You are on page 1of 9

Computational Materials Science 112 (2016) 161–169

Contents lists available at ScienceDirect

Computational Materials Science


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

Study of cohesion and adhesion properties of asphalt concrete with


molecular dynamics simulation
Guangji Xu, Hao Wang ⇑
Department of Civil and Environmental Engineering, Rutgers University, Piscataway, NJ 08854, USA

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

Article history: The objectives of this study is to develop a molecular modeling approach for studying cohesive and
Received 26 June 2015 adhesive properties of asphalt concrete and evaluate the accuracy of modeling through comparisons with
Received in revised form 21 September experimental data. Fully atomistic models were built for molecular dynamics (MD) simulation consider-
2015
ing two representative asphalt models and two types of aggregate mineral. MD simulations were per-
Accepted 20 October 2015
Available online 19 November 2015
formed to study thermodynamic and cohesive properties of asphalt binder, such as density, solubility
parameter, cohesive energy density, and surface free energy. The adhesion properties were investigated
by calculating the interaction energy and the work of adhesion at asphalt–aggregate interface for the first
Keywords:
Asphalt–aggregate interface
time. The bond energy parameters in dry and wet conditions were used to evaluate moisture sensitivity
Molecular dynamics simulation of interface adhesion. The results show that van der Waals force plays critical role for cohesive properties
Interaction energy of asphalt binder; while the adhesion bonding between asphalt to aggregate is largely dependent on the
Work of adhesion type of aggregate mineral (silica or calcite) in both dry and wet surface conditions. The effect of asphalt
Moisture susceptibility type was found significant for the adhesion between asphalt and silica at the relatively small moisture
content. The simulation results agree well with experimental measurements reported in the literature.
This work illustrates MD can help in understanding fundamental chemo-mechanics relationship of
asphalt concrete at an atomistic scale, which can be used as a useful tool for material design and perfor-
mance prediction.
Ó 2015 Elsevier B.V. All rights reserved.

1. Introduction mental theories including mechanical theory, chemical theory,


weak boundary theory, and thermodynamic theory have been pro-
Asphalt concrete has been widely used for roadway pavements posed [2]. Meanwhile, moisture effect has been found being the
due to its ability to provide structural capacity and smooth surface. major driving force causing the loss of adhesion between asphalt
In asphalt concrete, aggregate forms the skeleton of mixture; while and aggregate and further deterioration [3].
asphalt binder serves as the primary binding material. However, A number of studies have been conducted to evaluate adhesion
several types of pavement distresses, such as premature rutting, properties of the asphalt–aggregate interface and moisture suscep-
raveling, and cracking are commonly observed during the early life tibility of asphalt concrete using bond strength measurements
of asphalt pavements. from experimental and phenomenological engineering aspects
It has been found that three factors are responsible for the fail- [4–7]. Nanoscale experiments have also been used to measure sur-
ure of asphalt concrete, including cohesion loss within asphalt, face free energy and adhesive/cohesive force–displacement rela-
strength reduction of aggregate particles, and breakdown of adhe- tionships between asphalt molecular groups and aggregates at
sive bonding between aggregate and asphalt [1]. Cohesion proper- dry and wet conditions [8–11]. These experimental studies have
ties of asphalt binder and adhesion properties of asphalt to produced acceptable and reasonable results for better understand-
aggregate (i.e. bonding strength) are largely dependent on chemi- ing cohesion and adhesion properties of asphalt concrete.
cal compositions of asphalt and aggregate. To better understand Recently, atomistic modeling with molecular dynamic (MD)
the adhesive bond between asphalt and aggregates, four funda- simulation has been employed to study the link between chemical
structures and physical and mechanical behavior of asphalt con-
crete and durability, such as density, thermal expansion coefficient,
⇑ Corresponding author at: Department of Civil and Environmental Engineering,
and viscosity [12], adhesive strength [13], aging effect [14], curing
Rutgers, The State University of New Jersey, Piscataway, NJ 08854, USA.
Tel.: +1 848 4452874. [15] and self-healing [16]. Molecular dynamics simulation is a pow-
E-mail address: hwang.cee@rutgers.edu (H. Wang). erful and feasible approach for material design and performance

http://dx.doi.org/10.1016/j.commatsci.2015.10.024
0927-0256/Ó 2015 Elsevier B.V. All rights reserved.
162 G. Xu, H. Wang / Computational Materials Science 112 (2016) 161–169

prediction because the atomistic structures of asphalt and aggre- averages of a system, as shown in Eq. (1). Molecular dynamics sim-
gate and their interfaces determine the behavior of asphalt concrete ulations require the definition of a potential function, or a descrip-
at the continuum scale. The significant advantage of MD simulation tion of the terms by which the particles in the simulations will
is to provide fundamental observations in the initiation and evolu- interact. The force on an atom i can be directly calculated by the
tion of material damage at the nanoscale that are very difficult to derivative of the potential energy, E, with respect to the coordinate
obtain from the traditional experiments conducted at the ri.
laboratory.
While a number of studies have used MD simulation to study @E @ 2 ri
 ¼ mi 2 ð1Þ
asphalt concrete from a more fundamental and nature perspective, @ri @t
little work has been focused on deriving cohesion and adhesion
where mi is the mass of atom i. In a system consisting of N atoms, all
properties of asphalt concrete from the atomistic scale to the con-
pairs can result in N(N  1)/2 interactions, that is to say, even mod-
tinuum scale. Limitations exist for in-depth investigation of mois-
erate numbers of atoms N will generate a very large number of
ture susceptibility of asphalt concrete at the nanoscale. The
calculations.
interaction between asphalt and aggregate is governed by various
attractive intermolecular and intramolecular forces across the
3.2. Force field
interface. This necessitates investigation of the fundamental cohe-
sion and adhesion properties of asphalt concrete using a system-
Force field, sometimes used as potentials, refers to a mathemat-
atic atomistic modeling approach.
ical function used to compute the energies of a system of atoms
with various system conformations in the context of molecular
2. Objective and scope modeling. In general, the force field parameters can be obtained
from the experiments, together with quantum mechanical calcula-
The objectives of this study is to develop an atomistic modeling tions. The goal of a force field is to describe entire classes of mole-
approach for studying cohesive and adhesive properties of asphalt cules with reasonable accuracy. Basically, the total potential
concrete and evaluate the accuracy of modeling through compar- energy E can be comprised of a number of bonded and non-
isons with experimental data. In order to achieve this objective, bonded interaction terms, as shown in Eq. (2).
fully atomistic models were built for MD simulation considering
two representative asphalt models and two types of mineral aggre- E ¼ Ebonded þ Enon-bond ð2Þ
gate. Bulk asphalt models and confined layer models were used to where the bonded terms, Ebonded is for atoms interactions con-
derive thermodynamic material properties of asphalt binder, tributed by covalent bonds including 0062ond stretching, angle
including density, surface free energy, cohesive energy density, bending and dihedral and improper interaction, etc., while the
and solubility parameter. The work of adhesion between asphalt non-bond interaction term describes non-covalent contributions
binder and aggregate were derived from interaction energy with that mainly contain van der Waals energy, Coulomb electrostatic
different definitions of surface area. The effects of chemical struc- energy, and hydrogen bond energy. A number of force fields have
tures of asphalt and aggregate and moisture contents on the adhe- been developed over the years. The energy terms describing differ-
sion between asphalt and aggregate were investigated. ent kinds of deformations were added or changed in Eq. (2) for dif-
The fact that material failure originates from particle dislocation ferent force fields. In all cases using an appropriately parameterized
and bond rupture at the nanoscale necessitates the importance of force field is an important issue in molecular simulation.
atomistic modeling approach in material design. The physical and Molecular dynamics (MD) simulations were performed using
chemical compatibility of asphalt binder and aggregate affects fati- commercially available simulation software, Materials Studio
gue cracking and moisture damage in asphalt concrete. Molecular [18]. COMPASS II (Condensed-Phase Optimized Molecular Poten-
dynamics simulations can provide valuable insights in understand- tials for Atomistic Simulation Studies) force field was used for
ing the interaction between asphalt and aggregate at the nanoscale describing atom-level interactions in molecular models. COMPASS
and relate chemical compositions of asphalt concrete to its macro- is the first parameterized and validated ab initio-based force field
scopic behavior. Currently, understanding how asphalt concrete that has a broad coverage in covalent molecules including most
deforms and breaks under mechanical and environmental loading common organics, small inorganic molecules, and polymers. COM-
is often limited to phenomenological engineering approaches, PASS II is a significant development extension to the COMPASS
neglecting the underlying failure mechanism in the atomistic struc- force field in terms of atom types and force field terms, which
ture. The nanoscale details of material failure are still not clear and enable us to make accurate predictions material properties for a
the experiments are time-consuming and expensive. The atomistic wide range of compounds in isolation and in condensed phases
approach developed from this study laid the ground work to predict [19]. The more realistic the potential is, the closer the MD simula-
failure potential of asphalt concrete in a computational testing tions results represent the real properties of materials. The amount
environment. This will lead to better design of asphalt mixture for of time needed for MD simulations is largely dependent on the
road pavements with longer service life and less repair. functional complexity of potential.

3. Molecular dynamics simulation 3.3. Molecule models for asphalt

3.1. Basic principle Asphalt material is a complex chemical mixture of molecules


that are predominantly hydrocarbons with a small amount of
Molecular dynamics simulation is a technique for calculating structurally analogous heterocyclic species and functional groups
the equilibrium and transport properties of a classic many-body containing sulfur, nitrogen and oxygen atoms. From solubility
system [17], in which the motion of the constituent particles obeys point of view, asphalt materials are composed of three main con-
the laws of classical mechanics. It computes the motions of a num- stituents, i.e., asphaltene that is the most viscous and polar compo-
ber of atoms and molecules in a system as a function of time. The nents, saturates, consisting of aliphatic molecules and are least
basis of MD simulation is Newton’s law of motions and statistical viscous and non-polar, and resin whose properties are in between
mechanics, where statistical ensemble averages are equal to time of above two. Due to the complex mixture nature of asphalt, many
G. Xu, H. Wang / Computational Materials Science 112 (2016) 161–169 163

efforts were made to construct model asphalt mixtures with liquids and polymeric systems. In this study a cubic cell with an
reasonable compositions. The average molecular structures for initial density of 0.1 g/cm3 in 3D periodical condition was built at
different asphalt binders were proposed based on Nuclear first, with a purpose to randomly distribute three types of mole-
Magnetic Resonance (NMR) spectra [20]. These average asphalt cules and prevent the molecule chains twisted with each other.
molecular models can achieve proper element concentrations After a geometry optimization process, an equilibrium run with
and aromatic/aliphatic ratios, but it cannot reflect the molecular constant volume and temperature (NVT) for 200 ps with a time
packing between different molecular phases and thus accurately step of 1 fs was performed to pre-equilibrate the system, which
capture the complexity of asphalt composition. brought the asphalt system from an initial state to a more equili-
In order to build a model more similar to the real asphalt, the brated state at the target temperature. Then another molecular
diversity of chemistries and polarities are required. Since the focus dynamic run with isothermal–isobaric (NPT) ensemble was per-
of this study is not to determine the representation of molecular formed for 300 ps of relaxation time to shrink the system volume
structure for asphalt binder, an existing three-component model and to approach a stable state of the real density. During this pro-
asphalt system was used [21]. In the three-component model, cess, Nose–Hoover–Langevin (NHL) thermostat and Andersen
n-C22 molecules were used to represent saturates and 1,7- barostat were applied for system to maintain a target temperature
dimethylnaphthalene was used to represent resin (naphthene aro- and pressure, respectively. The relaxation time used in the NVT and
matics). Two different asphaltene molecules were proposed, one is NPT processes was proved to be enough because it was observed
a moderate-size aromatic core with very small branches [22], and that the temperature and density reached the stable values in
the other is smaller aromatic core and much longer alkane side the process. The time step of 1 fs was selected considering the bal-
branches [23]. This relative simple model has been proved accept- ance between accuracy of simulation results and simulation time
able to predict physical and rheological properties of asphalt bin- [21].
der in a general sense [12]. Fig. 1 shows the molecular structures After the density results were stabilized, an additional NVT of
of two different asphaltene models. Table 1 shows the molecular 100 ps was performed to further equilibrate the model. The final
compositions of asphalt mixtures with two different asphaltene structure of asphalt model is shown in Fig. 2(a). One important fea-
models. The mass fraction was selected to resemble the total car- ture of the asphalt model is that molecules occupy not only within
bon/hydrogen ratio reported by previous researchers [24,25]. the cell but also across the cell boundaries due to the periodicity. A
colloidal asphalt structure can be observed in which asphaltene
3.4. Construction of bulk asphalt model molecules were dispersed in resins and saturates. Fig. 2(b) shows
the microstructure of asphalt bulk model with distributions of free
Amorphous cell module in Materials Studio was employed to and substantial volumes. The dispersion of asphaltene molecules in
construct three-dimensional (3D) periodic structures of molecular the maltene molecular network causes molecular motion that is

Fig. 1. Molecular structures of (a) asphaltene 1; (b) asphaltene 2; (c) 1,7-dimethylnaphthalene; and (d) n-C22 (Carbon atoms are shown in grey, sulfur atoms in yellow, and
hydrogen atoms in white). (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)
164 G. Xu, H. Wang / Computational Materials Science 112 (2016) 161–169

Table 1
Molecular composition of asphalt model (after [21,24,25]).

Asphalt model Asphaltene-1/2 1,7-Dimethyl naphthalene n-C22


Molecule number Mass fraction Molecule number Mass fraction Molecule number Mass fraction
Asphalt-1 5 20.7 27 19.7 41 59.6
Asphalt-2 5 21.1 30 19.8 45 59.1

Fig. 2. Illustration of asphalt-1: (a) 3D periodic molecular model showing arrangements of asphaltene 1 (red), resins (yellow), and saturates (green); and (b) cell volumetric
data showing substantial volumes (gray) and free volumes (blue). (For interpretation of the references to color in this figure legend, the reader is referred to the web version of
this article.)

temperature dependent. The free volumes provide spaces for dioxide (or silica) and calcium carbonate (or calcite) are two most
molecular movements between different asphalt components and common chemical compounds. For example, limestone consists
thus influence thermodynamic properties of asphalt. primarily of calcium carbonate, while silicon dioxide is common
in sand and granite with a high percentage. Therefore, in this study,
silica (SiO2) and calcite (CaCO3) were used to represent ideal aggre-
3.5. Construction of confined asphalt layer
gate in molecular dynamics simulation.
Confined layers of asphalt binders were built for the
asphalt–aggregate interface model. The only difference between
3.7. Creation of asphalt–aggregate Interface
the bulk asphalt model and the confined asphalt layer lies in the
periodical conditions in the z direction. In the confined layer, the
The confined asphalt layer was attached to the aggregate sub-
box boundary in the z direction functions as hard repulsive wall,
strate to form an asphalt–aggregate interface model. At first, a unit
and if asphalt atoms move close to the z boundary during model
crystalline silica, with lattice parameters of a = b = 4.913 Å,
construction, the atoms will rebound back into the box. Compared
c = 5.405 Å, a = b = 90°, and c = 120°, was imported from structures
to the bulk asphalt model, confined asphalt layer had a relatively
database in Materials Studio. As for the calcite, a unit cell with
flat surface in the z direction, which was used for the asphalt–
dimension of a = b = 4.990 Å, c = 17.061 Å, a = b = 90°, and
aggregate interface model as described below. For the purpose of
c = 120° was used. It is believed that during the simulations, the
calculating surface free energy, a free surface was created by arbi-
orthogonal geometry was preferred for the purposes of model con-
trarily elongating the dimension of the cell in the z direction and
struction and further calculation. Therefore, the unit silica or cal-
equilibrating the periodic system in isometric–isothermal (NVT)
cite cell were first cleaved in [0, 0, 1] direction, and then followed
conditions. The illustration of bulk asphalt and the confined
by geometry transformation to have an orthogonal shape with P1
asphalt layer with free surface is shown in Fig. 3.
space group, in which a = b = c = 90°. By repeating the unit cell in
x and y directions to generate the same size as asphalt models, a
3.6. Molecular models of mineral aggregate 2-D bulk silica or calcite surface was created. After that, a vacuum
slab was added to form silica or calcite blocks with 3D periodic
A mineral is defined as an element or chemical compound that boundary conditions. Confined asphalt layers were then placed
is normally crystalline and resulted from geological processes. A over the silica or calcite blocks. A vacuum of 30 Å was added above
crystal structure is the orderly geometric spatial arrangement of to avoid the interaction across the mirror image in the z direction.
atoms in the internal structure. This crystal structure is based on The final bi-material interface model between asphalt and aggre-
regular internal atomic or ionic arrangement in the geometric gate was built as shown in Fig. 4.
form. Chemistry and crystal structure together define a mineral. Once the interface model was obtained, geometry optimization
Crystal structure greatly influences the physical properties of for 5000 iterations was carried out followed by a dynamic equili-
mineral. Mineral aggregates that are widely used in construction bration run of 200 ps with canonical ensemble (NVT), to make sure
materials include sand, granite, limestone, quartz, etc. Among all that the model configuration was further optimized. An additional
the mineral aggregate materials used in asphalt concrete, silicon 50 ps NVT with 50 frames was performed for a product run with
G. Xu, H. Wang / Computational Materials Science 112 (2016) 161–169 165

a b

(a) (b)
Fig. 3. Asphalt thin film formation for surface energy calculation: (a) bulk model and (b) confined asphalt layer with free surface.

at the interface and the diffusion of water through asphalt binder


and aggregate adsorption.
In MD simulation, the three-phase interface models were geom-
etry optimized first to eliminate the highest energy overlap and
then followed by a NVT equilibration for 100 ps to dissipate high
energies and optimize the configuration. An additional 50 ps NVT
run with 50 frames was then used for energy calculation. Unless
otherwise stated, all MD simulations for interface interaction were
performed at 298 K with a time step of 1 fs in this study. A cutoff
distance of 10.5 Å was adopted to speed up the computation. This
means that the interactions between two atoms separated by a dis-
tance greater than this pre-defined cutoff distance are ignored.

4. Results and discussion

4.1. Density of asphalt

Density is an important thermodynamic property and it is a


direct indicator to see whether the model size and force field could
yield reasonable and accurate simulation results. Therefore, density
Fig. 4. Asphalt-1/aggregate interface atomistic model: (a) silica and (b) calcite.
(Carbon atoms are grey, sulfur atoms are yellow, hydrogen atoms are white, red is
values at different temperatures were calculated for asphalt-1 and
oxygen, orange is silicon, and green is calcium.) (For interpretation of the references asphalt-2 models, respectively. With the simulation process stated
to color in this figure legend, the reader is referred to the web version of this above, the density results of asphalt models are shown in Fig. 5. The
article.) results show that the highest predicted density is 0.92 g/cm3
at 10 °C and the densities reduces to 0.85–0.88 g/cm3 at 60 °C.
As expected, a steady increase of density with decreasing tempera-
the subsequent calculation purposes. Trajectory files after molecu-
ture was observed. The density predictions are slightly smaller than
lar dynamic simulations were used for calculation with statistical
the reported experiment data from 0.99 to 1.33 g/cm3 at 60 °C [27].
mechanics. A modified script based on Perl language was devel-
This is properly because the higher hydrogen, more carbon, and less
oped to calculate the interaction energy between asphalt and
sulfur were used in the molecular model of asphalt, together with a
aggregate.
lack of heteroatoms besides sulfur [21]. An improved model asphalt

3.8. Consideration of moisture effect

Based on the asphalt–aggregate interface model, the moisture


effect was simulated by adding water molecules between asphalt
and aggregate to form a three-phase interface system. This was
assumed to represent the final stage of the moisture diffusion pro-
cess at the asphalt–aggregate interface, which was regarded as the
major mechanism for moisture-induced damage in asphalt con-
crete [26]. Three different numbers (100, 150, and 200) of water
molecules were considered to represent different levels of mois-
ture contents, which correspond to 7–14% mass fraction of asphalt.
The diffusion process and distribution pattern of water molecules
will affect simulation results and experimental observations.
Future should be conducted to simulate the asphalt–aggregate
interfacial behavior considering the moisture distribution pattern Fig. 5. Comparison of density results for two asphalt models.
166 G. Xu, H. Wang / Computational Materials Science 112 (2016) 161–169

system with 12 chemical components has been proposed recently results agree well with the measured values reported in the litera-
to provide more accurate density predictions [28], which will be ture, although the simulation results were slightly greater than
used in future study. most experimental results. The discrepancies could be caused by
the approximation of asphalt molecular models. Previous experi-
4.2. Surface free energy of asphalt mental measurements have shown that the concentration of satu-
rates or asphaltene could lower the value of surface free energy,
Surface free energy is theoretically defined as the magnitude of while the resins and the ratio of resin/asphaltene could increase
work required to create a unit area of a new surface of the material the value of surface free energy [33]. This indicates that further
under vacuum conditions. From an atomistic point of view, mole- study is needed to evaluate sensitivity of surface free energy of
cules in the bulk material are surrounded by other molecules from asphalt with different chemical compositions and elemental
all sides; the surface free energy equals the work needed to extract contents.
the molecules from the bulk and create a new area of surface mole-
cules. Therefore, surface free energy can be used to measure the 4.3. Cohesive energy density and solubility parameter of asphalt
disruption of intermolecular interaction that occurs when a surface
is created. Surface free energy has been proposed for selecting the Cohesive energy density (CED) is used as a measure of the
combination of asphalt binders and aggregate types that have the mutual attractiveness of molecules to assess intermolecular inter-
necessary compatibility to form strong bonds and resist moisture action inside an asphalt molecule model [36]. A solubility parame-
damage [29] and it was found related to the cohesive bond energy ter can be further defined from the CED using Eq. (5). The solubility
of asphalt materials [8]. parameter quantifies the solvent–solute interactions to separate
In this study, the surface free energy was calculated as the dif- solvent-solvent and solute–solute molecules as a means of mini-
ference between the potential energies of the thin asphalt layer mizing the Gibbs free energy of the entire system [37]. It has been
and its corresponding bulk asphalt divided by the surface area cre- used to study molecular attractions properties in the asphalt sys-
ated upon formation of the thin layer, as shown in Eq. (3) [30,31]. tem and new asphalt models and evaluate colloidal stability in
This definition calculates the contribution of internal potential asphalt [38,39]. Since the intermolecular interaction within asphalt
energy to the surface free energy but cannot consider the entropic is evaluated via van der Waals and electrostatic (including hydro-
contribution due to deformation and segregation of molecular gen bond) interactions, solubility parameter can be expressed in
chains to the surface energy. The Connolly surface area was used two components, as shown in Eq. (6).
to consider the surface unevenness due to molecular agglomera- pffiffiffiffiffiffiffiffiffiffi
tion in asphalt models. The Connolly surface area consist of all d¼ CED ð5Þ
the points of the van der Walls surface at which a solvent sphere qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
can touch and it is also known as accessible surface area [32]. d¼ d2v dw þ d2ele ð6Þ
The work of cohesion (Waa) can be calculated from the surface
free energy (ca) using Eq. (4), which is defined as the work required where d is solubility parameter; CED is cohesive energy density;
to separate a unit area of a liquid or solid into two unite areas. The and dvdw and dele are representing the contributions from van der
work of cohesion is used as an indicator for cohesion properties of Waals forces and electrostatic interactions, respectively.
asphalt when the failure within the asphalt is concerned. Table 3 shows the calculated CED and solubility parameter of
ca ¼ ðEfilm  Ebulk Þ=2A ð3Þ bulk asphalt model. The values of CED and solubility parameter
were found in good agreement between simulation results and
W aa ¼ 2ca ð4Þ previous results from experiments or simulations. It was found
that for solubility parameter values, van der Waals component
where ca is surface free energy; Waa is work of cohesion; Efilm and was much greater than electrostatic component. This means that
Ebulk are the potential energy of the confined asphalt layer and bulk cohesion is largely dependent on van der Waals forces which held
asphalt, respectively; and A is area of the new surface to be created. the molecules of asphalt binder together. On the other hand, the
Table 2 summarizes the calculated surface free energy and work results show that the asphalt-1 model has slightly greater values
of cohesion for two asphalt models. The difference between the of CED and solubility parameter than those of the asphalt-2 model.
dimensional area and the real surface area of asphalt model can It indicates that asphalt-1 has stronger mutual attractive interac-
be clearly observed. It can be seen that in general the calculated tions among its molecules fractions, and presents better cohesion

Table 2
Surface free energy and work of cohesion of asphalt.

Properties Asphalt-1 Asphalt-2 Experiment values


Dimensions (Å) 34.11  34.11  34.11 35.64  35.64  35.64 N/A
Surface area of new surface (Å2) 2258 2604 N/A
Surface free energy ca (mJ/m2) 38.81 37.01 13–47.6 [33,45]
Work of cohesion Waa (mJ/m2) 77.62 74.02 27–65 [34,35]

Table 3
Cohesive energy density and solubility parameter of bulk asphalt.

Properties Asphalt-1 Asphalt-2 Experiment or previous simulation values


CED (108 J/m3) 3.22 3.08 3.19–3.22 [40]
Solubility parameter d ((J/cm3)1/2) 17.95 17.56 13.30–22.50 [40,41]
Electrostatic d ((J/cm3)1/2) 1.11 0.90 N/A
Van der Waals d ((J/cm3)1/2) 17.92 17.54 N/A
G. Xu, H. Wang / Computational Materials Science 112 (2016) 161–169 167

properties as compared to asphalt-2. This finding is consistent with Although different theories (physical, chemical or mechanical)
the work of cohesion of two asphalt models that is derived from have been proposed to explain the adhesion between asphalt bin-
surface free energy as shown in Table 2. der and aggregate, the actual mechanism is not fully explained. The
Strategic Highway Research Program (SHRP) studies found that the
polar constituents in the asphalt adhere to the aggregate surface
4.4. Work of adhesion between asphalt and aggregate
through electrostatic forces, hydrogen bonding, or Van der Waals
interactions. Several different methods of measuring the energy
The work of adhesion is described as the work required to sepa-
of adsorption indicated that physisorption rather than chemisorp-
rate asphalt from aggregate at the interface. The work of adhesion
tion is occurring [1].
between asphalt and aggregate relates significantly to the resis-
Fig. 7 shows the calculated work of adhesion between different
tance of asphalt mixture to fracture-related failure behavior and
types of asphalt and aggregate models. Three definitions of inter-
durability. The work of adhesion between asphalt and aggregate
face areas, including the layer dimensional area, the Connolly area
can be calculated using Eq. (7), with the interaction energy (Einter)
of asphalt surface, and the Connolly area of aggregate surface, were
derived from Eq. (8). This approach has been used in the literature
used in the calculation. It was found that the selection of interface
to successfully calculate work of adhesion between polymer–metal,
contact area could cause significant variations in calculating the
polymer–polymer and more general solid–liquid systems [42,43],
work of adhesion from MD simulation. It is believed that the actual
which considers the pure physicochemical interaction between
interface contact area is dependent on the morphological feature of
two interacting materials. Based on the definition in Eqs. (7) and
asphalt and aggregate, which could be larger than the layer dimen-
(8), a negative value of Einter or Wadhesion indicates attraction
sional area and stay between the rough surface areas of asphalt and
between these two components, while a positive value of Einter or
aggregate. In general, the range of work of adhesion values
Wadhesion indicates repulsion. The calculated Einter or Wadhesion from
calculated using the Connolly area agree well with the experimen-
MD simulation are negative values and the absolute values were
tal results reported in the literature, which range from 60 to
used in the analysis for indicating the attraction between asphalt
150 mJ/m2 with different sources of asphalt and aggregate [35,44].
and aggregate.
The work of adhesion results show that for the same type of
W adhesion ¼ DEinter aggregate, the difference between the work of adhesion of
aag =A ð7Þ
asphalt-1 and asphalt-2 models is small. On the other hand, the
effect of aggregate mineral on the work of adhesion is much more
DEinter aag ¼ Etotal  ðEasphalt þ Eaggregate Þ ð8Þ obvious. In particular, for the same type of asphalt, the work of
adhesion between asphalt and calcite is much greater than that
where DEinter_aag is interaction energy between asphalt and aggre-
between asphalt and silica. Therefore, calcite is more prone to
gate; Etotal is total potential energy of asphalt and aggregate in con-
adsorb asphalt than silica at dry condition. This is in agreement
tact at equilibrium; and Easphalt and Eaggregate are the potential
with the experimental finding that a mineral with an alkaline
energies of asphalt and aggregate (i.e., silica and calcite) separated
(basic) nature such as calcite, results in a stronger adhesion with
in vacuum at equilibrium, respectively; Wadhesion is work of adhe-
asphalt, compared to quartz (acidic) with a rather acidic nature [5].
sion between asphalt and aggregate; and A is the interface contact
area between asphalt and aggregate.
Fig. 6 shows the calculated interaction energy between asphalt 4.5. Moisture effect on asphalt–aggregate adhesion
and aggregate and the contributions of non-bond components (van
der Waals and electrostatic energy). The results indicate that the Due to the hydrophobic nature of asphalt binder, water has a
interaction energy between asphalt and aggregate varies depend- detrimental impact on the bonding between asphalt and aggregate.
ing on the combination of material types. The values of interaction The energy associated with the displacement of asphalt by water
energy were found equal to the summation of van der Waals and from the asphalt–aggregate interface or debonding is referred to
electrostatic components, which indicated that adhesion between as the work of debonding, as expressed in Eq. (9) [45]. The calcu-
asphalt and aggregate was mainly the non-bond interaction. This lated values of the work of debonding were found negative that
is because no chemical bond was formed during the interaction was consistent with experimental findings. This indicates that
and therefore covalent components were negligible. On the other there exists a thermodynamic potential for water to disrupt the
hand, it was found that van der Waals energy constituted the asphalt–aggregate interface. In other words, no external energy is
major contribution to the non-bond energy terms. The Coulomb required to be added into the system in order to separate the
interaction was found negligible in the asphalt–silica interaction asphalt–aggregate interface due to the water-loving nature of
and serving as a minor role in the asphalt–calcite interaction. aggregate.

Fig. 7. Work of adhesion between asphalt and aggregate predicted by MD


Fig. 6. Interaction energy and its component predicted by MD simulation. simulation.
168 G. Xu, H. Wang / Computational Materials Science 112 (2016) 161–169

W debonding water ¼ ðDEinter aw þ DEinter agw  DEinter aag Þ=A ð9Þ plicated than the pure silica and calcite. Previous research has
found that ER values higher than 1.5 indicates that asphalt mixture
where Wdebonding_water is work of debonding when water displaces
has relatively high resistance to moisture damage; while asphalt
asphalt from the asphalt–aggregate interface; DEinter_aw is the inter-
mixtures with high susceptibility to moisture damage are related
action energy between asphalt and water; DEinter_agw is the interac-
to ER values smaller than 0.5 [47].
tion energy between aggregate and water; DEinter_aag the interaction
The results show that large differences were observed for the ER
energy between asphalt and aggregate; and A is the interface con-
values when the same type of asphalt interacts with silicate and cal-
tact area.
cite. This indicates that moisture susceptibility is strongly depen-
The energy ratio (ER) has been used in previous experimental
dent on aggregate chemistry. In particular, silica shows much less
studies to evaluate moisture effect on the adhesion between
water susceptibility than calcite. It is noted that mixed findings
asphalt and aggregate, which is based on the hypothesis that mois-
have been reported for the moisture susceptibility among different
ture sensitivity is directly proportional to the dry adhesive bond
mineral aggregates (such as limestone, granite, sandstone) due to
energy, and inversely proportional to the work of debonding
differences in chemical composition, aggregate adsorption, poros-
[8,45]. A higher value of ER is more likely indicating the less sus-
ity, contaminants, and surface texture [8,44–46]. It was found that
ceptibility to moisture damage. In this study, the ER was calculated
the ER value between asphalt and silica decreased as the moisture
using the ratio of the work of adhesion in dry condition as com-
content increased; while the ER value between asphalt and calcite
pared to the work of adhesion in wet condition, as shown in Eq.
were very small regardless of moisture contents. This indicates that
(10). It was assumed here that the introduction of water would
the asphalt–calcite bond could be susceptible to moisture damage
not change the interface area between asphalt and aggregate. In
at relatively small moisture content. Previous studies have
other words, water molecules at the interface were assumed in full
measured bond strength between asphalt binder and aggregate at
contact with asphalt and aggregate at the interface. Therefore, the
different moisture conditions and found that the pull-off strength
selection of interface contact area would not affect the results of ER
between limestone and asphalt binder dropped significantly after
values.
6–24 h moisture conditioning [7].
ER ¼ jW adhesion =W debonding water j ð10Þ On the other hand, for the two asphalt models interacting with
silica, the ER value was smaller for the asphalt-2 model at the
where Wadhesion is the work of adhesion between asphalt and aggre- relatively small moisture content. The ER values of two asphalt
gate in dry condition, and Wdebonding_water is work of debonding models became close when the moisture content increased. This
when water displaces asphalt from the asphalt–aggregate interface. indicates that the effect of asphalt composition on moisture sus-
Fig. 8 shows the calculated ER values for asphalt–silica and ceptibility with silica may be only significant at the relatively small
asphalt–calcite models with three scenarios of moisture contents. moisture content. However, when two asphalt models interact
The simulation results are toward to the lower ends as compared with calcite, the effect of asphalt on moisture susceptibility is neg-
to the range of ER values (0.3–30) observed from experiments ligible. This observation should be further studied by changing the
[44–46]. The discrepancy is properly because the chemical compo- chemical composition of asphalt in different ways, such as
sition of aggregate (such as quartz and limestone) are more com- polymer-modified binder or the addition of anti-stripping agent.
It is noted that the classical MD simulations conducted in this
study cannot account for the formation and breakage of bonds in
the reaction. For example, the dissociation of water molecules at
the surface of quartz can form hydrogen bonds and affect the cal-
culation of surface energy and work of debonding. Future study
with the reactive interatomic potential (such as ReaxFF) will be
conducted to conquer this limitation. The MD simulations could
be further improved by considering the effect of aggregate adsorp-
tion (such as hydroxylated silica or calcite surface), porosity, and
surface texture.

5. Conclusions

This study developed a thermodynamics-based molecular mod-


eling approach to investigate cohesion and adhesion properties of
asphalt concrete for the first time. Molecular dynamic simulation
was employed to derive thermodynamic properties of amorphous
asphalt models (such as surface free energy and cohesive energy
density) and work of adhesion between asphalt and aggregate at
dry and wet conditions. The simulation results agree well with
the experimental measurements reported in the literature. The
results show that van der Waals force plays critical role for cohe-
sive properties of asphalt binder; while the adhesion bonding
between asphalt to aggregate is largely dependent on the type of
aggregate mineral (silica or calcite) in both dry and wet surface
conditions. The effect of asphalt type was found significant for
the adhesion between asphalt and silica at the relatively small
moisture content.
The developed approach sheds light on the complex interaction
Fig. 8. Moisture sensitivity of (a) asphalt–silica and (b) asphalt–calcite interaction between different components of asphalt and at the organic–inor-
with ER values calculated from MD simulation. ganic bi-material interface in asphalt concrete. It promotes devel-
G. Xu, H. Wang / Computational Materials Science 112 (2016) 161–169 169

opment of computational experiments for material performance [21] L.Q. Zhang, M.L. Greenfield, Energy Fuels 21 (2007) 1712–1716.
[22] L. Artok, Y. Su, Y. Hirose, et al., Energy Fuels 13 (2) (1999) 287–296.
prediction from the atomistic scale to the continuum scale. This
[23] H. Groenzin, O.C. Mullins, Energy Fuels 14 (3) (2000) 677–684.
will provide knowledge into design of asphalt concrete using a [24] D.A. Storm, S.J. DeCanio, M.M. DeTar, V.P. Nero, Fuel 69 (6) (1990) 735–738.
‘‘bottom-up” approach to achieve performance requirements of [25] D.A. Storm, J.C. Edwards, S.J. DeCanio, E.Y. Sheu, Energy Fuels 8 (1994) 561–
roadway pavements. 566.
[26] N. Kringos, T. Scarpas, C. Kasbergen, P. Selvadurai, Int. J. Pavement Eng. 9 (2)
(2008) 115–128.
Acknowledgement [27] R.E. Robertson, J.F. Branthaver, P.M. Harnsberger, et al., Fundamental
Properties of Asphalt and Modified Asphalts, vol. I, Interpretive Report,
Technical Report FHWA-RD-99-212, McLean, VA, US Department of
The authors would like to acknowledge the support provided by Transportation, Federal Highway Administration, 2001.
China Scholarship Council (CSC) to Guangji Xu for study abroad. [28] D.D. Li, M.L. Greenfield, Fuel 115 (2014) 347–356.
[29] D. Little, A. Bhasin, Using surface energy measurements to select materials for
asphalt pavement, in: Final Report for NCHRP Project 9–37, 2006.
References [30] M. Deng, V.B.C. Tan, T.E. Tay, Polymer 45 (2004) 6399–6407.
[31] S. Verenich, S. Paul, B. Pourdeyhimi, J. Appl. Polym. Sci. 108 (2008) 2983–2987.
[1] C.W. Curtis, K. Ensley, J. Epps, Fundamental Properties of Asphalt–Aggregate [32] M.L. Connolly, J. Am. Chem. Soc. 107 (5) (1985) 1118–1124.
Interactions Including Adhesion and Absorption, SHRP-A-341, Washington, DC, [33] J. Wei, F. Dong, Y. Li, Y. Zhang, Constr. Build. Mater. 71 (2014) 116–123.
1993. [34] A. Bhasin, J. Howson, E. Masad, et al., Trans. Res. Rec. 2007 (1998) 29–37.
[2] K. Kanitpong, H.U. Bahia, Asph. Paving Technol. 72 (2003) 502–528. [35] Y. Tan, M. Guo, Constr. Build. Mater. 47 (2013) 254–260.
[3] S. Caro, E. Masad, A. Bhasin, D.N. Little, Int. J. Pavement Eng. 9 (2) (2008) 81–98. [36] E.K. Ensley, Thermodynamics of asphalt intermolecular interactions and
[4] M.J. Khattak, G.Y. Baladi, L.T. Drzal, J. Mater. Civil Eng. 19 (5) (2007) 411–422. asphalt–aggregate interactions, in: G.V. Chilingarian, T.F. Yen (Eds.),
[5] Z.W. Chen, J. Xie, Y. Xiao, J.Y. Chen, S.P. Wu, Constr. Build. Mater. 64 (2014) 60– Asphaltenes and Asphalts, first ed., Elsevier, Los Angeles, 1994, pp. 401–426.
66. [37] A.F.M. Barton, CRC Handbook of Solubility Parameters and Other Cohesion
[6] F. Canestrari, F. Cardonea, A. Graziania, et al., Road Mater. Pavement Des. 11 Parameter, second ed., CRC Press, Boca Raton, Florida, 1991.
(2010) 11–32. [38] J.R. Lin, T.F. Yen, The study of molecular attractions in the asphalt system by
[7] R. Moraes, R. Velasquez, H.U. Bahia, Trans. Res. Rec. 2209 (2011) 70–81. solubility parameter, in: M.K. Sharma, T.F. Yen (Eds.), Asphaltene Particles in
[8] A. Bhasin, D. Little, K.L. Vasconcelos, E.A. Masad, Trans. Res. Rec. 2007 (2001) Fossil Fuel Exploration, Recovery, Refining, and Production Processes, Plenum
37–45. Press, New York, 1994, pp. 171–184.
[9] R.A. Tarefder, A. Zaman, J. Mater. Civil Eng. 22 (7) (2010) 714–725. [39] P.G. Redelius, Fuel 79 (1) (2000) 27–35.
[10] H.R. Fischer, E.C. Dillingh, C.G.M. Hermse, Appl. Surf. Sci. 265 (2013) 495–499. [40] P. Wang, Z. Dong, Y. Tan, Z. Liu, Energy Fuels 29 (2015) 112–121.
[11] A.L. Lynea, V. Wallqvistb, B. Birgisson, Fuel 113 (2013) 248–256. [41] P.C. Painter, The Characterization of Asphalt and Asphalt Recyclability, SHRP-
[12] L.Q. Zhang, M.L. Greenfield, J. Chem. Phys. 127 (19) (2007) 194502. A-675, Washington, DC, 1993.
[13] Y. Lu, L. Wang, Int. J. Pavement Eng. 12 (4) (2011) 311–323. [42] S. Kisin, J.B. Vukić, P.G.T. van der Varst, et al., Chem. Mater. 19 (4) (2007) 903–
[14] R.A. Tarefder, I. Arisa, Energy Fuels 25 (2011) 2211–2222. 907.
[15] X. Zhou, S. Wu, G. Liu, P. Pan, Mater. Struct. (2014). [43] M.A.A.R. Quddus, O.J. Rojas, M.A. Pasquinelli, Biomacromolecules 15 (4) (2014)
[16] A. Bhasin, R. Bommavaram, M.L. Greenfield, D. Little, J. Mater. Civil Eng. 23 (4) 1476–1483.
(2011) 485–492. [44] A.E. Alvarez, E. Ovalles, S. Caro, Constr. Build. Mater. 28 (2012) 599–606.
[17] M.P. Allen, Introduction to molecular dynamics simulation, in: N. Attig, K. [45] A. Bhasin, Development of Methods to Qualify Bitumen–Aggregate Adhesion
Binder, H. Grubmüller, K. Kremer (Eds.), Computational Soft Matter: From and Loss of Adhesion Due to Water, Texas A&M University, Texas, 2006.
Synthetic Polymers to Proteins, Lecture Notes, John von Neumann Institute for [46] Y. Liu, A. Apeagyei, N. Ahmad, et al., Int. J. Pavement Eng. 15 (7) (2014) 657–
Computing, Julich, 2004, pp. 1–28. 670.
[18] Materials Studio, Biovia Software Inc., San Diego, 2015. [47] A. Bhasin, E. Masad, D. Little, R. Lytton, Limits on adhesive bond energy for
[19] H. Sun, J. Phys. Chem. B 102 (1998) 7338–7364. improved resistance of hot mix asphalt to moisture damage, Trans. Res. Rec: J.
[20] P.W. Jennings, J.A. Pribanic, M.A. Desando, M.F. Raub, Binder Characterization Trans. Res. Board 1970 (2006) 3–13.
and Evaluation by Nuclear Magnetic Resonance Spectroscopy, SHRP-A-341,
Washington, DC, 1993.

You might also like