You are on page 1of 13

Construction and Building Materials 190 (2018) 1295–1307

Contents lists available at ScienceDirect

Construction and Building Materials


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

Effect of moisture in aggregate on adhesive properties of warm-mix


asphalt
Shih-Hsien Yang a,⇑, Firmansyah Rachman b, Hery Awan Susanto a
a
Department of Civil Engineering, National Cheng Kung University, 1 University Rd, Tainan City 70101, Taiwan
b
Department of Civil Engineering, Muhammadiyah University, 91 Muhammadiyah Rd., Banda Aceh, Aceh 23245, Indonesia

h i g h l i g h t s

 Work of adhesion of may drop up to 50% when aggregate moisture content of 10%.
 Bond energy of WMA mixes diminish less with increasing aggregate moisture contents.
 PM binder bond energy showed about 6.7 times greater than that of AC20 binders.
 Quadrant plot can effectively rank the asphalt-aggregate adhesion properties.

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

Article history: Warm-mix asphalt (WMA) is one technology that has been found to have potential in minimizing the
Received 18 May 2016 environmental harm caused by road construction activities. Its low mixing and compacting temperatures
Received in revised form 28 August 2018 have opened great opportunities for more sustainable development. However, because WMA is mixed at
Accepted 30 August 2018
low temperatures, the drying process of the aggregate may fail to dry the WMA completely. The binder-
Available online 24 October 2018
aggregate adhesion property can be damaged by moisture remaining from incomplete drying. The present
study investigated the effects of aggregate moisture on the binder-aggregate interfacial characteristics of
Keywords:
WMA by evaluating its surface free energy and adhesive bond strength. The results showed that for the
Warm-mix asphalt
Moisture damage
Pen 60-70 binder type, the aggregate AG in general showed the higher bond energy and aggregate SD
Work of adhesion showed the lowest bond energy among three aggregate types. Also, the bond energy diminishes with
Bond energy increasing aggregate moisture contents but the percent drops in bond energy is less when WMA additives
Moisture aggregate was included. For the PM binder, SL aggregate group showed the highest bond energy and SD aggregate
Interfacial adhesion group showed the lowest bond energy. For the work of adhesion, the results showed that with aggregate
with 10% moisture contents has destructive effects on the work of adhesion which may decrease to 50%
compare to dry aggregate condition. Finally, the four quadrant diagrams can be used to combine two
distinct property of the bond energy and the work of adhesion as the hybrid performance indicator to rank
the effect of moist aggregate to the adhesion properties at binder-aggregate interface.
Ó 2018 Elsevier Ltd. All rights reserved.

1. Introduction and characteristics are equal or superior to those of pavement


produced with HMA [1–3]. The lower production temperature of
Warm-mix asphalt (WMA) is produced at a temperature WMA may leave residual moisture inside aggregates because the
approximately 30 °C lower than the temperature used in the pro- low heat of the aggregate may be insufficient to vaporize all water.
duction of hot-mix asphalt (HMA). According to the technology In addition, several WMA technologies introduce additional mois-
used to reduce the production and compaction temperature, ture in the production process, which may lead to the weakening of
WMA can be divided into three categories: (1) foaming process the adhesion bond between asphalt binder and aggregates [26,27].
[18,28,29]; (2) organic additives [1,14,18]; (3) chemical additives If asphalt concrete is weakened by moisture or insufficient
[1,12]. The fundamental principles of WMA technologies are adhesion, various pavement distresses, such as raveling, potholing,
believed to guarantee that the pavement’s performance, durability, rutting, and cracking, may eventually develop [4,5]. As the conse-
quence, adhesion promoters such as hydrated lime of liquid anti-
stripping additive are commonly used to enhance the adhesion
⇑ Corresponding author.
property between asphalt and mineral aggregate [1,13,14]. Khodaii
E-mail address: shyang@mail.ncku.edu.tw (S.-H. Yang).

https://doi.org/10.1016/j.conbuildmat.2018.08.208
0950-0618/Ó 2018 Elsevier Ltd. All rights reserved.
1296 S.-H. Yang et al. / Construction and Building Materials 190 (2018) 1295–1307

et al. investigate the effect of hydrated lime percentage to reduce SonneWarmix and Evotherm in their study and adopted pull-off
the moisture susceptibility of HMA. The results showed that with bond test to investigate the moisture susceptibility of mixtures.
dense grade mixture, saturated ITS values was affected as the per- The bond strength test result showed that only Sasobit had a sig-
centage of hydrated lime increased and reached a peak by adding nificant effect on the pull-off tensile strength of the binder [15].
2% of hydrated lime in WMA mixtures [13]. Xiao et al. investigated Alavie et al. performed the bitumen bond strength test to evaluate
the effect of hydrated lime to improve the performance of WMA the pull-off tensile strength of binder with three types of WMA
additives of Asphamin and Sasobit contained moist aggregate additives and moist aggregates. The results showed that produc-
and the results showed that the addition of hydrated lime tion at reduced temperatures (WMA case) can increase moisture
improved both wet ITS, TSR values and rutting resistance of mix- susceptibility because of lower adhesion between the asphalt and
tures [14]. aggregate [16]. For the surface free energy and work of adhesion
One of the major consequence of moisture damage to asphalt approaches, Wasiuddin et al. measured the contact angles of vari-
concrete is mix tenderness. The escape of residual moisture ous binder-WMA additive blends to evaluate their SFE levels. The
reduces asphalt binder film stiffness and thus reduces the stiffness study concluded that SasobitÒ in general reduces the SFE of asphalt
of the mix. The lower mix stiffness increases the cracking potential binders, but Aspha-Min did not affect SFE significantly [18]. Yu and
along the roller edge and causes unevenness at the end of rolling his co-workers applied surface free energy approach and compare
that increases pavement roughness. In some cases, mixes may with the tensile strength ratio (TSR) of to investigate the moisture
remain softer than normal for several weeks, which may lead to sensitivity of the warm rubberized mixtures. The results indicated
other pavement distresses [6]. In general, three types of that the ratio (RSFE) of work of adhesion of the aggregate–water-
approaches to evaluating WMA technologies have often been used: asphalt binder system to work of adhesion of the asphalt-
(1) field performance investigation [7–10]; (2) mechanical charac- aggregate system can be an potential index that represents the
terization of WMA binder or mixtures in the laboratory [1,11–14]; residual potential for the asphalt binder to spontaneously bond
and (3) analytical studies based on fundamental material proper- the aggregate. A good linear relationship was reported between
ties [17–19]. Field performance data collected from Nebraska and RSFE with the TSR results [19].
Michigan studies have shown no significant difference between As stated previously, the low mixing temperature of WMA may
WMA and HMA test sections [7,8]. Estakhri monitored the perfor- harm binder-aggregate adhesion by leading to moisture-induced
mance of more than 10 WMA projects in Texas. The results showed damage. The cited studies have clearly demonstrated that WMA
that the field performance of WMA is comparable to that of HMA may increase moisture susceptibility, especially in the early stage
regardless of the type of WMA; however, if the mix has a tendency of the pavement service life. Although some studies have investi-
toward moisture sensitivity, some WMA additives can exacerbate gated the effects of aggregate moisture conditions on mixture per-
instability or stripping of mixes [9]. Yang and Keita compared the formance, no direct observation has been performed for analyzing
binder aging rates for WMA and HMA after 2 years of field aging the effect of asphalt binder-aggregate interfacial bond energy and
and the results from the extracted binders showed that the aging work of adhesion in various WMA binders and moist aggregates.
rate of the WMA binder was lower than that of the HMA binder Thus, this study demonstrated experimental techniques which
[10]. allowed to directly observe the effects of various aggregate mois-
Methods used to quantify the moisture susceptibility of asphalt ture levels on the binder-aggregate interfacial adhesion character-
mixtures in the laboratory can be categorized into tests that char- istics of WMA samples by evaluating their work of adhesion and
acterize bulk mixture mechanical properties and tests that charac- bond strength.
terize binder-aggregate interfacial properties. Regarding bulk
mixture mechanical properties, test methods and performance 2. Materials and methodology
indicators that are commonly used include the mixture resilient
modulus [1,11], indirect tensile strength [1,11,12,14], tensile stress 2.1. Materials
ratio [1,12–14], creep compliance [11], Hamburg wheel tracking
2.1.1. Aggregate and binder
test [1,12], fracture energy ratio [11] and fatigue cracking resis- The aggregates used in this study were crushed river gravels collected from
tance. Martin et al. performed an extensive study on the moisture local aggregate quarries in Chang Hua and Ping Tong, Taiwan. Argillite (AG), sand-
susceptibility of WAM using Evotherm, Sasobit and Rediset addi- stone (SD), and slate (SL) were found to be the major aggregate types in these two
tives with specimen from both laboratory and field. Their results quarries. Table 1 shows the average chemical composition of selected aggregates.
These three aggregates have relatively similar chemical compositions; SiO2 and
showed that WMA can be more moisture susceptible in its early Al2O3 constitute more than 80% of all aggregates in this study. Based on the
life [1]. Gong et al. indicated that the fracture energy-based param-
eters appeared to be more distinctive moisture effect than other
parameters. They reported that a decreasing trend of energy ration Table 1
was observed in the moisture condition and incompletely drying Average Chemical Composition of Selected Aggregates.

aggregate mix specimen [11]. Benert et al. performed study on Constituent (%) Argillite (%) Sandstone (%) Slate (%)
WMA additives of Sasobit, Rediset, Evotherm using Hamburg SiO2 58.82 78.33 60.64
wheel tracking and Texas overlay tester. Their results showed that TiO2 0.73 0.25 0.73
the mixing temperature of the WMA and initial moisture content Al2O3 16.46 4.77 17.32
of the aggregate blend will have a significant impact on the final Fe2O3 1.1 1.07 2.25
FeO 7.2 0.33 3.66
mixture moisture damage potential performance [12]. Xiao et al.
MnO 0.09 – –
study effect of moist aggregate in WMA using Asphamin and MgO 4.92 1.16 2.6
Sasobit additives. The results showed that the rutting resistance CaO 0.76 5.5 1.54
of mixtures decreases when contained moist aggregate [14]. The Na2O 4.03 0.45 1.19
binder-aggregate interfacial adhesive properties have also been K2O 1.6 1.31 3.69
H2O 3.84 1.63 4.13
used to characterize the effects of moisture susceptibility in P2O5 0.17 0.08 –
WMA. Binder-aggregate interfacial properties such as mechanical CO2 0.01 5.03 1.47
bond strength [15,16], thermodynamic surface free energy (SFE) SO3 0.02 0.07 –
[17–19], and work of adhesion [17–19] are commonly used BaO – 0.05 0.38
S 0.05 – –
parameters. Mogawer et al. used WMA additives of Advera, Sasobit,
S.-H. Yang et al. / Construction and Building Materials 190 (2018) 1295–1307 1297

Z x2
percentage of SiO2 present in an aggregate, it can be classified as acidic (63–100%), 1
Gf ðDT Þ ¼ f ðxÞdx ð1Þ
basic (45–52%) or intermediate (52–63%) in nature. The silica content of Argillite pr 2 x1
and Slate was found to be 58.82% and 60.64% respectively. Thus it falls under the
category of intermediate meaning they are neither acidic nor basic. On the other where Gf ðDT Þ = bond energy (kN/m) and r = the radius of the contact area (mm).
hand, the silica content of Sandstone was found to be 78.33% that falls under the
category of acidic which means the aggregate is hydrophilic in nature. Two types
of asphalt binder were used in this study. One was penetration grade asphalt 60– 2.2.2. Sessile drop method for work of adhesion
70 (Pen 60-70) and the other was type III SBS polymer-modified asphalt (PM). Another way to characterize the asphalt-aggregate interfacial adhesive property
The basic properties of the binders are shown in Table 2. is to measure the work of adhesion between asphalt and aggregate that was calcu-
lated from the individual SFE values of asphalt and aggregate [22–25]. The SFE of
asphalt or aggregate was obtained by measuring the contact angle between asphalt
2.1.2. WMA additives
or aggregate and standard probe liquids through the sessile drop method by using
In addition, two WMA additives, namely EvothermÒ and SasobitÒ, were used in
First Ten Angstroms 125. The sessile drop method is performed by directly measur-
this study. EvothermÒ DAT (Dispersed Asphalt Technology) is a chemical additive
ing the contact angle between liquid and substrate. This method has been widely
technology that can improve coating, compaction, and mixture workability. The
reported in the literature and used in various fields such as chemistry, coatings,
recommended proportion for EvothermÒ is 0.4–0.8% of the weight of binder. Saso-
mining, painting, and geology [22,23]. Sample preparation for the sessile drop test
bitÒ is an organic additive manufactured through the Fischer–Tropsch process.
is described as follows. For 15 min, the P67 and Type III PM binders with or without
SasobitÒ has a fine crystalline morphology and is soluble in asphalt above 115 °C.
WMA additives were heated to 135 ± 5 °C and 165 ± 5 °C, respectively. To ensure a
Once melted into the asphalt, SasobitÒ forms a homogeneous solution that can
flat asphalt film surface, the melted binders were poured onto a glass microscopy
reduce the viscosity of the binder. The recommended proportion for SasobitÒ is
slide and then placed on top of three 1-mm-thick glass slides, which were sand-
0.8–4% of the weight of binder. For this study, in weight, 0.5% of the asphalt was
wiched between two 5-mm-thick pieces of glass. The excess binder was then
EvothermÒ DAT and 2.5% of the asphalt was SasobitÒ.
trimmed using a hot razor blade. Fig. 2 shows the setup of a sessile drop test for bin-
ders. On the other hand, for the aggregate substrate sample preparation, a sample
2.1.3. Preparation of moist aggregate preparation procedure which was similar to that of the pull-off test sample was
The effects of the moisture content of aggregate on the binder-aggregate inter- followed.
facial adhesion properties of the WMA were investigated by simulating various Next, to calculate the SFE of binder and aggregate, the procedure below was fol-
aggregate drying levels during the mixing process. Three aggregate moisture condi- lowed [23,25]. From the well-known Young equation describes the balance at the
tions were simulated, namely 0% (completely dry conditions), 5%, and 10%. The three phase contact of solid-liquid and gas.
aggregates were first cut into 2 cm  2 cm  1 cm cuboids. To simulate completely
dry conditions, the aggregates were heated in an oven for 6 h at a temperature of csv ¼ csl þ clv cos hc ð2Þ
135 °C and the sample weight was measured (W0). For the 5% and 10% aggregate
moisture conditions, the aggregate samples were first submerged in water within where csv is surface tension of solid in contact with vapor (mN/m); csl is interfacial
a vacuum chamber for 30 min and then the weight of the fully saturated samples tension between solid and liquid; clv is surface tension of liquid in contact with
was measured (W100). The weight of 5% (W5) and 10% (W10) moist aggregates can vapor (mN/m); and hc is contact angle. To solve the above equation, the work of
be determined according to the interpolation between W0 and W100. Aggregates adhesion concept was used. First, the work of adhesion between solid phase and liq-
with 5% and 10% moisture content were obtained by heating the aggregate samples uid face can be written as equation (3).
until the weight of the samples reached W5 and W10 with the precision of
1  105 g. W sl ¼ clv þ csv  cls ð3Þ

2.2. Methods where W ls is the work of adhesion of the liquid-solid interface (mN/m). Combine
equations (2) and (3) into Young’s-Dupre equation as equation (4).
2.2.1. Pull-off test for bond strength
The characteristics of the interface between asphalt and aggregate were W sl ¼ ð1 þ cos hÞclv ð4Þ
explored by measuring interfacial bond energy using pull-off test. This method is According to Van Oss et al.’s three-component theory of surface energy, the
effective for comparing pairs of adhesives and adherents, but since the test results total surface energy consists of an acid, a base, and dispersive components and
are geometry dependent, these tests can be used only for qualitative assessment of can be formulated as equation (5) [25].
the bond [20,21]. Soltesz et al. suggested that the maximum adhesive strength of
asphalt and aggregate is defined as the average stress, ravg, which can be applied W ls ¼ cLW
ls þ cls
þ
ð5Þ
in a normal direction to the surface without damaging the material [21]. In this
study. the pull-off test was performed using an MTS Tabletop 858 system with a where c is Lifshitz-van der Waals (LW) component of the total surface free
LW
ls
specially designed sample fixture to measure the interfacial bond energy between energy (mN/m) and cþls = Lewis acids and base components of the total surface free
asphalt and aggregate. First, the aggregate samples were cut into energy (mN/m). Furthermore, in terms of individual surface free energy compo-
2 cm  2 cm  1 cm blocks. After the cutting, the aggregate samples were polished nents, equation (5) can be expressed as equation (6) [25]:
using two abrasive papers, numbers 600 and 1200. The samples were then washed
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffiffiffiffiffi
using distilled water and ethanol (analytical regent). Next, aggregate samples was
W ls ¼ 2 cLWs þ cLW l þ 2 cþs cl þ 2 cs cþl ð6Þ
glued to the bottom alumina sample holder (Fig. 1b) and was heated in oven at
135 °C for 1 h to completely dry any remaining liquid or moisture. Third, the aggre-
gate sample was conditioning to various moisture level as described in previous where cLW
s=l = Lifshift–Van der Waals surface energy component of liquid/solid (mN/

section. Fourth, the liquid binder was then dripped on top of the aggregate and m); cþ
s=l = Lewis acid component of surface energy component of liquid/solid (mN/
the special designed top circular pull stub (Fig. 1a) used to compress the asphalt m); and c 
s=l = Lewis base component of surface energy component of liquid/solid
binder to form the asphalt-aggregate adhesion film with dimension of 50 lm in (mN/m).
thick and 10 mm in diameter (Fig. 1c). The excess binder was removed using a Finally, the Young-Dupre equation to calculate the surface free energy of solid
hot razor blade. Lastly, the fixture was placed in MTS machine (Fig. 1d) in an envi- material by measuring the contact angle are obtained by substituting equation
ronmental chamber with test temperature of 25 °C. A constant strain rate of (5) to equation (6) as shown in equation (7) [25].
0.05 mm/s was applied on the pull stub to measure the force required to separate
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffiffiffiffiffi 
the asphalt film and aggregate. The unit bond energy of the asphalt aggregate inter- pffiffiffiffiffiffiffiffiffiffiffi
ð1 þ coshÞcl ¼ 2 cLWs cl
LW
þ cþl cs þ cl cþs ð7Þ
face can be calculated using the area underneath the force-displacement curve
divided by the contact area, as shown in Eq. (1).
In equation above, there are three unknown surface free energy components of
solid which are cLWs , cs , and cs . In order to quantify those unknown surface energy
þ 

values, at least three probe liquids with known surface free energy is needed. Here,
water, glycerol, and ethylene glycol were used for contact angle measurement. The
surface tension components for these liquids are shown in Table 3 [22]. The images
Table 2
of three probe liquid drops were captured by an installed camera and the contact
Properties of Pen 60-70 and PM Binders.
angles were measured as shown in Fig. 3. A minimum of three drops for each probe
Test Pen 60-70 PMB liquid were applied and the standard deviation of each measured contact angle was
not more than 5°. The mean contact angles as presented in Tables 4 and 5 of each
Viscosity at 135 °C, Pa-s (ASTM D-4404-02) 0.38 2.2
probe liquid were used to calculate the SFE of binder or aggregate samples. The SFE
Viscosity at 150 °C, Pa-s (ASTM D-4404-02) 0.2 1.9
components of binder and aggregate used in this study are shown in Tables 6 and 7.
Penetration, 25 °C, 0.1 mm (ASTM D-5) 66 43
Once the surface free energy components were obtained then equation (6) was used
Softening point, °C (ASTM D-36) 53.5 69
to calculate the work of adhesion between asphalt and aggregate.
1298 S.-H. Yang et al. / Construction and Building Materials 190 (2018) 1295–1307

Fig. 1. Design of (a) Top and (b) Bottom Sample Molds (mm) (c) assembly of top mold and aggregate substrate and (d) pull-off test setup in the MTS machine.

binder type, WMA additive group and aggregate moisture level. It


is evident that aggregate type is a statistically significant source of
variability for the bond energy between binder and aggregate
except for the aggregate contained 10% moisture level. The results
were as expected since the aggregate surface chemistry plays a
vital role in adsorbing the asphaltic components and retaining
them on its surface. Thus, the aggregate type variable causes signif-
Fig. 2. Schematic Setup of the Sessile Drop Test for Asphalt Binders. icant differences in bond energy between the binder groups. The
high SiO2 contents sandstone as acidic aggregate demonstrates
the weakest bond energy among three types of aggregate used in
3. Results and discussion
the study. In addition, although the AG and SL has similar SiO2 con-
tent and were classified as intermediate aggregate in nature. The
3.1. Bond energy from direct pull-off test
bond energy of AG and SL for Pen 60-70 and PM showed com-
pletely opposite trend. In the Pen 60-70 group, AG showed the
3.1.1. Bond energy for different aggregate types
highest performance with an average bond energy of 1.58 kN/m,
The average bond energy values for different aggregate types
whereas in the PM group, SL had the highest bond energy
are shown in Fig. 4. There appears to be differences of bond energy
(7.27 kN/m). In addition, SD had the lowest performance in both
when different type of aggregate used within each binder/WMA
the Pen 60-70 and PM groups, with average values of 0.48 kN/m
additive/aggregate moisture level. In order to analyze this, statisti-
and 6.26 kN/m, respectively. In Table 10, ANOVA for each aggre-
cal analysis was done according to 95% confident level to examine
gate type was conducted to examine which variability factors are
if there were statistically significant differences of bond energy
significant in affecting the bond energy. For all three aggregate
when used different types of aggregate for each binder type mod-
type, binder type and aggregate moisture level are statistically sig-
ified with three WMA additive group at three aggregate moisture
nificant source of variability by themselves while WMA additive
levels. To be a statistical significant source of variability, the
group is not statistically significant source of variability when
p-value must be less than or equal to 0.05. Three-way analysis of
use of sandstone aggregate. The interaction of WMA additive group
variance (ANOVA) results presented in Tables 8, 9, and 11 for each
and aggregate moisture level is not statistically significant for all
three aggregate types.

Table 3
3.1.2. Bond energy for different aggregate moisture contents
Surface Tension Components of Probe Liquids (mN/m) [22].
Fig. 5 shows the interfacial bond energy with respect to various
Liquid Probe cTotal cLW c+ c aggregate moisture levels within each binder/WMA additive/
Water 72.8 21.8 25.5 25.5 aggregate type. The trends in Fig. 5 for the bond energy results
Glycerol 64 34 3.92 57.4 show that as aggregate moisture level increases, the bond energy
Formamide 58 39 2.28 39.6
decreases. Three-way ANOVA results presents in Tables 8, 9, and
Methyline Iodide (MI) 50.8 50.8 0 0
Ethylene Glycol (EG) 48 29 1.92 47 10 for each binder type, WMA additive group and aggregate type.
It appears that aggregate moisture level is statistically significant
S.-H. Yang et al. / Construction and Building Materials 190 (2018) 1295–1307 1299

aggregate type and moisture level group. In Table 8, ANOVA for


each binder type was conducted to examine which variability fac-
tors are significant in affecting the bond energy. The results
showed that WMA additive group, aggregate type, aggregate mois-
ture level as well as interaction of WMA additive group and aggre-
gate type are statistically significant source of variability.

3.1.4. Bond energy for different WMA additives


Fig. 8 shows the bond energy with respect to different WMA
additive group within each binder/aggregate type/aggregate mois-
ture level. For the Pen 60-70 binder type, samples without WMA
additives and samples with EvothermÒ showed an average bond
energy of 1.12 kN/m, which was higher than that of samples with
SasobitÒ. For the Polymer modified binder, some samples
with SasobitÒ showed higher bond energy than those of samples
with EvothermÒ additives and the control group. ANOVA was per-
Fig. 3. Sample Sessile Drop Image. formed and the results were presented in Tables 8, 10, and 11 for
each binder type, aggregate type and aggregate moisture level. It
appears that WMA additive group is statistically significant source
source of variability by itself but not when it interacts with other of variability for different binder type (Table 8) and aggregate type
variables such as aggregate type and WMA additive group. On (Table 10) but not statistically significant source of variability for
average, the bond energy decreased by 1.4% and 1.01% for every different level of aggregate moisture (Table 11). In Table 9, ANOVA
1% moisture content increase for the Pen 60-70 and PM groups, for each WMA additive group was conducted to examine which
respectively. It was suspect that the drops of bond energy maybe variability factors are significant in affecting the bond energy. For
due to some of the aggregate surface pores were occupied by the no WMA additive group, binder type, aggregate type, aggregate
water and reduce the mechanical interlock, the absorption of moisture level, WMA additive group and aggregate type interac-
asphalt into surface pores of aggregate surface, as well as asphalt tion and binder type and aggregate moisture level interaction are
surface coating area to the aggregate. In Table 11, ANOVA for each statistically significant source of variability. For the group using
aggregate moisture level was conducted to examine which vari- EvothermÒ additive, interactions of binder type  aggregate moist
ability factors are significant in affecting the bond energy. For all level and aggregate type  aggregate moist level are not statisti-
three aggregate moisture level, binder type is the only statistically cally significant source of variability. For the group using SasobitÒ
significant source of variability for all aggregate moisture level. additive, interactions of binder type  aggregate type and aggre-
Aggregate type is statistically significant source of variability for gate type  aggregate moist level are not statistically significant
aggregate moisture level of 0% and 5%. All the interactions (Binder source of variability.
type  WMA additive group, Binder type  Aaggregate type, WMA
additive group  Aaggregate type) are not statistically significant 3.2. Work of adhesion from sessile drop test
sources of variability for bond energy.
3.2.1. Effect of aggregate type
3.1.3. Bond energy for different binder types Fig. 9 shows the results of work of adhesion ordered by aggre-
Fig. 6 shows a very obvious difference between the pull-off gate type. SD performed the highest with 0% moisture content,
results of Pen 60-70 samples and those of PM binder samples. showing an average work of adhesion of 59.07 mN/m. The result
The average bond energy value of the Pen 60-70 samples was was as expected since SD has highest silica content which appears
1.02 kN/m, whereas the PM binder samples showed higher bond negative charge on the mineral surface. The negative charged
energy values, the average of which was 6.73 kN/m. Fig. 7(a) and aggregate surface attracts and orients bitumen polar molecules
(b) shows force-displacement plots for samples with the Pen 60- which is a major contributor to provide adhesion force between
70 and PM binders. Although the PM binder samples had slightly binder and aggregate. On the other hand, at aggregate moisture
different values of maximum force, the energy required to cause levels of 5% and 10%, AG outperformed the other two, showing val-
failure of the material, which is the area under the curve, was much ues of 50.97 mN/m and 33.97 mN/m (respectively) for work of
more than the corresponding value for the Pen 60-70 binder sam- adhesion. According to the overall testing results, AG has the high-
ples. On average, PM binder showed approximately 6.7 times est work of adhesion value, followed by SD and SL (48.00, 47.18,
greater bond energy than the Pen 60-70 binders. Furthermore, and 36.48 mN/m respectively). Again, the results of ANOVA were
the results of three-way ANOVA analysis were presented in Tables shown in Tables 12, 13, and 15 for each binder type, WMA additive
9–11 for each WMA additive group, aggregate type and moisture group and aggregate moisture level. The aggregate type is statisti-
level. Binder type appears to be a significant source of variable cally significant source of variable to effect work of adhesion.
by itself but not when it interacts with other variable such as Table 14 showed the ANOVA analysis for each aggregate type

Table 4
Contact Angle Measurement Results for Asphalt Binders.

Asphalt Binder Ethylene Glycol Glycerol Water


Angle (°) COV Angle (°) COV Angle (°) COV
P67 77.37 2% 95.68 1% 91.03 2%
P67EV 72.23 2% 89.75 1% 81.29 1%
P67SS 78.30 2% 96.02 1% 100.65 1%
PM 85.42 2% 90.81 3% 81.11 2%
PMEV 82.77 2% 91.48 2% 88.61 1%
PMSS 83.46 1% 91.51 2% 97.92 1%
1300 S.-H. Yang et al. / Construction and Building Materials 190 (2018) 1295–1307

Table 5
Contact Angle Measurement Results for Aggregates.

Aggregate Ethylene Glycol Glycerol Water


Angle (°) COV Angle (°) COV Angle (°) COV
AG0% 29.42 4% 83.53 3% 75.32 4%
AG5% 32.47 6% 79.43 5% 68.30 6%
AG10% 36.33 2% 82.16 2% 60.92 3%
SD0% 28.86 4% 74.86 3% 57.13 4%
SD5% 23.68 6% 74.49 5% 20.40 7%
SD10% 23.45 5% 83.65 7% 44.75 5%
SL0% 53.84 4% 101.64 2% 87.19 5%
SL5% 65.90 3% 107.40 2% 81.48 2%
SL10% 43.29 3% 84.90 4% 68.68 4%

Table 6
Surface Free Energy Components of Aggregate in mN/m.

No. Sample cLW c+ c c


mN/m mN/m mN/m
1 0%AG 50.09 0 3.1 53.19
2 5%AG 37.3 0 17.13 37.3
3 10%AG 16.57 0 24.13 35.98
4 0%SD 61.13 0 0 61.13
5 5%SD 35.71 0 30.06 35.71
6 10%SD 10.01 0 44.2 54.21
7 0%SL 29.53 0 21.69 29.53
8 5%SL 19.44 0 16.54 35.98
9 10%SL 10.69 0 13.79 24.49

Table 7
Surface Free Energy Components of Asphalt in mN/m.

No. Sample cLW c+ c c


mN/m mN/m mN/m
1 P67 18.19 0 9.438 18.19
2 P67EV 20.46 0 15.87 20.46
3 P67SS 22.09 0 2.294 22.09
4 PM 18.82 0 20.79 18.82
5 PMEV 16.53 0 13.57 16.53
6 PMSS 9.912 0 5.457 15.37

10
AG SD SL presence of water or moisture on the surface of the aggregate
8
can reduce the mechanical interlock, thereby increasing the sus-
Bond Energy (kN/m)

6 ceptibility of asphalt mixtures to stripping [24]. Thus, the work


of adhesion between asphalt and aggregate is decreased by the
4
increase of moisture content. Three-way ANOVA results were pre-
2 sented in Tables 12–14 for each binder type, WMA additive group
and aggregate type. The residual moisture level is a significant
0
source of variable in all cases. It is understandable since asphalt
binder is known with weak chemical affinity for aggregate,
whereas the aggregate is characterized by a strong affinity for
water. Thus, with increasing level of residual moisture content,
Binder Type/ WMA Additive Group/ Agg. Moisture Content
the lower the work of adhesion is expected.
Fig. 4. Pull-Off Test Results by Aggregate Type.

and both binder type and aggregate moisture level are significant 3.2.3. Effect of binder type
source of variables for all three aggregate types while WMA addi- Fig. 11 shows the results regarding work of adhesion for the Pen
tive group is not statistically significant source of variability when 60-70 and PM binders. The work of adhesion for the PM binder was
use of sandstone aggregate. This results coincide to the results of observed to be almost the same as the work of adhesion for the Pen
bond energy. 60-70 binder within control group. However, when WMA additives
were used, the work of adhesion of Pen 60-70 binder was signifi-
cant higher than work of adhesion of PM binder. It also observed
3.2.2. Effects of aggregate moisture content that the work of adhesion difference between Pen60-70 and PM
Fig. 10 clearly shows that moisture in aggregates reduces work binder is greater in the SS WMA additive group than that in the
of adhesion for asphalt and aggregate. The work of adhesion can be EV and control group. This implied that SS additive improve the
reduced to approximately 50% of its standard value by the pres- wetting between Pen60-70 binder and aggregate especially when
ence of 10% moisture content. This result is expected because the AG and SD aggregate were used. However, the improvement was
S.-H. Yang et al. / Construction and Building Materials 190 (2018) 1295–1307 1301

Table 8
ANOVA results for bond energy when use of Pen 60-70 and Polymer Modified binder.

P67 Binder
SS df MS F p-value sig
Additive 0.467 2 0.233 140.392 5.889E07 Yes
Aggregate 5.391 2 2.695 1621.747 3.665E11 Yes
Moist% 0.113 2 0.056 33.990 1.229E04 Yes
Additive  Aggregate 0.344 4 0.086 51.689 9.341E06 Yes
Additive  Moist% 0.017 4 0.004 2.576 1.187E01 No
Aggregate  Moist% 0.012 4 0.003 1.844 2.139E01 No
PM Binder
SS df MS F p-value Sig
Additive 0.281 2 0.141 10.288 6.142E03 Yes
Aggregate 4.504 2 2.252 164.824 3.151E07 Yes
Moist% 2.337 2 1.169 85.545 3.982E06 Yes
Additive  Aggregate 1.581 4 0.395 28.934 8.284E05 Yes
Additive  Moist% 0.121 4 0.030 2.217 1.570E01 No
Aggregate  Moist% 0.197 4 0.049 3.596 5.826E02 No

Table 9
ANOVA results for bond energy with different WMA additive group.

No WMA Additive
SS df MS F p-value sig
Binder 143.651 1 143.651 15002.742 2.66E08 Yes
Aggregate 2.722 2 1.361 142.124 1.93E04 Yes
Moist% 0.907 2 0.454 47.370 1.64E03 Yes
Binder  Aggregate 0.828 2 0.414 43.243 1.95E03 Yes
Binder  Moist% 0.310 2 0.155 16.172 1.21E02 Yes
Aggregate  Moist% 0.062 4 0.016 1.624 3.25E01 No
EV WMA Additive
SS df MS F p-value sig
Binder 135.082 1 135.082 8250.682 8.81E08 Yes
Aggregate 2.744 2 1.372 83.804 5.43E04 Yes
Moist% 0.392 2 0.196 11.984 2.05E02 Yes
Binder  Aggregate 3.688 2 1.844 112.616 3.04E04 Yes
Binder  Moist% 0.122 2 0.061 3.723 1.22E01 No
Aggregate  Moist% 0.020 4 0.005 0.308 8.59E01 No
SS WMA Additive
SS df MS F p-value sig
Binder 162.060 1 162.060 9928.795 6.082E08 Yes
Aggregate 1.692 2 0.846 51.836 1.380E03 Yes
Moist% 0.520 2 0.260 15.934 1.244E02 Yes
Binder  Aggregate 0.146 2 0.073 4.465 9.571E02 No
Binder  Moist% 0.337 2 0.169 10.336 2.629E02 Yes
Aggregate  Moist% 0.080 4 0.020 1.224 4.248E01 No

quickly decrease when PM binder was used with SS additive. The tension of binder. The wax-based additive, in the other hand,
results of ANOVA suggested that the asphalt binder type played a reduce the friction within the asphalt molecular to improve the
significant role to affect work of adhesion between binder and flow ability of binder. By contrast, for samples with the PM binder,
aggregate for different WMA additive group, different aggregate relative to the control group, work of adhesion was decreased by
type, and different aggregate moisture level as shown in Tables 3% and 16% with the addition of EvothermÒ and SasobitÒ, respec-
13–15. tively. The results of ANOVA suggested that the WMA additive type
played a significant role to affect work of adhesion between binder
3.2.4. Effect of WMA additives and aggregate for different asphalt binder, different aggregate type
Fig. 12 showed the work of adhesion trends for the Pen 60-70 except for SD, and different aggregate moisture level as shown in
and PM binders regarding addition of WMA additives. The Pen Tables 12, 14 and 15.
60-70 binder shows a positive trend with addition of WMA addi-
tives; on average, relative to binder samples with no additive, 3.3. Ranking of interfacial adhesion properties using four quadrant
the work of adhesion was increased by 3% and 5% with the addition diagrams
of EvothermÒ and SasobitÒ, respectively. The increasing in work of
adhesion implies better wetting ability of asphalt binder. The EV is Each sample combination subjected to the sessile drop and pull-
a surfactant-based chemical additive while the SS is a wax-based off tests was paired individually regarding aggregate moisture con-
additive. The surfactant-based additive form oriented monolayers tent. SFE measurement uses thermodynamic theory whereas bond
at the surface of asphalt binder and act by decreasing the surface energy measurement considers physical force. Moreover, the
1302 S.-H. Yang et al. / Construction and Building Materials 190 (2018) 1295–1307

Table 10
ANOVA results for bond energy with various aggregate types.

Aggregate-AG
SS df MS F p-value sig
Binder 115.571 1 115.571 11053.518 4.908E08 Yes
Additive 0.390 2 0.195 18.667 9.365E03 Yes
Moist% 0.488 2 0.244 23.328 6.235E03 Yes
Binder  Additive 1.443 2 0.722 69.029 7.928E04 Yes
Binder  Moist% 0.099 2 0.050 4.757 8.761E02 No
Additive  Moist% 0.029 4 0.007 0.699 6.316E01 No
Aggregate-SD
SS df MS F p-value sig
Binder 151.670 1 151.670 27465.418 7.952E09 Yes
Additive 0.016 2 0.008 1.422 3.417E01 No
Moist% 0.351 2 0.175 31.766 3.508E03 Yes
Binder  Additive 0.293 2 0.146 26.512 4.920E03 Yes
Binder  Moist% 0.185 2 0.093 16.752 1.138E02 Yes
Additive  Moist% 0.024 4 0.006 1.099 4.648E01 No
Aggregate-SL
SS df MS F p-value sig
Binder 176.156 1 176.156 15384.843 2.534E08 Yes
Additive 0.209 2 0.105 9.136 3.226E02 Yes
Moist% 0.992 2 0.496 43.325 1.947E03 Yes
Binder  Additive 0.321 2 0.161 14.022 1.558E02 Yes
Binder  Moist% 0.544 2 0.272 23.751 6.032E03 Yes
Additive  Moist% 0.098 4 0.024 2.133 2.406E01 No

10 Fig. 14 shows a similar trend: for each 1% increase in mois-


0% Moisture ture, the pull-off test results decreased by 1.1% on average.
8 5% Moisture
Bond Energy (kN/m)

Fig. 15 shows that for the Pen 60-70+SasobitÒ group, the pull-
10% Moisture
6 off test results decreased by only 1%. This is expected because
sessile drop tests use SFE calculation theory based on individual
4
SFE values.
2 Regarding aggregate type, Figs. 13–15 show that SD had the
lowest performance on pull-off tests, and SL had the lowest aver-
0
age performance on sessile drop tests.
Regarding the PM binder groups, Fig. 16 indicates that the pull-
off test results declined by 1.1% for each 1% increase in moisture
Binder Type/ WMA Additive Group/ Aggregate Type content. The group with the PM binder and EvothermÒ, shown in
Fig. 17, had similar pull-off test values for AG and SD; the decreases
Fig. 5. Pull-Off Test Results by Aggregate Moisture Content. of bond energy for AG, SD, and SL were 0.6%, 0.9, and 0.8%, respec-
tively. In the SasobitÒ group, shown in Fig. 18, SL was stronger than
AG and SD in the pull-off test. Review of Figs. 16–18 shows that for
results are far different in magnitude; therefore, the results were
the pull-off test, the average decreases of AG, SD, and SL were
normalized to dimensionless values by dividing each sample result
0.82%, 0.89, and 1.27% respectively. This suggests that, in this
by the control set of the sample. In this study, the two control sets
study, samples incorporating SL in the PM groups were more sen-
were employed. The control set for all Pen 60-70 samples was the
sitive to moisture than samples incorporating AG and SD were. For
SD sample with no binder additives and 0% moisture content. The
the sessile drop test, SD showed more sensitivity to moisture than
control set for all PM samples was the SD sample with no additives
AG and SL showed; the decreases were 4.24%, 5.95%, and 3.98% for
and 0% moisture content. An example of normalization for pull-off
AG, SD, and SL respectively. Thus, it can be concluded that in the
testing is shown in Eq. (8).
PM groups, sessile drop and pull-off tests showed different values
1:75 kN and different trends.
NORM:P67EV  AG5% ¼ m
¼ 3:3 ð8Þ
0:53 kN
m

4. Summary and conclusions


3.3.1. Aggregate moisture content
Because moisture levels are correlated with the type of aggre- The study aimed to demonstrate experimental techniques to
gate, moisture is here discussed in relation to binder and additive directly observe the effects of various aggregate moisture levels
groups. For both the pull-off test and sessile drop test, Fig. 13 on the binder-aggregate interfacial adhesion characteristics of
shows some decreasing trends caused by increasing moisture con- WMA samples by evaluating their work of adhesion and bond
tent in the Pen 60-70 group. However, the decrease values are not strength. Interfacial adhesion properties were investigated using
the same. For each 1% moisture increase, the pull-off test and ses- pull-off tests to measure the physical bond characteristics and ses-
sile drop test results decreased on average by 2.2% and 4.7%, sile drop test to measure the work of adhesion at the interface of
respectively. WMA and aggregate with present of water.
S.-H. Yang et al. / Construction and Building Materials 190 (2018) 1295–1307 1303

Table 11
ANOVA results for bond energy with different aggregate moisture level.

Aggregate Moist-0%
SS df MS F p-value sig
Binder 160.683 1 160.683 1451.114 2.836E06 Yes
Additive 0.179 2 0.090 0.809 5.069E01 No
Aggregate 2.640 2 1.320 11.920 2.064E02 Yes
Binder  Additive 0.307 2 0.154 1.388 3.484E01 No
Binder  Aggregate 1.401 2 0.700 6.325 5.771E02 No
Additive  Aggregate 0.110 4 0.027 0.248 8.972E01 No
Aggregate Moist-5%
SS df MS F p-value sig
Binder 148.379 1 148.379 1178.651 4.295E06 Yes
Additive 0.001 2 0.001 0.005 9.951E01 No
Aggregate 2.337 2 1.168 9.280 3.144E02 Yes
Binder  Additive 0.262 2 0.131 1.042 4.323E01 No
Binder  Aggregate 1.303 2 0.651 5.175 7.770E02 No
Additive  Aggregate 0.249 4 0.062 0.494 7.446E01 No
Aggregate Moist-10%
SS df MS F p-value sig
Binder 131.815 1 131.815 986.042 6.130E06 Yes
Additive 0.020 2 0.010 0.074 9.300E01 No
Aggregate 1.778 2 0.889 6.649 5.347E02 No
Binder  Additive 0.116 2 0.058 0.434 6.751E01 No
Binder  Aggregate 0.645 2 0.323 2.414 2.053E01 No
Additive  Aggregate 0.208 4 0.052 0.389 8.088E01 No

10 10
Pen 60/70 Polymer Modified Control (No Additive)
Bond Energy (kN/m)

8 8 Evotherm
Bond Energy (kN/m)

Sassobit
6 6
4 4
2 2
0 0
EV-AG10%
EV-SD0%
EV-SD5%

SS-SD0%
SS-SD5%

SS-SL0%
SS-SL5%
Cont-AG0%
Cont-AG5%

EV-AG0%
EV-AG5%
Cont-AG10%
Cont-SD0%
Cont-SD5%

Cont-SL0%
Cont-SL5%
Cont-SD10%

Cont-SL10%

EV-SD10%
EV-SL0%
EV-SL5%
EV-SL10%
SS-AG0%
SS-AG5%
SS-AG10%

SS-SD10%

SS-SL10%

WMA Group/ WMA Aggregate Type/ Agg. Moisture Level Binder Type/ Aggregate Type/ Agg. Moisture Level

Fig. 6. Pull-Off Test Results by Asphalt Binder Type. Fig. 8. Pull-Off Test Results by WMA Additive.

1000 1000
Pen 60-70 Polymer Modified 100
Work ofAadhesion (mN/m)

90 AG
800 800 80 SD
70 SL
Force (N)

60
Force (N)

600 600 50
40
400 400 30
20
200 10
200
0

0 0

Displacement (mm) Displacement (mm)


Binder Type/ WMA Additive Group/ Agg. Moisture Content
(a) (b)
Fig. 9. Sessile Drop Test Results by Aggregate Type.
Fig. 7. Force-Displacement Curves (a) Pen 60-70 and (b) PM.

d The results from pull-off test showed that the bond energy for bond energy was observed in the P67EV-0% as 1.92 kN/m and
the PM binder group showed superior performance (approxi- the lowest bond energy showed in the P67-10%. Also, the bond
mately 6 times larger) than that for the Pen60-70 binder group. energy diminishes with increasing aggregate moisture contents,
In the Pen 60-70 binder group, the aggregate AG in general the percent drops in bond energy is less when WMA additives
showed the higher bond energy and aggregate SD showed the was included. Last, samples without WMA additives and with
lowest bond energy among three aggregate types. The highest EvothermÒ showed an average bond energy of 1.12 kN/m,
1304 S.-H. Yang et al. / Construction and Building Materials 190 (2018) 1295–1307

Table 12
ANOVA results for Work of Adhesion when use of Pen 60-70 and Polymer Modified binder.

P67 Binder
SS df MS F p-value sig
Additive 95.296 2 47.648 586.312 2.108E09 Yes
Aggregate 863.512 2 431.756 5312.790 3.204E13 Yes
Moist% 3965.877 2 1982.939 24400.199 7.217E16 Yes
Additive  Aggregate 1.363 4 0.341 4.192 4.036E02 Yes
Additive  Moist% 6.261 4 1.565 19.261 3.620E04 Yes
Aggregate  Moist% 411.356 4 102.839 1265.441 3.096E11 Yes
PM Binder
SS df MS F p-value sig
Additive 773.638 2 386.819 587.840 2.087E09 Yes
Aggregate 633.741 2 316.871 481.541 4.606E09 Yes
Moist% 2909.133 2 1454.566 2210.471 1.065E11 Yes
Additive  Aggregate 11.076 4 2.769 4.208 3.998E02 Yes
Additive  Moist% 50.845 4 12.711 19.317 3.583E04 Yes
Aggregate  Moist% 301.644 4 75.411 114.600 4.269E07 Yes

Table 13
ANOVA results for Work of Adhesion with different WMA additive group.

No WMA Additive
SS df MS F p-value sig
Binder 2.634 1 2.634 603.175 1.631E05 Yes
Aggregate 527.100 2 263.550 60363.044 1.098E09 Yes
Moist% 2420.178 2 1210.089 277156.643 5.207E11 Yes
Binder  Aggregate 0.039 2 0.020 4.515 9.423E02 No
Binder  Moist% 0.176 2 0.088 20.144 8.157E03 Yes
Aggregate  Moist% 250.921 4 62.730 14367.600 1.453E08 Yes
EV WMA Additive
SS df MS F p-value sig
Binder 104.353 1 104.353 588.032 1.716E05 Yes
Aggregate 525.335 2 262.667 1480.137 1.821E06 Yes
Moist% 2412.162 2 1206.081 6796.293 8.655E08 Yes
Binder  Aggregate 1.484 2 0.742 4.181 1.047E01 No
Binder  Moist% 6.844 2 3.422 19.283 8.830E03 Yes
Aggregate  Moist% 250.199 4 62.550 352.469 2.397E05 Yes
SS WMA Additive
SS df MS F p-value sig
Binder 1198.998 1 1198.998 586.976 1.722E05 Yes
Aggregate 438.609 2 219.304 107.362 3.344E04 Yes
Moist% 2014.017 2 1007.009 492.987 1.633E05 Yes
Binder  Aggregate 17.125 2 8.563 4.192 1.043E01 No
Binder  Moist% 78.740 2 39.370 19.274 8.838E03 Yes
Aggregate  Moist% 208.897 4 52.224 25.567 4.144E03 Yes

which was higher than that of samples with SasobitÒ. In the PM tively). Aggregate moisture content has a significant effect on
binder group, SL aggregate group, in general, showed the high- work of adhesion, aggregates with 10% moisture content can
est bond energy and SD aggregate group showed the lowest reduce the work of adhesion to 50% compare to dry condition.
bond energy. The highest bond energy was observed in PMEV- Regarding the asphalt binder type, no statistically significant
0% as 7.84 kN/m and the lowest bond energy was observed in differences were found between the Pen 60-70 and PM binders.
PMEV-10% as 5.78 kN/m. The bond energy decreases with For binders without WMA additives, the work of adhesion for
increases aggregate moisture content. Last, samples with the PM binder was slightly higher than the work of adhesion
EvothermÒ with AG and SD aggregates showed lowest bond for the Pen 60-70 binder. However, when WMA additives were
energy but EvothermÒ with SL aggregates showed the highest used, the work of adhesion for the Pen 60-70 binder with WMA
bond energy among all moisture contents. additives was higher than the work of adhesion for the PM bin-
d For the work of adhesion, for absolutely dry conditions, SD had der with WMA additives. Regarding the type of WMA additive,
an average value of 59.07 mN/m, which was superior to the val- the work of adhesion for the Pen 60-70 binder mixed with
ues of AG and SL. For conditions of 5% and 10% moisture con- EvothermÒ and SasobitÒ increased by 3% and 6%, respectively,
tent, AG had work of adhesion values of 50.97 mN/m and compared with the control sample. By contrast, the work of
33.97 mN/m at 5% and 10% moisture content, respectively. For adhesion for the PM binder mixed with EvothermÒ and SasobitÒ
overall testing, AG had the highest work of adhesion value fol- decreased by 3% and 6%, respectively, compared with the con-
lowed by SD and SL (48.00, 47.18, and 36.48 mN/m, respec- trol case.
S.-H. Yang et al. / Construction and Building Materials 190 (2018) 1295–1307 1305

Table 14
ANOVA results for Work of Adhesion with various Aggregate types.

Aggregate-AG
SS df MS F p-value sig
Binder 248.155 1 248.155 76.566 9.401E04 Yes
Additive 74.051 2 37.026 11.424 2.220E02 Yes
Moist% 1969.048 2 984.524 303.766 4.278E05 Yes
Binder  Additive 272.598 2 136.299 42.054 2.061E03 Yes
Binder  Moist% 11.758 2 5.879 1.814 2.750E01 No
Additive  Moist% 3.525 4 0.881 0.272 8.824E01 No
Within 12.964 4 3.241
Total 2592.099 17 152.476
Aggregate-SD
SS df MS F p-value sig
Binder 239.827 1 239.827 31.806 4.866E03 Yes
Additive 71.456 2 35.728 4.738 8.810E02 No
Moist% 4593.678 2 2296.839 304.606 4.255E05 Yes
Binder  Additive 263.175 2 131.588 17.451 1.057E02 Yes
Binder  Moist% 27.477 2 13.739 1.822 2.738E01 No
Additive  Moist% 8.186 4 2.047 0.271 8.827E01 No
Aggregate-SL
SS df MS F p-value sig
Binder 344.610 1 344.610 168.623 2.029E04 Yes
Additive 102.715 2 51.358 25.130 5.434E03 Yes
Moist% 1245.284 2 622.642 304.668 4.253E05 Yes
Binder  Additive 378.303 2 189.152 92.555 4.474E04 Yes
Binder  Moist% 7.423 2 3.712 1.816 2.747E01 No
Additive  Moist% 2.218 4 0.554 0.271 8.828E01 No

Table 15
ANOVA results for Work of Adhesion with different aggregate moisture level.

Moist-0%
SS df MS F p-value sig
Binder 344.610 1 344.610 168.623 2.029E04 Yes
Additive 102.715 2 51.358 25.130 5.434E03 Yes
Aggregate 1245.284 2 622.642 304.668 4.253E05 Yes
Binder X Additive 378.303 2 189.152 92.555 4.474E04 Yes
Binder X Aggregate 7.423 2 3.712 1.816 2.747E01 No
Additive x Aggregate 2.218 4 0.554 0.271 8.828E01 No
Moist-5%
SS df MS F p-value sig
Binder 226.788 1 226.788 184.659 1.698E04 Yes
Additive 67.652 2 33.826 27.542 4.583E03 Yes
Aggregate 746.275 2 373.138 303.821 4.277E05 Yes
Binder  Additive 249.003 2 124.502 101.373 3.743E04 Yes
Binder  Aggregate 4.465 2 2.233 1.818 2.744E01 No
Additive  Aggregate 1.337 4 0.334 0.272 8.823E01 No
Moist-10%
SS df MS F p-value sig
Binder 92.081 1 92.081 273.697 7.818E05 Yes
Additive 27.409 2 13.704 40.734 2.190E03 Yes
Aggregate 205.581 2 102.791 305.529 4.230E05 Yes
Binder  Additive 100.958 2 50.479 150.040 1.730E04 Yes
Binder  Aggregate 1.225 2 0.613 1.821 2.740E01 No
Additive  Aggregate 0.365 4 0.091 0.272 8.827E01 No

d The bond energy and the work of adhesion represent two dis- 60-70 binder, most scenarios fallen in quadrant II except for
tinct properties as the indication to the adhesion properties of SD aggregate groups. For the PM binder, the results mainly scat-
asphalt-aggregate system. Using the four quadrant diagrams, ter up and down alone the x-axis in the quadrant II and III.
the results from the bond energy and the work of adhesion
can be normalized and ranked to show the effect of various Although further improvements are still needed, the direct
factors in the study. It was clearly showed that the for the Pen pull-off test combine with work of adhesion are found to be the
1306 S.-H. Yang et al. / Construction and Building Materials 190 (2018) 1295–1307

100 400%
0% Moisture AG0% SD0%
Work of Adhesion (mN/m)

5% Moisture

Normalized Pull-Off Test


80 SL0% AG5%
10% Moisture SD5% SL5%
60 AG10% SD10%
SL10%
40 200%

20

0%

Binder Type/ WMA Additive Group/ Aggregate Type Normalized Sessile Drop Test

Fig. 10. Sessile Drop Test Results by Aggregate Moisture. Fig. 14. Quadrant Diagram for Pen 60-70 + EvothermÒ.

100 400%
Work of Adhesion (mN/m)

Pen 60/70 Polymer Modified Binder AG0% SD0%

Normalized Pull-Off Test


80 SL0% AG5%
SD5% SL5%
60 AG10% SD10%
40 SL10%
200%
20
0
EV-AG10%
EV-SD0%
EV-SD5%
EV-AG0%
EV-AG5%

SS-SD0%
SS-SD5%

SS-SL0%
SS-SL5%
Cont-AG0%
Cont-AG5%

Cont-SD0%
Cont-SD5%

Cont-SL0%
Cont-SL5%
Cont-SL10%

EV-SD10%
EV-SL0%
EV-SL5%
EV-SL10%
SS-AG0%
SS-AG5%
SS-AG10%

SS-SD10%
Cont-AG10%

SS-SL10%
Cont-SD10%

0%

WMA Group/ WMA Aggregate Type/ Agg. Moisture Level Normalized Sessile Drop Test

Fig. 11. Sessile Drop Test Results by Asphalt Binder Type. Fig. 15. Quadrant Diagram for Pen 60-70 + SasobitÒ.

100 Control (No Additive) 200%


90 AG0% SD0%
Evotherm (EV)
Normalized Pull-Off Test

80 SL0% AG5%
Work of Adhesion (mN/m)

70 Sassobit (SS)
150% SD5% SL5%
60
AG10% SD10%
50
40 SL10%
30 100%
20
10
0 50%

0%

Binder Type/ Aggregate Type/ Agg. Moisture Level Normalized Sessile Drop Test

Fig. 12. Sessile Drop Test Results by WMA Additive Type. Fig. 16. Quadrant Diagram for PM Binder.

400% 200% AG0% SD0%


AG0% SD0% SL0% AG5%
SL0% AG5%
Normalized Pull-Off Test
Normalized Pull-Off Test

SD5% SL5%
SD5% SL5% 150% AG10% SD10%
AG10% SD10% SL10%
SL10%
200% 100%

50%

0% 0%

Normalized Sessile Drop Test Normalized Sessile Drop Test

Fig. 13. Quadrant Diagram for Pen 60-70. Fig. 17. Quadrant Diagram for PM Binder + EvothermÒ.
S.-H. Yang et al. / Construction and Building Materials 190 (2018) 1295–1307 1307

200% [9] Cindy K. Estakhri, Laboratory and field performance measurements to support
AG0% SD0% the implementation of warm mix asphalt in Texas. No. FHWA/TX-12/5-5597-
Normalized Pull-Off Test

01-1, Texas Transportation Institute, College Station, TX, 2012.


SL0% AG5%
150% [10] Shih-Hsien Yang, Anthony Keita, Preliminary evaluation of field aging
SD5% SL5% characteristic of warm mix asphalt, Advanced Materials Research, vol. 723,
Trans Tech Publications, 2013, pp. 149–156, https://doi.org/10.4028/
100% www.scientific.net/AMR.723.149.
[11] Wenyi Gong, Mingjiang Tao, Rajib Mallick, Tahar El-Korchi, Investigation of
moisture susceptibility of warm-mix asphalt mixes through laboratory
50% mechanical testing, Transp. Res. Rec.: J. Transp. Res. Board 2295 (2012) 27–
34, https://doi.org/10.3141/2295-04.
[12] Thomas Bennert, Ali Maher, Robert Sauber, Influence of production
0% temperature and aggregate moisture content on the initial performance of
0% 50% 100% 150% 200% warm-mix asphalt, Transp. Res. Rec.: J. Transp. Res. Board 2208 (2011) 97–107,
https://doi.org/10.3141/2208-13.
Normalized Sessile Drop Test [13] A. Khodaii, H. Kazemi Tehrani, H.F. Haghshenas, Hydrated lime effect on
moisture susceptibility of warm mix asphalt, Constr. Build. Mater. 36 (2012)
Fig. 18. Quadrant Diagram for PM Binder + SasobitÒ. 165–170.
[14] Feipeng Xiao, Jayson Jordan, Serji N. Amirkhanian, Laboratory investigation of
moisture damage in warm-mix asphalt containing moist aggregate, Transp.
Res. Rec. 2126 (1) (2009) 115–124.
promising methods to investigate the bonding performance [15] Mohammad Alavi, Elie Hajj, Andrew Hanz, Hussain Bahia, Evaluating adhesion
between aggregate and asphalt. properties and moisture damage susceptibility of warm-mix asphalts:
Bitumen bond strength and dynamic modulus ratio tests, Transp. Res. Rec.:
J. Transp. Res. Board 2295 (2012) 44–53, https://doi.org/10.3141/2295-06.
Acknowledgements [16] Walaa S. Mogawer, Alexander J. Austerman, Hussain U. Bahia, Evaluating the
effect of warm-mix asphalt technologies on moisture characteristics of asphalt
binders and mixtures, Transp. Res. Rec. 2209 (1) (2011) 52–60.
The study is supported by Ministry of Science and Technology
[17] Cindy K. Estakhri, Joe W. Button, Allex E. Alvarez, Field and laboratory
(MOST) of Taiwan to National Cheng Kung University under investigation of warm mix asphalt in Texas. No. FHWA/TX-10/0-5597-2, Texas
MOST-103-2221-E-006-173. Transportation Institute, Texas A & M University System, 2010.
[18] Nazimuddin M. Wasiuddin, Musharraf M. Zaman, Edgar A. O’Rear, Effect of
sasobit and aspha-min on wettability and adhesion between asphalt binders
Conflict of interest and aggregates, Transp. Res. Rec. 2051 (1) (2008) 80–89, https://doi.org/
10.3141/2051-10.
[19] J. Yu, C. Xiong, X. Zhang, Z. Ge, G. An, Assessing moisture sensitivity of
The authors declare that there is no conflict of interest regard- rubberized warm mix asphalt mixtures using the surface free energy method
ing the publication of this paper. and dynamic water pressure tester, J. Testing Evaluat. 46 (2) (2018) 580–592,
https://doi.org/10.1520/JTE20170241. ISSN 0090-3973.
[20] E. Fini, H. Elham, Adhesion Mechanisms of Bituminous Crack Sealant to
References
Aggregate and Laboratory Test Development Doctoral Dissertation, University
of Illinois at Urbana-Champaign, IL, USA, 2008.
[1] A.E. Martin, E. Arambula, F. Yin, L.G. Cucalon, A. Chowdhury, R. Lytton, J. Epps, [21] U. Soltesz, E. Baudendistel, R. Schäfer, Stress analyses of pull-off tests for
C. Estakhri, E.S. Park, NCHRP Report 763: Evaluation of the Moisture strength measurements of coatings, in: Bioceramics and the Human Body,
Susceptibility of WMA Technologies, Transportation Research Board of the Springer, Dordrecht, 1992, pp. 504–509.
National Academies, Washington, D.C., 2014. [22] C.J. Van Oss, Interfacial Forces in Aqueous Media, CRC Press, New York, 2006.
[2] Ramon Francis Bonaquist, Mix Design Practices for Warm Mix Asphalt, vol. [23] Frederick M. Fowkes, Determination of interfacial tensions, contact angles, and
691, Transportation Research Board, 2011. dispersion forces in surfaces by assuming additivity of intermolecular
[3] Z. Hossain, A. Bhudhala, M. Zaman, E. O’Rear, Evaluation of the use of warm interactions in surfaces, The J. Phys. Chem. 66 (2) (1962). 382-382.
mix asphalt as a viable paving material in the United States, Federal Highway [24] A.R. Tarrer, Vinay Wagh, The effect of the physical and chemical characteristics
Administration. DTFH61-06-H-00044: Task 3, 2009. <http://www.oktc.org/ of the aggregate on bonding, No. SHRP-A/UIR-91-507, Strategic Highway
otc/final_report_wma-01dec09.pdf>. Research Program, National Research Council, Washington, DC, USA, 1991.
[4] Baoshan Huang, Xiang Shu, Qiao Dong, Junan Shen, Laboratory evaluation of [25] Van Oss, Carel Jan, Use of the combined Lifshitz–van der Waals and Lewis
moisture susceptibility of hot-mix asphalt containing cementitious fillers, J. acid–base approaches in determining the apolar and polar contributions to
Mater. Civ. Eng. 22 (7) (2010) 667–673, https://doi.org/10.1061/(ASCE) surface and interfacial tensions and free energies, J. Adhes. Sci. Technol. 16 (6)
MT.1943-5533.0000064. (2002) 669–677.
[5] Young Choi, Case study and test method review on moisture damage, No. AP- [26] Muhammad Rafiq Kakar, Meor Othman Hamzah, Jan Valentin, A review on
T76/07, 2007. moisture damages of hot and warm mix asphalt and related investigations, J.
[6] J. William, Pine, Performance issues caused by residual moisture, in: Moisture Cleaner Prod. 99 (2015) 39–58.
Damage to Hot-Mix Asphalt Mixtures. Synopsis a Work, January 22, 2012, [27] Agnieszka Woszuk, Wojciech Franus, A review of the application of zeolite
Washington, D.C., Transportation Research Board, Washington, D.C., 2015, pp. materials in Warm Mix Asphalt technologies, Appl. Sci. 7 (3) (2017) 293.
19–39, https://doi.org/10.17226/22126. [28] A. Woszuk, A. Zofka, L. Bandura, W. Franus, Effect of zeolite properties on
[7] Yong-Rak Kim, Jun Zhang, Hoki Ban, Moisture damage characterization of asphalt foaming, Constr. Build. Mater. 139 (2017) 247–255.
warm-mix asphalt mixtures based on laboratory-field evaluation, Constr. [29] A. Woszuk, R. Panek, J. Madej, A. Zofka, W. Franus, Mesoporous silica material
Build. Mater. 31 (2012) 204–211, https://doi.org/10.1016/ MCM-41: novel additive for warm mix asphalts, Constr. Build. Mater. 183
j.conbuildmat.2011.12.085. (2018) 270–274.
[8] Graham C. Hurley, Brian D. Prowell, Andrea N. Kvasnak, Michigan Field Trial of
Warm Mix Asphalt Technologies: Construction Summary, National Center for
Asphalt Technology, Auburn, AL, 2009.

You might also like