Davis, C., Castorena, C., 2015. Implications of Physicoechemical Interactions in Asphalt Mastics On Asphalt Microstructure

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Construction and Building Materials 94 (2015) 83–89

Contents lists available at ScienceDirect

Construction and Building Materials


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

Implications of physico–chemical interactions in asphalt mastics on


asphalt microstructure
Chelsea Davis 1, Cassie Castorena ⇑
North Carolina State University, 2501 Stinson Dr., Raleigh, NC 27695, United States

h i g h l i g h t s

 Asphalt mastics are evaluated using Atomic Force Microscopy.


 Physico–chemical interactions in mastics alter asphalt morphology.
 Filler specific surface area affects extent of physico–chemical interaction.
 Physico–chemical interactions alter asphalt mastic rheology.

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

Article history: Asphalt binder and filler blend to form asphalt mastic, which constitutes the effective adhesive film in
Received 7 January 2015 asphalt concrete. Pavement performance can be improved through better engineering of the mastic,
Received in revised form 10 June 2015 which requires a fundamental understanding of the interaction between asphalt and filler. Physico–
Accepted 12 June 2015
chemical interactions result in adsorption of polar fractions of the asphalt onto filler surfaces, leading
to the formation of an interphase layer on the surface of particles and modifying the asphalt binder
matrix. This study seeks to investigate the effects of physico–chemical interaction on binder matrix
Keywords:
microstructure using Atomic Force Microscopy (AFM) and qualitatively relate microstructural findings
Asphalt
Mastic
to macroscopic rheology.
Physico–chemical interaction Ó 2015 Elsevier Ltd. All rights reserved.
Microstructure
Atomic Force Microscopy

1. Introduction fracture resistance, and moisture susceptibility of asphalt concrete


[3]. Current Superpave asphalt concrete specifications include pro-
Asphalt concrete is a composite material consisting of aggre- visions for filler only through specifying an allowable range of
gates of varying size, asphalt binder, and air voids. Coarse aggre- mass ratios between filler and effective binder [2]. These specifica-
gates in asphalt mixtures are effectively coated by a blend of tions were developed based on empirical observations. While a
asphalt binder and filler, termed asphalt mastic [1]. Fillers consist great deal of research has been conducted to understand the per-
of particulate matter less than 0.075 mm in diameter [2]. Fillers formance of binders and mixtures, relatively little attention has
used in asphalt concrete include both natural and manufactured been given to the asphalt mastic and correspondingly, filler. An
origins [3]. Natural fillers are the dust portion of mineral aggregate. improved understanding of the interaction between filler and
Manufactured fillers, on the other hand, are produced as asphalt binder could lead to improved engineering and specifica-
by-products of industrial processes (e.g., fly ash). The mastic tions of fillers in asphalt concrete and hence, improved perfor-
constitutes the weakest phase of asphalt concrete and therefore mance of pavements.
performance of asphalt pavements is highly correlated to the prop- Various studies have shown there are three primary mecha-
erties of the mastic. It has been demonstrated that filler can signif- nisms by which fillers reinforce asphalt binder [4–6]: volume fill-
icantly influence constructability, oxidative aging, stiffness, ing, particle structuralization, and physico–chemical interactions.
Volume filling and particle structuralization are both means of
⇑ Corresponding author. mechanical reinforcement. Volume filling increases asphalt mastic
E-mail addresses: daviscr2@email.appstate.edu (C. Davis), cahintz@ncsu.edu (C. stiffness simply as a result of the replacement of asphalt binder
Castorena). volume with rigid particles. Filler particles start to form an inter-
1
Address: Appalachian State University, 287 Rivers St, Boone, NC 28608, United connected network at a filler volumetric concentration of roughly
States.

http://dx.doi.org/10.1016/j.conbuildmat.2015.06.026
0950-0618/Ó 2015 Elsevier Ltd. All rights reserved.
84 C. Davis, C. Castorena / Construction and Building Materials 94 (2015) 83–89

40%, leading to a rapid increase in the rate of stiffening with surfaces along with interfacial region with asphalt was measured
increasing volume fraction [6–8]. At lower filler concentrations, using AFM. It was demonstrated that a 2–5 lm gap existed
particle contact is not established and the mastic will behave as between asphalt film and filler which the authors attributed to sur-
a dilute suspension, with volume filling constituting the dominant face tension between the filler and asphalt. In addition, nanoscale
mechanical reinforcement mechanism. The third type of reinforce- adhesion tests were conducted which demonstrated an increase
ment, physico–chemical interaction, involves the adsorption of in adhesion with addition of filler up to filler volumetric concentra-
polar fractions of the asphalt binder onto the surface of filler parti- tions ranging from 0.1 to 0.2 at which point subsequent addition of
cles [5,9,10]. Physico–chemical interaction is illustrated in Fig. 1. filler resulted in a decrease in adhesion. However, no AFM experi-
Physico–chemical interaction leads to formation of an interphase ments were conducted directly on asphalt mastics. Nazzal et al.
adsorbed layer of the polar fractions of asphalt on the surface of fil- [20] studied the dispersion of nanoclay modifiers in asphalt binder
ler particles. In addition, loss of certain components of the asphalt using AFM, which the authors concluded indicated the nanoclay
to adsorption modifies the chemistry and morphology of the was well dispersed within the asphalt. Despite no observed differ-
non-adsorbed, ‘‘effective’’ binder matrix. ences in morphology, the inclusion of the nanoclay significantly
While mechanical reinforcement mechanisms can be readily altered the adhesive forces of asphalt materials based on nanoin-
inferred from observed mechanical behavior, understanding phy- dentation experiments. Nanoclay dosages studied were 2% and
sico–chemical interactions is more challenging using macroscopic 4% by weight of binder, which is substantially lower than typical
measures. Several researchers have inferred physico–chemical filler concentrations in asphalt mastics. Also, nanoclay particles
interactive effects through rheological measurements coupled are much finer than typical mineral fillers and thus, further study
with micromechanical models assumed to accurately reflect of morphology implications of fillers on asphalt binders is needed.
macroscopic mastic behavior with emphasis on the thickness and This study seeks to investigate the effects of physico–chemical
effective properties of the adsorbed layer [11,12]. Others have interaction between asphalt and filler on ‘‘effective’’ binder matrix
inferred physico–chemical effects through thermodynamic microstructure using AFM and qualitatively relate microstructural
measures. Craus et al. [10] investigated physico–chemical findings to macroscopic rheology.
interactive effects through measuring heat release using a differen-
tial microcalorimeter. In addition, researchers have investigated
changes in glass transition temperatures between binder and mas- 2. Materials and methods

tics of varying concentrations to investigate physico–chemical 2.1. Materials


interaction effects on matrix properties [5] and adsorbed
interphase layer thickness [13]. The majority of past research An unmodified binder with a PG grade of 64–22 was used in this study. The bin-
efforts have focused on identifying interphase layer effects with lit- der was mixed with three different fillers: granite, Portland cement (PC), and
hydrated lime to produce mastics. The granite is a natural filler whereas PC and
tle consideration to the modified asphalt matrix characteristics.
hydrated lime are both manufactured fillers. The specific gravity and specific sur-
Understanding the physico–chemical interaction between asphalt face area (SSA) of each filler are presented in Table 1. Specific surface area was mea-
and filler on the resultant asphalt matrix is critical to enable sured using a Quantachrome Monosorb single-point BET surface area analyzer. The
improved design of mastics. Brunauer–Emmett–Teller (BET) theory was used as the analysis technique for SSA
Asphalt is comprised of a complex, heterogeneous blend of ali- measurements. Specific surface area is an important factor governing physico–
chemical interaction intensity as greater specific surface areas indicate greater
phatic and aromatic hydrocarbons with moderate amounts of sul- opportunity for adsorption [5]. Each filler was blended with binder at a volumetric
fur and trace amounts of oxygen, nitrogen, and other elements concentration of approximately 24%, leading to dust to binder mass ratios ranging
including metals [14]. This complex chemistry gives rise to com- from 0.8 to 1.0, within the specified range in Superpave mix design specifications
posite structure in asphalt binders on the order of nanometers to [2]. Use of higher concentrations was avoided to minimize reinforcement by
particle-to-particle contact which becomes most prevalent at volume concentra-
micrometers. Hence, understanding the effects of physico–chemi-
tions greater than 40% [6,7].
cal interaction between asphalt binder and filler on the effective All of the mastics in this study were mixed with a laboratory stand mixer with
asphalt matrix properties is critical to improved engineering of supplemental hand mixing to initially agitate mastics. Before mixing, the binder
mastics. Recently, there has been significant advancement in and filler were heated at 150 °C for 30 min and then poured into pre-weighed
understanding asphalt binder microstructure through use of pint-sized containers. The binder was then transferred to a hot plate to keep the
binder fluid during mixing. The oven-heated filler was slowly poured into the bin-
Atomic Force Microscopy (AFM) [15–18]. AFM is a type of scanning
der while agitating. The mastic containers were then covered to avoid dust contam-
probe microscopy in which a very fine cantilever tip is rastered ination and allowed to cool at room temperature. In order to prevent variations in
across the surface of a specimen to map the distribution and prop- thermal history the binder sample source was also placed in the oven at 150 °C for
erties of constituent phases. The opacity of asphalt does not lend 30 min prior to creating test specimens.
well to optical microscopy methods, making scanning probe
microscopy an attractive method for studying binder microstruc- 2.2. Experimental methods
ture. Non-contact or tapping mode, where very small forces are
applied as the cantilever is rastered across a specimen’s surface, Each mastic as well as the bulk binder was tested using the Dynamic Shear
has proven useful in studying soft materials, like asphalt, for Rheometer (DSR) to determine rheological properties. In addition, AFM in
Tapping Mode to investigate microstructural implications of physico–chemical
obtaining images of surface topography and phase contrast which
interactions between filler and binder on the effective binder matrix.
can be used to identify different microstructural phases in a mate-
rial. In addition, spectroscopy mode, in which the cantilever tip
moves vertically into a specimen rather than across its surface, 2.2.1. Dynamic Shear Rheometer
can be used to perform nanoindentation experiments for measure- Rheological properties of each binder and mastic sample were determined
using an ARG-2 Dynamic Shear Rheometer (DSR) from TA Instruments. Frequency
ment of stiffness and adhesive characteristics. Recent research has
sweep tests at multiple temperatures were used to generate the dynamic shear
shown that asphalt binders consist of a variety of microdomains modulus master curve for all binders and mastics. The frequency sweep test applies
[16,17] with varying rheological properties [18]. cyclic strain at constant amplitude over a range of loading frequencies and temper-
Limited research has also been conducted to understand the atures. Loading was applied at frequencies ranging from 0.15 Hz to 25 Hz. The fre-
effect of filler on binder morphology using AFM. Tan and Guo quency sweeps in this experiment were run at 64, 50, 35, 20, and 5 °C. The 8 mm
parallel plate DSR geometry was used for test temperatures of 5, 20, and 35 °C
[19] used AFM to study the interaction between asphalt and filler whereas the 25 mm parallel plate geometry was utilized for testing at 50 and
by preparing asphalt droplets which were placed in contact with 64 °C. Time–temperature superposition was utilized to construct dynamic shear
slices of aggregate. The relative roughness of different aggregate modulus master curves.
C. Davis, C. Castorena / Construction and Building Materials 94 (2015) 83–89 85

Fig. 1. Physico–chemical interaction in asphalt mastics.

Table 1 microdomains. In addition, Allen et al. [18] studied the microrheology of the
Filler properties. three microdomains and demonstrated that the paraphase is the stiffest phase
and the periphase is the softests with the ‘‘bee’’ phase having intermediate stiff-
Filler type Specific gravity Specific surface area (m2/g) ness and that the stiffness of the periphase and paraphase.
Baghouse fines (Granite) 2.58 1.639 The image processing protocol developed calculated the relative proportion of
Portland cement dust 3.15 2.240 each microdomain in terms of area occupied. This was accomplished by isolating
Lime 2.30 14.408 the different microdomains based on color and determining the relative number
of pixels each microdomain occupied. Example output of the developed image pro-
cessing protocol is shown in Fig. 4. The ‘‘bee’’ structures are denoted by the blue
area, paraphrase by the green area, and periphase by the yellow–red area. The
2.2.2. Atomic Force Microscopy developed image analysis protocol allowed for determining changes in the relative
To prepare AFM test samples, the binder or mastic was poured into a Pyrex Petri quantities of the different microstructures as a result of adsorption on filler
dish to produce a thickness of approximately 5 mm. Samples were covered and then surfaces.
stored at room temperature for 24 h prior to testing to allow the sample to anneal,
allowing for the microstructure to form as suggested by Fischer et al. (2013).
A Veeco D3000 Scanning Tip AFM (Fig. 2) with a wafer stage was used in 3. Results and discussion
Tapping Mode to gather height (topographic) and phase contrast images in this
study. In Tapping Mode, the cantilever is oscillated at its resonant frequency. A
piezo drive is altered using feedback control to maintain a constant tip to sample 3.1. AFM
spacing, termed set point. The amplitude of the resultant oscillation varies with
changes in surface domains, which are converted into a false color image depicting AFM results are presented in Fig. 5 for the base binder and the
various heights as a function planar position. In addition, the phase lag between the
mastics prepared with granite, PC, and lime. Note that for each
piezo signal and cantilever response will vary with surface features and can also be
used to construct a map of surface variations. material in Fig. 5, the topographic image is shown on the left and
In this study, each sample was scanned at three different surface locations. In phase image on the right. The bright spots within the topographic
addition, multiple samples were scanned to ensure results were representative of images for granite and lime mastics are speculated to correspond
the material. The scanning area at each location was constant at 25 lm by to points where very fine filler particles are near the surface and
25 lm. A single crystal silicon cantilever probe with a resonance frequency of
325 kHz and force constant of 40 N/m was used for each sample based on recom-
hence, a drastic change in surface topology occurs. However, phase
mendations by Fischer et al. [21]. A scanning rate of 1 Hz and speed of 2 lm per sec- images on the right do not show these features, indicating they
ond were utilized. The initial data scale for the height image was set to 100 nm and represent the surface features of the effective binder matrix.
phase image data scale set to 40° but subject to adjustment based on the range of The images in Fig. 5 demonstrate clear identification of the
roughness of the sample surface. Once all the initial parameters were set, the sur-
three microdomains: paraphase, periphase, and bees, previously
face and tip were located and scanning was initiated. The parameters were adjusted
until a clear and detailed image was output. Nanoscope III software was used in discussed. However, it is apparent that the relative proportion
order to view and process images. and size of the features changes with the addition of filler and
Images indicated probing of the binder phase in all samples (i.e., no filler parti- appears to be filler specific, providing evidence of physico–chemi-
cles) as all filler particles were coated with asphalt binder. This observation will be cal interaction induced changes to the effective binder matrix.
further discussed later. It should be noted that a small amount of filler settlement is
expected and hence, the filler concentration at the surface of the mastics may not be
Fig. 6 displays the results of image analysis to determine the rela-
representative of the bulk volume of the mastic. However, since each mastic sample tive surface area each microdomain occupies for the materials
experienced the same thermal and storage history, settlement effects are assumed studied.
to be similar among samples. One notable feature in Fig. 5 is that the size decreases and rel-
Qualitative comparison between AFM images from different mastics and bin-
ative number of bee structures increases with the addition of filler.
der were used to assess the effects of physico–chemical interaction on effective
binder microstructure. In order to quantify the phase images for further compar- It can be seen that the relative change in the size, relative quantity,
ison, a combined Matlab – Photoshop CS5 protocol was developed to identify the and special distribution of bee features is greatest for the mastic
relative area of AFM images comprised of the different microdomains. In all prepared with lime, indicating a higher degree of physico–chemi-
specimens (both mastic and binder) evaluated, three distinct microdomains were cal interaction intensity. This is expected since the lime filler has
observed, which have also been observed in many asphalts in past studies [16–
18]. These microdomains are termed the Catana (or ‘‘bee’’ phase), the Periphase,
a much higher specific surface area than the other fillers,
and the Paraphase [16] and are illustrated in Fig. 3. Note that other microdo- (14.408 m2/g compared to 1.639 m2/g and 2.24 m2/g for granite
mains have been observed in asphalt binders as well [16,18] but were not pre- and PC, respectively), and hence, given the same volume concen-
sent in the binder utilized in this study. The ‘‘bee’’ features have been given a tration, provides more opportunity for adsorption. Recall, the bee
significant amount of attention in past investigations. Pauli et al. [17] demon-
structures correspond to the crystallization of waxes and thus,
strated that the ‘‘bees’’ form a result of wax crystallization, and are contained
within the saturate and naphthene aromatic fractions of binder. Das et al. [22] results indicate a change in wax crystallization and spatial distri-
demonstrated that the wax crystallization plays a critical role in fracture resis- bution with addition of filler. In addition, as polar components of
tance as fractures tend to initiate and propagate at the interface between asphalt are lost to adsorption, an increase in the relative
86 C. Davis, C. Castorena / Construction and Building Materials 94 (2015) 83–89

Fig. 2. Veeco D3000 scanning tip AFM.

However, one also must consider that the stiffness of each phase
may change as a result of physico–chemical interaction. Changes
in the microrheology of individual microdomains upon oxidative
aging was reported by Allen et al. [18].

3.2. DSR

Dynamic shear modulus master curves for the binder and


three mastics evaluated in this study are presented in Fig. 7. It
can be seen that the mastics all demonstrate higher moduli than
the binder, as expected. It can also be observed that the different
fillers have differing stiffening effects on the binder, despite
being added to the binder in equal volumetric concentrations.
Recall, volumetric concentrations of fillers were relatively low
at 24% and hence not anticipated to lead to particle structural-
ization. Since volumetric concentrations of fillers are consistent
for all mastics, the effects of volume filling on rheology are
anticipated to be equivalent for all three filler types. Hence, dis-
crepancies in stiffening effects of different fillers can largely be
attributed to physico–chemical interaction. Trends in stiffening
effects of the fillers are consistent with microstructural observa-
tions: Lime has the greatest stiffening effect and induced the
greatest change in binder microstructure whereas granite and
PCC show lower stiffening effects (and also have much lower
SSAs). Thus, microstructural findings support the hypothesis that
Fig. 3. Asphalt binder microdomains. differences in stiffening effects can be attributed to physico–
chemical interactions. It should be noted that the effective,
non-adsorbed binder matrix is hypothesized to soften as a result
concentration of waxes is expected as waxes are primarily nonpo- of physico–chemical interaction. Thus, the increase in stiffening
lar [17]. Fig. 6 indicates a slight increase in the relative concentra- effect with increasing degree of physico–chemical interaction
tion of bees for the lime mastic compared to the base binder. indicates that the role of the adsorbed layer significantly affects
Results in Figs. 5 and 6 also indicate a decrease in the relative mastic rheology.
concentration of the paraphase. Note that three replicate AFM In addition to comparing the master curves of different binders
images were collected and analyzed for each material studied. and mastic, the effect of temperature on the stiffening effect of the
The largest change in the concentration of the paraphase occurs fillers was also investigated through evaluating |G⁄| ratios at each
in the lime mastic, which is consistent with expectations given temperature of frequency sweep testing at a single loading fre-
its relatively high specific surface area. However, granite has a quency of 10 Hz. |G⁄| ratios are computed as the ratio of the mastic
lower specific surface area than PC but leads to a greater reduction modulus to the base asphalt. Results are presented in Fig. 8. Results
in the concentration of paraphase domain so trends are not demonstrate little change in stiffening ratios with temperature for
entirely consistent and merit further investigation. No significant granite and PC mastics. However, for the lime mastic, the stiffening
change in the relative concentration of the periphase domain was ratio is found to increase with temperature, indicating potential
observed to occur as a result of physico–chemical interaction. influence of particle structuralization, which could be induced as
Note that based on Allen et al.’s [18] study, it was found that the a result of an increase in ‘‘effective’’ filler content due the forma-
paraphase is the softest microdomain and the periphase is the stiff- tion of a nontrivial semi-rigid adsorbed interphase layer on filler
est. Thus, one might infer that a loss in paraphase concentration particles, further displaying physico–chemical interaction intensity
would indicate an overall increase in macroscopic stiffness. compared to other mastics [11].
C. Davis, C. Castorena / Construction and Building Materials 94 (2015) 83–89 87

Fig. 4. Illustration of image processing output.

Binder Granite Mastic

PC Mastic Lime Mastic

Fig. 5. AFM images. For each material, left image corresponds to topographic and right image corresponds to phase.

4. Conclusions & recommendations (3) Changes in binder microstructure provide an indicator of


physico–chemical interaction intensity. Mastics that exhib-
(1) Effective (i.e., non-adsorbed) binder in mastics has signifi- ited the most significant change in microstructure upon filler
cantly different microstructure than the bulk binder as a addition also demonstrated the greatest change in rheology.
result of physico–chemical interactions. (4) A greater stiffening effect is observed for fillers which induce
(2) The extent of change in binder microstructure as a result of greatest microstructural changes as a result of physico–
physico–chemical interaction is affected by the filler specific chemical interaction. Physico–chemical interaction is antici-
surface area. pated to lead to a softening of the effective asphalt binder
88 C. Davis, C. Castorena / Construction and Building Materials 94 (2015) 83–89

Fig. 6. Effect of physico–chemical interaction on different microdomains.

Fig. 7. Dynamic shear modulus (|G⁄|) master curves.

Fig. 8. |G⁄| Ratios at 10 Hz loading frequency.

matrix since polar components adsorb to the surface of (5) It is should be noted that findings of this study are limited to
the filler. Thus, results suggest the formation of adsorbed a single binder blended with three fillers at a single volumet-
interphase layer plays a more critical role in determining ric concentration. It is recommended that findings be con-
macroscopic rheology of mastics than softening of the firmed by studying additional binder – filler combinations
binder matrix. using a similar methodology. It is also recommended that
C. Davis, C. Castorena / Construction and Building Materials 94 (2015) 83–89 89

nanoindentation experiments be performed to determine [5] Clopotel C, Velasquez RA, Bahia HU. Measuring physico–chemical interaction
in mastics using glass transition. Road Mater Pavement Des 2012;13(Suppl.1):
the microrheology of the different binder microphases with
304–20.
and without inclusion of filler to allow for improved under- [6] Underwood BS, Kim YR. Experimental investigation into the multiscale
standing of the implication of physico–chemical interaction behavior of asphalt concrete. Int J Pavement Eng 2011;12(4):357–70.
on the rheology of the modified binder matrix. It is also [7] Rigden PJ. The use of fillers in bituminous road surfacings: a study of filler–
binder systems in relation to filler characteristics. J Soc Chem Ind Lond
recommended that AFM imaging be conducted at varying 1947;66(9):299–309.
temperatures to investigate thermal effects on binder [8] Heukelom W. The role of filler in bituminous mixes. J Assoc Asphalt Paving
morphology. Technol 1965;34:396–429.
[9] Anderson DA, Goetz WH. Mechanical behavior and reinforcement of mineral
(6) The results of this study were limited to morphological and filler–asphalt mixtures. J Assoc Asphalt Paving Technol 1973;42:37–66.
linear viscoelastic characterization of asphalt mastics. It is [10] Craus J, Ishai I, Sides A. Some physico-chemical aspects of the effect and role of
recommended that future work also address the effects of the filler in bituminous paving mixtures. J Assoc Asphalt Paving Technol
1978;46:558–88.
physico–chemical interactions between asphalt binder and [11] Buttlar W, Bozkurt D, Al-Khateeb G, Waldhoff A. Understanding asphalt mastic
filler on chemistry and molecular structure. It is also sug- behavior through micromechanics. In: Transportation research record, no.
gested that performance implications of interactions 1681; 1999. pp. 157–169.
[12] Underwood BS, Kim YR. Microstructural association model for upscaling
between filler and binder further be investigated with prediction of asphalt concrete dynamic modulus. J Mater Civil Eng 2013;25:
respect to viscoplasticity, fracture resistance, and adhesion 1153–61.
to aggregate. [13] Tan Y, Guo M. Interfacial thickness and interaction between asphalt and
mineral fillers. Mater Struct 2014;47:605–14.
[14] Lesuer D. The colloidal structure of bitumen: consequences on the rheology
and on the mechanisms of bitumen modification. Adv Colloid Interface Sci
2009;145:42–82.
Acknowledgments [15] Loeber L, Sutton O, Morel J, Valleton JM, Muller G. New direct observations of
asphalts and asphalt binders by scanning electron microscopy and atomic
force microscopy. J Microsc 1996;182(1):32–9.
This research was funded by North Carolina State University’s [16] Masson JF, Leblond V, Margeson J. Bitumen morphologies by phase-detection
Faculty Research and Professional Development Program (FRPD). atomic force microscopy. J Microsc 2006;221(1):17–29.
This support is greatly appreciated. [17] Pauli AT, Grimes RW, Beemer AG, Turner TF, Branthaver JF. Morphology of
asphalts, asphalt fractions and model wax-doped asphalts studied by atomic
force microscopy. Int J Pavement Eng 2011;12(4):291–309.
[18] Allen RG, Little DN, Bhasin A, Lytton RL. Identification of the composite
References relaxation modulus of asphalt binder using AFM nanoindentation. J Mater Civil
Eng 2013;25:530–9.
[1] Underwood BS, Kim YR. Microstructural investigation of asphalt concrete for [19] Tan Y, Guo M. Micro- and nano-characterization of interaction between
performing multiscale experimental studies. Int J Pavement Eng asphalt and filler. ASTM J Test Eval 2014;42(5):1–9.
2013;14(5):498–516. [20] Nazzal MD, Kaya S, Gunay T, Ahmedzade P. Fundamental characterization of
[2] Asphalt Institute. Superpave mix design. Superpave series no. 2 (SP- asphalt clay nanocomposites. J Nanomech Micromech 2013;3(1):1–8.
02). Lexington, KY: Asphalt Institute; 2001. [21] Fischer H, Poulikakos LD, Planche JP, Das P, Grenfell J. Challenges while
[3] Bahia HU, Faheem AF, Hintz C, Yang SH, Al-Qadi I, Reinke G, Glidden S. Test performing AFM on bitumen. In: Kringos N. et al., editor. Multi-scale modeling
methods and specification criteria for mineral filler used in HMA. In: National characterization of infrastructure materials, RILEM 8; 2013. pp. 89–98.
cooperative highway research program final report. National Research Council, [22] Das PK, Jelagin D, Birgisson B, Kringos N. Micro-mechanical investigation of
Washington DC; 2010. low temperature fatigue cracking behaviour of bitumen. In: 7th RILEM
[4] Kim M, Buttlar W. Stiffening mechanisms of asphalt – aggregate mixtures. In: international conference on cracking in pavements, RILEM book series. vol.
Transportation research record, no. 2181; 2010. pp. 98–108. 4; 2012. pp. 1281–1290.

You might also like