication that the binder is approaching a value that will lead to pavement cracking.
31 Please note that the authors did not considered the climate influence while setting these limit
32 values. Anderson et al. (5) showed with actual field data that the threshold values identified by
33 Glover et al. for the G′/(η′/G′) parameter at 15°C and 0.005 rad/s require more validation work.
34 Therefore, setting limits for LAS fatigue parameter B based on the 0.0009 MPa/s at 15 °C and
35 0.005 rad/s was not performed, and instead a correlation with the 10 rad/s frequency used in
36 current specifications was investigated.
37 To propose a new limit for the G′/(η′/G′) parameter calculated with the alternative
38 approach suggested in this paper (i.e., from LAS frequency sweep results collected at
39 intermediate PG temperature and 10 rad/s) a value for the LAS B parameter was selected from a
40 field study performed in Wisconsin (33). On the mentioned study, the researchers found that the
41 LAS test related very well to field performance when performed at the required Superpave
42 intermediate temperature grade of the project location, and the fatigue life parameter (Nf) was
43 determined at the peak stress. To a specific project section, it was also found that a value of the
44 LAS B parameter approximately = -3.43 corresponds to 29%-50% of cracked segments (Table
45 3). Considering this range of cracked segments as a maximum limit of allowed cracking, the
46 curve in Figure 1(c) can be used to calculate the limit value for the G′/(η′/G′) parameter, at
47 intermediate temperature and 10 rad/s. Thus, the G′/(η′/G′) parameter value at cracking limit is
Moraes and Bahia 10

1 11.0 MPa/s, with this limit being suggested for binders with up to 50% of RAP. Please note that
2 Figure 1(c) shows the binders that presented LAS B parameter between -3.17 and -4.5, to mimic
3 the values obtained in the field and to include the heavily aged binders from recycled pavements.
4
5 Correlation Between ΔTc and Fatigue Parameter B
6 Anderson et al. (5) introduced the ΔTc parameter as a means of indexing the non-load associated
7 cracking potential of asphalt binders (Equation 6).
8
9 ΔTc = Tc,S – Tc,m (6)

10 Where,

11 ΔTc = Difference in critical low temperature PG grade


12 Tc,S = Critical low temperature grade predicted using the BBR Stiffness (S)
13 Tc,m = Critical low temperature grade predicted using the BBR m-slope (m-value)
14
15 In this study, the Equation 6 was followed for the ΔTc calculation of the selected
16 binders. Please note that, for all binders, the critical low temperature grade predicted using the
17 BBR is listed in Table 1, for stifiness and m-value. ΔTc results were correlated with the G′/(η′/G′)
18 parameter calculated with the alternative approach suggested in this paper and described above
19 (i.e., from LAS frequency sweep results collected at intermediate PG temperature and 10 rad/s).
20 As it can be seen in Figure 2(a), a strong correlation indicated by R2 = 0.87 is seen for
21 the relationship between ΔTc and the G′/(η′/G′) parameter (calculated from LAS frequency
22 sweep results collected at intermediate PG temperature and 10 rad/s). As ∆Tc decreases and
23 becomes more negative, indicating that the binder has lost some of its relaxation properties and is
24 not able to recover to its initial performing stage, higher values of the durability parameter
25 G′/(η′/G′) are obtained. Therefore, both parameters appear to have potential to relate to binder
26 durability, being used as a surrogate quantifying the loss of relaxation properties as an asphalt
27 binder ages.
28 Figure 2(b) shows the relationship between the LAS B parameter (measured at the
29 binders’ intermediate PG temperature) and ΔTc. As it can be seen, a very good correlation is
30 indicated by R2 = 0.72. Please note that the binders included in Figure 2(b) presented LAS B
31 parameter between -3.17 and -4.5, to mimic the values obtained in the field and to include the
32 heavily aged binders. Binders with ΔTc out of the allowable range (i.e., -5 to 2.5 °C) were
33 removed before the comparison.
34 A limit to reduce the risk of crack initiation was set at ΔTc = -2.5 °C by Anderson et al.
35 (5), at which point a preventive maintenance was suggested to avoid the pavement reaching a
36 critical stage. Same authors also suggested a limit of ΔTc minimum of -5.0 °C to be used as an
37 indication of the need for immediate pavement rehabilitation. The AASHTO PP78 “Design
38 Considerations When Using Reclaimed Asphalt Shingles (RAS)” in Asphalt Mixtures includes a
39 criteria for ΔTc minimum of -5.0 °C, meaning that a recovered and aged binder with a ΔTc of
40 -6.0 °C is not acceptable. Based on the minimum limit of ΔTc = -5.0 °C, the curve in Figure 2 (b)
41 can be used to calculate the limit value for the LAS B parameter. Thus, the calculated LAS B
42 parameter value at cracking limit is -4.69. Please note that this value of LAS B parameter = -4.69
43 was not found in the field study (33), even when considering project sections with 100% of
44 cracked segments (Table 3). However, two considerations are important to explain this value: (1)
45 the field data did not include sections containing RAP; and (2) the graph in Figure 2 (b) includes
46 100% RAP binder, which in reality does not happen in the field.
Moraes and Bahia 11

1 DETERMINATION OF LIMITS AT LOW TEMPERATURE TO COVER FRACTURE


2 OF ASPHALTS
3
4 Correlation Between G’/(η’/G’) and Fracture Energy
5 Research studies performed by Marasteanu at al. (34) showed that the SENB test results correlate
6 well with the observed field thermal cracking pavement condition index (PCI), indicating the
7 high promise of using this binder fracture test as an unmodified and modified binder low
8 temperature characterization test to complement the current BBR test. According to the authors,
9 the SENB test is capable of clearly discriminating between modified and unmodified binders in a
10 repeatable fashion. Because unmodified asphalt binders generally have very similar failure
11 stresses, the SENB fracture energy parameter (Gf) of the modified binder will be the focus of this
12 section.
13 Figure 3(a) shows a correlation between failure energy at fracture (calculated from
14 SENB at the low temperature performance grade) and G′/(η′/G′) parameter (calculated from LAS
15 frequency sweep results measured at intermediate PG temperature and 10 rad/s). Given the very
16 good R2 = 0.87, it clearly appears that both parameters are directly correlated. Figure 3(b) shows
17 that the correlation still remains even when considering severely aged materials (i.e. RAP), with
18 these materials showing much higher values of failure energy at fracture.
19 Results in Figure 3 clearly indicates that the SENB fracture energy parameter (Gf) is
20 strongly and inversely correlated with the durability parameter G′/(η′/G′). This observation
21 seems logical since the durability parameter increases with aging due to increased |G*| while the
22 fracture energy decreases due to brittleness.
23 Therefore, higher values of fracture energy which have been correlated to a lower
24 propensity for cracking can be considered as an indication of higher duralibity of the binders, or
25 vice-versa. This good correlation between rheological measurements and fracture properties can
26 lead the to a complete characterization of the effect of aging on binder properties and cracking.
27
28 Correlation Between ΔTc and Fracture Energy
29 Figure 4(b) shows the correlation between failure energy at fracture (calculated from SENB at
30 the low temperature performance grade) and the ΔTc parameter. The fairly good correlation
31 indicated by R2 = 0.62 shows that both parameters appear to quantify the loss of relaxation
32 properties as an asphalt binder ages. Higher values of fracture energy (i.e., lower propensity for
33 cracking) are seen at higher positive values of ΔTc. To confirm that this correlation is in fact
34 related to the molecular differences between binders, Gel Permeation Chromatography (GPC)
35 analysis were performed and are going to be discussed next.
36
37 Correlation Between ΔTc and Molecular Size Distribution
38 In a previous study, Moraes et al. (13) have showed that changes in the colloidal structure of
39 asphalt binders due to changes in the degree of association of the different asphalt fractions after
40 aging are reflected in the molecular size distribution as obtained by Gel Permeation
41 Chromatography (GPC). The authors divided the GPC chromatogram into 13 equal elution time
42 areas, and the molecules eluted during the first third of the elution period were classified as large
43 molecular size (LMS), those eluted during the second third as medium molecular size (MMS),
44 and those eluted in the last third as having small molecular size (SMS) (Figure 4(a)).
45 By applying the same approach, the present study divided in three molecular size ranges
46 the GPC chromatogram of a select group of binders. To simulate the same aging conditions that
Moraes and Bahia 12

1 the binders faced prior to the low temperature testing performed in this study, all samples were
2 aged in RTFO followed by PAV aging for 20 h at 100 °C before being analyzed through GPC.
3 Figure 4(c) shows the relationship between the large molecular size (LMS) molecules
4 and ΔTc. As it can be seen, a good explanation for the ΔTc can be obtained when looking to the
5 molecular distribution of the binders, since aging results in a shift towards the higher molecular
6 weight molecules. Therefore, asphalts with higher content of LMS are likely to crack. Figure
7 4(d) validates this finding, since it shows that asphalt binders with higher content of small
8 molecular size (SMS) molecules are less prone to crack, as showed by higher positive values of
9 ΔTc.
10 To summarize, Figure 4(e) confirms the above findings showing that an increase in
11 large molecular sizes molecules at the expense of the small molecular sizes happens as the
12 asphalts ages.
13 The changes in the colloidal structure of the asphalts binders due to changes in the
14 asphalt fractions (i.e. asphaltenes and maltenes) upon aging are clearly reflected in the molecular
15 weight distribution as obtained by GPC. The LMS is the percentage of high molecular weight
16 molecules in a binder. In asphalt binders, since oxidative aging causes formation of more polar
17 molecules at the expense of the lower weight molecules, aging results in a shift towards the
18 higher molecular weight molecules. Therefore, the transformation of the molecular size
19 distribution seems to be the reason behind the changes in the performance properties of asphalts.
20
21
22 SUMMARY AND CONCLUSIONS
23 This study was conducted to evaluate various durability parameters and the effects of oxidation
24 on the properties of asphalt binders, measured at intermediate and low pavement temperatures.
25 Based on the results collected using modified testing and analysis procedures to cover durability
26 and cracking of binders, it is found that a number of useful interrelationships exist between new
27 parameters, allowing a choice when selecting simpler measures that could be used for
28 specification tests. The following points summarize the main findings of this study.
29 § The current DSR procedure for determination of the durability parameter G′/(η′/G′) is
30 time consuming since it requires a temperature-frequency sweep test with mastercurve
31 generation. Also, the suggested testing temperature of 15 °C is not applicable for specific
32 binder’s grades selected for diverse climates.
33 § An alternative approach is proposed to directly calculate the durability parameter
34 G′/(η′/G′) developed by Glover et al. (4). In the proposed approach, G′/(η′/G′) is
35 calculated from DSR test results using the frequency sweep step of the LAS test
36 procedure, performed at intermediate performance grade temperature and 10 rad/s.
37 o The suggested approach allows a quicker (30 minutes test) determination of the
38 durability parameter, and utilizes a small amount of material (8 mm DSR sample).
39 Also, the test is performed at the actual intermediate performance grade
40 temperature of the binder.
41 o With this approach, the G′/(η′/G′) paramenter can be calculated without
42 mastercurve data, and disregarding complicated shifts.
43 § By applying the LAS B parameter = -3.43, collected from a field sections showing
44 between 29%-50% of cracked segments, a limit value for the G′/(η′/G′) at intermediate
45 temperature and 10 rad/s was determined. The calculated G′/(η′/G′) parameter value at
46 cracking limit is 11.0 MPa/s, with this limit being suggested for binders with up to 50 %
47 of RAP.
Moraes and Bahia 13

1 § Failure energy at fracture (calculated from SENB at the low temperature performance
2 grade) and G′/(η′/G′) parameter (calculated from LAS frequency sweep results collected
3 at intermediate PG temperature) were found to inversely correlate very well. Higher
4 values of fracture energy, which have been correlated to a lower propensity for cracking,
5 can be an indication of higher durability of the binders, or vice-versa. This good
6 correlation between rheological measurements and fracture properties can lead to a
7 complete characterization of the effect of aging on binder properties and cracking.
8 § Failure energy at fracture also showed correlation with the ΔTc parameter. Thus, both
9 parameters appear to quantify the loss of relaxation properties as an asphalt binder ages.
10 Higher values of fracture energy (i.e., lower propensity for cracking) are seen at higher
11 positive values of ΔTc.
12 § From GPC results it was found through a correlation between the large molecular size
13 (LMS) molecules and ΔTc that asphalts with higher content of large molecular size
14 (LMS) molecules are likely to crack. It was also found that asphalt binders with higher
15 content of small molecular size (SMS) molecules are less prone to crack, as showed by
16 higher positive values of ΔTc. Therefore, the transformation of the molecular size
17 distribution seems to be the reason behind the changes in the performance properties of
18 asphalts.
19
20 It is recommended that the durability parameter G′/(η′/G′), calculated from simple
21 frequency sweep at intermediate temperature performance grade, and the DTc be used to control
22 cracking potential at intermediate and low temperatures, respectively. The limits should be
23 based on testing binders with and without recycled materials, and binder additives (e.g., oils and
24 polymers).
25
26 The authors confirm contribution to the paper as follows: study conception and design:
27 Raquel Moraes, Hussain Bahia; data collection: Raquel Moraes; analysis and interpretation of
28 results: Raquel Moraes, Hussain Bahia; draft manuscript preparation: Raquel Moraes, Hussain
29 Bahia. All authors reviewed the results and approved the final version of the manuscript.
30
31
32 REFERENCES
33 1. Bahia. H. Critical Evaluation of Asphalt Modification Using Strategic Highway Research
34 Program Concepts. Transportation Research Record: Journal of the Transportation Research
35 Board, No. 1488, 1995, pp. 82-88.
36 2. Read, J., and D. Whiteoak. The Shell Bitumen Handbook. Thomas Telford Publishing, Fifth
37 Edition, 2003.
38 3. Kandhal, P. S. Low-Temperature Ductility in Relation to Pavement Performance.
39 Low-Temperature Properties of Bituminous Materials and Compacted Bituminous Paving
40 Mixtures, ASTM STP 628. C. R. Marek, Ed. American Society for Testing and Materials,
41 1977, pp. 95-106.
42 4. Glover, C., R. R. Davison, C. H. Domke, Y. Ruan, P. Juristyarini, D. B. Knorr, and S. H.
43 Jung. Development of a New Method for Assessing Asphalt Binder Durability with Field
44 Validation. Publication FHWA/TX-05/1872-2, Report 0-1872-2, U.S. Department of
45 Transportation, 2005.
Moraes and Bahia 14

1 5. Anderson, M., G. N. King, D. I. Hanson, and P. B. Blankenship. Evaluation of the


2 Relationship between Asphalt Binder Properties and Non-Load Related Cracking. Journal of
3 the Association of Asphalt Paving Technologists, Vol. 80, 2011, pp. 615-664.
4 6. Rowe, G. M. Evaluation of the Relationship between Asphalt Binder Properties and
5 Non-Load Related Cracking (prepared discussion). Journal of the Association of Asphalt
6 Paving Technologists, Vol. 80, 2011, pp. 615-664.
7 7. AASHTO TP 101. Estimating Damage Tolerance of Asphalt Binders Using the Linear
8 Amplitude Sweep. American Association of State Highway and Transportation Officials,
9 2014.
10 8. Hintz, C., R. Velasquez, Z. Li, and H. Bahia. Effect of Oxidative Aging on Binder Fatigue
11 Performance. Journal of Association of Asphalt Paving Technologists, Vol. 80, 2011, pp.
12 527-548.
13 9. Ortega-Rodrıguez, A., S. A. Cruz, A. Gil-Villegas, F. Guevara-Rodrıguez, and C.
14 Lira-Galeana. Molecular View of the Asphaltene Aggregation Behavior in Asphaltene-Resin
15 Mixtures. Energy & Fuels, Vol. 17, 2003, pp. 1100-1108.
16 10. Ruan, Y., R. R. Davison, and C. J. Glover. Oxidation and Viscosity Hardening of
17 Polymer-Modified Asphalts. Energy & Fuels, Vol. 17, 2003, pp. 991-998.
18 11. Petersen, J. C. A Review of the Fundamentals of Asphalt Oxidation - Chemical,
19 Physicochemical, Physical Property, and Durability Relationships. Transportation Research
20 Circular E-C140, Transportation Research Board, 2009.
21 12. Tarefder, R. A., and I. Arisa. Molecular Dynamic Simulations for Determining Change in
22 Thermodynamic Properties of Asphaltene and Resin Because of Aging. Energy & Fuels, Vol.
23 25, 2011, pp. 2211-2222.
24 13. Moraes, R., and H. Bahia. Effect of Mineral Filler on Changes in Molecular Size Distribution
25 of Asphalts during Oxidative Ageing. Road Materials and Pavement Design, Vol. 16,
26 Supplement 2, 2015, pp. 55-72.
27 14. Dehouche, N., M. Kaci, and K. A. Mokhtar. Influence of Thermo-Oxidative Aging on
28 Chemical Composition and Physical Properties of Polymer Modified Bitumens. Construction
29 and Building Materials, Vol. 26, 2012, pp. 350-356.
30 15. Khattak, M. J., and G. Y. Baladi. Engineering Properties of Polymer-Modified Asphalt
31 Mixtures. Transportation Research Record: Journal of the Transportation Research Board,
32 No. 1638, 2014, pp. 12–22.
33 16. Bahia, H., D. I. Hanson, M. Zeng, H. Zhai, and M. A. Khatri. Characterization of Modified
34 Asphalt Binders in Superpave Mix Design. National Cooperative Highway Research Program
35 (NCHRP) Report 459, Transportation Research Board, National Research Council, National
36 Academies, Washington, D.C., 2001.
37 17. Collins, J. H., M. G. Bouldin, R. Gelles, R., and A. Berker. Improved Performance of Paving
38 Asphalt by Polymer Modification. Journal of Association of Asphalt Paving Technologists,
39 Vol. 60, 1991, pp. 43-79.
40 18. Rostler, F. S., and R. M. White. Composition and Changes in Composition of Highway
41 Asphalts, 85-100 Penetration Grade. Journal of Association of Asphalt Paving Technologists,
42 Vol. 31, 1962, pp. 35-89.
43 19. Stastna, J., L. Zanzotto, and O. J. Vacin. Viscosity Function in Polymer-Modified Asphalts.
44 Journal of Colloid and Interface Science, Vol. 259, 2003, pp. 200–207.
45 20. G. Polacco, A. Muscente, D. Biondi, and S. Santini. Effect of composition on the properties
46 of SEBS modified asphalts. European Polymer Journal, Vol. 42, Issue 5, 2006, pp. 1113-
47 1121.
Moraes and Bahia 15

1 21. McDaniel, R. S., and R. M. Anderson. Recommended Use of Reclaimed Asphalt Pavement in
2 the Superpave Mix Design Method: Technician's Manual. National Cooperative Highway
3 Research Program (NCHRP) Report 452, Transportation Research Board, National Research
4 Council, National Academies, Washington, D.C., 2001.
5 22. Shen, J., S. Amirkhanian, and J. A. Miller. Effects of Rejuvenating Agents on Superpave
6 Mixtures Containing Reclaimed Asphalt Pavement. Journal of Materials in Civil
7 Engineering, Vol. 19, Issue 5, 2007; pp. 376-384.
8 23. D’Angelo, J. A., K. Grzybowski, S. Lewis, R. Walker. Evaluation of the Performance
9 Properties of Asphalt Mixes Produced with Re-refined Heavy Vacuum Distillate Bottoms.
10 Canadian Technical Asphalt Association, November, 2013.
11 24. Golalipor, A. Investigation of the Effect of Oil Modification on Critical Characteristics of
12 Asphalt Binders. Ph.D. Thesis, University of Wisconsin-Madison, Madison, Wisconsin,
13 2013.
14 25. Bahia H, E. Lyngdal, R. Varma, D. Sweirtz, P. Teymourpour. Modified Binder (PG+)
15 Specifications and Quality Control Criteria. Wisconsin Highway Research Program (WHRP)
16 0092-14-22, Transportation Pooled Fund Program (TPF-5/302), 2016.
17 26. AASHTO M 320. Standard Specifications for Transportation Materials and Methods of
18 Sampling and Testing, Part 2B. American Association of State Highway and Transportation
19 Officials, 2016.
20 27. AASHTO T 240. Effect of Heat and Air on a Moving Film of Asphalt (Rolling Thin-Film
21 Oven Test). American Association of State Highway and Transportation Officials, 2013.
22 28. AASHTO R 28. Accelerated Aging of Asphalt Binder Using a Pressurized Aging Vessel
23 (PAV). American Association of State Highway and Transportation Officials, 2009.
24 29. AASHTO T 313. Standard Method of Test for Determining the Flexural Creep Stiffness of
25 Asphalt Binder Using the Bending Beam Rheometer (BBR). American Association of State
26 Highway and Transportation Officials, 2005.
27 30. AASHTO T XXX. Determining the Fracture Properties of Asphalt Binders Using the Single
28 Edge Notched Bending Test. American Association of State Highway and Transportation
29 Officials, 2011.
30 31. Bahia, H., R. Velasquez, H. Tabatabaee, and S. Puchalski. The Role of Asphalt Binder
31 Fracture Properties in Thermal Cracking Performance of Mixtures and Pavements. Canadian
32 Technical Asphalt Association, 2012.
33 32. Velasquez, R., H. Tabatabaee, H. Bahia. Low Temperature Cracking Characterization of
34 Asphalt Binders by Means of the Single-Edge Notch Bending (SENB) Test. Journal of the
35 Association of Asphalt Pavement Technologists, Vol. 80, 2011, pp. 583-614.
36 33. Bahia, H., H. Tabatabaee, and T. Mandal. Field Validation of Wisconsin Modified Binder
37 Selection Guidelines – Phase II. Wisconsin Highway Research Program Project: 0092-13-02
38 Final Report, December, 2013.
39 34. Marasteanu, M., W. Buttlar, H. Bahia, and C. Williams. Investigation of Low Temperature
40 Cracking in Asphalt Pavements. National Pooled Fund Study - Phase II, Minnesota
41 Department of Transportation Research Services, August, 2012.
42
43
44
45
46
Moraes and Bahia 16

1 LIST OF FIGURES
2
3 FIGURE 1 (a) Absolute value of the fatigue parameter B, for unmodified and modified binders.
4 (b) Absolute value of the fatigue parameter B, for recycled binders (unmodified and modified
5 with polymers). (c) Comparison of absolute value of LAS parameter B to G′/(η′/G′) calculated
6 from LAS frequency sweep results collected at intermediate PG temperature and 10 rad/s.
7
8 FIGURE 2 (a) Log G′/(η′/G′) versus ΔTc, for unmodified and modified binders. (b) Absolute
9 value of LAS parameter B versus ΔTc, for unmodified, modified and artificialy recycled binder.
10
11 FIGURE 3 (a) Correlation between failure energy at fracture and G′/(η′/G′), for modified
12 binders. (b) Correlation between failure energy at fracture and G′/(η′/G′) for artificialy aged
13 binder.
14
15 FIGURE 4 (a) Typical GPC chromatogram of an asphalt binder showing three elution time
16 fractions (13). (b) Correlation between failure energy at fracture (calculated from SENB at the
17 low temperature performance grade) and ΔTc. (c) Correlation between LMS and ΔTc. (d)
18 Correlation between SMS and ΔTc. (e) Molecular size distribution of asphalt binders, before and
19 after oxidative aging.
20
21
22
23
24
25
26
27 LIST OF TABLES
28
29 TABLE 1 Performance Grade of the selected unmodified and modified binders.
30 TABLE 2 Selected testing for materials evaluation.
31 TABLE 3 LAS B parameter and field performance information (33).
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
Moraes and Bahia 17

1
2 (a)

3
4
5 (b)

6
7 (c)
8 FIGURE 1 (a) Absolute value of the fatigue parameter B, for unmodified and modified binders.
9 (b) Absolute value of the fatigue parameter B, for recycled binders (unmodified and modified
10 with polymers). (c) Comparison of absolute value of LAS parameter B to G′/(η′/G′) calculated
11 from LAS frequency sweep results collected at intermediate PG temperature and 10 rad/s.
Moraes and Bahia 18

1
2 (a)
3
4

5
6 (b)
7
8 FIGURE 2 (a) Log G′/(η′/G′) versus ΔTc, for unmodified and modified binders. (b) Absolute
9 value of LAS parameter B versus ΔTc, for unmodified, modified and artificialy recycled binder.
10
11
12
13
14
15
16
17
18
19
20
Moraes and Bahia 19

1
2
3 (a)
4
5

6
7
8 (b)
9
10
11 FIGURE 3 (a) Correlation between failure energy at fracture and G′/(η′/G′), for modified
12 binders. (b) Correlation between failure energy at fracture and G′/(η′/G′) for artificialy aged
13 binder.
14
15
16
17
18
19
20
Moraes and Bahia 20

1
2 (a) (b)

3
4 (c) (d)
5

6
7 (e)
8 FIGURE 4 (a) Typical GPC chromatogram of an asphalt binder showing three elution time
9 fractions (13). (b) Correlation between failure energy at fracture (calculated from SENB at the
10 low temperature performance grade) and ΔTc. (c) Correlation between LMS and ΔTc. (d)
11 Correlation between SMS and ΔTc. (e) Molecular size distribution of asphalt binders, before and
12 after oxidative aging.
Moraes and Bahia 21

1 TABLE 1 Performance Grade of the selected unmodified and modified binders.


2
High Temp. Intermediate Low Temp.
Modifier (Unaged & Temp. True Grade PG Grade
Asphalt Binder RTFO aged) (PAV aged) (PAV aged), (°C)
True True High-Low
Type and Content (%) S m-value
Grade (°C) Grade (°C) Temp.
C N/A 61.8 18.4 -25.8 -26.3 58-22
(Russian crude source)
D N/A 62.5 15.9 -28.9 -28.7 58-28
(Russian crude source)
E N/A 69.9 12.6 -34.2 -33.1 70-28
(Russian crude source)
G N/A 52.3 14.5 -28.4 -30.0 52-28
(Russian crude source)
M-A 5% Oil-A 58.3 10.8 -29.9 -32.2 58-28
(Venezuelan crude source)
M-B 10% Oil-B + 2%
Plastomer
58.5 6.7 -38.1 -38.5 58-34
(Russian crude source)
M-C 8% Oil-B + 3% Elastomer 59.7 7.9 -36.6 -37.6 58-34
(Russian crude source)
M-D 8% Oil-B + 2% Plastomer 62.5 6.6 -38.4 -38.1 58-34
(Russian crude source)
M-F 8% Oil-B + 4% Plastomer 59.7 7.9 -36.8 -39.8 58-34
(Venezuelan crude source)
M-G 8% Oil-B + 5% Elastomer 59.2 9.1 -38.7 -39.9 58-34
(Russian crude source)
A-RAP1 N/A 90.1 N/A -15.4 -12.2 88-12
A-RAP2 N/A 105.3 N/A -21.4 -14.6 100-12
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
Moraes and Bahia 22

1 TABLE 2 Selected testing for materials evaluation.


2
Property Test Type Standard Testing Conditions
AASHTO M 320 (26) and
Dynamic Shear @ High PG and Intermediate PG
PG Grading associated laboratory testing
Rheometer (DSR) Temperature
procedures
Rolling Thin-Film Oven @ 325 °F (163 °C) for 85
AASHTO T 240 (27)
(RTFO) minutes
Aging
Pressure Aging Vessel @ 2.10 MPa and 100 °C for 20
AASHTO R 28 (28)
(PAV) hours
Intermediate Frequency Sweep
Proposed revised
Temperature Linear Amplitude Sweep Amplitude Sweep
version of the
Cracking (LAS) Tested @ Intermediate PG
AASHTO TP 101-12 (7)
Resistance Temperature
Bending Beam
Low Temperature AASHTO T 313 (29) @ Low PG Temperature
Rheometer (BBR)
Cracking
Single Edge Notched
Resistance AASHTO T XXX-11 (30) @ Low PG Temperature
Beam (SENB)
Molecular Size
Gel Permeation Flow rate of 1 mL/min @ 40 °C
Distribution ---
Chromatography (GPC)
(MSD)
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
Moraes and Bahia 23

1 TABLE 3 LAS B parameter and field performance information (33).


2
Project ID Cracked Segments LAS B parameter @ IT
PG
1190-14-70 67% -3.39
9049-09-70 14% -3.61
1130-12-71 100% -3.06
9140-07-70 0% -3.17
1177-10-70 29% -3.41
6590-00-70 50% -3.44
31-20-06-70 100% -3.34
3

You might also like