You are on page 1of 10

Construction and Building Materials 264 (2020) 120685

Contents lists available at ScienceDirect

Construction and Building Materials


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

Adhesion between steel slag aggregates and bituminous binder based on


surface characteristics and mixture moisture resistance
Bárbara Luiza Riz de Moura a, Jamilla Emi Sudo Lutif Teixeira a,⇑, Renata Antoun Simão b,
Mahdieh Khedmati c, Yong-Rak Kim d, Patrício José Moreira Pires a
a
Civil Engineering Pos-Graduate Program, Universidade Federal do Espírito Santo, Vitória, ES, Brazil
b
Department of Metallurgical and Materials Engineering, Universidade Federal do Rio de Janeiro, Rio de Janeiro, Brazil
c
Whittier Research Center, University of Nebraska-Lincoln, Lincoln, NE, USA
d
Zachry Department of Civil & Environmental Engineering, Texas A&M University, College Station, TX, USA

h i g h l i g h t s

 Binder-slag adhesion assessment based on small scale characterization.


 Air-cooled blast furnace (ACBFS) slag and Linz-Donawitz (LD) slag were examined.
 Chemical and surface energy characteristics affects binder-slag adhesion.
 The AFM appeared as a promising method to evaluate binder-aggregate adhesion.
 Small-scale surface characteristics agreed well with the large-scale observations.

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

Article history: Steel slag aggregates (SSA) have shown promising mechanical and physical properties for asphalt con-
Received 13 May 2020 crete application. However, SSA surface characteristics vary depending on the type of slag, which results
Received in revised form 12 August 2020 in different binder-SSA adhesion that affects moisture damage of asphalt mixtures containing SSA. This
Accepted 19 August 2020
study evaluated the adhesion in the binder-SSA system considering two typical types of slags (i.e. air-
Available online 22 September 2020
cooled blast furnace slag (ACBFS) and Linz-Donawitz steel slag (LD)) using various tests: morphological,
physical, chemical, and surface energy measurements. Also, the moisture damage susceptibility of
Keywords:
asphalt mixtures containing SSA were assessed based on two standard methods: ASTM D3625 and
Steel slag
Blast furnace slag
AASHTO T283. Test results showed strong influence of the SSA chemical composition and surface energy
Adhesion characteristics on the SSA-binder adhesion. LD showed a better adhesion with the asphalt binder than
Surface energy ACBFS due to its particular chemical and surface energy characteristics. The diminished adhesive bonding
Moisture damage from ACBFS could be improved with proper surface treatment using additives such as hydrated lime and
electrostatic precipitator powder, which change surface chemistry of ACBFS.
Ó 2020 Elsevier Ltd. All rights reserved.

1. Introduction use of air-cooled blast furnace slag (ACBFS, also referred as iron
slag) and Linz-Donawitz steel slag (LD). The former, obtained in
Industrial by-products from the steel industry have already the initial steps of the process to produce molten iron, is a result
been studied and applied as infrastructure materials in the various of reduced calcium carbonate which is associated with silicon
layers of pavement structures. Among the different types of slags oxide, as well as other remaining process compounds such as alu-
produced in the steelmaking process, it is highlighted herein the minum and magnesium oxides. LD slag, on the contrary, is
obtained in the latter steps of the steel fabrication, and it is mainly
composed of calcium and iron oxides, followed by silicon, magne-
⇑ Corresponding authors at: Centro Tecnológico, Dept. de Engenharia Civil
sium and manganese oxides [1,2]. The ACBFS has attractive proper-
(CT-DEC). Universidade Federal do Espírito Santo (UFES). Av. Fernando Ferrari,
514 - CT I, Goiabeiras, CEP: 29060-970, Vitória, ES, Brazil. ties for pavement application such as solidity, abrasion resistance
E-mail addresses: barbaralrmoura@gmail.com (B.L.R. de Moura), jamilla.teixeira@ and high angle of friction, due to the angular shape, which justify
ufes.br (J.E.S.L Teixeira), renata@metalmat.ufrj.br (R.A. Simão), mahdieh.khedmati@ its application [3]. Similarly, LD has favorable characteristics such
huskers.unl.edu (M. Khedmati), yong-rak.kim@tamu.edu (Y.-R. Kim), patricio. as high stability, promising frictional properties and stripping
pires@ufes.br (P.J.M. Pires).

https://doi.org/10.1016/j.conbuildmat.2020.120685
0950-0618/Ó 2020 Elsevier Ltd. All rights reserved.
Bárbara Luiza Riz de Moura et al. Construction and Building Materials 264 (2020) 120685

resistance [4,5]. However, those materials, when combined with chemical interactions between binder and SSA components.
asphalt binder to produce asphalt mixtures, present significant Toward that end, two different types of SSA (i.e. ACBFS and LD)
variances on the binder-aggregate surface adhesion, which subse- were used and characterized using the following methods:
quently affects the mixture’s moisture resistance. Chemically, the
types of steel slag aggregates (SSA) present changes in the percent-  Laser scanning microscopy (LSM) for analysis of the surface
ages of SiO2 and Fe2O3 [6]. These two oxides play an important role roughness and aggregate image measurement system (AIMS)
in the asphalt binder and SSA adhesion. Thus, it is important to for shape properties (e.g. texture, angularity);
identify the main differences among SSA that can notably affect  Contact angle measurements and atomic force microscopy
the overall asphalt mixture performance, in particular the moisture (AFM) test to investigate the free energy of the surface and
damage resistance. the adhesive work between binder and SSA, respectively;
Related to the asphalt mixture performance, the adhesion can  X-ray fluorescence (XRF) test to identify the main oxides pre-
be defined as the ability of the aggregate to be enveloped by the sented in each SSA; and
asphalt binder, and it is usually dependent on the synergistic effect  ASTM D3625 [19] which is empirical but can easily assess adhe-
of chemical interactions, physical adhesion and mechanical inter- sive bonding of binder to aggregates.
locking. A good binder-aggregate adhesion is essential to avoid
damage due to fatigue and moisture [7]. Cracks in this interface In addition, the moisture resistance of HMA mixtures with par-
are indicative of low adhesiveness between the compounds of tial replacement of natural aggregates by SSA was evaluated using
the bituminous mixtures. The interaction between binder and AASHTO T283-14 [20]. The HMA mixtures were designed using
aggregate is much more influenced by the surface characteristics 75% of natural aggregates (NA) and 25% of SSA (either ACBFS or
of the aggregates than by those of the binder. Aggregate with same LD). Also, additional HMA containing ACBFS were evaluated by
physical and mechanical characteristics may present variation in adding different additives, such as hydrated lime (HL), natural
adhesiveness due to differences in the chemical composition which aggregate passing #200 sieve (NA’), and two steel dust by-
is relevant to the interaction forces formed between these materi- products (FGD and EP).
als [8–11].
The cohesive (within the asphalt binder itself) and adhesive
2. Materials and methods
interactions (at the binder-aggregate interface) are related to the
surface energy of the asphalt mixture components. The surface
Fig. 1 shows a flowchart to summarize the selected materials
energy concepts can be used to quantify the work of adhesion
and the procedures presented in the study. The HMA mixtures
between materials. The thermodynamic changes in the surface
were named according to the type of slag used followed by the
energy are related to the disassociation of the bonded materials
type of filler addition.
and, consequently, the cracking potential in the asphalt composite
[12]. Zhou et al. [13] comment that SSA has greater interface
energy comparing to natural aggregates, such as basalt and ande- 2.1. Materials
site, when they are in association with asphalt binder, and SSA is
more susceptible to cohesive failure while the natural aggregates Two different SSA (ACBFS and LD) were evaluated in this study.
are more prone to adhesive failures. Aguiar-Moya et al. [14] com- Samples from those materials were obtained from a local steel
ment that the work of adhesion generated in binder-aggregate company. For the morphological testing methods and HMA pro-
interface directly affects the resistance of asphalt mixture to mois- duction, SSA particles were obtained after the company internal
ture damage because this parameter quantifies the ease with crushing process, as shown in Fig. 2 (a and b). The SSA’s well-
which water can displace asphalt binder from the aggregate’s distributed aggregate gradations with NMAS of 19 mm are pre-
surface. sented in Fig. 3. For the thermodynamic analysis, cylindrical sam-
Regarding the chemical composition, aggregates that are nega- ples with about 23 mm in diameter as shown in Fig. 2 (c and d)
tively charged in the presence of water, such as silica-based aggre- were extracted from large prismatic blocks of slag (approximately
gates, have low adhesiveness to the asphalt binder [15]. In parallel, 100 mm  100 mm  200 mm). These samples were polished to
those that present compounds such as iron, calcium, magnesium, minimize the effects of aggregate roughness on the results.
and aluminum, tend to have high affinity to the bituminous binder In some experimental procedures employed in this study,
[16]. Studies by Cala et al. [17,18] proposed a moisture damage asphalt cement with a penetration grade of 50/70 was used. Table 1
index based on an evaluation of aggregates with different litho- presents its characterization, which is in accordance with ASTM
graphs. They suggested that SiO2 is one of the compounds that standard limits.
can contribute negatively to moisture damage, while other oxides
such as Al2O3, Fe2O3 and MgO are important compounds to mini- 2.2. Binder-Slag adhesion methods
mize such damage.
Existing studies demonstrate that the binder-aggregate interac- 2.2.1. Morphological properties
tion is directly related to the small-scale chemical and thermody- a) Shape, angularity and texture by Aggregate Image Measurement
namic internal characteristics. Thus, better understanding of System (AIMS)
those characteristics of SSA is essential for an adequate application. It is believed that mechanical interlocking is a mechanism of
As an attempt, this study aims to evaluate the adhesive character- physical adhesion that influences the ability of the aggregate to
istics of the binder-slag system through different micro-surface be covered by asphalt cement. According to several studies, this
analyses in order to better understand the adhesion mechanisms binder-aggregate adhesion mechanism is directly related to the
between SSA and asphalt binder. Resulting findings can help better aggregate roughness [21]. Besides this property, other morpholog-
selection and engineering of SSA in asphalt mixtures if used as sup- ical properties affect the asphalt mixture’s performance, such as
plemental recycled aggregates. the aggregate’s shape and angularity.
This study aims to identify the small-scale characteristics of two To assess SSA’s morphology, AIMS equipment was used. It is
different slags that might influence the binder-slag adhesion, con- capable of obtaining different morphological properties of particles
sidering the mechanical interlocking (physical adhesion), surface through digital imaging technology. The equipment captures
free energy and work of adhesion (thermodynamic adhesion) and images of reasonable number of aggregate particles in different
2
Bárbara Luiza Riz de Moura et al. Construction and Building Materials 264 (2020) 120685

Fig. 1. Flowchart of the experimental plan.

Fig. 2. ACBFS a) granular and c) cylindrical samples; LD b) granular and d) cylindrical samples.

100
90
80
70
Passing (%)

60
50
40
30
20
10
0
0.0001 0.001 0.01 0.1 1 10 100 1000

Particle size (mm)

ACBFS LD

Fig. 3. Particle size distribution curve of ACBFS and LD.

resolutions and angles of view providing an analysis of the shape, In addition to AIMS evaluation, the SSA surface roughness was
angularity and texture in a short amount of time [22–24] and the also assessed using LSM. This equipment is able to scan multiple
results were obtained according to AASHTO TP-81 [25]. Al Rousan locations on the surface of a sample using a laser scanning micro-
[26] proposed a classification system after gathering AIMS results scope. The microscope collects optical images and high-resolution
of 13 types of coarse aggregates and 5 types of fine aggregates with surface data by combining laser light with white light. The high-
known shape characteristics. For this study, ACBFS and LD samples resolution image of the sample is obtained by the reflection inten-
were evaluated using AIMS procedure and based on the results, sity of the laser beams; thus, the topography is plotted by analyz-
they were classified according to the classification established by ing the intensity of the return of the laser beam to the z position.
Al Rousan [26] and compared. A statistical representation of the particle roughness, the root
b) Roughness by Laser-Scanning Microscopy (LSM) mean square (RMS) roughness, was obtained by measuring the

3
Bárbara Luiza Riz de Moura et al. Construction and Building Materials 264 (2020) 120685

Table 1
Asphalt binder characterization results.

Characteristics ASTM Method Specification Result Unit


Penetration D5 50 to 70 51 0.1 mm
Softening Point D 36 46 min 50 °C
Brookfield Viscosity 135° C-SP21 20 RPM D 4402 274 min 323 cP
Brookfield Viscosity 150 °C-SP21 D 4402 112 min 163 cP
Brookfield Viscosity 177 °C-SP21 D 4402 57 to 285 min 61 cP
RTFOT - Residual Penetration D5 55 min 61 %
RTFOT – Softening Point Increase D 36 8 max 5.0 °C
RTFOT - Ductility at 25 °C D 113 20 min >150 cm
RTFOT - % Weight Range D 2872 0.50 to 0.50 0.305 %
Ductility at 25 °C D 133 60 min >150 cm
Solubility in trichloroethylene D 2042 99.5 min 99.9 % mass
Flash Point D 92 235 min 344 °C
Thermal Suscetibility Index X 018 1.5 to 1.7 1.1 N/A
Relative Density at 20/4 °C D 70 (1) 1.011 N/A
Heating at 177 °C X 215 (2) (2) N/A

magnitude of a set of collected points in the surface and squaring measured. First, three probe liquids with known surface energies
each number, then taking their average and the square root of and not capable of dissolving or chemically reacting with the
the average. The RMS roughness of the studied SSA was calculated SSA, such as distilled water, glycerol (polar liquids) and diiodo-
based on scanning measurements of five different areas (1430 mm methane (non-polar liquid), were used to quantify the SSA’s sur-
by 1000 mm) per sample. Three aggregate particles were scanned face free energy. Second, an asphalt binder drop was used to
for each type of SSA. The output data were also represented in a verify the difference in wettability of the asphalt binder on differ-
3D topographic image. ent SSA solid surfaces; similar analyses are found in [31–33]
(Fig. 4b). The binder was previously heated to 80 °C.
2.2.2. Thermodynamic properties b) Atomic Force Microscopy (AFM)
a) Contact Angle Assessment The AFM has been used in many studies to extend the study of
Wettability can be defined as the acceptance of a liquid to the micro-rheological properties of asphalt binders, and to evalu-
spread over a solid surface. The wetting phenomena can be ate the binder’s microstructural changes when subjected to load
observed when a drop of a liquid rests on a solid level surface [34–36]. Also, its scanning technique allows gathering information
and interacts with it. The more spread the drop, the higher the sur- about the substrate surface. AFM has recently become an interest-
face energy of the material’s interface. This concept has been ing alternative for the study of material’s adhesion [37]. AFM is
applied in many studies to verify the surface energy of the materi- able to measure the fundamental attraction and retraction forces
als [14,27–29]. A way to experimentally determine wetting is to between the material from the AFM tip and the selected substrate.
look at the contact angle (h), which is the angle connecting the Each time the tip approximates and withdraws from a substrate,
solid–liquid interface (Fig. 4a). It is established between the tan- instantaneous attraction and retraction forces are developed, as
gent to a drop of liquid and the solid where the liquid was depos- illustrated in Fig. 5a [38,39]. As a result, and the deflective curve
ited [30]. If h = 0°, the liquid completely wets the substrate. If 0° is generated. This outcome curve allows the measurement of the
< h < 90°, high wetting occurs. If 90° < h < 180°, low wetting occurs. adhesive force between different materials with different atomic
If h = 180°, the liquid does not wet the substrate at all. compositions. The work of the adhesion force can be obtained by
Equation (1) (from the acid-base theory) is recognized as a ther- calculating the area between the retraction curve and the x-axis
modynamic equilibrium between materials and it relates the mea- (Fig. 5b).
sured angle h to the surface energy components: nonpolar In this study, the AFM was employed to verify the adhesion
component (cLW), acidic component (c+) and basic component forces between binder and SSA aggregates. Thus, the slag aggre-
(c-). The S and L are related to the solid and liquid materials, gates were considered, the substrate and two different procedures
respectively. were established. In the first one, polished slag samples were sub-
jected to a load applied by a silicon tip of undetermined elastic
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffiffiffiffiffi
constant - ranging between 20 and 80 kN/m and a frequency
cL ð1 þ coshÞ ¼ 2 cLW
S cL þ 2
LW
cþS cL þ 2 cL cþS ð1Þ
between 319 and 367 Hz. In the second procedure, a modified
In this study, an automated goniometer (Ramé-Hart, 100–50) AFM test was performed by covering another silicon tip of the
was used coupled with a CCD camera to obtain several images of AFM with the same specifications by an asphalt binder film prior
the drop of different liquids on SSA’s surface. The obtained images to contact on the slag substrate. In this way, the work of adhesion
are treated and the contact angles in the two ends of the drop are between binder-slag could be quantified.

Fig. 4. a) 2020 Ramé-Hart instrument co. image, b) Illustration of a drop of asphalt binder in the SSA surface and the surface free energy components.

4
Bárbara Luiza Riz de Moura et al. Construction and Building Materials 264 (2020) 120685

Fig. 5. a) AFM resulting curves during the tip approximation and retraction from the substrate and b) Generic AFM curve with highlighted work of adhesion.

2.2.3. Chemical analysis by X-ray fluorescence (XRF) no. 200 sieve (NA’), included in the experimental plan to evaluate
To determine the main oxides components of the ACBFS and LD the filling effect only since the granitic aggregate is mainly com-
samples, XRF analyses were performed in the SSA samples using a posed of silicon oxide and it is not expected to improve the mois-
WDS spectrometer and a Rh anode tube. Representative samples of ture damage. Fig. 6 presents the studied HMA mixtures, including
dry SSA containing difference particle sizes were crushed to obtain the main differences on their composition.
XRF samples with particles passing in the n° 200 sieve (75 mm). The studied HMA were subjected to AASHTO T 283 procedure.
Six specimens of each mixture were compacted using Marshall
2.2.4. Visual observation according to ASTM D3625 impactor to obtain cylindrical specimens with air voids of 7 ± 1%.
The ASTM D3625 [19] is a rapid and commonly used procedure According to the AASHTO T 283 [20] procedure, three specimens
to visualize the loss of adhesiveness in non-compacted asphalt (conditioned set) were partially saturated, then submitted to a
mixtures under the action of boiling water. For this study, initially freezing condition at 18 ± 3 °C and subsequently to a water bath
two samples of 250 g of SSA (LD and ACBFS) were mixed with at 60 ± 1 °C before subjected to indirect tensile strength (ITS) test.
12.5 g of binder content (binder content of 5.0%). Since LD is denser The other three specimens (dry set) were submitted to ITS without
than ACBFS, a notable difference in the total volume of aggregates conditioning. The ITS test was conducted at a 0.8 mm/s displace-
was observed. Thus, a third LD sample (with the same volume of ment rate at 25 °C until the rupture and the tensile strength ratio
aggregate particles used for ACBFS test) was also mixed with (TSR) (Equation (2)), which represents the fraction of tensile
12.5 g of asphalt binder and evaluated. After mixing the aggregates strength remaining after moisture conditioning, was calculated.
with asphalt binder, each mixture was placed in a beaker contain- S2
ing 500 ml of distilled water and placed in an oven under a temper- TSR ¼ ð2Þ
S1
ature of 115 °C for 10 min. At the end of the process, the mixture
was drained and poured into a white ceramic container for the S1: average tensile strength of the dry subset (kPa);
visual analysis of adhesion loss between binder and SSA. S2: average tensile strength of the conditioned subset (kPa).

2.3. HMA moisture damage evaluation based on AASHTO T283 3. Results and discussion

Initially, two asphalt mixtures were produced using 75% of nat- 3.1. Morphology results
ural aggregates and 25% of SSA, named HMA_ACBFS and HMA_LD,
according to the type of slag used in the mixture composition. The Table 2 shows the mean results obtained in AIMS and the clas-
aggregate gradation for both mixtures was kept the same and it sification according to the criteria established by Al Rousan [26]. It
was based on the results from Teixeira et al. [4], representing a typ- can be noted that the morphological characteristics are very simi-
ical Superpave dense graded mixture. The optimum binder content lar between both SSA. Considering the texture results, which is
(5.14%) for the HMA_LD was also selected based on the referenced assumed to be a characteristic that will greatly affect the mechan-
study. For the HMA_ACBFS mixture, cylindrical Marshall com- ical interlocking between the binder and the SSA, it can be seen
pacted samples with 75 blows on each side were produced with that either ACBFS or LD slags presented particles ranging from
different binder contents until a 4 ± 1% was obtained. Thus, the smooth to high roughness. Considering these results, it is possible
optimum binder content for this mixture was found and equal to to assume that the differences on the binder-slag adhesion behav-
5.84%. ior were not related to their morphological properties.
In addition, the HMA_ACBFS received the addition of four chem-
ically different powders to improve its moisture resistance. The 3.2. RMS roughness
selected powders were: 1) hydrated lime (HL), a typical additive
used to improve moisture resistance of asphalt mixtures; 2) fluid Fig. 7 shows representative 3D topographic image of each SSA.
gas desulfurization powder (FGD), a steel by-product powder with The average RMS surface roughness was 41.60 ± 13.27 mm and
high content of CaO, 3) electrostatic precipitator powder (EP), 28.80 ± 9.07 mm, for ACBFS and LD samples, respectively. From
another steel by-product powder that contains high percentages the presented values, the roughness of the ACBFS is notoriously
of Fe2O3 in its composition, and 4) natural aggregate passing in superior to the LD. Although the ACBFS presents higher values of
5
Bárbara Luiza Riz de Moura et al. Construction and Building Materials 264 (2020) 120685

Fig. 6. Studied HMA mixtures for moisture damage evaluation.

3.3. Contact angle results


Table 2
Aggregates classification according to Al Rousan [26]. The contact angles formed with each aggregate with the probe
Morphological Limit values ACBFS LD liquids are presented in Table 3. With these values, the surface free
Characteristic energies for the ACBFS and LD were calculated and equivalent to
2D Form < 6.5 (Circular) 45.7% 45.5% 34.65 ergs/cm2 and 36.75 ergs/cm2, respectively. For comparisons,
6.5 – 8.0 (Semi-Circular) 27.6% 26.2% Moraes, Velasquez and Bahia [40] found surface free energy values
8.0 – 10.5 (Semi-Elongated) 20.4% 24.6% of 19.3 and 31.3 ergs/cm2 for granite and limestone aggregates,
> 10.5 (Elongated) 6.3% 3.7% respectively. It can be observed that both SSA presented compara-
Angularity < 2100 (Rounded) 16.9% 22.0%
2100 – 4000 (Sub-Rounded) 58.4% 56.3%
ble values of surface free energy of those found for limestone
4000 – 5400 (Sub-Angular) 19.2% 16.9% aggregate samples, which implies a better adhesive characteristic.
> 5400 (Angular) 5.5% 4.6% However, the difference was not as significant between the two
Sphericity < 0.6 (Flat elongated) 10.2% 9.0% slags as it was expected.
0.6 – 0.7 (Low sphericity) 33.1% 23.6%
Table 3 also shows the values of contact angle between aggre-
0.7 – 0.8 (Moderate sphericity) 34.4% 37.1%
> 0.8 (High sphericity) 22.3% 30.2% gate and binder. As it can be seen, the results for ACBFS and LD
Texture < 165 (Polished) 5.4% 8.8% are similar, which corroborates with the surface energy results.
165 – 275 (Smooth) 24.9% 21.9% ACBFS presented slightly greater value of contact angle, denoting
275 – 350 (Low Roughness) 22.7% 17.2% a lower adhesiveness to the asphalt binder. Since LD is better
350 – 460 (Moderate 23.6% 27.3%
Roughness)
adhered by the binder, it was expected that the adhesion work
> 460 (High Roughness) 23.4% 24.7% for this type of slag would be considerably higher than of ACBFS,
resulting in contact angle values considerably lower than of ACBFS
ones. It is believed that this method could not properly identify dif-
ference in thermodynamic properties, especially, by the hetero-
roughness, this is not enough to guarantee a satisfactory adhesion. geneity of the composition of the material and the distribution of
According to Yoon and Tarrer [11], the physical properties of the pores on the surface [40,41].
aggregate - such as surface area and roughness - influence the ease
of bonding between binder and aggregate, but there is no strong 3.4. Work of adhesion
correlation between these factors. In their study, the authors con-
clude that the chemical and electrochemical properties are deter- As mentioned before, the work of adhesion was measured with
minants for the adhesiveness in the bituminous matrix. and without having the tip covered by asphalt, which means that

Fig. 7. 3D image of the surface in a) ACBFS and b) LD.

6
Bárbara Luiza Riz de Moura et al. Construction and Building Materials 264 (2020) 120685

Table 3
Contact angle values for each liquid for ACBFS and LD characterization and values of contact angle between asphalt binder and aggregates.

Substrate h(°)
Distilled water Glycerol Diiodomethane Asphalt binder
ACBFS 63.38 93.59 43.48 58.41 ± 0.11
LD 65.41 81.55 48.04 56.94 ± 0.06

Atraction Retraction Atraction Retraction

Deflective force [nN]


Deflective force [nN]

0.3 0.3
0.2 0.2 Interatomic interactions
0.1 0.1
0 0
-0.1 0 100 200 -0.1 0 50 100 150 200
-0.2 -0.2
Distance tip-sample [nm] Distance tip-sample [nm]
(a) (b)

Atraction Retraction
Atraction Retraction

Deflective force [nN]


Deflective force [nN]

0.3 0.3
0.2 Interatomic interactions
0.2
0.1 0.1
0 0
-0.1 0 50 100 150 200 -0.1 0 50 100 150 200
-0.2 -0.2
Distance tip-sample [nm] Distance tip-sample [nm]
(c) (d)
Fig. 8. Maps 1 - ACBFS work of adhesion for: a) tip without binder e b) tip with binder and LD work of adhesion for: c) tip without binder e d) tip with binder.

the results of the tip without the binder measure the interaction of the unsatisfactory bonding of EFRA to the binder indicates that
the aggregate with the silicon tip, while those of the binder tip CaO and Al2O3 (important oxides that are related to good moisture
evaluate the interaction of the binder with the slag. In Fig. 8, the resistance of certain aggregates) are not able to overcome the neg-
average values of the maps 1 for the ACBFS and the LD, without ative effects of SiO2 on aggregate’s moisture resistance. Parallel to
and with the binder, are presented. Fig. 9 shows the comparative this, LD slag is mainly composed of CaO and Fe2O3. It is possible to
for all maps in the samples for the same tip, without and with evaluate the good adhesiveness of this aggregate due to the pres-
asphalt binder at the tip. ence of Fe2O3 and the low content of SiO2, since calcium oxide con-
From Fig. 8, it can be noticed that when the tip is not covered by tent are similar on both SSA’s.
asphalt, the retraction forces between silicon tips and the different As already mentioned, and well known by experience, silica
slags do not seem significant, and the pulling force is reduced aggregates have low adhesiveness to the asphalt binder. Corrobo-
quickly. However, the tests performed with the tip covered by rating with this, the works developed by Cala et al. [17–18] con-
asphalt (Fig. 8b and 8d), shows that the pulling force is not reduced cluded that felsic (silicon rich) aggregates have less resistance to
immediately as the tip-binder assembly moves away from the moisture damage than mafics, whose main composition is magne-
aggregate. The calculated work of adhesion presented in Fig. 9 indi- sium and iron and with low silicon content. According to the
cates that the LD slag presented a higher interaction with the tip in authors, SiO2 is the main oxide responsible for damage due to
the three regions tested, in relation to the values of the ACBFS slag. moisture. Also, other oxides, such as Al2O3, Fe2O3, MnO and MgO
But what stands out in these results is the high values of the work - typical of mafic rocks - are important to avoid moisture damage
of adhesion for LD slag when tested with tip covered by asphalt in asphalt mixtures.
binder; the same does not occur for the ACBFS aggregate. Further- In addition to SSA, the selected fillers added in HMA for mois-
more, it can be seen that the distances where the interatomic ture damage evaluation were also subjected to XRF tests, and the
bonds between binder and SSA components are different between results are shown on Table 5. As expected, the FGD powder has a
binder-ACBFS and binder-LD. For binder-LD, the interatomic inter- high calcium content similarly to the HL and its application is jus-
actions start to occur at greater distances. Also, once the inter- tified by the action of basic fillers to increase the base component
atomic adhesive forces are established, they need long distances and, consequently, the adhesion binder-aggregate of the mastic as
to be null again. This result reveals a stronger interaction between found by Arbabpour Bidgoli, Naderi and Moghadas Nejad [42]. The
LD SSA with the asphalt binder in comparison with ACBFS SSA . EP powder has about 63% of iron content and the NA’ is mostly
composed of silicon oxide.
3.5. Chemical composition
3.6. ASTM D3625 results
Table 4 presents the chemical composition of the ACBFS and LD
slag. It can be observed that ACBFS is rich in silicon, calcium and Fig. 10 shows the results from ASTM D3625. It is clear the visual
aluminum oxides, presenting a binary basicity (BB = %CaO/%Si2O) differences between the adhesion of asphalt binder with different
higher than 1, indicating that it is a basic compound. However, SSA’s. While LD showed great adhesion with the asphalt binder,
7
Bárbara Luiza Riz de Moura et al. Construction and Building Materials 264 (2020) 120685

Tip Tip with binder Tip Tip with binder


90

Work of Adhesion
90

Work of Adhesion
75 75

[10-19 J]

[10-19 J]
60 60
45 45
30 30
15 15
0 0
1 2 3 1 2 3 Map
Map

(a) (b)
Fig. 9. Work of adhesion for a) ACBFS e b) LD.

ACBFS notably presented unsatisfactory adhesion results, present-


Table 4 ing a considerable number of particles not completely covered by
Chemical composition of the studied slags by XRF. asphalt binder. This result confirms the low binder-EFRA adhesion
Oxides ACBFS (%) LD (%)
based on work of adhesion (thermodynamic adhesion) and chem-
ical interactions between binder and SSA components.
Na2O 0.28 0.17
MgO 5.20 5.00
Al2O3 9.30 4.90
3.7. HMA moisture damage results
SiO2 32.40 12.70
P2O5 0.10 1.30
SO3 4.50 0.16 Fig. 11 presents the obtained values for tensile strength for the
K2O 0.32 <0.10 conditioned and unconditioned groups – and the respective stan-
CaO 41.2 43.40
dard deviations - for all the mixtures and their respective TSR.
TiO2 0.49 0.36
MnO 0.55 3.20 According to many roadway agencies specifications, a minimum
Fe2O3 2.70 26.4 TSR of 70% is recommended.
SrO 0.11 0.10 From Fig. 11, only the HMA_ LD reached a TSR above 70% while
mixtures with ACBFS did not. These results corroborate with the
advanced surface adhesion tests, where LD presented a significant
amount of Fe2O3 and low percentage of SiO2 (compounds that are
related to positive and negative asphalt mixture moisture resis-
Table 5
tance, respectively) and improved work of adhesion and wettabil-
Chemical composition of the studied additives.
ity according to AFM and SFE results. Regarding the effects of
Oxide HL FGD EP NA’ additional fillers, the mixture with natural aggregate powder
Al2O3 0.41 <0.10 1.76 14.52 (HMA_ACBFS_NA’) behaved similarly to the reference mixture
CaO 66.83 56.40 12.14 7.46 (HMA_ACBFS), which indicates that the addition of a silicon-rich
Cr2O3 <0.01 <0.01 0.03 <0.01
filler did not contribute to resistance to moisture conditioning. Fur-
Fe2O3 0.16 0.12 62.94 8.82
K2O 0.19 0.04 0.33 3.56 thermore, it should be noted that the addition of any filler, just to
MgO 1.35 0.55 1.95 2.61 physically filling the asphalt mixture, does not guarantee a good
MnO <0.01 <0.01 0.53 0.12 performance against resistance to moisture damage.
Na2O <0.1 <0.1 <0.10 2.86 The mixture HMA_ACBFS_HL presented a TSR of 65%, showing
P2O5 0.17 0.17 0.08 2.35
SiO2 1.28 0.38 5.55 54.65
the improvement on the moisture resistance due to HL addition
TiO2 0.02 <0.01 0.15 1.63 (from 54% to 65%). The positive effect due to the addition of HL
ZrO2 <0.01 – – 0.13 on the mixture’s moisture resistance was expected, as other stud-
SO3 – 25.09 1.11 – ies showed similar results [21,43]. Although the addition of HL
ZnO – 0.02 0.08 –
improved the moisture resistance of asphalt mixture in compar-
ison with the mixture without any additive (i.e. HMA_ACBFS), it
did not reach the minimum recommended TSR value (i.e. 70%).

Fig. 10. ASTM D3625 - Boiling Test: a) LD (same vol. as ACBFS); b) LD (same mass as ACBFS), and c) ACBFS (250 g in mass and 94 cm3 in volume).

8
Bárbara Luiza Riz de Moura et al. Construction and Building Materials 264 (2020) 120685

Unconditioned Conditioned TSR

1.8 74% 80%


1.6 65% 63% 70%
58%
Tensile Strength (MPa) 1.4 52% 60%
1.2 54% 50%

TSR (%)
1.0
40%
0.8
30%
0.6
0.4 20%

0.2 10%
0.0 0%
HMA_ACBFS HMA_ACBFS_HL HMA_ACBFS_FGD HMA_ACBFS_EP HMA_ACBFS_NA' HMA_LD

Fig. 11. Moisture induced damage results.

Positive effects on moisture resistance gain were observed with the  In general, it was shown that LD can potentially improve mois-
addition of EP powder (which is rich in Fe2O3). HL and EP powders ture damage resistance of asphalt mixtures. However, other
were more beneficial to the increase of moisture damage resis- performance tests need to be carried out to confirm the overall
tance than other fillers. The mixture with FGD powder presented effects of LD due to its expansive nature. ACBFS did not show
a TSR of 58%. Although FGD presents a high content of CaO (as satisfactory adhesiveness to asphalt binder in this study. How-
HL), its effect was much less than HL. This could be related to ever, it can be engineered with proper antistripping agents, as
two aspects. First, the available free lime content in FGD might this study presented the effects of HL and EP.
be low. Dick et al. [44] showed that FGD free lime content can be  This study showed that there is indeed much difference among
decreased when this material is stockpiled due to carbonation of slags, and the differences affect bonding between slags and bin-
Ca(OH)2 by forming CaCO3 and ettringite and thaumasite forma- der due to surface characteristics. It is important to note that
tion. Second, the high content of SO3 in FGD could be a negative properties of slags can vary due to the raw materials used and
factor. The two aspects require further investigation, which is in the steel fabrication process. A careful application is thus
progress by the authors. needed for practical purposes, and further studies to investigate
the case-specific challenges are necessary. The authors are cur-
4. Conclusions rently working on follow-up studies and results will be pre-
sented when available.
This study evaluated the adhesion characteristics in the binder-
slag system through physical-geometrical-chemical surface char- CRediT authorship contribution statement
acterization to better understand the adhesion mechanisms that
may affect the asphalt mixture performance. From the test results, Bárbara Luiza Riz Moura: Conceptualization, Investigation,
the following conclusions can be drawn: Methodology, Validation, Formal analysis, Writing - original draft,
Writing - review & editing. Jamilla Emi Sudo Lutif Teixeira: Con-
 Considering the three mechanisms that contribute to binder- ceptualization, Investigation, Methodology, Validation, Formal
aggregate adhesion, the physical adhesion (due to surface analysis, Writing - original draft, Writing - review & editing, Fund-
roughness) was not predominant to explain the differences ing acquisition. Renata Antoun Simão: Conceptualization,
found between SSA-binder systems. On the other hand, the bet- Methodology, Validation, Formal analysis, Writing - original draft,
ter bonding potential of LD was explained based on the chemi- Writing - review & editing. Mahdieh Khedmati: Investigation, Val-
cal and surface energy measurements. This, although somewhat idation, Formal analysis, Writing - original draft, Writing - review &
indirect, indicates that the adhesion mechanism of slag aggre- editing. Yong-Rak Kim: Conceptualization, Validation, Writing -
gates with asphalt binder is more influenced by aggregate sur- original draft, Writing - review & editing. Patrício José Moreira
face chemistry/energy than roughness. Pires: Conceptualization, Methodology, Validation, Writing - orig-
 The AFM appeared as a promising method to evaluate the inter- inal draft, Writing - review & editing, Funding acquisition.
atomic bonds between aggregates and binder, and it can be
used to quantify the work of adhesion between those materials.
 Chemical composition along with other measurements indi- Declaration of Competing Interest
cated that the high content of silica oxide approximates the
behavior of the iron slag (ACBFS) to that of a felsic rock, which The authors declare that they have no known competing finan-
presents low adhesiveness to the binder. Parallel to this, iron cial interests or personal relationships that could have appeared
oxide seems important for the good adhesiveness of LD slag. to influence the work reported in this paper.
 The observation from the small-scale surface characteristics
agreed well with the large scale observations from the mixture Acknowledgements
tests: ASTM D3625 and AASHTO T283.
 It was observed that the additives can affect moisture resis- This study was financed in part by ArcelorMittal Tubarão. The
tance. In particular, HL and EP could improve moisture resis- second author also thanks the CAPES (Pos-Doctoral Program Grant
tance with ACBFS. This indicated the positive effects of CaO no. 88881.171157/2018-01) and CNPq (303845/2017-1) for the
and Fe2O3 on the asphalt mixture’s moisture damage scholarship grants. The authors also thank COPPE (UFRJ) and UFC
resistance. Federal Universities on the name of D.Sc. Thaisa Ferreira Macedo,
9
Bárbara Luiza Riz de Moura et al. Construction and Building Materials 264 (2020) 120685

M.Sc. Mieka Arão and Ph.D. Verônica Castelo Branco, for their tech- Sustain Mater. Technol. (2018) 35p, https://doi.org/10.1016/j.susmat.2018.
e00071.
nical support on surface free energy and AIMS tests.
[22] I.S. Bessa, V.T.F.C. Branco, J.B. Soares, J.A.N. Neto, Aggregate Shape Properties
and Their Influence on the Behavior of Hot-Mix Asphalt, J. Mater. Civ. Eng. 27
References (7) (2015) 8p, https://doi.org/10.1061/(ASCE)MT.1943-5533.0001181.
[23] Y. Xiao, F. Wang, P. Cui, L. Lei, J. Lin, M. Yi, Evaluation of Fine Aggregate
[1] H.A. Rondón-Quintana, J.C. Ruge-Cárdenas, M.M. de Farias, Behavior of Hot- Morphology by Image Method and Its Effect on Skid-Resistance of Micro-
Mix Asphalt Containing Blast Furnace Slag as Aggregate: Evaluation by Mass Surfacing, J. Materials. 11 (2018), https://doi.org/10.3390/ma11060920.
and Volume Substitution, J. Mater. Civ. Eng. 31 (2019) 04018364, https://doi. [24] V.M.C. Araujo, I.S. Bessa, V.T.F. Castelo Branco, Measuring Skid Resistance of
org/10.1061/(ASCE)MT.1943-5533.0002574. Hot Mix Asphalt Using the Aggregate Image Measurement System (AIMS), J
[2] M.M.A. Aziz, M.R. Hainin, H. Yaacob, Z. Ali, F.-L. Chang, A.M. Adnan, Constr. Build. Mater. 98 (2015) 476–481p, https://doi.org/10.1016/
Characterisation and Utilisation of Steel Slag for the Construction of Roads j.conbuildmat.2015.08.117.
and Highways, J. Mater. Res. Innov. 18 (2014) 255–259p, https://doi.org/ [25] AASHTO TP,, 81: Standard Method of Test for Determining Aggregate Shape
10.1179/1432891714Z.000000000967. Properties by Means of Digital Image Analysis, American Association of State
[3] G. C. Wang, Slag Use in Asphalt Paving. In: The Utilization of Slag in Civil Highway and Transportation Officials, Washington, 2012.
Infrastructure Construction. Woodhead Publishing, (2016) 201-238p. [26] T.M. Al Rousan, Characterization of Aggregate Shape Properties Using a
https://doi.org/10.1016/B978-0-08-100381-7.00010-0. Computer Automated System, Dissertation (PhD), Texas A&M University,
[4] J.E.S.L. Teixeira, A.G. Schumacher, P.J.M. Pires, V.T.F.C. Branco, H.B. Martins, 229p, 2004.
Expansion Level of Steel Slag Aggregate Effects on Both Material Properties and [27] A. Bhasin, E. Masad, D. Little, R. Lytton, Bituminous – Aggregate Surface
Asphalt Mixture Performance, J Transp. Res. Rec. (2019) 1–10p, https://doi.org/ Requirements Limits on Adhesive Bond Energy for Improved Resistance of
10.1177/0361198119835513. Hot-Mix Asphalt to Moisture Damage, J Transp. Res. Rec. (2006) 3–13p.
[5] K. Kim, S. Haeng Jo, N. Kim, H. Kim, Characteristics of Hot Mix Asphalt [28] J. Howson, E. Masad, A. Bhasin, D. Little, R. Lytton, Comprehensive Analysis of
Containing Steel Slag Aggregate According to Temperature and Void Surface Free Energy of Asphalts and Aggregates and the Effects of Changes in
Percentage, J Constr. Build. Mater. 188 (2018) 1128–1136p, https://doi.org/ pH, J. Constr. Build. Mater. 25 (2011) 2554–2564p, https://doi.org/10.1016/
10.1016/j.conbuildmat.2018.08.172. j.conbuildmat.2010.11.09.
[6] J. Lizarazo-marriaga, P. Claisse, E. Ganjian, Effect of Steel Slag and Portland [29] A. Bhasin, E. Masad, D. Little, R. Lytton, Limits on Adhesive Bond Energy for
Cement in the Rate of Hydration and Strength of Blast Furnace Slag Pastes, J. Improved Resistance of Hot-Mix Asphalt to Moisture Damage, J Transp. Res.
Mater. Civ. Eng. 23 (2011) 153–160p, https://doi.org/10.1061/(ASCE)MT.1943- Rec. (2006) 3–13, https://doi.org/10.3141/1970-03.
5533.0000149. [30] D.Y. Kwok, A.W.U. Neumann, Contact Angle Measurement & Contact Angle
[7] Y. Hou, X. Ji, J. Li, X. Li, Adhesion Between Asphalt and Recycled Concrete Interpretation, J. Adv. Colloid Interface Sci. 81 (1999) 167–249p.
Aggregate and Its Impact on the Properties of Asphalt Mixture, J. Materials 11 [31] A. Bhasin, D.N. Little, K.L. Vasconcelos, E. Masad, Surface Free Energy to
(2018) 15p, https://doi.org/10.3390/ma11122528. Identify Moisture Sensitivity of Materials for Asphalt Mixes, J Transp. Res. Rec.
[8] C.W. Curtis, K. Ensley, J. Epps, Fundamental Properties of Asphalt- (2007) 37–45p, https://doi.org/10.3141/2001-05.
AggregateInteractions Including Adhesion and Absorption, Strateg. Highw. [32] A.W. Hefer, A. Bhasin, D.N. Little, Bitumen Surface Energy Characterization
Res. Progr. SHRP-A-341 (1993). Using a Contact Angle Approach, J. Mater. Civ. Eng. (2006) 759–767p.
[9] J. Zhang, A.K. Apeagyei, G.D. Airey, J.R.A. Grenfell, Influence of Aggregate [33] M.R. Kakar, M.O. Hamzah, M.N. Akhtar, D. Woodward, Surface Free Energy and
Mineralogical Composition on Water Resistance of Aggregate-Bitumen Moisture Susceptibility Evaluation of Asphalt Binders Modified with
Adhesion, Int. J. Adhes. Adhes. 62 (2015) 45–54p, https://doi.org/10.1016/j. Surfactant-Based Chemical Additive, J. Clean. Prod. 112 (2016) 2342–2353p,
ijadhadh.2015.06.012. https://doi.org/10.1016/j.jclepro.2015.10.101.
[10] J.P. Aguiar-Moya, A. Baldi-Sevilla, J. Salazar-Delgado, J.F. Pacheco-Fallas, L. [34] T.F. Macedo, G.A. Badilla-Vargas, P.H. Osmari, A.D. Oliveira, R.A. Simão, L.F.M.
Loria-Salazar, F. Reyes-Lizcano, N. Cely-Leal, Adhesive Properties of Asphalts Leite, F.T.S. Aragão, An Experimental Testing and Analysis Procedure to
and Aggregates in Tropical Climates, Int. J. Pavement Eng. 19 (2016) 738–747p, Determine Linear Viscoelastic Properties of Asphalt Binder Microstructural
https://doi.org/10.1080/10298436.2016.1199884. Components, J. Constr. Build. Mater. 230 (2020) 116999, https://doi.org/
[11] H.H. Yoon, A.R. Tarrer, Effect of Aggregate Properties on Stripping, J Transp. Res. 10.1016/j.conbuildmat.2019.116999.
Rec. (1988) 37–43p. [35] R.G. Allen, D.N. Little, A. Bhasin, R.L. Lytton, Identification of the Composite
[12] A. Bhasin, D.N. Little, Application of Microcalorimeter to Characterize Relaxation Modulus of Asphalt Binder Using AFM Nanoindentation, J. Mater.
Adhesion Between Asphalt Binders and Aggregates, J. Mater. Civ. Eng. 21 Civ. Eng. 25 (2013) 530–539p, https://doi.org/10.1061/(ASCE)MT.1943-
(2009) 235–243p, https://doi.org/10.1061/(ASCE)0899-1561(2009)21:6(235). 5533.0000615.
[13] X. Zhou, G. Zhao, S. Tighe, M. Chen, S. Wu, S. Adhikari, Y. Gao, Quantitative [36] Y. Veytskin, C. Bobko, C. Castorena, Nanoindentation and Atomic Force
Comparison of Surface and Interface Adhesive Properties of Fine Aggregate Microscopy Investigations of Asphalt Binder and Mastic, J. Mater. Civ. Eng.
Asphalt Mixtures Composed of Basalt, Steel Slag, and Andesite, J. Constr. Build. 28 (2016) 04016019, https://doi.org/10.1061/(asce)mt.1943-5533.0001532.
Mater. 246 (2020) 118507. [37] M. Xu, J. Yi, D. Feng, Y. Huang, D. Wang, Analysis of Adhesive Characteristics of
[14] J.P. Aguiar-Moya, J. Salazar-Delgado, A. Baldi-Sevilla, F. Leiva-Villacorta, L. Asphalt Based on Atomic Force Microscopy and Molecular Dynamics
Loria-Salazar, Effect of Aging on Adhesion Properties of Asphalt Mixtures with Simulation, J. Appl. Mater. Interfaces 8 (2016) 12393–12403p, https://doi.
the Use of Bitumen Bond Strength and Surface Energy Measurement Tests, J. org/10.1021/acsami.6b01598.
Transp. Res. Rec. 2505 (2015) 57–65p, https://doi.org/10.3141/2505-08. [38] H. Fischer, H. Stadler, N. Erina, Quantitative Temperature-Depending Mapping
[15] A. W. Hefer, Adhesion in Bitumen-Aggregate Systems and Quantification of the of Mechanical Properties of Bitumen at the Nanoscale Using the AFM Operated
Effects of Water on the Adhesive Bond, (2004). Dissertation (PhD), Texas A&M with Peakforce Tapping Mode, J. Microsc. 250 (3) (2013) 210–217.
University, 222p. [39] S. Nahar, A. Schmets, G. Schitter, A. Scarpas, Quantitative Nanomechanical
[16] I.L. Jamieson, J.S. Moulthrop, D.R. Jones, SHRP Results on Binder – Aggregate Property Mapping of Bitumen Micro-Phase by Peak-Force Atomic Force
Adhesion and Resistance to Stripping, Asphalt Yearbook, Institute of Asphalt Microscopy, ISAP Con 30 Asphalt Pave (2014) 1397–1404p.
Technology, UK, 1995. [40] R. Moraes, R. Velasquez, H. Bahia, Using Bond Strength and Surface Energy to
[17] A. Cala, S. Caro, M. Lleras, Y. Rojas-Agramonte, Impact of The Chemical Estimate Moisture Resistance of Asphalt-Aggregate Systems, J. Constr. Build.
Composition of Aggregates on the Adhesion Quality and Durability of Asphalt- Mater. 130 (2017) 156–170p, https://doi.org/10.1016/
Aggregate Systems, J. Constr. Build. Mater. 216 (2019) 661–672p, https://doi. j.conbuildmat.2016.10.043.
org/10.1016/j.conbuildmat.2019.05.030. [41] A. Bhasin, Development of Methods to Quantify Bitumen-Aggregate Adhesion
[18] A. Cala, S. Caro, M. Lleras-Jacbosen, Y. Rojas-Agramonte, Understanding the and Loss of Adhesion Due to Water, Dissertation (PhD), Texas A&M University,
Role of the Chemical Composition of Aggregates on the Moisture Susceptibility 158p, 2006.
of Asphalt Mixtures, Transportation Research Board - 98th Annual Meeting, Paper [42] M.A. Bidgoli, K. Naderi, F.M. Nejad, Effect of Filler Type on Moisture
number 19–04533 (2019) 7p. Susceptibility of Asphalt Mixtures Using Mechanical and Thermodynamic
[19] ASTM D3625 / D3625M-12: Standard Practice for Effect of Water on Properties, J. Mater. Civ. Eng. 31 (4) (2019) 04019024.
Bituminous-Coated Aggregate Using Boiling Water, ASTM International, [43] D.N. Little, J.A. Epps, P.E. Sebaaly, The Benefits of Hydrated Lime in Hot Mix
West Conshohocken, PA, USA (2012). Asphalt, National Lime Association (2006).
[20] AASHTO T,, 283: Resistance of Compacted Bituminous Mixture to Moisture- [44] W. A. Dick, Y. Hao, R. C. Stehouwer, J. M. Bigham, W. E. Wolfe, D. Adriano, J. H.
Induced Damage, American Association of State Highway and Transportation Beeghly, R. J. Haefner, Beneficial Uses of Flue Gas Desulfurization By-Products:
Officials, Washington, 2014. Examples and Case Studies of Land Application. In: Land Application of
[21] P. Jitsanigam, W.K. Biswas, M. Compton, Sustainable Utilization of Lime Kiln Agricultural, Industrial, and Municipal By-Products (SSSA Book Series), 6-18
Dust as Active Filler in Hot Mix Asphalt with Moisture Damage Resistance, J. (2000) 505-536p. https://doi.org/10.2136/sssabookser6.c18.

10

You might also like