Download as pdf or txt
Download as pdf or txt
You are on page 1of 65

MICROMECHANICAL FINITE ELEMENT MODELING OF ASPHALT CONCRETE

MATERIALS CONSIDERING MOISTURE PRESENCE

BY

ZIAD GEORGES GHAUCH

THESIS

Submitted in partial fulfillment of the requirements


for the degree of Master of Science in Civil Engineering
in the Graduate College of the
University of Illinois at Urbana-Champaign, 2014

Urbana, Illinois

Advisers:

Professor Imad L. Al-Qadi


Research Assistant Professor Hasan Ozer
ABSTRACT

Asphalt Concrete (AC) is a composite material consisting of natural or recycled


aggregates blended with petroleum-based binder. The majority of pavements in the U.S. include
AC materials which are often exposed to the adverse effects of moisture. Moisture damage is one
of the major factors that decrease the service life of pavements by causing and/or facilitating the
development of several distresses. In this context, this study numerically investigates the effect
of moisture presence on the micro, meso, and macroscale responses of AC materials. A
Micromechanical modeling framework based on the Finite Element Method (FEM) was
developed to examine the potential of moisture damage in AC materials. The microstructure of
the material was characterized using the non-destructive X-ray Computed Tomography (CT)
technique. Images obtained from X-ray CT scans were used to generate FEM-based
micromechanical models. Preliminary analyses were performed to identify the Representative
Volume Element (RVE) of the composite AC material. It was observed that relatively small
window sizes, as low as 15 mm, were able to reasonably capture the bulk and shear moduli of the
AC mixture. A hydro-micromechanical approach for studying moisture damage was followed.
Moisture fields throughout the microstructure were generated in a mass diffusion procedure
followed by mechanical loading with the properties of AC constituents evolving as a function of
moisture state. Results obtained quantified the contribution of cohesive and adhesive damage on
the overall mixture response to moisture presence.

ii
ACKNOWLEDGEMENTS

My first acknowledgement goes to my advisor, Professor Imad Al-Qadi. This work would not
have been possible if it was not for his valuable guidance, caring, constant support, and
understanding. I am also thankful to all my colleagues at the Advanced Transportation Research
& Engineering Laboratory (ATREL) here at the Department of Civil & Environmental
Engineering at the University of Illinois at Urbana-Champaign. I would also like to aknowledge
all those who assisted in the experimental part of this work, and all the staff members at ATREL.
My second acknowledgment goes to all those who helped me reach this point and shape me into
the person that I am today.

My final and deepest acknowledgment goes to my parents, whose love and support I treasure, to
each and everyone of my wonderful family, and to Nour, of course. To my old brother, for his
never-ending love, I dedicate this work.

iii
TABLE OF CONTENTS

CHAPTER 1 - INTRODUCTION ........................................................................................................ 1


1. 1 Moisture Damage.......................................................................................................................... 1
1. 2 Micromechanical Models .............................................................................................................. 8
1.3 Moisture Damage and Micromechanical Models .......................................................................... 13
1.4 Critical Issues and Computational Challenges .............................................................................. 15
CHAPTER 2 – OBJECTIVE, SCOPE, AND RESEARCH PLAN .................................................... 17
2.1 Objective and Scope .................................................................................................................... 17
2.2 Research Plan .............................................................................................................................. 17
CHAPTER 3 – EXPERIMENTAL PROCEDURE ............................................................................ 19
3.1 Overview of Experimental Approach ........................................................................................... 19
3.2 Laboratory Specimen Preparation ................................................................................................ 19
3.3 X-ray CT Imaging........................................................................................................................ 20
3.4 Dynamic Shear Rheometer Testing .............................................................................................. 22
CHAPTER 4 – MICROMECHANICAL FINITE ELEMENT MODEL .......................................... 23
4.1 Image Segmentation and Filtering ................................................................................................ 23
4.2 Finite Element Mesh Development .............................................................................................. 24
4.3 Material Properties....................................................................................................................... 25
4.4 Air Voids Distribution ................................................................................................................. 28
4.5 Hydro-Mechanical Analysis of Moisture Damage ........................................................................ 29
CHAPTER 5 – REPRESENTATIVE VOLUME ELEMENT ANALYSIS ....................................... 31
5.1 Overview ..................................................................................................................................... 31
5.2 Definition of Representative Volume Element.............................................................................. 31
5.3 Representative Volume Element of Asphalt Concrete Materials ................................................... 32
CHAPTER 6 – ANALYSIS OF MOISTURE-INDUCED DAMAGE ............................................... 38
6.1 Adhesive Degradation .................................................................................................................. 38
6.2 Moisture Concentration Fields ..................................................................................................... 40
6.3 Hydro-Mechanical Analysis ......................................................................................................... 42
6.4 Relaxation Test Simulation .......................................................................................................... 46
CHAPTER 7 – CONCLUSIONS AND FUTURE WORK ................................................................. 49
7.1 Summary ..................................................................................................................................... 49
7.2 Conclusions ................................................................................................................................. 49
7.3 Recommendations for Future Work.............................................................................................. 50
REFERENCES .................................................................................................................................... 52

iv
CHAPTER 1

INTRODUCTION
1. 1 Moisture Damage
1.1.1 Overview of Moisture Damage
The presence of moisture in pavements causes or facilitates the development of several
distresses. AC materials, commonly used in the surface courses of pavement structures, are
constantly exposed to the adverse effects of moisture, whether in liquid and/or vapor state. This
results in a distress known as raveling, which can be defined as the physical separation of the
aggregates from the asphalt binder in the presence of water (1-6). Raveling in general results
from the deterioration of the asphalt binder, a process known as cohesive damage, and/or
weakening of the aggregate-matrix interface, which is refered to as adhesive damage. A
schematic illustration of the characteristics of cohesive and adhesive failure is shown in Figure
1.1.

Figure 1.1 Cohesive and Adhesive failure in AC materials (7)


Research on moisture damage in AC materials can be traced back to the early studies of
Nicholson in 1932 (8), Field and Phang in 1967 (9), Fromm in 1974 (10), and Lottman in 1978
(11). The last three studies were the first to underline stripping phenomenon in AC in the field.

Until the last couple of decades, the study of moisture damage in AC has been essentially limited
to experimental work on a macroscale level. Little information was gained on the fundamental
mechanisms that lead to moisture damage in AC materials, particularly from a microstructural
point of view. Most of the research in that period has focused on proposing and validating
experimental procedures for controlling and mitigating moisture damage in AC, with the
AASHTO T283 (90) being an example. The limitations of laboratory tests proposed in that period
for controlling and reducing moisture damage are well known. In addition to being empirical in
1
nature, such laboratory tests presented unsatisfactory predictions of field performance. Also,
results obtained from those experimental tests were strongly dependent on the specimen
conditionning methodology.

1.1.2 Moisture Transport


Moisture present in the surface layers of the pavement originates from various sources, ranging
from rainfall and humid environments to wet subgrade soils. Incomplete drying of the aggregates
during plant production also contributes to the presence of moisture in AC.

Moisture from various sources listed above moves in the AC structure through various
mechanisms. Masad et al. (12) underlined three main mechanisms of moisture transport: first,
infiltration of surface water, second, permeation of water vapor, and third, capillary rise of
subsurface water. Earlier research has mostly focused on the first moisture transport mechanism.
Kringos (13) noted that not only moving water causes damage to AC materials, but also static
water was found to contribute to moisture damage in AC. The study concluded that moisture
damage is associated with the presence of dynamic and/or static water, which will be separately
discussed in the following sections.

1.1.2.1 Flow Field and Air Voids Structure


Moisture infiltrates throught the air voids structure of the AC material in a process known as
advective flow. In this process, mastic films composed of binder and fine particles, are gradually
washed out, and new layers of mastic are consequently exposed (13). This phenomenon is
directly related to the permeability of the AC material.

Significant experimental and analytical research related to characterizing the permeability of AC


can be found in the literature (14-16). Traditionally, research efforts were focused on expressing
the permeability of AC materials as a function of the air voids (15, 17). Figure 1.2 visualizes the
relationships found in the literature between the permeability and the air voids content of
common AC materials. Permeability values for AC were found to range from more than 10-6
cm/s for relatively impervious AC materials to less than 10 cm/s for highly porous materials. The
wide range of permeability values found from earlier studies, as shown in Figure 1.2, is mainly
attributed to the fact that AC specimens with the same percent air voids can have completely
different air voids structures.

2
10

0.1
Permeability (cm/s)

0.01

0.001 Dense graded


Stone Mastic Asphalt
0.0001 Porous
9.5mm NMAS coarse graded
12.5mm NMAS coarse graded
0.00001
19mm NMAS coarse graded
25mm NMAS coarse graded
0.000001
2 4 6 8 10 12 14 16 18 20
Air voids content (%)

Figure 1.2 Summary of major correlations between the permeability of common AC


materials and the corresponding total air voids content (15, 17)

Analytical approaches for determining the permeability of AC materials used early models for
flow in porous media. The first analytical model was proposed by Ergun (19) which presented a
closed-form solution for the absolute permeability of a porous media consisting of a set of
packed spheres of equal diameter. Another approach for determining the permeability of a
porous media was originally proposed by Kozeny (20) and later modified by Carman (21, 22).
Masad et al. (16, 18) further presented empirical approaches for predicting the permeability of
AC based on, first, the Kozeny-Carman equation for flow in a saturated porous media and,
second, a probabilistic analysis of the distribution of the air voids sizes.

1.1.2.2 Diffusion Due to Moisture Gradient


Another important mechanism of moisture transport in AC materials is diffusion. Diffusion,
which refers to the transport of moisture at the atomic or molecular level, depends on material
properties (i.e. storage rate, storage capacity, and diffusion coefficients) as well as environmental
factors (i.e. relative humidity) (23). The diffusion of moisture within the mastic, which was
found to be directly dependent on the water flow field, reduces the cohesive strength of the
mastic (13). Ultimately, diffusing moisture reaches the mastic-aggregate inteface, thus causing
adhesive failure.

3
Moisture diffusion occurs in response to a concentration gradient. If we assume that, first, the
mastic has negligible porosity and is continuous, free of any defects/cracks, and second, the
aggregates are dry, then molecular diffusion is associated with a moisture concentration gradient
between the inner dry side of the mastic in touch with the aggregate and the outer mastic surface
exposed to moisture (13, 24, 25).

The process of moisture diffusion can be characterized using Fick’s first and second laws. For
isotropic non-steady state diffusion, where the fluid concentration C changes with time t, Fick’s
second law can be expressed as follows:

= (1.1)

Where D is the diffusion coefficient of the material, assumed constant and isotropic, and C is the
concentration of the diffusing substance. For isotropic steady state diffusion, C/ t=0, equation
(1.1) reduces to the equation below, known as Fick’s first law, where the rate of mass transfer is
proportional to the concentration gradient of the diffusing fluid (26):

=− (1.2)

Where J represents the mass flux. Several studies related to moisture diffusion in AC can be
found in the literature. Kassem et al. (28) examined moisture diffusion in AC and presented
experimental measurements of diffusion coefficients for different mastics with predetermined
field performance. Arambula et al. (27, 29) performed experimental and numerical analysis of
water vapor diffusion in AC materials. X-ray CT images were used to characterize the internal
microstructure of the AC materials. Diffusion coefficients for the individual components were
input in the numerical algorithm to estimate the diffusion coefficient of the whole composite
material.

Limited research has been performed in the past to determine the diffusion coefficient of AC
materials, particularly for the mastic phase. In addition, reported measurements of diffusion
coefficients are strongly dependent, among other things, on the source of the materials in
question and the testing procedure adopted (27). Figure 1.3 summarizes experimental attempts
for measuring the diffusion coefficients reported in major earlier studies for various aggregates.

4
1.E+00 Ossian Granite (80)
Scottish lowland granite (80)
Aggregate diffusion coefficient

1.E-01 Montana waste rock (79)


1.E-02 Dolomitican Marble (78)
Pietra Serena sandstone (78)
1.E-03 Pietra de Lecce limestone (78)
(cm2/s)

1.E-04 Rock wool (77)


Basaltic rock (76)
1.E-05 Norway sandstone (32)
1.E-06 Scottish granite (32)
Rapakivi granite (75)
1.E-07 India granite (67)
1.E-08 Denmark chalk (67)
Israel chalk (67)
1.E-09 Kozushima rhyolite rock (35)

Figure 1.3 Experimental diffusion coefficients reported in the literature for various
aggregates
For the diffusion of moisture in the air voids structure, the diffusion coefficient of water in air
has been estimated by several researchers. Geankoplis (31) reported a diffusion coefficient of
0.26 cm2/s at 25ºC. Arambula et al. (29) reported a value of the diffusion coefficients in air equal
to 0.264 cm2/s at 25ºC, relatively close to the earlier prediction by Geankoplis at the same
temperature.

1.1.3 Adhesive & Cohesive Moisture Damage Mechanisms


The transport of moisture in the surface layers induces both cohesive and adhesive damage to the
AC materials used in those layers. Table 1.1 summarizes the main mechanisms behind cohesive
and adhesive processes. A detailed review of the theories behind adhesive and cohesive damage
is beyond the scope of this work and can be found in other studies.

1.1.4 Factors Affecting Moisture Damage


Factors affecting moisture transport include, among other things, the physical and chemical
properties of the materials used, the air voids distribution and connectivity, the aggregates shape,
roughness, and texture, as well as the environmental conditions (i.e. freeze-thaw cycles, relative
humidity, rainfall patterns) and construction practices (i.e. compaction methods). Lu and Harvey
(36) concluded that the pavement structure and age, the air voids content, and the rainfall
patterns represent the most influencing factors behind moisture damage.

5
Table 1.1 Mechanisms behind adhesive and cohesive failures in AC materials
Failure Depends on Can be mitigated by Process
o Aggregate-mastic interface o Increased aggregate
bonding strength angularity and o Thermodynamic
Adhesive o Diffusion characteristics of roughness o Chemical
mastic o Additives o Mechanical
o Thickness of mastic film o Increased film thickness
o Moisture flow field o Modified binders
o Mastic desorption o Additives
Cohesive o Physical
characteristics o Reduced film thickness
o Thickness of mastic film

The strong influence of the distribution and connectivity of air voids on the moisture
susceptibility of AC is well documented in the literature. The configuration of the air voids
structure depends on several factors, including the properties of the aggregates, the mix design
approach, and the compaction method (37-39).

Previous research has further highlighted the idea that the traditional simple measure of percent
air voids is not enough to characterize the complexity of the air voids structure within AC
materials. Arambula et al. (40) observed that the percent air voids parameter is insufficient to
fully characterize the connectivity and distribution of the air voids. This was evidenced by the
fact that AC composites with equal air voids percentages had different air voids configurations.
Following this line of though, the practice of imposing an upper limit on the percent air voids
(i.e. 8% for dense-graded materials) does not provide the best approach for controlling moisture
damage. Water was still found to infiltrate at relatively low air voids percentages (30).

This led to the emergence of other, more advanced ways to characterize the distribution and
connectivity of the air voids structure. Such approaches were mostly based on 2D and 3D
scanning techniques that were used to characterize the internal microstructure of AC. Masad et
al. (30) examined the relationship between moisture damage and the distribution of air voids
within different AC materials, as shown in Figure 1.4. The air voids size parameter was
characterized using the average diameter, while moisture damage was characterized as the ratio
in fatigue life found from Superpave Indirect Tensile test (IDT) of moisture conditioned
specimens over that of dry specimens. Figure 1.4 clearly highlights the existence of a pessimum
air voids size range, at which moisture damage inflicted to the material is highest (30, 41). The
study explained that for AC specimens with air voids sizes higher than the pessimum range,

6
water easily drains out of the mixture, while for specimens with air voids lower than the
pessimum range, the potential for water infiltration into the mixture is relatively low.

Figure 1.4 Moisture damage measured in terms of fatigue life ratio Nfc/Nfu as a function of
the average air voids diameter for an AC specimen with granite aggregates (30)
1.1.5 Moisture Damage Control Methods
Several treatments to address moisture damage in AC materials have been proposed in the past,
with each treatment having it’s own advantages and drawbacks. Current methods for addressing
moisture damage include, but are not limited to, laboratory testing to assess moisture
susceptibility of AC materials, anti-strip additives to increase the aggregate-binder interface
bonding strength, and quality control to avoid material exposure to moisture before construction.
A detailed review of the various control methods for mitigating moisture damage is beyond the
scope of this study.

7
1. 2 Micromechanical Models
The behavior of AC has been examined both on macroscale and microscale levels. From a
macroscale perspective, a unique constitutive relation is used to characterize the overall material
response (i.e. linear elastic, linear viscoelastic, elasto-visco-plastic, etc). From a microscale
perspective, the microstructure of AC is discretized and individual constituents are explicitly
modeled (42). Micromechanical models are particularly attractive in AC applications due to the
high level of heterogeneity and complex microstructure of AC materials.

The overall behavior of AC is governed by the microstructural properties of each of the


constituents of the composite, namely aggregates and mastic, where the mastic phase includes
the asphalt binder, fine aggregates, and fines. The overall behavior of AC is also affected by the
cohesive properties of the aggregate-mastic interface and the air voids structure (43). For the
aggregates, important microstructural features include mineralogy, size, shape, texture, stiffness,
and packing configuration. For the mastic phase, important microscale features include bonding
strength, viscoelastic or viscoplastic response, and volume percentage (44).

Micromechanical models are increasingly being used to extract the fundamental material
properties of heterogeneous materials such as AC using the properties of individual constituents.
Such micromechanical models provide a valuable tool for predicting the fundamental material
properties of the complex composite material given the input properties of each of the
constituents.

Micromechanical models in general can be divided into three main categories: statistical,
analytical, and computational models. Each category will be separately discussed in the
following sections.

Statistical Models
Numerous statistical micromechanical models based on experimental data can be found in the
literature (45, 46, 48). Based on experimental measurements, Heukelom and Klomp (45)
established a statistical relationship between the elastic modulus of dense-graded materials with
air voids in the range 3-5% and the experimental mix design parameters. One study (50)
underlined the limitations of statistical models presented by Heukelom and Klomp (45) and

8
Heukelom (46) in particular, and statistical models based on experimental data in general. The
main limitations of statistical micromechanical models include, among other things (50):

o Statistical equations are only applicable when the parameter values are within the range
used in the experiments.
o Aggregate gradation and maximum aggregate size are not included in statistical models.
o Difficulty to establish a universal statistical model for all asphalt and aggregate types.

Analytical Models
A plurality of analytical models were presented in the past for determining the elastic modulus of
particulate-filled composite materials in general. Hashin and Shtrikman (47) used elastic theory
to present a two-phase micromechanical model. Hashin (52) proposed a micromechanical model
for spherical particles embedded in a matrix phase. Christensen and Lo (53) proposed a three-
phase model of a spherical particle, coated with matrix, and integrated into an equivalent infinite
composite medium. While the model of Christensen and Lo has the advantage of taking into
account particle interaction and size distribution, it still presents several drawbacks (50). Mainly,
the assumed particle gradation is not representative of the actual aggregate gradations usually
encountered in AC applications. Christensen (55) evaluated and compared three types of
analytical micromechanical models of composite materials consisting of a continuous matrix
phase with a high concentration of rigid spherical inclusions in suspension. The study found that
results from the three analytical micromechanical models investigated vary widely.

Other analytical micromechanical models were used specifically for AC applications. Using
elastic theory, Kerner (49) proposed a model for a composite material consisting of a matrix
phase with spherical particle inclusions. Li et al. (50) presented a two-layer built-in
micromechanical model for predicting the elastic modulus of AC. The model assumes an
asphalt-coated circular aggregate embedded into an equivalent AC medium. Similarly, Hirsh
(51) presented a theoretical micromechanical model for predicting the elastic modulus of AC
materials assuming a two-phase material.

Christensen et al. (54) further modified Hirsch’s model for predicting the complex modulus of
AC using mixture properties of Voids in Mineral Aggregates (VMA), Voids Filled with Asphalt
(VFA), and binder modulus.

9
While the aforementioned analytical micromechanical models discussed above share the main
advantage of being relatively simple to implement, several drawbacks persist. The main
limitations of analytical micromechanical models include the following (56, 57):

o Poor prediction of the fundamental properties, particularly for AC materials where the
percentages of inlcusions and the contact between inclusions are high.
o Oversimplification of the complex internal microstructure.
o Inapplicability of the analytical models for materials with inelastic constitutive properties
with plasticity and damage.

1.2.1 Computationally-Based Micromechanical Models


In the last couple of decades, a growing number of micromechanical models based on
computational approaches have emerged. Such numerical micromechanical models, whether
developed specifically for AC applications or later adapted for such applications, were mostly
based on either the Finite Element Method (FEM) or the Discrete Element Method (DEM). The
following sections present a review of both the DEM-based and FEM-based micromechanical
models.

DEM-Based Micromechanical Models


DEM computational micromechanical models are commonly used for AC applications. The
DEM is built on solving Newton’s second law for individual discrete particles in a time stepping
approach. Initially introduced by Cundall in 1971 (58) for applications in rock mechanics, the
DEM was later used for the analysis of soils (59). Since then, several computational studies
based on the DEM have emerged. In general, the DEM is successfully applied for problems
involving large displacements, crack growth, complex microstructural features, and varying
contact conditions (56).

Chang & Meegoda (60) developed a micromechanical model called ASBAL based on the DEM.
The viscoelastic behavior of the asphalt binder is modeled with a mechanical analog model
(Burgers element). The proposed model accomodates both 2D and 3D micromechanical
structures. In the case of 2D models, aggregates are assumed circular, elliptical, or polygonal,
while for 3D models, aggregates are assumed spherical, ellipsoidal, or polyhedral solids. Ullidtz
(61) used 2D DEM micromechanical models to predict rutting and failure potential of cohesive

10
particulate media using the Dem2D code, where the particulate sample is digitally fabricated.
Input properties for the particles include density, contact stiffness, damping, friction coefficient,
particle size, shape, angularity, and particle size distribution. Particles are first created to loosely
fill a box, and then the sample is compacted to achieve a desired pore volume. Rothenburg et al.
(62) performed DEM modeling of idealized elastic aggregates with viscoelastic contact model to
assess the rutting potential in pavements. You et al. (63, 64) performed 2D and 3D
micromechanical DEM modeling to predict the stiffness of AC at various temperatures and
frequencies. Their studies concluded that the stiffness of AC composites was underpredicted by
2D models, while 3D DEM models yielded relatively accurate stiffness predictions over the
range of frequencies and temperatures investigated. You & Buttlar (65) presented a 2D
microfabric DEM approach for predicting the complex modulus of AC using the PFC-2D code.
Using the same DEM code, Buttlar & You (56) developed a 2D micromechanical model of a
Superpave Indirect Tension test (IDT) specimen. The study used an extension of the DEM
method, namely microfabric-DEM, where each material phase is modeled with a cluster of
discrete elements, thus allowing for improved modeling of the complex aggregate shapes. Liu et
al. (66) presented 2D DEM micromechanical models to predict the dynamic modulus and phase
angle of AC materials, where the mastic was modeled with the Burgers model. The study
concluded that the proposed 2D DEM model accurately predicts the viscoelastic properties of
AC within a 90% confidence range.

FEM-Based Micromechanical Models


A large portion of computational micromechanical models is based on the FEM. Dai & You (44)
presented a micromechanical 2D FE framework for determining the viscoelastic properties of
AC. The mastic material was modeled with viscoelastic material properties, and the mastic-
aggregate nodes were fully bonded. Dai (68, 69) developed a microstructural 2D and 3D FE
computational model to predict the complex modulus and phase angle of stone-based materials
under cyclic loading and with varying frequencies. The study concluded that 3D FE models
presented better predictions of complex modulus and phase angle than 2D FE models. The study
further noted that the local maximum strain found from micromechanical finite element analysis
can be 3.2 to 13.6 times higher than the nominal maximum strain, which underlines the
importance of micromechanical analysis of AC materials. Kose et al. (70) used X-ray CT images
to develop FE micromechanical models of AC using elastic material properties for the

11
aggregates and the asphalt binder. Papagiannakis et al. (71) used a 2D micromechancial FEM
model to determine the dynamic shear modulus and phase angle of AC materials tested in the
Superpave Shear Tester. Dai et al. (43) presented 2D FEM and DEM-based micromechanical
models to predict the creep stiffness of AC based on optical scanning of the face of an AC
specimen. Coleri et al. (72, 73) developed 2D and 3D micromechanical FEM models to evaluate
the shear modulus of AC. The study similarly observed that 2D models underpredicted the shear
modulus at all investigated temperatures and loading frequencies. Dai et al. (74) proposed 2D FE
micromechanical models for predicting the creep stiffness using linear and damage-coupled
viscoelastic properties for the mastic phase. The study used computer-generated digital AC
samples that were based on actual sectionned surface images of the specimens. The proposed
micromechanical model was able to predict the AC master curve within a margin of 11.7%.

Several studies have compared the predictions of 2D and 3D micromechanical computational


models. Results from 2D numerical analysis were found to essentially underestimate the overall
response of the material, while results from 3D numerical analysis were generally found to be
more accurate predictors of the response of AC (44, 66).

1.2.2 Image Acquisition and Digital Samples


The first step in the development of a micromechanical model consists of developing a digital
sample that characterizes the microstructure of the composite AC material. Earlier attempts for
creating a digital sample can be mainly divided into two categories: first, based on a virtual
fabrication of a digital AC specimen, and second, based on 2D and 3D image acquisition of the
actual specimen. Figure 1.5 summarizes the main approaches for creating digital samples. In
addition, commonly-used 2D and 3D scanning techniques are also shown in Figure 1.5. In the
first approach, the sample is virtually fabricated either by idealizing the geometry of the
aggregate particles with predetermined shapes (disks in 2D or spheres in 3D), or by inputting the
distribution and shape of the aggregates (66). The advantages of this approach is that it does not
require time-consuming laboratory and experimental work to characterize the microstructure of
AC. This being said, the main drawback of the virtual approach is it’s inability to capture
complex aggregate shapes and distributions, which is common in AC materials. In the second
approach, the digital sample is created by characterizing the internal microstructure of the AC
material through either 2D or 3D images. This approach requires complex experimental settings

12
and efforts to translate 2D/3D images into digital samples. However, as opposed to the virtual
technique, this image-based approach accurately characterizes the distribution and shape of the
aggregate particles within the AC specimen. This is one of the main reasons behind the
popularity of the image-based approach for micromechanical modeling purposes. In particular,
the X-ray CT technique has been successfully used in the past to characterize the internal
microstructure of AC materials.

Predetermined (idealized)
Virtually particles
fabricated
User-defined particles

Nuclear Magnetic
Microstructural Resonance (NMR)
model
3D Transmission Elect.
scanning Microscopy
X-ray Computed
Image- Tomography (CT)
based
Scanning Elect.
2D Microscopy (SEM)
scanning Atomic Force Microscopy
(AFM)

Figure 1.5 Common approaches for generating digital samples for composite materials
1.3 Moisture Damage and Micromechanical Models
Micromechanical models related to moisture damage in AC can be essentially grouped into two
broad categories. The first type of microcmehanical models simulates moisture transport in AC,
while the second type simulates both moisture transport and the resulting deterioration of the
constituents of the AC material. Each micromechanical modeling approach will be discussed in
detail in the following sections.

1.3.1 Fluid Flow Micromechanical Models


A plurality of fluid-flow micromechanical models can be found in the literature (83-85). Fluid-
flow micromechanical models were used to predict the water flow field in terms of velocity and
pressure within the AC specimen. Since an accurate prediction of the flow field within the air

13
voids structure is dependent on the internal microstructure of the AC material, fluid flow models
found in the literature are mostly based on actual imaging of the specimen.

Fluid-flow micromechanical models were based on different computational approaches. Al-


Omari & Masad (83) followed a 3D computational approach to simulate fluid flow in the
microstructure of AC by solving the governing equation of steady state incompressible flow
using the Finite Difference Method. The computational scheme was validated with closed-form
solutions and then extrapolated to predict the permeability of an actual AC microstructure
obtained from X-ray CT imaging. Using micromechanical simulations of fluid flow in several
AC materials, Kutay et al. (84) observed that the horizontal permeabilty can be several orders of
magnitude higher than the vertical permeabilty, as shown in Figure 1.6. This was attributed to the
anisotropic and heterogeneous nature of the air voids structure. The numerical fluid flow model
was based on the Lattice-Boltzmann (LB) technique. The LB method provides an approximation
to the continuous Boltzmann equation by discretizing the physical space with a set of uniformly
spaced Lattice nodes and the velocity space with a discrete set of microscopic velocity vectors
that characterize the flow of fluid particles (86).

Figure 1.6 Distribution of (a) vertical velocity (mm/s), and (b) horizontal velocity (mm/s) in
the air voids structure of a 19 mm NMAS SMA composite for steady state flow (84)
1.3.2 Moisture Damage Micromecahnical Models
The second type of micromechanical models simulates both moisture transport and the resulting
deterioration of AC properties (7, 23). Caro (23) formulated a 2D FEM-based micromechanical
model for predicting moisture-induced damage in AC. The developed computational model
14
coupled the effect of moisture transport with the mechanical response of the microstructure.
Moisture transport was modeled via moisture diffusion only; infiltration and capillary rise of
moisture were neglected in the coupled numerical model. The impact of moisture presence was
examined in terms of both cohesive and adhesive degradation. Cohesive damage was modeled by
changing the viscoelastic bulk properties of the mastic as a function of the moisture
concentration level, while adhesive damage was accounted for via the Cohesive Zone Model,
where cohesive elements were inserted along the aggregate-mastic interface. Kringos (81) first
incorporated both physical and mechanical damage mechanisms associated with moisture
transport in a finite element code called RoAM (Raveling of Asphalt Mixes). In the RoAM code,
which is based on the model of a circular aggregate surrounded by a film of constant thickness,
both mechanical loading as well as moisture diffusion, water flow, advective degradation of the
mastic-aggregate interface, and mastic dispersion are included.

1.4 Critical Issues and Computational Challenges


Moisture damage is a major cause of raveling experienced in AC materials exposed to moisture
and a catalyst of other distresses commonly found in flexible pavements. Even though moisture
damage constitutes a major concern for transportation agencies in the U.S. and worldwide, a
comprehensive understanding of moisture damage and the effectiveness of common treatments
to mitigate moisture damage is still lacking. This lack is mostly expressed by the absence of a
unique standardized effective laboratory approach to assess the moisture susceptibility of AC.
This is further amplified by the fact that existing laboratory approaches for evaluating moisture
susceptibility of AC are significantly time-consuming, which reduces the capacity of
transportation agencies to conduct moisture sensitivity tests during production and quality
control stages.

The emergence of powerful computational platforms has paved the way for more advanced
numerical modeling. In the last couple of decades, computationally-based micromechanical
studies that examine the microstructural behavior of composite materials have emerged.
Micromechanical models provide an attractive and promising approach for mechanistically
evaluating the behavior of composite materials such as AC. Within the context of moisture
damage, such micromechanical models can be used, in a first stage, to improve our

15
understanding of the mechanisms behind moisture damage. Such models can ultimately help in
developing more effective treatments that mitigate moisture damage in AC materials.

However, the micromechanical modeling approach still poses several challenges. The use of
micromechanical modeling for predicting the potential of moisture damage in AC requires a
detailed discretization of all the features at the microscale level. This task becomes even more
challenging when dealing with a highly heterogeneous material such as AC.

With the limitations of traditional analytical and statistical micromechanical models in predicting
the response of AC materials, computational micromechanical models offer a viable alternative.
However, due to the fact that such numerical approaches take into consideration all the
microstructural features of the heterogeneous AC material, some methods can be
computationally challenging. Some challenges that have to be addressed include (1) the proper
characterization of the constituent material properties in the AC composite, (2) adequate
representation of the aggregate-aggregate and mastic-aggregate interfaces, (3) the development
of large-scale computational models, and (4) the bridging of the gap between the local and global
scale properties.

16
CHAPTER 2

OBJECTIVE, SCOPE, AND RESEARCH PLAN


Existing computational micromechanical models have demonstrated large benefits in improving
our understanding of composite materials. Such micromechanical models provide a valuable tool
for understanding the mechanisms behind AC distresses in general, and moisture damage in
particular.

2.1 Objective and Scope


The pbjective of this study is to develop a micromechanical FE model for predicting the response
of AC materials with moisture presence. With a better understanding of the microstructural
behavior of AC, the impact of several moisture damage treatments used for pavements can be
quantified. This will be achieved by bridging the gap between the overall response of AC
materials and the corresponding microscale properties.

The development of a FEM-based micromechanical model requires overcoming several


challenges. Some of the major challenges that have to be addressed include:

o Highly heterogeneous nature and complex internal microstructure of the aggregates and
the air voids.
o Time and temperature-dependent behavior of the asphalt binder.
o High computational cost associated with performing high-resolution 3D modeling.
o Interaction between the mastic phase and the aggregate inclusions
o Large mismatch between the aggregate inclusions and the mastic stiffnesses.

2.2 Research Plan


The research approach for developing the micromechanical FE model is summarized below and
shown in Figure 2.1. Initially, AC specimens were prepared following the Superpave mix design
procedure. The prepared specimens were scanned using non-destructive X-ray Computed
Tomography to characterize the internal microstructure. Image processing and segmentation
techniques were used to identify and isolate individual constituents within the heterogenous AC
material. The images were then used to generate meshes that were imported to a FE code.
Viscoelastic material properties for the mastic phase were determined and input to the FE model.

17
Analysis of statistical homogeneity was performed to numerically determine the corresponding
Representative Volume Element (RVE) dimensions for the specimen being studied. Based on
this analysis, a reduced micromechanical FE model was established and subsequently used.
Finally, a hydro-mechanical analysis was performed on the reduced micromechanical FE model.
In the first stage, moisture concentration fields were generated based on predetermined boundary
conditions, after which, mechanical loading was applied with input properties dependent on the
moisture concentration fields found from the first stage.

Preparation of AC specimen
Mastic
testing
X-ray CT scanning

Image processing
& segmentation

Finite Element
meshes

Rheological Micromechanical
data FE model

Reduced RVE analysis


micromechanical
model

Moisture diffusion Hydro-


Mechanical
Analysis
Mechanical loading

Figure 2.1 Flow chart highlighting the research plan

18
CHAPTER 3

EXPERIMENTAL PROCEDURE

3.1 Overview of Experimental Approach


A Stone Mastic Asphalt (SMA) mixture, commonly used in the surface courses of pavements,
was prepared. Two different laboratory tests were performed, as shown in Figure 3.1. First, the
non-destructive X-ray CT technique was used to capture the internal microstructure of the
prepared SMA specimen. Second, the binder used in the mixture was tested in the Dynamic
Shear Rheometer (DSR) in order to extract the complex shear modulus G* and phase angle ϕ
that characterize the viscoelastic domain. A brief overview of the preparation of the SMA
specimen is presented in the next section.

Mastic Complex shear


Superpave modulus G* &
phase DSR phase angle ϕ
Experimental
tests
AC NDT Digital sample
of AC micro-
mixture X-Ray CT structure

Figure 3.1 Overview of experimental tests

3.2 Laboratory Specimen Preparation


The AC specimen was prepared using a PG76-22 binder and steel slag aggregates. The binder
and the aggregates were mixed at a temperature of 165ºC, after which, the mixture was oven-
aged for 2 hours at a temperature of 165ºC. The SuperPaveTM gyratory compaction approach was
used to compact the cylindrical specimen to a height of 170 mm. The specimen was compacted
in the Superpave Gyratory Compactor (SGC) at a temperature of 150ºC. The properties of the
SMA composite are shown in Table 3.1.

The air voids content of the specimen was measured using the Corelok® device. Steel slag
aggregates used had a 12.5 mm Nominal Maximum Aggregate Size (NMAS), where the NMAS
is defined as one sieve size larger than the first sieve to retain more than 10% of the aggregates.
The gradation of the steel slag aggregates used in the SMA material are shown in Figure 3.2.

19
Table 3.1 Properties of the SMA mixture
NMAS (mm) 12.5
Asphalt binder type PG76-22
Mass of aggregates (g) 8066
Mass of binder (g) 524
Maximum specific gravity (Gmm) 2.972
Bulk specific gravity (Gmb) 2.764
Asphalt binder (% by weight of mixture) 6.0
Target air voids (%) 7.0

100

80

60
% Passing

Sieve 0.45 %
d
No. Opening, d (mm) Passing
1" 25.0 4.3
40 3/4" 19.0 3.8 100
1/2" 12.5 3.1 100
3/8" 9.5 2.8 81.6
#4 4.75 2.0 64.3
#8 2.36 1.5 30.5
20 #16 1.18 1.1 18.2
#30 0.6 0.8 14.8
#50 0.3 0.6 12.9
#100 0.15 0.4 11.1
#200 0.075 0.3 10.1

0
0.0 1.0 2.0 3.0 4.0
Sieve size0.45

Figure 3.2 Aggregates gradation for the 12.5 mm NMAS SMA specimen. Pictures of the
SMA material after compaction in the SGC to a height of 170 mm are also shown

3.3 X-ray CT Imaging


The non-destructive X-ray CT technique is commonly used to characterize the internal structure
of common heterogeneous, opaque, composite materials. In particular, X-ray CT is increasingly

20
being used for AC applications, as the benefits of this technique are well understood and applied
(68, 88, 89).

Figure 3.3 (a) Schematic illustration of the process of horizontal slice stacking to generate a
full 3D CT image, and (b) front view of the reconstructed 3D digital sample
In this study, the internal microstructure of the SMA material was characterized using the X-ray
CT technique. Once slices (i.e. sequence of horizontal planar images of the specimen) are
obtained, a 3D image can be created by combining the slices spaced at predetermined intervals s,
as shown in Figure 3.3. In terms of image resolution, the vertical spacing of the stacked slices
was approximately equal to s = 1 mm. Within the planar horizontal slices, the pixel size was
calculated by dividing the field of view of the X-ray CT machine, which was approximately
equal to the size of the specimen (D = 150 mm) by the number of pixels in each direction, which
was equal to 512x512 pixels. This resulted in a horizontal slice resolution slightly less than 0.2 x
0.2 mm. A picture of the prepared SMA specimen, 100 mm in diameter and 150 mm in height, is
shown in Figure 3.4, along with the reconstructed 3D image from the X-ray CT data.
21
Figure 3.4 SMA cylindrical specimen, (a) picture of laboratory specimen, and (b)
reconstructed 3D image based on X-Ray CT data

Given that the smallest pixel size, or image resolution, equal to 0.2 mm, was actually higher than
some portion of the fine particles, then a mastic phase, which accounts for the asphalt binder and
fine particles smaller than the pixel size, was subsequently used to characterize both the binder
phase and the fine particles.

3.4 Dynamic Shear Rheometer Testing


The Dynamic Shear Rheometer (DSR) test, presented in AASHTO TP5, was used to characterize
the linear viscoelastic behavior of the bituminous binder used in the SMA material. The
complex modulus G* measures the stiffness of the viscoelastic phase when cyclically sheared at
varying temperatures and loading frequencies. The complex modulus G* can be divided into two
parts: an elastic recoverable part characterized by the storage modulus G’, and the viscous part
characterized by the loss modulus G’’. The phase angle ϕ, which ranges from 0 to 90º, represents
the time lag between the applied shear stress and/or strain and the recorded shear response.
Obviously, an increase in the phase angle ϕ towards 90º translates into an increased viscosity,
while a drop in ϕ towards 0º underlines a more elastic behavior.

22
CHAPTER 4

MICROMECHANICAL FINITE ELEMENT MODEL

4.1 Image Segmentation and Filtering


The first step in the development of the micromechanical FE model involves the processing,
segmentation, and filtering of X-ray CT images. In X-ray CT, a given pixel possesses a grayscale
intensity that ranges from 0 to 255, depending on the density of the material at the point where
the pixel is located. The high end of the grayscale intensity spectrum is essentially associated
with the aggregates, which have a higher density than the mastic phase. On the low end of the
spectrum, air voids present the lowest grayscale intensities.

Figure 4.1 Overview of image processing, segmentation, and mesh creation


Once Xray CT images were acquired, image segmentation and filtering was performed using the
Simpleware package (93). Segmentation of the different phases was essentially based on a
threshold for the grayscale intensity of each pixel. For the aggregate part, a grayscale spectrum

23
ranging from 155 to 255 was adopted. During image segmentation, some aggregates that share
close boundaries were found to be connected to each other and consequently, the aggregates
were separated manually. The entire process of image processing and segmentation is visualized
in Figure 4.1.

4.2 Finite Element Mesh Development


After the image processing phase was done, the segmented images were then converted into 2D
microstructural meshes. Two domains were generated, one for the aggregate inclusions and one
for the mastic phase. The air voids were assumed to be embedded in the mastic phase. A slight
smoothing algorithm was implemented in order to obtain better mesh qualities without
significantly affecting the shape of each domain. The micromechanical FE models of the SMA
composite were developed in the Abaqus FE code (34).

Figure 4.2 2D FE mesh of a square micromechanical model with 70 mm side

24
For the 2D plane strain model shown in Figure 4.2, a total of 202,344 quadratic plane strain
elements of type CPE6 were used. The mastic phase contained a total of 115,922 elements, and
the remaining elements were part of the aggregate inclusions. The 2D plane strain model consists
of a square window with a side equal to 70 mm. The smallest element size used in the model was
of the order of 0.04 mm.

4.3 Material Properties


The selection of appropriate material properties for each of the constituents of the
micromechanical FE model is crucial for accurately prediciting the overall response of the
material. A Linear Viscoelastic (LVE) material model was used for the mastic phase, as
discussed in the following section. The aggregates were assumed to behave as a linear elastic
material, with a Young’s Modulus of 60 GPa, and a Poisson’s Ratio of 0.15. A linear elastic
material model was found appropriate for the aggregates due essentially to the fact that the
modulus of the aggregates was many orders of magnitude higher than that of the mastic phase.
Finally, the mastic-aggregate interface was discretized through a series of bilinear 2-node spring
elements, as discussed later.

4.3.1 Cohesive Properties


The presence of inlcusions within the binder phase has an influence on the overall response of
the AC specimen. Following this line of thought, the constitutive material properties for the
viscoelastic phase should take into account the presence of these inclusions. The threshold level
that separates the larger particles considerd part of the aggregate inclusions from the smaller
particles embedded within the mastic phase is directly affected by the CT images resolution.
Given a pixel size of 0.2 mm, particles smaller than this size will not be essentially capturd by
the images and, consequently, will be considered as part of the mastic phase.

The model presented by Buttlar (98) was used to extract the viscoelastic properties of the mastic
phase given the viscoelastic properties of the binder as well as the concentration of inclusions
within the mastic phase. The volume concentration of inclusions within the mastic phase was
estimated at approximately 44%. Results obtained for the mastic phase in terms of complex shear
modulus and phase angle are shown in Figure 4.3.

25
10000 90
Complex modulus G* (MPa)

1000 75

Phase angle (deg)


60
100
45
10
30
1 15
Complex Modulus
Phase Angle
0.1 0
-5.0 -4.0 -3.0 -2.0 -1.0 0.0 1.0 2.0 3.0 4.0 5.0
Log(freq)

Figure 4.3 Complex modulus and phase angle results for the viscoelastic phase of the SMA
material
Table 4.1 Prony Series parameters for the mastic phase

i τi (sec) gi
1 3.328E+04 -3.164E-06
2 5.565E+03 1.166E-05
3 9.305E+02 -4.017E-06
4 1.555E+02 2.774E-04
5 2.601E+01 4.708E-03
6 4.349E+00 3.208E-02
7 7.271E-01 9.920E-02
8 1.215E-01 1.744E-01
9 2.032E-02 2.084E-01
10 3.398E-03 1.961E-01
11 5.682E-04 1.542E-01
12 9.500E-05 1.304E-01
E0(GPa) 13.81
v 0.3

The LVE model was used to capture the response of the mastic phase. Complex Modulus G* and
phase angle ϕ results obtained from DSR testing decribed earlier were converted into Prony
Series parameters that characterize the LVE model. The Prony Series parameters, described in
terms of the reduced time τi (sec) and the dimensionless shear parameter gi are shown in Table
4.1.

26
In order to model the degradation of the viscoelastic properties of the mastic phase with moisture
accumulation, the relaxation modulus values at any instant were shifted downward by
predetermined amounts. The relaxation modulus values characterized by the Prony Series
parameters shown in Table 4.1 were used for the case where the mastic was in a dry condition.
When the mastic phase becomes fully saturated, the relaxation modulus values at any time were
shifted by 15 and 50 percent downward. This degradation in the relaxation modulus of the mastic
phase with moisture accumulation essentially accounts for the cohesive damage that takes place
within the mastic phase. In this approach, the damage inflicted to the mastic phase by the
presence of moisture was assumed to be fully recoverable.

4.3.2 Adhesive Properties


The aggregate-mastic interface was discretized through a series of bilinear spring elements. Each
set of adjacent nodes on the aggregate and mastic parts were connected by two bilinear springs
acting in the global horizontal and vertical directions. The interface springs were characterized
by two different stiffnesses in tension and in compression. In tension, after a predetermined value
of spring opening is reached, the tensile force developed in the springs becomes constant and
independent of the spring opening.

Furthermore, in tension, the set of interface springs had a variable stiffness depending on the
moisture content. The interface springs stiffness in tension was in fact relaxed depending on the
moisture concentration level. For a dry condition, an adequately high tensile stiffness of 106
N.mm-1 was used, and the value was linearly reduced to a value of 103 N.mm-1 at full saturation.

In compression, the interface springs had a sufficiently high stiffness in order to eliminate any
potential interpenetration between the adjacent mastic and aggregate surfaces. The interface
spring stiffness in compression was kept constant and independent of the moisture concentration
level.

4.3.3 Moisture Diffusion


The choice of the diffusion coefficients of the individual components of the AC material is
crucial for the numerical study of moisture damage. The mastic phase was assumed to have a
constant diffusion coefficient of 0.01 mm2/s, while a window of diffusion coefficients ranging
from 0.0001 mm2/s to 0.01 mm2/s was adopted for the aggregate inclusions. This is due to the

27
fact that limited data can be found in the literature on the diffusion of moisture in bituminous
materials and mastic phases.

4.4 Air Voids Distribution


Using the X-ray CT data obtained, and after appropriate image processing and segmentation, the
air voids were isolated, and the distribution of the air voids along the SMA specimen was
obtained. Results were compared with the air voids percentage experimentally measured using
the standard Corelok method. Figure 4.4 shows the distribution of air voids along the depth of
the specimen. The value of the air voids at each depth along the specimen height was determined
by evaluating the percent air voids at discrete slices. The air voids content along the ith slice
(AVi) was calculated by dividing the area of air voids along that slice by the corresponding total
area of the slice. Figure 4.4 also shows an isometric view of the distribution of air voids along
the SMA specimen. The total air voids for the whole specimen (AVT), can be simply calculated
by dividing the sum of the air voids along each slice (AVi) by the total number of slices, n, as
follows (12):

∑ AV
AV = (5.1)

The prediction of air voids within the SMA material is dependent on several factors, including
the resolution of the X-ray CT images, and the grayscale intensity threshold adopted to seperate
the air voids from the other phases of the material. For the first point, any air voids smaller than
the image resolution (or pixel size), which in this case, was around 0.2 mm, will not be detected
and consequently will not be taken as part of the air voids structure. For the second point, to find
the proper threshold for the grayscale spectrum characterizing the air voids, a trial-and-error
procedure was adopted. The threshold intensity was varied until the average air voids percent
measured from the digital sample matches the laboratory-measured air void percent of 7.0 %.

28
This resulted in a grayscale intensity range of 0 to 103 for characterizing the air voids.

Air voids (%)


0.0 2.0 4.0 6.0 8.0 10.0
0
Specimen depth (mm)

50

100

150
Figure 4.4 Air voids distribution along the depth of the SMA specimen with moving
average filter

4.5 Hydro-Mechanical Analysis of Moisture Damage


The FE model presented in this study examines the reponse of a microscale specimen to
combined moisture diffusion and mechanical loading. Moisture diffusion fields were generated
in an initial step, followed by a mechanical loading step where the moisture fields of the previous
step are used as predefined input fields.

4.5.1 Moisture Diffusion


In this study, a unit moisture concentration BC was applied all-around the specimen. Obviously,
a dry condition corresponds to a moisture content of 0.0, while a fully saturated condition
represents a moisture concentration value of 1.0. Initially, a transient mass diffusion step was
used to examine moisture diffusion within the microscale specimen. The linear 3-node and 4-
node 2D mass diffusion elements were used in the simulation. The concentration of moisture
within the window was measured through the CONC output variable. Since a unit moisture

29
concentration was applied along the boundaries of the window, the variable CONC consequently
ranges from 0.0 to 1.0. Output was generated every 2 minutes. The obtained moisture fields were
communicated to the next stage where the mechanical load was applied.

4.5.2 Mechanical Analysis


The moisture conditioning stage was followed by a mechanical analysis step. Moisture diffusion
fields generated from the first conditioning stage were communicated to the second mechanical
analysis step. In this mechanical step, the dependency of the mastic phase LVE properties on the
moisture concentration was activated in order to model cohesive degradation. In addition, the
dependency of the mastic-aggretage interface spring properties on the moisture concentration
level was also activated to account for adhesive degradation.

30
CHAPTER 5

REPRESENTATIVE VOLUME ELEMENT ANALYSIS:

FROM MACRO TO MESO SCALE

5.1 Overview
In this chapter, the concept of Representative Volume Element, or RVE, is reviewed. The
chapter summarizes the efforts that were implemented to reduce the large macro-scale
micromechanical model to smaller meso-scale sizes close to the RVE. The main motivation here
is to numerically investigate potentially smaller meso-scale micromechanical models that offer
the advantage of being computationally cheaper while still being statistically representative of
the macroscale properties.

5.2 Definition of Representative Volume Element


The RVE can be defined as the smallest volume of material that can be used to represent the
corresponding global properties. Continuum mechanics and homogenziation theories are all
based on the RVE concept. Theoretically, the RVE represents a length scale limit in which the
composite material can be modeled as a uniform continuum medium. The homogenization
theory has been extensively used in the analysis of composite materials with microstructural
features contributing to the bulk material response.

The concept of RVE was first presented by Hill in 1963 (94) and later generalized by Huet in
1990 (95). Hill defined RVE as a sample of composite material that is structurally entirely
typical of the whole composite, and that contains a sufficient number of inclusions for the overall
moduli to be effectively independent of surface values of traction and displacement. Another
study (91) defined the RVE as the model of material used to determine the corresponding
effective properties of the homogeneous macroscopic representation. While the RVE should
certainly be smaller than the macroscopic body, it should be large enough to contain sufficient
details related to the microstructure.

Another study (92) highlighted that the RVE is clearly defined if, first, it constitutes a unit cell of
a periodic microstructure, and second, it includes a volume large enough to contain a set of
microscale elements, and thus possess statistical homogeneity. The size of the RVE of a
composite with periodic microstructure is essentially smaller than the corresponding RVE of a

31
composite with random microstructure. Hashin (91) identified three scales in composite
materials, as shown in Figure 5.1.

Micro Dimension of inhomogeneity

Composite
RVE: intermediate dimension
materials Meso
for statistical homogeneity
scales

Effectively homogeneous
Macro
composite material

Figure 5.1 Characterization of different scales in composite materials (91)


It is important to note at this point that homogenization methods are based on the assumptions of
spatial periodicity of the RVE as well as the local uniformity of macroscopic fields within each
RVE (96). The assumptions can be easily violated for the majority of composite materials.
Progressing damage and cracking are some of the factors that can hamper the use of
homogenization theories.

5.3 Representative Volume Element of Asphalt Concrete Materials


Due to the fact that AC is a highly heterogeneous material, an accurate measurement of any
fundamental property should be based on specimen geometries that satisfy the RVE requirement.
In fact, heterogeneous materials such as AC exhibit homogeneous behavior only when the
specimen volume is higher than a certain RVE size. Once the specimen is above the RVE size,
material homogeneity will minimize the variability in the physical and mechanical properties.

Determining the RVE of AC is a challenging task, since the size of the RVE depends on many
factors (33, 87). Among other things, the size of the RVE in AC was found to be highly
dependent on the maximum size of the aggregates in the composite. Using 2D plane strain FE
analysis, Weissman et al. (82) examined the dimensions of the RVE of AC specimens tested in
the simple shear test and concluded that the RVE dimensions depend on, among other things, the
material homogeneity, the testing temperature, the loading rate, and the stress state (82).
Therefore, it becomes evident that an accurate determination of the RVE is possbile only if the
RVE size requirement is made dependent on the factors listed above.

32
Property

RVE

Window size

Figure 5.2 Illustration of the convergence to RVE size


5.3.1 Convergence Analysis to Representative Volume Element Size
From a computational point of view, the model size should be as close as possible to the RVE.
While models with sizes larger than the RVE may require increased computational time,
numerical models with sizes less than the RVE will essentially yield random results that do not
actually measure any fundamental properties.

To determine the RVE size of the SMA material, the following approach was adopted.
Fundamental properties obtained from models with decreasing window sizes are compared. The
minimum window size that causes no further drop in the property is selected as the RVE. A
schematic illustration of this approach is shown in Figure 5.2. Four different window sizes were
investigated in order to predict the RVE dimension of the SMA material. A 70 mm window
plane strain model was developed, from which, three additional windows, of sizes 50, 30, and 15
mm were generated for the analysis of RVE. The smaller windows were randomly selected from
the large 70 mm model. This process is illustrated in Figure 5.3 that also shows the
characterisitcs of each window.

With decreasing window size, the apparent properties associated with each window were
calculated. Then the RVE was determined as the smallest window size at which the calculated
apparent properties converge to a constant effective property.

5.3.2 Parametric Study


Two fundamental properties were calcualted, the apparent isotropic bulk and shear moduli, Kapp
and µ app, respectively. A displacement-controlled unit BC was applied all around the window.

33
The approach for determining the apparent bulk and shear moduli for each 2D domain size was
based on (97). For the apparent bulk moduli, Kapp, a uniform all-around extension equal to unity
was applied. To determine the apparent shear moduli, µ app, a uniform unit extension was applied
in the horizontal direction, while a uniform unit compression was applied in the vertical
direction. For each case, the total strain energy for the whole model was calculated as the sum of
the element’s inidividual strain energies, as shown in Figure 5.4. A dimensionless window size
ratio ξ/ξ0 was defined to characterize the size of the window, where ξ represents the dimension of
the square window, and ξ0 represents the dimension of the smallest window, equal to 15 mm.

Window size (mm x mm) 70x70 50x50 30x30 15x15


Number of pixels in horizontal direction 350 250 150 75
Pixel size along horizontal direction (mm) 0.2 0.2 0.2 0.2
Normalized window size (ξ/ξo) 4.7 3.3 2.0 1.0

Figure 5.3 Reconstructed 3D image of the SMA sample, and illustration of the decreasing
window size approach for determining the RVE of the AC material (ξo = 15 mm)

34
Once the strain energy values UK and Uµ corresponding to the bulk and shear loading cases are
obtained, the apparent isotropic instantaneous bulk and shear moduli can be calculated as
follows:

! !
= 6.1
2"ԑ

! !
&
% 6.2
2"ԑ
Where UKtot and Uµ tot represent the total strain energy for the whole window for bulk and shear
loading cases, respectively, V is the volume of the window, and ԑ represents the strain applied
along the edge.

Uniform extension in 1
Uniform in-plane all- direction and uniform
around extension compression is 2 direction
(ԑ11 = ԑ22 = 1) (ԑ11 = - ԑ22 = 1)

Total strain energy = sum Total strain energy = sum


strain energy of all strain energy of all
elements in window elements in window
UK = ∑UKe Uµ = ∑Uµe

Apparent isotropic bulk Apparent isotropic shear


moduli, Kapp moduli, µapp

ԑ11 ԑ11

2
1 ԑ22
ԑ22

Figure 5.4 Schematic illustration of the approach for determining the apparent
instantaneous bulk and shear moduli

35
Window Size Effect

The effect of the window size on the RVE dimension is examined in this section. Eight slices
were selected at random throughout the depth of the SMA specimen. Along each slice, four
models with window sizes equal to 70, 50, 30, and 15 mm were generated. Thus, the generated
windows correspond to dimensionless window size ratios ξ/ξ0 equal to 4.7, 3.3, 2.0, and 1.0,
respectively. This sums up to a total of 32 computational runs that were performed for the bulk
loading case, at a testing temperature of 25ºC, and a phase modulus mismatch between the
aggregate (E) and the mastic (Eo) equal to Eo/E = 0.23.

28
Bulk moduli Ko (GPa) for ξ / ξo > 1

26

24

22

20
20 22 24 26 28
Bulk moduli Ko (GPa) for ξ / ξo = 1

Figure 5.5 Effect of window size on instantaneous bulk moduli

Figure 5.5 shows the instantaneous bulk moduli K0 values found from models with window sizes
ξ/ξ0 greater than 1.0 (i.e 2.0, 3.3, and 4.7) versus the bulk moduli values found from the models
with the smallest window size (ξ/ξ0 = 1.0). Even though some scatter can be observed in Figure
5.5, for our current purposes, the domain with the smallest window size was able to reasonably
predict the bulk response of the specimen. This allowed us to develop numerical
micromechanical frameworks with a resonable computational cost.

Effect of Modulus Mismatch


Figure 5.6 shows the distribution of strain energy within the micromechanical domain with ξ/ξ0 =
4.7 and at a temperature of 25ºC. Results shown for both bulk and shear loading cases

36
correspond to modulus mismatch ratios Eo/E of 0.12 and 0.35 (we recall that Eo is the
instantaneous modulus of the mastic and E is the Young’s modulus of the aggregates part,
assumed constant and equal to 60 GPa). The increase in strain energy with increasing phase
mismatch (i.e. increasing mastic stiffness) can be clearly observed in those figures.

(a) (c)

(b) (d)

Figure 5.6 Strain energy contour plot (ELSE) for bulk loading with a phase modulus
mismatch E/E0 of (a) 0.12, and (b) 0.35, and for shear loading with a phase modulus
mismatch E/E0 of (c) 0.12, and (d) 0.35

37
CHAPTER 6

ANALYSIS OF MOISTURE-INDUCED DAMAGE


Based on prelimary RVE analysis, the micromechanical model with the smallest window size
(ξ/ξ0 = 1.0) is selected for generating the results presented in the following sections. First, the
effect of adhesive degradation, modeled via the series of bilinear interface springs, on the bulk
and shear response of the composite AC material is examined. Second, the moisture
concentration fields generated from the diffusion analysis are presented. Finally, the moisture
fields are communicated to a mechanical loading step where, again, diplacement-controlled unit
BC’s are applied to predict the bulk and shear moduli of the composite AC material with
moisture presence.

6.1 Adhesive Degradation


The deterioration in instantaneous bulk and shear moduli with decreasing interface stiffness is
shown in Figure 6.1. On the horizontal axis, the interface springs tensile stiffness k, is decreased
from an initial value of ko = 106 N.mm-1 down to a value as low as 103 N.mm-1. We recall that in
compression, the aggregate-mastic interface springs were assinged a significantly high stiffness
in order to eliminate any compressive displacements within the springs, which would result in
interpenetration of the mastic and aggregate surfaces.

25
9

20 8
Bulk moduli K0 (GPa)

Shear moduli µ0 (GPa)

7
15

6
10
Eo/E=0.35 Eo/E=0.35
5
Eo/E=0.23 Eo/E=0.23
Eo/E=0.12 Eo/E=0.12
5 4
1.000 0.100 0.010 0.001
(a) k/k0 (b) 1.000 0.100
k/k0
0.010 0.001

Figure 6.1 Effect of mastic-aggregate interface stiffness ratio (k/k0) on instantaneous (a)
bulk and (b) shear moduli for different mastic properties (k0 = 106 N.mm-1)

38
Results for bulk and shear moduli were obtained for three mastic viscoelastic properties. The
different viscoelastic properties for the mastic phase were achieved by decreasing the relaxation
modulus values at any time by 0, 15, and 50%. In Figure 6.1, the ratio E0/E corresponds to the
instantaneous modulus of the viscoelastic mastic phase (E0) normalized by the aggregate’s
Young’s Modulus (E), equal to 60 GPa. This results in three different E0/E ratios of 0.12, 0.23,
and 0.35, corresponding to drops in relaxations of 50, 15, and 0%, respectively. It should be
noted that the stiffness ratio k/ko is plotted on a logarithmic scale on the horizontal axis. Results
shown in Figure 6.1 correspond to a window size of ξ/ξ0 = 1.0 and a temperature of 25ºC.

It can be observed from Figure 6.1 that both the instantaneous bulk and shear moduli K0 and µ 0
are essentially constant as the interface stiffness ratio k/ko drops from a value of 1.0 to around
0.1. Beyond this value, any further drop in the interface stiffness significantly reduces both K0
and µ 0. For the bulk case, a drop in the interface stiffness ratio k/ko from 0.1 to 0.001 results in a
decrease in the bulk moduli K0 of 45.3%, 41.6%, and 33.5%, corresponding to Eo/E = 0.35, 0.23,
and 0.12, respectively. The same plunge in shear moduli µ 0 values was observed for interface
stiffness ratios k/ko below 0.1.

The curves presented in Figure 6.1 showing the drop in bulk and shear moduli with interface
spring stiffness will be later used in the moisture analysis part of this study. In particular, as
moisture accumulates through the micromechanical model, the mastic-aggregate interface
stiffness will have to be accordingly reduced. In other words, when the bilinear spring element
connects 2 nodes that are completely dry, an interface stiffness ratio k/ko equal to 1.0 will be
adopted. As the nodes connected by the spring element become fully saturated, the interface
stiffness ratio k/ko will be linearly decreased to values as low as 0.001. This reduction in the
stiffness of the interface springs simulates the weakening of the mastic-aggregate interface
attributed to moisture presence.

It should be also noted that the numerical simultations involving low interface stiffness ratios
required relatively higher computational time. This can be explained by the fact that lower
interface stiffness ratios increase the level of nonlinearities within the spring elements along the
mastic-aggregate interface.

39
6.2 Moisture Concentration Fields
The model with the smallest window (ξ/ξ0 = 1.0), at a temperature of 25ºC, was used in a mass
diffusion step to generate the moisture concentration fields along the specimen. Unit moisture
concentration boundary conditions were applied all-around the specimen.

Figure 6.2 shows the contour plot of moisture concentration within the smallest 15-mm window
model for the case of a high mastic diffusion coefficient of 0.01 mm2/s. The moisture
concentration was measured in the mass diffusion step through the normalized concentration
NNC11 output variable, which ranges from 0 to 1. Obviously, the dry condition corresponds to a
normalized concentration value NNC11 equal to 0.0, while a fully saturated condition
corresponds to a NNC11 value of 1.0.

For the smallest micromechanical model with a high diffusion coefficient, a short conditioning
period of just 2 hours was needed to essentially saturate the specimen. As shown in Figure 6.2, at
the end of the 2-hours moisture conditioning stage, some large aggregates were not fully
saturated on the inside. For lower mastic diffusion coefficients, the conditioning time had to be
increased in order to significantly saturate the specimen.

Figure 6.3 plots the evolution of moisture concentration with conditioning time at four discrete
points within the micromechanical model.

o Point A represents the midpoint of an aggregate located towards the edge of the window.
It can be observed that point A becomes only 30% saturated after 2 hours of conditioning.
o Point B corresponds to the midpoint of an aggregate located more towards the center of
the window. After 2 hours of moisture conditioning, point B is less than 10% saturated.
o Point C and D are located within the mastic phase towards the window edge and center,
respectively. At the end of conditioning stage, both points are essentially fully saturated.

40
(a) (d)

(b) (e)

(c)

Figure 6.2 Moisture concentration fields for moisture conditioning times of (a) 2 minutes,
(b) 10 minutes, (c) 20 minutes, (d) 40 minutes, and (e) 2 hours. The mesh used for the
smallest window is shown in (a)

41
1

0.8
A
Moisture concentration

B
0.6 C
D
0.4

0.2

0
0 20 40 60 80 100 120
Conditioning time (min)
Figure 6.3 Variation of moisture concentration with time at four discrete points

6.3 Hydro-Mechanical Analysis


The moisture concentration fields generated from the previous analysis were communicated to a
mechanical step at every 2 minutes. In this mechanical step, the dependency of the mastic phase
LVE properties and the mastic-aggretage interface stiffness on the moisture concentration were
activated in order to model both cohesive and adhesive damage, respectively.

Figure 6.4 plots the strain energy contours in the mechanical loading step for the dry case and for
a mechanical loading step following a 75 minutes conditioning stage. Results were obtained
using the smallest window size ξ/ξo = 1 and a mastic diffusion coefficient of 0.01 mm/s2. Again,
two types of unit displacement boundary conditions were applied in the mechanical step
corresponding to uniform all-around extension (bulk case) and uniform extension in the
horizontal direction and compression in the vertical direction (shear case).

Figures 6.4a and 6.4c show the distribution of strain energy after the bulk loading of the dry and
wet specimens, respectively. The wet specimen was subjected to a 75-minutes moisture
conditioning period. By comparing those Figures, it can be observed that the strain energy

42
distribution in the wet specimen is significantly lower than the corresponding strain energy in the
dry specimen. Given that the same unit displacement was applied, a drop in strain energy
magnitude characterizes the degradation that occurs within the material resulting from both
cohesive and adhesive damage.

The same observation can be made by comparing Figures 6.4b and 6.4d that show the
distribution of strain energy after the shear loading of the dry and wet specimens, respectively. It
can be observed that the mechanical loading of the wet specimen produces lower strain energy
distributions as compared to the mechanical loading of the dry specimen. The drop in strain
energy is again associated with the degradation of the adhesive and cohesive properties of the
material following moisture conditioning.

By comparing Figures 6.4a and 6.4b that correspond to the bulk and shear loading of the dry
specimen to Figures 6.4c and 6.4d that correspond to the bulk and shear loading of the wet
specimen, the increased degradation of the wet specimen following mechanical loading can be
clearly visualized. In fact, in Figures 6.4c and 6.4d, significant “gaps” along the mastic-aggregate
interface can be observed. These “gaps” are essentially associated with extensive openings of the
interface springs simulating adhesive degradation.

Figure 6.5 shows the degradation of the instantaneous bulk and shear moduli with conditioning
time. The moduli degradation resulting from the isolated effect of cohesive and adhesive
damage, as well as the combined effect of cohesive and adhesive damage is examined. Using
these results, attempts will be made to assess the relative contribution of each damage
mechanism (i.e. cohesive or adhesive) to the overall damage accumulated at the mixture level.

Figure 6.5a and 6.5b correspond to, first, a relatively high diffusion coefficient for the mastic
phase, and second, a 50% reduction in the cohesive properties of the mastic phase as the model
becomes fully saturated. Figures 6.5c and 6.5d are similar to Figures 6.5a and 6.5b, with the only
exception that in this case, the cohesive properties of the mastic were only degraded by 15% at
full saturation. By comparing Figures 6.5a to 6.5c and 6.5b to 6.5d, it can be observed that for
the case of a 15% reduction in cohesive properties, the contribution of adhesive and cohesive
damage to moduli degradation is approximately the same. When the reduction in cohesive
properties increases to 50%, obviously, the contribution of cohesive damage to moduli

43
degradation becomes dominant. Similar to Figures 6.5a and 6.5b, Figures 6.5e and 6.5f plot the
evolution of bulk and shear moduli with conditioning time, with the only difference that the
diffusion coefficient of the mastic phase was reduced from 0.01 to 0.0001 mm2/s, with all other
parameters equal. With the lower diffusion rate in the mastic phase, the conditioning time was
increased from 2 hours to approximately 4 hours in order to achieve significant levels of
saturation.

(a)

(b)

(c)

(d)

Figure 6.4 Comparison of strain energy contour along the deformed shape resulting from
mechanical loading for dry conditions: (a) bulk case, and (b) shear case, and following a
conditioning time of 75 minutes: (c) bulk case, and (d) shear case (scale factor = 2)

44
(a) (b)
20 10.0

Shear Moduli (GPa)


Bulk Moduli (GPa)

18 9.0

16 Adh failure 8.0 Adh failure


Coh failure Coh failure
14 Adh+Coh failure Adh+Coh failure
7.0

12 6.0

10 5.0
0 30 60 90 120 0 30 60 90 120
Conditioning time (min) Conditioning time (min)

(c) (d)
20 10.0
Shear moduli (GPa)
Bulk moduli (GPa)

19
9.5
18
9.0
17

Adh failure 8.5


16 Adh failure
Coh failure Coh failure
Adh+Coh failure Adh+Coh failure
15 8.0
0 30 60 90 120 0 30 60 90 120
Conditioning time (min) Conditioning time (min)

(e) (f)
20 10.0
Shear moduli (GPa)

9.6
Bulk moduli (GPa)

19
9.2
18
8.8
17 Adh failure Adh failure
8.4
Coh failure Coh failure
Adh+Coh failure Adh+Coh failure
16 8.0
0 50 100 150 200 0 50 100 150 200
Conditioning time (min) Conditioning time (min)

Figure 6.5 Relative contribution of cohesive and adhesive damage to the deterioration of
bulk and shear moduli with conditioning time. For Dm = 0.01 mm2/s and 50% cohesive
degradation, (a) bulk moduli, and (b) shear moduli. For Dm = 0.01 mm2/s and 15% cohesive
degradation, (c) bulk moduli, and (d) shear moduli. For Dm = 0.0001 mm2/s and 50%
cohesive degradation, (e) bulk moduli, and (f) shear moduli

45
From Figures 6.5e and 6.5f, it can be observed that a decrease in the diffusion coefficient of the
mastic phase reduced the amount of adhesive failure. This can be explained by the fact that with
a lower diffusion coefficient, moisture requires more time to reach the mastic-aggregate interface
where adhesive failure occurs. Consequently, for the conditioning period examined, the majority
of the degradation of bulk and shear moduli originated from cohesive failure of the mastic phase
where water diffuses, and limited adhesive degradation along the mastic-aggregate interface
where adhesive damage occurs.

6.4 Relaxation Test Simulation


In this section, the response of the micromechanical model was examined through a 100-seconds
relaxation simulation. A unit compressive displacement was applied along the top side of the
window. The amplitude of the top compressive displacement was linearly increased to unity,
which corresponds to a vertical compressive strain of 0.067. In this step, the dependency of (1)
the mastic phase LVE properties and (2) the mastic-aggretage interface stiffness on the moisture
concentration fields was also activated in order to model both cohesive and adhesive damage,
respectively.

Using the reaction forces at the bottom of the micromechanical model, the vertical stress
resulting from the application of a vertical compressive strain was determined. The stress-strain
histories were used to back-calculate the Young’s relaxation moduli curves. The degradation in
relaxation moduli at each conditioning stage for cohesive, adhesive and combined cohesive and
adhesvie damage, are presented in Figures 6.6a, 6.6b, and 6.6c, respectively. It is evident that the
drop in relaxation moduli with increasing moisture conditioning time is more pronounced in the
initial phase of relaxation.

At this point, we define the following terms:

o θ, the conditioning time in the moisture diffusion phase.


o ωa, the adhesive damage parameter, equal to the relaxation moduli resulting from pure
adhesive degradation at a certain conditioning time θ normalized by the relaxation moduli
of the dry case.
o ωc, the cohesive damage parameter, equal to the relaxation moduli from pure cohesive
degradation divided by the initial dry moduli.

46
o ω, the total moisture damage parameter, as the ratio of relaxation moduli resulting from
combined cohesive and adhesive degradation to the corresponding dry relaxation moduli.

16 16
Young's Relaxation Modulus

Young's Relaxation Modulus


12 12
(GPa)

(GPa)
8 8

4 0.0 hrs 4 0.0 hrs


0.5 hrs 0.5 hrs
1.0 hrs 1.0 hrs
2.0 hrs 2.0 hrs
0 0
0.0001 0.01 1 100 10000 0.0001 0.01 1 100 10000
(a) Relaxation Time (s) (b) Relaxation Time (s)

16
Young's Relaxation Modulus

12
(GPa)

4 0.0 hrs
0.5 hrs
1.0 hrs
2.0 hrs
0
0.0001 0.01 1 100 10000
(c) Relaxation Time (s)

Figure 6.6 Effect of moisture conditioning time (θ) on Young’s relaxation moduli curves
considering (a) adhesive degradation, (b) cohesive degradation, and (c) combined cohesive
and adhesive degradation
From Figure 6.6a, the following observations can be made. Adhesive damage ratio ωa, was found
to follow a power function with respect to the conditioning time θ, where the rate of adhesive
damage dωa/dθ constantly increases with condioning time for the examined 2-hr conditioning
window. From Figure 6.6b, that plots the ratios ωa/ω and ωc/ω as a function of θ, it is evident that
the contribution of adhesive damage to total moisture damage increases with θ. This can be
explained by the fact that moisture requires some time to reach the mastic-aggregate interface
47
where adhesive damage occurs, which underlines the idea that adhesive damage occurs at a
relatively later stage.

The opposite trend was observed for cohesive damage, ωc, where the rate of cohesive damage
accumualtion, dωc/dθ, decreases wihtin the examined conditining window. In fact, the adhesive
damage increases rapidly to a peak value at around θ=1.0 hr, after which, it becomes constant
along a plateau. From Figure 6.7b, it can be observed that the contribution of cohesive damage to
total moisture damage drops with θ. In fact, as moisture dissipates into the mastic, the latter
becomes more and more saturated, the role cohesive damage plays with respect to total moisture
damage drops gradually.

0.18 1.0
Adhesive (ωa) Normalized Damage Parameter Adhesive (ωa/ω)
Cohesive (ωc) Cohesive (ωc/ω)
Moisture (ω) 0.8
Damage Parameter

0.12
0.6

0.4
0.06
0.2

0.00 0.0
0 1 2 3 0 1 2 3
Conditioning Time θ (min) (b) Conditioning Time θ (hrs)
(a)

Figure 6.7 Evolution of moisture damage parameters with conditioning time


In terms of total moisture damage ratio, ω, an inflection point was observed at around θ = 1.5
hours and at appriximately a damage level of 12%, as shown in Figure 6.7(a). Before the
inflection point, cohesive damage dominates as moisture has not yet reached the mastic-
aggregate interface. After the inflection point, adhesive damage kicks-in while cohesive damage
stabilizes along a plateau.

48
CHAPTER 7

CONCLUSIONS AND FUTURE WORK

7.1 Summary
This study numerically investigated the effect of moisture presence on the micro, meso, and
macroscale responses of a SMA material commonly used in surface layers of pavement
structures. A micromechanical modeling framework based on the FEM was developed to predict
the response of the AC material with moisture presence. The X-ray CT imaging technique was
used to characterize the microstructure of the SMA material. The X-ray CT scans were then used
to generate the FEM-based micromechanical models. A hydro-micromechanical approach for
studying moisture damage was followed. Moisture fields throughout the microstructure of the
AC composite were generated in a mass diffusion procedure followed by a mechanical loading
step with the properties of the constituents of the AC material evolving as a function of the
moisture state. Preliminary analyses were performed to identify the Representative Volume
Element of the composite AC specimen.

7.2 Conclusions
The following observations can be made based on the computational results of this study:

o For AC materials, strain energy-based domain characterization proves to be an efficient


way of extracting apparent isotropic elastic properties (Bulk, Shear, and Young’s Moduli)
at the meso-scale level.
o Based on RVE analysis for the 12.5mm NMAS SMA specimen, a micromechanical
model with a 15 mm window size was able to reasonably capture the bulk and shear
moduli of the AC material, as compared to models with larger window sizes. This is
deemed acceptable for the purpose of this study in which the relative contribution of
cohesive and adhesive damage with moisture presence was investigated.
o The maximum principal strain within the micromechanical model is many orders of
magnitude higher than the applied nominal strain, which underlines the importance of
micromechanical analysis.
o The evolution of moisture damage parameter with conditioning time can be essentially
divided into two phases. An initial phase where moisture diffuses throught the mastic and

49
has not yet reached the mastic-aggregate interface and consequently, cohesive damage
dominates and a relatively late stage where moisture reaches the mastic-aggregate
interface and consequently adhesive damage dominates while cohesive damage stabilizes.
o The rate of adhesive damage increases with conditioning time, while the opposite is true
for cohesive damage.

One advantage of the developed computationally-based micromechanical model is that it can


accommodate complex microstructures as well as the presence of more elaborate features on the
microscale, something that is not offered by earlier statistical and analytical models. This
micromechanical model can be forward-integrated to replace complex and time-consuming
laboratory tests for characterizing moisture susceptibility of common AC materials such as
SMA. This being said, the proposed micromechanical model has several drawbacks. Mainly, the
current micromechanical model does not take into consideration the effect of aggregate interlock
and/or interaction on the AC behavior.

7.3 Recommendations for Future Work


Based on the computational micromechanical FE analysis conducted and the results obtained, the
following recommendations can be suggested:

First, while this study examined the effect of moisture transport on the micromechanical
response of AC material, future work can provide a fully coupled approach for micromechanical
analysis of moisture damage. This can be achieved not only by imposing mechanical properties
dependent on the moisture field, but also moisture transport properties that are dependent on the
stress field within the AC material. In fact, the mechanical loading of the material may initiate
cracks which may strongly affect the transport of moisture within the material.

Second, future work should be also oriented towards extracting appropriate cohesive and
adhesive properties for accurately characterizing the mastic and the mastic-aggregate interface
degradation, respectively. Such degradation properties, which can be only extracted from
experimental approaches, can be integrated with micromechanical models, which will provide
better predictions of the performance of AC materials.

Third, based on the homogenization theories, moisture damage propagation is one of the
problems violating the assumptions these theories hinge on. Future studies have to examine the

50
effect of adhesive and cohesive damage progression in the RVE and macroscale domains on the
local and spatial periodicity.

Finally, it has been concluded from this study that computationally-based micromechanical
models, in particular 3D models, are computationally expensive. This is mainly due to the fact
that such models require a high resolution in order to capture all the microstructural features of
the heterogeneous AC materials. In the future, the problem of micromechanical analysis of
heterogeneous materials such as AC can be tackled using more efficient approaches such as the
use of Multiscale Finite Element Methods that surpass the problem of explicitly modeling the
microstructural features encountered with traditional micromechanical models. Such Multiscale
approaches offer the advantage of accomodating the effect of small-scale features on the overall
response without the need to explicitly model all the microstructural features, which would be
excessively time-consuming. An implementation of multi-scale approaches should be
investigated to capture the microstructural and macrostructural effects of moisture damage.

51
REFERENCES

(1) Riedel, W., & Weber, H., (1953). “On the adhesiveness of bituminous binders on
aggregates”, Asphalt and Teer.

(2) Majidzadeh, K., & Brovold, F., (1968). “State-of-the-art: effect of water on bitumen-
aggregate mixtures”, HRB Special Report No. 98.

(3) Fromm, H.J., (1979). “The mechanism of asphalt stripping from aggregate surfaces”, Journal
of the Association of Asphalt Paving Technologists, Vol.43.

(4) Stuart, K., (1990). “Moisture damage in asphalt mixtures, a state-of-the-art report”, FHWA-
RD-90-019, VA 22101-2296, 1990.

(5) Hicks, R.G., (1991). “Moisture damage in asphalt concrete”, NCHRP Synthesis of Highway
Practice, 175, TRB, NRC.

(6) Lytton, R., (2002). “Mechanics and measurement of moisture damage”, Proceedings of
Moisture Damage Symposium, WRI, Wyoming.

(7) Kringos, N., (2007). “Modeling of combined physical-mechanical moisture induced damage
in asphaltic mixes”, PhD Thesis, TU Delft.

(8) Nicholson, V., (1932). “Adhesion tension in asphalt pavements, its significance and methods
applicable in its determination”, Journal of the Association of Asphalt Paving Technologists
(AAPT), 3, 28-48.

(9) Field, F., & Phang, W. A., (1967). “Stripping in asphaltic concrete mixes,observations and
test procedures”, Canadian Technical Asphalt Association, 12, 61-68.

(10) Fromm, H. J., (1974). “The mechanisms ofasphalt stripping from aggregate surfaces”,
Journal of the Association of Asphalt Paving Technologists (AAPT), 43, 191-219.

(11) Lottman, R. P., (1978). “Predicting moisture-induced damage to asphaltic concrete”, Rep.
No. NCHRP 192, National Cooperative Highway Research Program, Washington, D.C.

52
(12) Masad, E., Arambula, E., Ketcham, R.A., Abbas, A.R., & Martin, A.E., (2007).
“Nondestructive measurements of moisture transport in asphalt mixtures”, Journal of the
Association of Asphalt Paving Technologists (AAPT), 76, pp. 919-952.

(13) Kringos, N., Scarpas, A., (2005). “Raveling of asphalt mixes due to water damage:
computational identification of controlling parameters”. Journal Transport Res Rec, 1929, pp.
79-87.

(14) Al-Omari, A., Tashman, L., Masad, E., Cooley, A., & Harman, T., (2002). “Proposed
methodology for predicting HMA permeability”, Journal of the Association of Asphalt Pavement
Technologists (AAPT), 71, pp. 30-58.

(15) Cooley Jr., L. A., Prowell, B. D., & Brown, E.R., (2002). “Issues pertaining to the
permeability characteristics of coarse-graded Superpave mixes”, Rep. No. 02-07, National
Center for Asphalt Technology, Auburn University.

(16) Masad, E., Birgisson, B., Al-Omari, A., and Cooley, A., (2004). “Analytical derivation of
permeability and numerical simulation of fluid flow in hot-mix asphalt”, Journal of Materials in
Civil Engineering (ASCE), 16(5), 487-496.

(17) Chen, J.S., Lin, K.Y., & Young, S.Y., (2004). “Effects of crack width and permeability on
moisture-induced damage of pavements”, Journal of Materials in Civil Engineering (ASCE),
16(3), pp. 276-282.

(18) Masad, E., Al-Omari, A., & Lytton, R., (2006). “A simple method for predicting laboratory
and field permeability of hot mix asphalt”, Transportation Research Board 85th Annual Meeting,
Washington, D.C.

(19) Ergun, S., (1952). “Fluid flow through packed columns”, Chemical Engineering Progress,
48, pp. 89-94.

(20) Kozeny, J., (1927). “Ueber kapillare Leitung des Wassers im Boden”, Sitzungsberichte
Wiener Akademie, 136(2), pp. 271–306.

(21) Carman, P.C., (1937). “Fluid flow through granular beds”, Transactions, Institution of
chemical Engineers, London, 15, pp. 150–166.

53
(22) Carman, P.C., (1956). “Flow of gases through porous media”, Butterworths, London

(23) Caro, S.S., (2009). “A coupled micromechanical model of moisture-induced damage in


asphalt mixtures: formulation and applications”, PhD Thesis, Texas A&M University.

(24) Masad, E., Lytton, R., & Little, D., (2005). “Moisture-induced damage in asphaltic mixes”,
Proceedings of the International Workshop on Moisture Induced Damage of Asphaltic Mixes.

(25) Cheng, D., Little, D.N., Lytton, R., & Holste, J.C., (2003). “Moisture damage evaluation of
asphalt mixtures by considering both moisture diffusion and repeated-load conditions”, Journal
of the Transportation Research Board, No. 1832, TRB, NRC, 2003.

(26) Crank, J., (1975). The mathematics of diffusion, 2nd edition, Oxford University Press, New
York.

(27) Arambula, E., Caro, S., and Masad, E. (2009). “Experimental measurements and numerical
simulation of water vapour diffusion through asphalt pavement materials”, Journal of Materials
in Civil Engineering (ASCE), 22(6), pp. 588-598.

(28) Kassem, E., Masad, E., Bulut, R., and Lytton, R., (2006). “Measurements of moisture
suction and diffusion coefficient in hot-mix asphalt and their relationships to moisture damage”,
Journal of the Transportation Research Board, 1970, pp. 45-54.

(29) Arambula, E., Garboczi, E. J., Masad, E., & Kassem, E. (2009). “Numerical analysis of
moisture vapor diffusion in asphalt mixtures using digital images”, RILEM Materials and
Structures Journal.

(30) Masad, E., Castelblanco, A., and Birgisson, B. (2006). “Effects of air void size distribution,
pore pressure, and bond energy on moisture damage”, Journal of Testing and Evaluation, 34(1),
1-9.

(31) Geankoplis, C.J., (1993). Transport processes and unit operations, 3rd edition, PTR Prentice
Hall, Englewood Cliffs.

54
(32) Romero, P., & Masad, E., (2001). “Relationship between the representative volume element
and mechanical properties of asphalt concrete.” Journal of Materials in Civil Engineering, 13(1),
pp. 77-84.

(33) Kim, Y.R., Sudo-Lutif, J., & Allen, D.H., (2009). “Determination of representative volume
elements of asphalt concrete mixtures and their numerical investigation through finite element
method.” Transportation Research Board 88th Annual Meeting, Washington, DC.

(34) ABAQUS, (2011). Abaqus/Standard User’s Manual, version 6.11, Dassault Systemes.

(35) Yokoyama , T., and Nakashima, S., (2005). “Diffusivity anisotropy in a rhyolite and its
relation to pore structure.” Eng. Geol. (Amsterdam), 80, pp. 328-335.

(36) Lu, Q., & Harvey, J.T., (2006). “Field Investigation of factors associated with moisture
damage in asphalt pavements”, 10th International Conference on Asphalt Pavements, Quebec,
Canada, pp. 691-700.

(37) Mohamed, E.H.H., Halim, A.O.A.E., & Kennepohl, G., J., (1993). “Assessment of the
influence of compaction method on asphalt concrete resistance to moisture damage”.
Construction & Building Materials, 7(3), pp. 149-156.

(38) Masad, E., Kassem, E., & Chowdhury, A., (2009). “Application of imaging technology to
improve the laboratory and field compaction of HMA”, Rep. No. FHWA/TX-09/0-5261-1, Texas
Transportation Institute (TTI), College Station, Texas.

(39) Brown, E. R., Haini, M. R., Cooley, A., & Hurley, G., (2004). “Relationship of air voids, lift
thickness, and permeability in hot mix asphalt pavements”, Rep. No. NCHRP 531,
Transportation Research Board, Washington, D.C.

(40) Arambula, E., Masad, E., & Martin, A. E., (2007). “The influence of air void distribution on
the moisture susceptibility of asphalt mixes”, Journal of Materials in Civil Engineering (ASCE),
19(8), pp. 655-664.

(41) Terrel, R. L., & Al-Swailmi, S. (1993). “The role of pessimum voids concept in
understanding moisture damage to asphalt concrete mixtures”, Journal of the Transportation
Research Board, 1386, pp. 31-37.

55
(42) Abbas, A., Masad, E., Paapagiannakis, T., Harman, T., (2007). “Micromechanical modeling
of the viscoelastic behavior of asphalt mixtures using the discrete-element method”, Int J
Geomech, 7, 131.

(43) Dai, Q., You, Z., (2007). “Prediction of creep stiffness of asphalt mixture with
micromechanical finite-element and discrete element models”, Journal Eng. Mech., 133(2), pp.
163-173.

(44) Dai, Q., You, Z., (2008). “Micromechanical finite element framework for predicting
viscoelastic properties of heterogeneous asphalt mixtures”. Mater Struct, 41(6), pp. 1025-1037.

(45) Heukelom, W., & Klomp, A.J.G., (1964). “Road design and dynamic loading”, Proc., Assn.
of Asphalt Paving Technologists, Vol. 33, Cushing-Malloy, Inc., Ann Arbor, Mich., pp. 92-165.

(46) Heukelom, W., (1973). “An improved method of characterizing asphaltic bitumens with the
aid of their mechanical properties”, Proc., Assn. of Asphalt Paving Technologists, Vol. 42,
Cushing-Malloy, Inc., Ann Arbor, Mich., pp. 67-98.

(47) Hashin, Z., & Shtrikman, S., (1963). “A variational approach to the theory of the elastic
behavior of multiphase materials”, J Mech Phys Solids, 11, pp. 127-140.

(48) Van der Poel, (1954). “A general system describing the viscoelastic properties of bitumens
and its relation to routine test data”, J Appl Chem, 4, pp. 221-236.

(49) Kerner, E.H., (1956). “The elastic and thermoplastic properties of composite media”, Proc.,
Physc. Soc., London, 69, pp. 808-819.

(50) Li, G., Li, Y., Metcalf, J.B., Pang, S.S., (1999). “Elastic modulus prediction of asphaltic
concrete”, Journal Mater Civil Eng, 11(3), pp. 236-241.

(51) Hirsch, T.J., (1962). “Modulus of elasticity of concrete affected by elastic moduli of cement
paste matrix and aggregate”, J Am Concr Inst, 59(3), pp. 427-452.

(52) Hashin, Z., (1965). “Viscoelastic behavior of heterogeneous media”, J Appl Mech, 32(9),
pp. 630-636.

56
(53) Christensen, R.M., & Lo, K.H., (1979). “Solutions for effective shear properties in three
phase sphere and cylinder models”. J Mech Phys Solids, 27, pp. 315-330.

(54) Christensen, D.W., Pellien, T., & Bonaquist, R.F., (2003). “Hirsch models for estimating the
modulus of asphalt concrete”. Asphalt Paving Technol., 72, pp. 97-121.

(55) Christensen, R.M., (1990). “A critical evaluation for a class of micro-mechanics models.” J
Mech Phys Solids, 38(3), pp. 379-404.

(56) Buttlar, W.G., You, Z., (2001). “Discrete element modeling of asphalt concrete: a
microfabric approach”, J Tansport Res Board, 1757, 111.

(57) Buttlar, W.G., & Roque, R., (1997). “Micromechanical modeling to predict the stiffness of
asphalt concrete”, Proc., 5th Pan American Congress on Applied Mechanics; Applied Mechanics
in the Americas, San Juan, Puerto Rico, Vol.4, pp. 175-178.

(58) Cundall, P.A., (1971). “A computer model for simulating progressive large scale
movements in blocky rock systems”, Symp. Of the International Society for Rock Mechanics,
Nancy, France, II-8.

(59) Cundall, P.A., & Strack, O.D.L. (1979). “Discrete numerical model for granular
assemblies”, Geotechnique, 29(1), pp. 47-65.

(60) Chang, G.K., Meegoda, J.N., (1997). “Micromechanical simulation of hot mixt asphalt”,
ASCE J Eng Mech, 123, 495.

(61) Ullidtz, P., (2001). “Distinct element method for study of failure in cohesive particulate
media”, Transport Res Rec, 127.

(62) Rothenburg, L., Bogobowicz, A., Haas, R., (1992). “Micromechanical modeling of asphalt
concrete in connection with pavement rutting problems”, Proc 7th Int Conf on Asphalt
Pavements, vol. 1, pp. 230-245.

(63) You, Z., Adhikari, S., Kutay, M.E., (2009). “Dynamic modulus simulation of the asphalt
concrete using the x-ray computed tomography images”, Mater Struct, 42(5), pp. 617-630.

57
(64) You, Z., Adhikari, S., Dai, Q., (2008). “Three-dimensional discrete element models for
asphalt mixtures” Journal Eng Mech, 2008, 134(12), pp. 1053-1063.

(65) You, Z., Buttlar, W.G., (2004). “Discrete element modeling to predict the modulus of
asphalt concrete mixtures”, J Mater Civil Eng – Am Soc Civil Eng (ASCE), 16, 140.

(66) Liu, Y., Dai, Q., You, Z., (2009). “Viscoelastic model for discrete element simulation of
asphalt mixtures”, J Eng Mech, 135, 324.

(67) Witthuser, K., Arnepalli, D., and Singh, D.N., (2006). “Investigations on diffusion
characteristics of granite and chalk rock mass”, Geotech. Geologic Eng., 24, pp. 325-334.

(68) Dai, Q. (2011). “Two- and three-dimensional micromechanical viscoelastic finite element
modeling of stone-based materials with X-ray computed tomography images”, Construction &
Building Materials, 25, pp. 1102-1114.

(69) Dai, Q., (2010). “Prediction of dynamic modulus and phase angle of stone-based composites
using micromechanical finite element approach”, Journal Mater Civil Eng, 22(6), pp. 618-627.

(70) Kose, S., Guler, M., Bahia, H.U., & Masad, E., (2000). “Distribution of strains within hot-
mix asphalt binders”, Journal Transport Res Rec, 1728, Washington, D.C., pp. 21-27.

(71) Papagiannakis, T., Abbas, A., Masad, E., (2002). “Micromechanical analysis of the
viscoelastic properties of asphalt concrete”, Journal Transport Res Rec, 1789, pp. 113-120.

(72) Coleri, E., Harvey, J.T., Yang, K., Boone, J.M. (2012). “Development of a micromechanical
finite element model from computed tomography images for shear modulus simulation of asphalt
mixtures”, Construction & Building Materials, 30, pp. 783-793.

(73) Coleri, E., Harvey, J.T., Yang, K., Boone, J.M. (2012). “A micromechanical approach to
investigate asphalt concrete rutting mechanisms”, Construction & Building Materials, 30, pp.
36-49.

(74) Dai, Q., Sadd, M.H., You, Z., (2006). “A micromechanical finite element model for linear
and damage-coupled viscoelastic behavior of asphalt mixtures”, Int J Numer Analyt Methods
Geomech, 30, 1135.

58
(75) Valkiainen, M., Aalto, H., Lehikoinen, J., and Uusheimo, K., (1996). “The effect of
thickness in the through-diffusion experiment.” Research Notes 1788 Final Rep., Technical
Research Center of Finland VTT, Espoo, Finland.

(76) Kodikara, J., and Chakrabarti, S., (2005). “Modeling of moisture diffusion in crushed
basaltic rock stabilized with cementitious binders.” J. Mater. Civ. Eng., 17(6), pp. 703-710.

(77) Jirickova, M., and Cerny, R., (2002). “Hygric and thermal properties of capillary active rock
wool thermal insulation.” Proc., Building Physics 6th Nordic Symp., Norwegian University of
Science and Technology, Trondheim, Norway.

(78) Henon, F.E., Carbonell, R.G., and DeSimone, J.M., (2002). “Effect of polymer coatings
from CO2 on water-vapor transport in porous media”, The American Institute of Chemical
Engineers Journal, 48(5), pp. 941-952.

(79) Dobchuk, B.S., Barbour, S.L., and Zhou, J., (2004). “Prediction of water vapor movement
through waste rock”, J. Geotech. Geoenviron. Eng., 130(3), pp. 293-302.

(80) Bradbury, M.H., Lever, D., and Kinsey, D., (1982). “Aqueous phase diffusion in crystalline
rock.” Proc., Material Research Society 5th Int. Symp. On the Schientific Basis for Nuclear Waste
Management: Scientific Basis for Radioactive Waste Management, New York, pp. 569-578.

(81) Kringos, N., Scarpas, A., Kasbergen, C., amd Selvadurai, , P., (2008). “Modeling of
combined physical-mechanical moisture-induced damage in asphaltic mixes, part 1: governing
processes and formulations”, International Journal of Pavement Engineering, 9(2), pp. 115-118.

(82) Weissman, S.L., Harvey, J., Sackman, J.L., and Long, F., (1999). “Selection of laboratory
test specimen dimension for permanent deformation of asphalt concrete pavements.” Journal
Transport Res Rec 1681, Washington, DC, pp. 113-120.

(83) Al-Omari, A., & Masad, E., (2004). “Three dimensional simulation of fluid flow in X-ray
CT images of porous media”, International Journal for Numerical and Analytical Methods in
Geomechanics, 28(13), pp. 1327-1360.

59
(84) Kutay, M. E., Aydilek, A. H., & Masad, E., (2007). “Estimation of directional permeability
of HMA based on numerical simulation of micro-scale water flow”, Journal Transport Res Rec,
2001, pp. 29-36.

(85) Masad, E., Al-Omari, A., & Chen, H.-C., (2007). “Computations of permeability tensor
coefficients and anisotropy of asphalt concrete based on microstructure simulation of fluid
flow”, Computational Materials Science, 40(4), pp. 449-459.

(86) Maier, R., Kroll, D., Kutsovsky, Y., Davis, H.T., and Bernard, R., (1997). “Simulation of
flow through bead packs using the Lattice Boltzmann method”, AHPCRC Preprint 97-034,
University of Minnesota, Minneapolis.

(87) Coleri, E., Harvey, J.T., (2011). “Analysis of Representative Volume Element for Asphalt
Concrete Laboratory Shear Testing”, J. Materials in Civil Eng., 23(12), pp. 1642-1653.

(88) Masad, E., (2004). “X-ray computed tomography of aggregates and asphalt mixes”, Mater
Eval, 62, 775.

(89) Wang, L., Paul, H., Harman, T., D’Angelo, J. (2004). “Characterization of aggregates and
asphalt concrete using X-ray computed tomogprahy”. J Assoc Asphalt Paving Technol, 73.

(90) AASHTO T 283-07, “Standard method of test for resistance of compacted hot mix asphalt
to moisture-induced damage”, American Association of State Highway & Transportation
Officials, Washington, D.C.

(91) Hashin, Z., (1983). “Analysis of composite materials: a survey”, J. Appl. Mechanics, 50, pp.
481-505.

(92) Ostoja-Starzewski, M., (2001). “The use, misuse, and abuse of stochastic random media”,
Proceeding of European Conference on Computational Mechanics, 2001.

(93) Simpleware, ScanIP, & +ScanFE, (2010), Simpleware Ltd., Innovation Center, Rennes
Drive, Exeter EX4 4RN, UK.

(94) Hill, R., (1963). “Elastic properties of reinforced solids – some theoretical principles”,
Journal of the Mechanics of Solids, 11, pp. 357-372.

60
(95) Huet, C., (1990). “Application of variational concepts to size effects in elastic
heterogeneous bodies”, Journal of the Mechanics and Physics of Solids, 38(6), pp. 813-841.

(96) Fish, J., & Shek, K., (2000). “Multiscale analysis of composite materials and structures”,
Composites Science and Technology, 60(12), pp. 2547-2556.

(97) Jiang, M., Jasiuk, I., & Ostoja-Starzewski, M., (2002). “Apparent elastic and elastoplastic
behavior of periodic composites”, International Journal of Solids & Structures, 39, pp. 199-212.

(98) Buttlar, W.G., Bozkurt, D., Ghazi, G.A., & Waldhoff, A.S., (1999). “Understanding asphalt
mastic behavior through micromechanics”, Journal Transport Res Rec, 1681, pp. 157-169.

61

You might also like