You are on page 1of 103

COMPARISON BETWEEN MECHANISTIC AND MECHANISTIC-EMPIRICAL

APPROACHES TO PREDICT ASPHALT PAVEMENT SERVICE LIFE

by

Emmeline Lemos Pereira

A THESIS

Presented to the Faculty of

The Graduate College at the University of Nebraska

In Partial Fulfillment of Requirements

For the Degree of Master of Science

Major: Engineering Mechanics

Under the Supervision of Professor David H. Allen

Lincoln, Nebraska

November, 2009
COMPARISON BETWEEN MECHANISTIC AND MECHANISTIC-EMPIRICAL

APPROACHES TO PREDICT ASPHALT PAVEMENT SERVICE LIFE

Emmeline Lemos Pereira, MS

University of Nebraska, 2009

Advisor: David Allen

This work proposes a comparison between two different asphaltic pavement

design models to predict the mechanical behavior and service life of asphalt pavements.

The first model is based on a mechanistic-empirical analysis and the second one is based

on an in-house finite element code. The mechanistic-empirical model applies

axisymmetric loads to calculate pavement service life. Damage is included in the form of

fatigue cracking and permanent deformation based on empiric equations. The material

property used is linear elastic and viscoelastic.

On the other hand, the fundamental difference between the two models is the way

damage is calculated. The mechanistic model uses viscoelastic cohesive-zone models to

predict crack propagation and accounts for energy dissipation. Additionally, a Mohr-

circle analysis was performed in order to predict zones with the highest stresses, therefore

prone to crack initiation.


Three different analysis were performed, two including damage, but different

material properties and one without damage. Mohr circle indicated that shear stresses

reach their maximum values under the tires, therefore the analysis including damage have

cohesive zones positioned at a 45 degrees plane. This enables growth of potential shear

cracks. A two-dimensional viscoelastic model has been formulated in the FEM code to

define combinations of design variables which may increase pavement service life.

The results show that the service life is increased by increasing the asphalt layer

thickness. For instance, a 20 cm thick asphalt layer provides a 25% better life than a 10

cm thick asphalt layer in cases in which damage is included. However, if damage is not

included, the life of a 20 cm thick asphalt layer goes up to 38% compared to a 10 cm

thick asphalt layer. Also, when the mechanistic-empirical and the finite element models

are compared, the mechanistic-empirical model results in a lower pavement service life.

Overall, results indicate that the FEM model proposed herein is more scientific

based than the mechanistic-empirical model, therefore it could play a fundamental role in

pavement design providing a more realistic approach to pavement life.


i

Dedication

to Mommy and Pino;


to Dan;
to Lord Almighty.
Acknowledgments

I would like to express my sincere acknowledgment first and foremost to my

advisor Dr. David Allen, for all his patience and valuable knowledge. He taught me more

than I could ask (and imagine), and help me to become a researcher. I am also extremely

grateful to the members of my committee, Dr. Yong Kim and Dr. Mehrdad Negahban for

their valuable input and interest in this research.

Thanks to Marilena Soares for helping the dream of graduate school in the States

becomes real, and her friendship throughout the years. Very special thanks to Roberto

Soares for his friendship, guidance, valuable thoughts, and for taking the time to read my

thesis over and over. I have no words to express how thankful I am to you.

I would also like to thank classmates who shared challenges brought by classes.

Special thanks to Jamilla Lutif and Thiago Aragao for the essential help in the first

semester. Also, I would like to thank James Nau for all the help with my computer.

I would like to express my acknowledgment to the Army Research Office with

the contract number W911NF-04-2-0011.

I want to sincerely thank Daniel Benício Albuquerque. Through each stage of

preparation, you shared the burdens, anxieties, tears and pleasures of this study. You

always stand by my side filling me with courage; pushing me to go further, praying with

me and for me, and spent innumerous hours helping with the thesis format. I could not do

without you. For my parents Fátima Lemos and Agripino Pereira. and my sisters Eveline

Fontgalland and Emiliane Lemos, for their constant incentive, unconditional love, and
i

always believing in me (even when I did not believe it). Besides the physical distance,

you were always very present and supportive of my work.

Thanks to my best friends Clarisse Bessa and Rebeka Colares. I am very thankful

for the friendship, care, and all the time devoted to me. You filled my days with

happiness and were always there in the hard times. My sincere gratitude to my host

family, in special to Margaret Sullivan, for all the help, trips, thanksgivings, and

introducing me to the American culture.

Finally, I want to express my immense gratitude to God for the wisdom and

perseverance that He has been bestowed upon me during this research. “ I am able to do

all things through Him who strengthens me.” (Philippians 4: 13)


ii

List of Tables

Table 1. Material Properties for Viscoelastic Layers........................................................ 71

Table 2. Material Properties for Elastic Layers ................................................................ 71


iii

List of Figures

Figure 1. Heavy Vehicle and Truck Traffic on I-20/59. ..................................................... 2


Figure 2. Fatigue Cracking ................................................................................................. 3
Figure 3. Flexible pavement ............................................................................................... 7
Figure 4. Concrete Pavement .............................................................................................. 7
Figure 5. Composites Pavement ......................................................................................... 8
Figure 6. Minimum and Maximum Stresses in the Tensile Zone ....................................... 9
Figure 7. Tension Stress at the Bottom of the Base Layer................................................ 10
Figure 8. Advanced Alligator Cracking Stage .................................................................. 10
Figure 9 . Pothole on Asphalt Pavement Surface ............................................................. 11
Figure 10. Rutting on Asphalt Pavement .......................................................................... 12
Figure 11. Permanent Deformation, Transversal View .................................................... 12
Figure 12. Compressive Strain at the Top of the Subgrade .............................................. 13
Figure 13. Top Down Cracking ........................................................................................ 14
Figure 14. Thermal Cracking ............................................................................................ 16
Figure 15. Bleeding ........................................................................................................... 17
Figure 16. Raveling ........................................................................................................... 18
Figure 17. Corrugation ...................................................................................................... 19
Figure 18. AASTHO Road Test sketch ............................................................................ 20
Figure 19. Multilayer System ........................................................................................... 21
Figure 20. Pavement Mesh Example in 2D ...................................................................... 23
Figure 21. Pavement Mesh Example in 3D ...................................................................... 23
Figure 22. Finite Element Method Used in Biomedical Field. Courtesy SCI Institute. ... 26
Figure 23. Finite Element Method – Mesh generation ..................................................... 26
Figure 24. Aloha Flight 243 .............................................................................................. 27
Figure 25. Crack Propagation ........................................................................................... 28
Figure 26. Mode I, Mode II and Mode III cracks ............................................................. 29
Figure 27. Cohesive Zone ................................................................................................. 30
Figure 28. KENLAYER Pavement Configuration ........................................................... 33
Figure 29. Stress Tensor in Cartesian Coordinates ........................................................... 33
Figure 30. Tire Configuration for Different Applied Loads. Courtesy Tekscan, Inc. ...... 35
Figure 31. Cylindrical Coordinates ................................................................................... 36
Figure 32. Stress Tensor in Cylindrical Coordinates ........................................................ 37
Figure 33. Boundary Conditions ....................................................................................... 43
Figure 34. General Three-Dimensional Body................................................................... 53
Figure 35. Cohesive Zone ................................................................................................. 58
Figure 36. Typical Two-Lane Asphalt Multilayer Roadway ........................................... 68
Figure 37. Pavement Boundary Conditions ..................................................................... 69
Figure 38. Pavement Cross-Section .................................................................................. 69
Figure 39. Generalized Maxwell Model ........................................................................... 70
Figure 40. Tandem axle truck ........................................................................................... 72
Figure 41. Tire and Axle Dimensions ............................................................................... 73
Figure 42. 3-D View of a Typical Two-Lane Asphalt Multilayer Roadway .................... 74
Figure 43. Tire Pressure Distribution................................................................................ 74
Figure 44. Pressure Distribution ....................................................................................... 75
iv

Figure 45. Load Distribution............................................................................................. 75


Figure 46. Roadway Mesh ................................................................................................ 76
Figure 47. Monotonic Load .............................................................................................. 77
Figure 48. Stress in xx direction for first and second layers ............................................. 77
Figure 49. Stress in xx direction for first and second layers ............................................. 78
Figure 50. Stress in xy direction for first and second layers ............................................. 78
Figure 51. Mohr Circle ..................................................................................................... 79
Figure 52. Maximum Stresses .......................................................................................... 80
Figure 53. Maximum Angles ............................................................................................ 81
Figure 54. Roadway Mesh with Cohesive Zones ............................................................. 81
Figure 55. Failure criterion ............................................................................................... 83
Figure 56. Displacement versus time ................................................................................ 84
Figure 57. Displacement versus time in log scale............................................................. 84
Figure 58. Different design lifes for HMA thickness of 10 cm. ....................................... 85
Figure 59. Different design lifes for HMA thickness of 20 cm ........................................ 86
v

Table of Contents

1 INTRODUCTION ...................................................................................................... 1

1.1 Study Objective .................................................................................................... 4


1.2 Study Scope .......................................................................................................... 4

2 BACKGROUND ........................................................................................................ 6

2.1 Pavement Structure .............................................................................................. 6


2.2 Type of Distresses ................................................................................................ 8
2.2.1 Alligator or Fatigue Cracking ....................................................................... 8
2.2.2 Rutting......................................................................................................... 11
2.2.3 Top-Down Cracking ................................................................................... 13
2.2.4 Thermal Cracking ....................................................................................... 15
2.2.5 Other Distresses .......................................................................................... 16
2.3 Pavement Design and Analysis Methods ........................................................... 19
2.3.1 Empirical Pavement Design ........................................................................ 19
2.3.2 Mechanistic-Empirical Pavement Design ................................................... 21
2.3.3 Mechanistic Pavement Design .................................................................... 22
2.4 Failure Criteria ................................................................................................... 24
2.5 Finite Element Methods ..................................................................................... 25
2.6 Cohesive Zone Model ........................................................................................ 27

3 KENLAYER ............................................................................................................. 31

3.1 Overview of the KENLAYER software............................................................. 31


3.2 Solution Method ................................................................................................. 32
3.3 Damage Analysis................................................................................................ 46

4 MECHANISTIC ROADWAY MODEL .................................................................. 52

4.1 Mechanistic Roadway Model ............................................................................. 52


4.2 Cohesive Zone Model ........................................................................................ 57

5 FINITE ELEMENT FORMULATION .................................................................... 59


vi

6 MECHANISTIC ROADWAY ANALYSIS............................................................. 66

6.1 Mechanistic Roadway Model Analysis .............................................................. 68


6.1.1 Roadway Geometry .................................................................................... 68
6.1.2 3-D and 2-D ................................................................................................ 73
6.1.3 Roadway Mesh............................................................................................ 76
6.1.4 Analysis....................................................................................................... 76
7 Results ....................................................................................................................... 82

8 Conclusion ................................................................................................................ 87
1

1 INTRODUCTION
A well designed and maintained transportation system is very important for any

economy. People and goods move through transportation networks; thus, the better

designed and maintained the system, the faster and the safer they may be transported

from one location to another. An efficient transportation system has not only to be safe

and provide comfort to its users, but it also has to be durable.

Building a transportation system requires large investments because of its

extensive infrastructure which may include streets, highways, parkings, bridges, airports,

and railways. Therefore, making all this infrastructure last longer is a must for sustaining

economic growth.

More specifically, highways have been extensively built and used in America. As

an example, interstates have been crucial for the economic growth in the United States.

Even though it is estimated that interstates correspond to about 11% of the total mileage

of America’s transportation network, more than 70% of the total travel passes through

them (AASHTO, 2004). Additionally, these highways have the highest percentage of

heavy-vehicle traffic (Figure 1). Therefore, it is imperative that the infrastructure devoted

to such high-speed arterials be durable; otherwise, maintenance costs would be extremely

high.
2

Figure 1. Heavy Vehicle and Truck Traffic on I-20/59.

However, increasing the durability of highways has never been simple. For

instance, it is not easy to increase durability of highways when pavements deteriorated in

premature age. Even though deterioration is a natural process of pavement aging,

pavement repair and reconstruction are major maintenance cost generators for

transportation agencies; but, obtaining optimal pavements has been a complex task. Such

task depends not only on the design, but also on factors that often are beyond designers’

control such as quality of construction materials and construction methods. Nevertheless,

roadway pavements are designed to last a pre-determined number of years ranging from

15 to 25 years, commonly called pavement service life. If the pavement deteriorates

sooner, not meeting its service life, it results in extra maintenance or even reconstruction

costs. The consequences of such deterioration will be a poor pavement infrastructure,

implying delays or accidents.


3

In order to minimize such occurrences, designers stablish a failure criterion,

meaning the pavement service life is over. A criterion relates to a particular distress on

the pavement structure, which causes a specific failure. There are five major problems

that must be concerned about in order to minimize the chances of a pavement structure

failure. These problems are: Alligator or Fatigue Cracking, Rutting, Top Down-Cracking,

Bottom to Top Cracking, and Thermal Cracking. Figure 2 illustrates fatigue cracking on a

one-lane roadway.

Figure 2. Fatigue Cracking

Designers have adopted three different methods to address such problems:

Empirical method, Mechanistic-Empirical method, and Mechanistic method. Historically,

empirical methods prevail in the pavement community. With the advance of pavement

mechanics, these methods gave place to Mechanistic-Empirical methods. These days

however, there are even more reliable ways to predict pavements behavior in terms of
4

stresses, strains and different material models. In this study, an investigation of a

Mechanistic-Empirical and a Mechanistic method is conducted using two distinct

softwares and a comparison between both methods is performed

1.1 Study Objective

The primary objective of this study is to compare two different asphalt pavement

design software. The first one is a computer software named KENLAYER, developed at

the Kentucky University (Huang, 2004) and it is based on a mechanistic-empirical

analysis. The second model is termed Mechanistic Roadway Model which uses an in-

house finite element code (Allen, 1994). This comparison will allow researchers to

investigate the observed result differences between these two design tools. Ultimately,

the findings of this study may provide direction for future roadway pavement design

methods in phase of implementation and to be developed; as well as motivation for

pavement designers to utilize the more efficient design tool.

1.2 Study Scope

This study is presented in eight chapters as follows. The first chapter is the

introduction. The second chapter provides a background on flexible pavement structure

and distresses, design and analysis methods and failure criteria commonly used. It also

gives an outline in finite element methods, and cohesive zones. The following chapter

presents an overview of KENLAYER (Huang, 2004) with emphasis on its solution

method. Chapter four presents an overview of the Mechanistic Roadway Model with

emphasis on its solution method. Chapter five is the finite element formulation used to

address the research problem implemented on the Mechanistic Roadway Model. The

sixth chapter describes the analysis, explains the roadway geometry and loading adopted
5

on the case study and discusses the results obtained from both programs and chapter

seven presents the final conclusions.


6

2 BACKGROUND
A summary of relevant information as well as findings from previous research

studies are presented below. This chapter is divided into six sections. The first section

gives a brief description about flexible pavement structure. The second and third sections

describe the type of distresses acting on flexible pavements and the methods used for

analysis and design of these structures, respectively. The concept and illustration of

failure criteria is introduced in Section 2.4. Relevant information about finite element

methods adopted in the field of pavement research is provided in Section 2.5. Finally,

Section 2.6 discusses information on cohesive zones.

2.1 Pavement Structure

Roadway pavements are complex structures composed of multiple layers with different

properties. Each layer is composed of a material that has a particular structural behavior which

makes pavement design and analysis a real challenge for engineers. Three pavement types are

commonly used: flexible or asphalt pavements (Figure 3), rigid or concrete pavements (Figure 4)

and composite pavements (Figure 5).

Flexible pavements are composed of multiple layers. The layers are set such that

the one with the highest load bearing capacity (e.g., asphalt layer) is placed on top and

the layer with the lowest load bearing capacity is the subgrade which is the last layer

from top to bottom. This type of pavement is widely used and it allows designers to have

certain flexibility in using different material combinations for each layer which may

allow construction costs to be reduced. The present research focuses on flexible

pavements only.
7

Figure 3. Flexible pavement

Figure 4. Concrete Pavement


8

Figure 5. Composites Pavement

2.2 Type of Distresses

Roadways are subjected to repeated moving loading in a relatively short of time.

This loading often has a slow but compromising impact on the pavement throughout the

years, which results in a variety of distresses on the pavement structure and surface. In

this section, some of these distresses are discussed as follows.

2.2.1 Alligator or Fatigue Cracking

Fatigue is the process of damage and failure due to cyclic loading. Such loading

can be applied during a long period of time, generating microscopic physical damage in

the structure that eventually leads to crack development and/or other macroscopic

damage of the structure. Even though the cyclic stresses may be below the ultimate

tensile strength of the material, the roadway tends to degrade after repeated loading even

at a stress level below the nominal strength of the material.

Figure 6 shows sinusoidal loading with the minimum and maximum stresses both

in the tensile range. This is a similar case to what occurs at the bottom of the asphalt layer

under loading. Such loading cycles may lead to cracking on the bottom of the asphalt

which eventually propagates to the top as the structure weakens.


9

Figure 6. Minimum and Maximum Stresses in the Tensile Zone

Fatigue cracking is a common roadway pavement distress. The loading applied by

the truck tire generates a critical tensile stress at the bottom of the asphalt layer under the

vehicle path. This stress tends to distend the bottom surface of the structure

perpendicularly to the direction of the load as shown in Figure 7. Once the bottom face of

the layer is under tension, cracks propagate from bottom to top until they reach surface

where longitudinal parallel cracks are formed. Ultimately, these cracks propagate at an

increasing pace until they connect to each other. At this stage, the pavement surface looks

like an alligator skin as shown in Figure 8. Potholes can eventually be formed due to

progressive fatigue cracking. In this case, whole pieces of asphalt layer are detached from

the pavement causing the formation of small-to-middle sized bowl-shaped holes which

can cause much vehicular damage and pose unsafe situations for drivers as they try to

deviate from these sharp depressions (Figure 9).


10

Figure 7. Tension Stress at the Bottom of the Base Layer

Figure 8. Advanced Alligator Cracking Stage


11

Figure 9 . Pothole on Asphalt Pavement Surface

2.2.2 Rutting

Rutting is characterized by a permanent deformation in the wheel path as shown

in Figure 10 and Figure 11. Rutting occurs due to deformations in the asphalt layer or in

the subgrade. These deformations may occur due to improper compaction and asphalt

mixture composition, or due to compressive cyclic loads on top of the subgrade (Figure

12). Excessive rutting may cause drainage difficulties since water may accumulate on

wheel path depressions, which may cause water infiltration into the pavement structure as

well as potential for hydroplaning. Furthermore, rutting can eventually cause water

bleeding which is characterized by water pumping through the cracks, if the pavement

has high porosity.


12

Figure 10. Rutting on Asphalt Pavement

Figure 11. Permanent Deformation, Transversal View


13

Figure 12. Compressive Strain at the Top of the Subgrade

2.2.3 Top-Down Cracking

Top-Down Cracks (Figure 13) are longitudinal and/or transverse cracks that

initiates at the surface of asphalt pavement and propagate downward. Climatic

conditions, traffic, ageing, structure and construction qualities are the main causes

pointed out to the initiation and propagation of this distress. Contrary to what occurs in

fatigue cracking, cracking originating at the top of asphalt layers is a degradation

mechanism which has not been largely studied because load-associated fatigue cracking

of hot-mix asphalt (HMA) that occurs in the wheel path, has been thought to always

initiate at the bottom of the HMA layer, where the tensile strain is highest under a wheel

load. It has been previously thought that such cracking would propagate to the surface,

forming one or more longitudinal parallel cracks that after many load repetitions are

connected and form many-sided, sharp-angled pieces that reassemble the skin of an

alligator. However, recent studies have determined that load-related HMA fatigue cracks
14

can also be initiated at the surface of the pavement and propagate downward through the

HMA layer. This mechanism has been called Top-Down Cracking (Harmelink, 2008).

Figure 13. Top Down Cracking

As in some other distresses, Top-Down Cracking should be avoided or at least

minimized by adopting some measures that are found in the literature. Among the

measures, laboratory tests to characterize the material that will be used in the project are

of extreme importance. There are some tests that are considered important and that

should be indispensable in any pavement design process. Some of these tests are:

Flexural Test and Tensile Strength Test. The Flexural Test may be used to estimate in-

place HMA fatigue properties. The flexural test determines the fatigue life of a small

HMA beam specimen (380 mm long × 50 mm thick × 63 mm wide) by subjecting it to

repeated flexural bending until failure. Results are usually plotted to show cycles to

failure versus applied stress or strain (AASHTO TP). The Tensile Strength Test is also a
15

good indicator of cracking potential. There are three tests that may be used to measure

tensile strength of an asphalt specimen. The Indirect Tension Test applies a constant rate

of vertical deformation until failure. This test can be seen in the AASHTO TP 9-96

(Determining the Creep Compliance and Strength of Hot Mix Asphalt Using the Indirect

Tensile Test Device). The Thermal Cracking Test determines the tensile strength and

temperature at fracture of an HMA sample by measuring the tensile load in a specimen

which is cooled at a constant rate while being restrained from contraction. The test is

terminated when the sample fails by cracking. This test can be seen in AASHTO TP 10

(Method for Thermal Stress Restrained Specimen Tensile Strength). The Stiffness test

determines an HMA's elastic or resilient modulus. Although these values are fairly well-

defined for many different mix types, these tests are still used to verify values, determine

values in forensic testing or determine values for new mixtures or at different

temperatures. Many repeated load tests can be used to determine resilient modulus as

well.

2.2.4 Thermal Cracking

Thermal cracking of asphalt pavements is a serious problem in Canada and the

northern parts of the United States. On many occasions, either at the design stage or

during service, highway agencies demand a forecast of pavement performance rating,

which is highly sensitive to the intensity of cracking. There are two failure modes for

thermal cracking: low-temperature cracking and thermal-fatigue cracking. Low-

temperature cracking is caused by accumulated thermal stresses in the pavement layer

during cold winters or spring thaws. Thermal-fatigue cracking (Figure 14) is caused by

daily cyclic thermal loading. Classical probabilistic approaches have been applied to
16

pavement design systems, including thermal cracking predictions, during the past three

decades. Advances in reliability analysis, however, have proven that classical reliability

methods are inconsistent, and mandate that current design procedures should be revised

accordingly.

Figure 14. Thermal Cracking

2.2.5 Other Distresses

There are other distresses that may not lead to a major structural failure such as

that caused by fatigue cracking and rutting. However, these types of distresses may still

be deleterious to the pavement life and, in some instances, may require immediate repairs

due to the excessive discomfort they generate or because of the potential accident threats

they pose. Bleeding, raveling, and corrugation or shoving are some of these other types of

distresses.
17

Bleeding is characterized by the presence of a shiny surface formed by application

of excessive asphalt in the mixture as shown in Figure 15. Although bleeding may not

lead to structural failure, it may cause loss of skid resistance which may make the

pavement unsafe for driving mainly when the surface is wet.

Figure 15. Bleeding

Raveling (Figure 16) is characterized by the disintegration of the asphalt layer as

a result of aggregation segregation and poor compaction. Raveling is a failure that may

cause excessive roughness and discomfort for users.


18

Figure 16. Raveling

Corrugation (Figure 17) is formed by plastic movements of the asphalt layer due to

starting and stopping vehicle movements associated with poor asphalt mixture design

and/or moisture in the subgrade. This distress can affect safety on high-speed facilities as

sharp bumps are formed.


19

Figure 17. Corrugation

2.3 Pavement Design and Analysis Methods

There are three major procedures used to design pavements: empirical design,

mechanistic-empirical design, and mechanistic design. Even though empirical designs

were widely adopted during the last century, recent research (Harichandran et al, 2001)

has pointed to the advantages of mechanistic-empirical and mechanistic methods. These

three methods are discussed as follows.

2.3.1 Empirical Pavement Design

In the early 1900s, the choice of input loads, geometry and material properties in

pavement design was based on empirical evidence, such as experience and experiments

(AASTHO, 1972). Because scientific theory was not utilized, this method is

appropriately called empirical design. Nevertheless, this method has been used most

often for the design of pavement design.


20

The most famous empirical method is the one developed by the American

Association of State Highway and Transportation Officials (AASHTO) and presented in

its design manual in 1961. AASTHO performed a full scale road test in Ottawa, IL from

1956 – 1961 for rigid and flexible pavements. An experimental road section was build

and eventually became part of I-80 (Figure 18). Thickness of the pavement and moving

load were controlled. A large amount of data was collected from this four year test and

empirical correlations were developed relating the service life of the pavement submitted

to different loads and material properties.

Figure 18. AASTHO Road Test sketch

However, this method only works for a predetermined set of load, geometry and

materials properties. When a new variable is introduced, a new pavement thickness and a

new set of experiments would be necessary. Despite this shortcoming, this method is still

in commonly use for pavement design.


21

2.3.2 Mechanistic-Empirical Pavement Design

In order to overcome the weaknesses of the empirical design, the mechanistic-

empirical pavement design methods arose in late 1980s. These methods are more

elaborate because they do not rely only on experiments and on practical experience, but

also on theoretical fundamentals such as stress predictions. It is based on the multi-

layered theory developed by Burmister in 1944. This theory was developed having in

mind problems in airport runways and the goal was to predict the expected magnitude of

the displacements and the stresses imposed by the loading.

The multilayer system (Figure 19) assumes a three layer system with circular load

area. The first two layers are considered to be of known thickness, but of infinite width;

and the third layer considers both thickness and width infinite. This method assumes that

materials for all layers are isotropic (same properties in all directions), homogeneous

(spatially uniform composition), and linear elastic. It also assumes that there are no

stresses and displacements in the third layer.

Figure 19. Multilayer System


22

In contrast to the empirical method, the mechanistic-empirical approach predicts

displacements and stresses on the materials. Material properties, road thicknesses and

loads can be changed without the necessity of extensive new road tests. Predictions can

be done in a faster way, thus saving not only time but also money.

Currently, the mechanistic-empirical method is the most used method in

pavement design and analysis (NCHRP, 2004). It has had a significant importance in the

field and has result in the ability to better understand the pavement behavior. However,

the assumptions present in this method, usually contribute to an under prediction of the

pavement service life. Considering the roadway as a half-space with circular load and

isotropic homogeneous elastic materials is not the best solution for problem. In fact, loads

applied by trucks are not circular but present sinusoidal shape, and many of the pavement

materials do not behave as elastic material but as viscoelastic and viscoplastic.

2.3.3 Mechanistic Pavement Design

More recently, researchers have developed theoretical procedures which rely

primarily on mathematical and physical principles, in an attempt to predict pavement

performance with more accuracy (C&T Research Records, 1998). Due to this fact,

researchers believe that a mechanistic pavement design may not only provide more

reliable results, but it can also be used over a broad range of, loading and material

characteristics (Soares, 2005). Most mechanistic pavement design procedures use mainly

numerical methods as backbones, and adopt the Finite Element Method (FEM)1 to obtain

numerical solutions.

1
A more detailed description of Finite Element Method can be found in section 2.5
23

FEM consists in dividing a complex geometry, in this case a pavement structure

into a large number of finite elements, connected by nodes. Such configuration is named

“mesh” and is illustrated in Figure 20 and Figure 21. This method allows engineers to

analyze the pavement structure in as many layers as desired, as well as analyze the

loading effect and displacements on a micro-scale. FEM can also account for crack

initiation and growth using cohesive zone models (CZM)2, which often leads to a more

accurate prediction of pavement service life because the crack behavior can be predicted

more precisely.

Figure 20. Pavement Mesh Example in 2D

Figure 21. Pavement Mesh Example in 3D

Besides its powerful capability to solve pavement problems with different

materials, loads and geometries, the mechanistic pavement method is still not widely

2
A more detailed description of Cohesive Zone Model can be found in section 2.6
24

used. Its complexity, computer requirements and processing time are not attractive to

pavement design engineers. Although it appears to have the potential to be superior when

compared to mechanistic-empirical methods, currently some approximations made by the

mechanistic method are usually introduced in order to solve the problem in a faster way,

and these approximations can compromise accuracy.

2.4 Failure Criteria

Over the years, numerous empirical correlations have been developed to predict

the number of load cycles necessary to cause failure of pavement (Yoder, 1975; Scrivner,

1968). Although these correlations are not physically based, they are widely used in

mechanistic-empirical pavement design. The two most common types of failure are

fatigue cracking and permanent deformation (or rutting). Most equations used to predict

the number of cycles to cause failure by fatigue use tension strain at the bottom of the

asphalt layer and the elastic modulus of the asphalt layer as inputs. A popular equation

was developed by Finn et al. (1977)

   E 
log N f  15.947  3.291log  t6   0.8541log  3  (2.1)
 10   10 

where N f is the number of cycles to cause failure in fatigue;

 t is the tension strain at the bottom of the asphalt layer;

E resilient modulus for elastic layer.

According to equation , the pavement fails when fatigue cracking covers 10% of the

wheel path area. Another failure criterion is due to permanent deformation. To predict

failure due to permanent deformation, a common method is based on the vertical strain at
25

the top of the subgrade layer, which is believed to be responsible for this type of failure.

The number of cycles to predict permanent deformation can be predicted by

4.4843
 106 
N d  1.077 x10 
18
 (2.2)
 v 

where N d is number of cycles to cause failure in permanent deformation;

 v is vertical strain at the top of subgrade.

In this particular equation, the pavement is considered failed when the

deformation reaches 0.5 inches on the surface in the wheel path. The failure criteria

presented here are specific for the two equations described. Although both equations are

not necessarily valid for a different set of inputs, most pavement designers rely on this

approach to predict pavement life.

2.5 Finite Element Methods

Multi-layer theory is the most popular method to predict pavement response to

vehicle loading. Besides being widely accepted due to its simplicity and capability to

obtain rapid results, this method utilizes numerous simplifying assumptions that make its

results questionable. Such assumptions are: homogeneous isotropic elastic layers; circular

loading area; and uniform pressure (Burmister, 1945). In order to overcome inaccuracies

introduced by these simplifying assumptions, numerical techniques such as finite element

method became quite popular in pavement applications

The term finite element method was first used in 1960 by R. W. Clough in his

paper The Finite Element Method in Plane Stress Analysis. The FEM method arose in

early 1940s with the necessity to solve problems with complex structure. Currently, due
26

to capability as an analysis tool, this method is used not only in the engineering field but

widely in many areas such as aeronautical, biomedical (Figure 22), and automotive

industries.

It is much acclaimed for its capability to give approximate solutions for complex

geometries that are normally not feasible by other methods. As it can be seen in Figure

23, this method consists of dividing the geometry of interest in finite elements and

connecting them by nodes. Approximate solutions to the problem of interest can then be

obtained by solving partial differential equations (PDE) using a numerical method.

Figure 22. Finite Element Method Used in Biomedical Field. Courtesy SCI Institute.

Figure 23. Finite Element Method – Mesh generation


27

However, besides its many positive features, FEM also presents some limitations.

For example, mesh generation of a large number of degrees of freedom often requires a

large amount of time to be completed. Also, accuracy is related to the size of the mesh

(number of elements and nodes), usually requiring a mesh convergence study. When the

problem presents multiple variables, such as pavement problems, the solution requires an

extensive amount of computer processing time.

2.6 Cohesive Zone Model

Fatigue cracking is one of the main causes of failure in structural components.

Over the years, a number of structures have collapsed due to fatigue cracking. One very

known example is the accident involving the Aloha Airlines flight 243 (Figure 24) in

Hawaii. Every take off and landing subjects the aircraft to cyclic pressure loads which

eventually leads to crack initiation and growth. In this particular case, the number of

cracks reached a limit and the structure collapsed in 1988 causing the accident.

Figure 24. Aloha Flight 243


28

Cracks are not only found in aircrafts but also bridges, concrete and pavements. In

pavement, cracks are present nearly in every layer. However, the number of cracks a

pavement can withstand is not the main concern, but when cracks coalesce sufficiently to

allow water penetration. When this point is reached, water weakens the foundation and

contributes to pavement deterioration and eventual failure.

To prevent or minimize these problems, it is important to accurately model

cracks. Cracks create new surfaces along the material and they require some dissipation

of energy in order to be created. If there is not enough energy available the crack will not

form. Griffith (Griffith, 1921) was the first one to formulate this concept for crack

growth. He stated that when the energy required for crack growth (G) is larger than the

material's resistance to crack growth (R), the crack will develop (Figure 25).

GRa0 (2.3)

where a is the crack length. This concept is the basic assumption of linear elastic fracture

mechanics (LEFM). Although LEFM has a significant importance in science, it is often

inaccurate for materials with rate-dependency.

Figure 25. Crack Propagation


29

In order to address the shortcomings of LEFM, the cohesive zone model (CZM)

was developed (Dugdale, 1960 and Barenblatt, 1962). In this powerful method, crack

nucleation, initiation and propagation can be modeled computationally to simulate mode

I, II, and mixed-mode cracks (Figure 26). CZM is based on the assumption of existence

of cohesive tractions acting in the fracture process ahead of the crack tip. This approach

has been used to model a variety set of materials, including viscoelastic asphaltic

materials (Allen and Searcy, 2000).

Figure 26. Mode I, Mode II and Mode III cracks

Ahead of the crack tip, there is a region termed as damage zone, where internal,

time-dependent boundaries are created by local void formation, fibrillation, and

coalescence. The cohesive zone (CZ) (Figure 27) is the mechanically equivalent two

dimensional surface of the damage zone. In this method, it is assumed that the damage

zone can be approximated by a region of zero thickness and cohesive tractions.

Cohesive zones play an important role in predicting fracture behavior of a

structural material. This enables pavement engineering to improve service life of the

structure, by simulating crack initiation and propagation.


30

Figure 27. Cohesive Zone


31

3 KENLAYER

3.1 Overview of the KENLAYER software

KENLAYER is a computer program developed by Huang at the University of

Kentucky, (Huang, 2004) designed to predict life in elastic multilayer systems subjected

to axisymmetric loads. This program is applicable to asphaltic pavements or rigid layers.

Single, dual, dual-tandem and dual-tridem wheels are among the options for truck load

configuration. Material properties for each layer include linear and nonlinear elastic and

viscoelastic. Each layer is initially specified as linear elastic, unless specified otherwise.

Damage is included in the form of fatigue cracking and permanent deformation and then,

the pavement design life is assessed by the accumulation of damage. There is also an

option to divide each year in up to twelve increments; in which each period can be set

with different material properties and different load groups. Therefore, variations suffered

by the pavement in volume of traffic and temperature along the year can be simulated.

The solution obtained from the elastic multilayer analysis is solved for each truck tire

individually. The damage caused by fatigue cracking and permanent deformation in each

period over all load groups is summed up to evaluate the design life.

The nonlinear elastic option requires the specification of nonlinear layers because

they are stress dependent. Also, an iterative procedure is used, in which while the moduli

of nonlinear layers are adjusted (as the stresses vary), the moduli of linear layers remain

the same. During each iteration, a constant set of moduli is computed from the stresses

obtained from the previous iteration, so that the problem is considered linear. The

software calculates the elastic moduli of nonlinear layers as well as determines the new
32

set of stresses after determining stresses due to single multiple wheels. The process is

repeated until the moduli converge to a specified tolerance.

When specifying a linear viscoelastic layer, KENLAYER uses creep compliance

as an input parameter. Solutions for moving or stationary loads are also included.

3.2 Solution Method

Roadway problems deal with mixed boundary conditions, a combination of

displacement and traction boundary conditions, and a complex geometry which makes

the solution sometimes only feasible by numerical computation. However, KENLAYER

makes some assumptions in order to simplify the problem. Those assumptions and how

they affect the results will be described later.

The first step in this study is to mathematically explain the method adopted by

KENLAYER, using the approach believed by this author to be the more appropriated for

solving mechanist-empirical problems. The next step is pose the boundary value problem

by deriving governing field equations and expressing all variables involved. Some

variables are known a priori while others will be determined by the field equations. The

formulation described herein is based on the continuum mechanics approach where the

body is considered as a continuum and neglects the structure of materials on a scale much

smaller than the characteristic scale of the problem.

Kinetics, kinematics and constitutive equations are first written in cartesian

coordinates, to describe some of the assumptions made. Then all equations are converted

to cylindrical coordinates since KENLAYER uses elastic multi-layer theory (Burmister,

1943), where the pavement configuration is assumed as a cylinder (see Figure 28) which

permits, by assuming axisymmetric response, to reduce the number of equations and


33

consequently facilitate the solution of the problem. After all equations are obtained from

the equations, additional constraints can be obtained from the boundary conditions.

The field variables (also called state variables) are represented by stress tensor,

 ij ( x , t ) , (Figure 29), strain tensor,  ij ( x , t ) , and displacement vector ui ( x , t ) . x is the

spatial coordinate within the body and t is the time of interest.

Figure 28. KENLAYER Pavement Configuration

Figure 29. Stress Tensor in Cartesian Coordinates


34

Assuming the problem is an elastic problem, the above fifteen field (unknown)

variables need to be found, requiring the same number of field equations in order to solve

the problem. The conservation of linear momentum is given by

 2ui
 ji , j   fi   2 (3.1)
t

where  is the mass per unit volume and f i is the body force per unit of volume.

Assuming a quasi-static problem equation (3.1) reduces to

 ji , j   fi  0 (3.2)

The strain-displacement solution is given by

 ij 
1
2
ui, j  u j ,i  (3.3)

assuming there is only small deformation.

The constitutive equation for a linear elastic material is given by:

 ij   kk ij  2 ij (3.4)

where  and  are Lamé constants and assuming isotropic media.

Conditions of compatibility, imposed on the components of strain are necessary

and sufficient to ensure a continuous single-value displacement field. Based on that, the

displacements inside kinematic equations are eliminated to produce equations with only

strain as unknowns. These interrelations are known as Saint-Venant’s compatibility

equations and are represented by:

 ij ,kl   kl ,ij   ik , jl   jl ,ik (3.5)


35

Constraints imposed by boundary conditions can be applied on a body as a

displacement, traction or a mixture of both. The first approximation made by

KENLAYER is to assume that, due to the fact that the roadway length in the lane

direction is larger than the other directions; its effect on the load is negligible. The second

approximation is to assume that the traction is known everywhere in the roadway. These

assumptions convert the problem to a traction only boundary condition given by:

ti   ji n j (3.6)

where t i is the traction per unit area on the surface and n j is the unit outer normal.

The tire-pavement contact area is a main concern when designing a pavement.

The tire configuration can vary with different applied loads, shown in Figure 30, and also

with inflation, pressure and tire type, which makes the problem difficult to solve. Based

on that, KENLAYER assumes the vehicle load to be circular and evenly distributed,

making the problem axisymmetric.

Figure 30. Tire Configuration for Different Applied Loads. Courtesy Tekscan, Inc.
36

Due to its axisymmetric characteristic, the problem need to be converted to

cylindrical coordinates. Also, the governing equations should be rewritten in cylindrical

coordinates (see Figure 31 and Figure 32).

The equilibrium equations (3.2) in cylindrical coordinates are given by:

 rr 1  r  rz 1
    rr      f r  0 (3.7)
r r  z r

 r 1     z 2
    r  f  0 (3.8)
r r  z r

 rz 1   z  zz 1
    rz  f z  0 (3.9)
r r  z r

where  rr ,   and  zz are the normal stress tensors in the r ,  and z directions

respectively;  r ,   z and  rz are the shear stresses and f r , f  and f z are the body

force per unit volume in the same directions.

Figure 31. Cylindrical Coordinates


37

Figure 32. Stress Tensor in Cylindrical Coordinates

The strain-displacement equations in cylindrical coordinates are given by

ur
 rr  (3.10)
r

1 u ur
   (3.11)
r  r

uz
 zz  (3.12)
z

1  1 u u u 
 r   r
    (3.13)
2  r  r r 
38

1  u 1 u 
 z     z
 (3.14)
2  z r  

1  u u 
 rz   z  r  (3.15)
2  r z 

where  rr ,   and  zz are the strain tensors in the r ,  and z directions respectively;

 r ,   z and  rz are the shear strains and ur , u and u z are the displacement vector in

the same directions.

The constitutive equations are given by:

 rr     rr     zz   2 rr (3.16)

      rr     zz   2 (3.17)

 zz     rr     zz   2 zz (3.18)

 r     rr     zz   2 r (3.19)

  z     rr      zz   2  z (3.20)
39

 rz     rr      zz   2 rz (3.21)

The compatibility equations in cylindrical coordinates are given by:

 2 rr  2 zz  2 rz
 2 2 0 (3.22)
z 2 r rz

 2 2  2 z 1  2 zz 1  zz 2  rz
    0 (3.23)
z 2 r z r 2  2 r r r z

 2 rr  rr  2  r r    2  2 
 r  2  r 0 (3.24)
 2 r r r  r 

 2 r  2  1  1  2 rz  2 zz  1 
 r   z      zz   0 (3.25)
z 2 rz  r  r z r  r 

1  2 rr 2  1    1   1   rz 2
2
 
 rz  
r z r  r  r  r r
 r 
z  
 r rz
2  r  r   0 (3.26)

 2 rz  rr 2 2 2
r   r z   r  r    r r   0 (3.27)
 2 z r rz z
40

Assuming axisymmetry, the functions are independent of the coordinate  .

Therefore,  r   z  0 and u  0 . Neglecting body forces, where f r  f  0 ,

several terms in equations (3.7) through (3.27) become negligible, which reduces the

number of equations and unknown variables.

Equations (3.7) through (3.9) are reduced to:

 rr  rz 1
   rr      0 (3.28)
r z r

 rz  zz 1
   rz  f z  0 (3.29)
r z r

Equations (3.10) through (3.15) are reduced to:

ur
 rr  (3.30)
r

uz
 zz  (3.31)
z

ur
  (3.32)
r
41

1  u u 
 rz   z  r  (3.33)
2  r z 

Equations (3.16) through (3.21) are reduced to

 rr     rr     zz   2 rr (3.34)

      rr     zz   2 (3.35)

 zz     rr     zz   2 zz (3.36)

 rz  2 rz (3.37)

The fifteen unknowns and field equations are thus reduced to ten unknowns and

ten field equations, represented by equations (3.28) through (3.37).

Equations (3.22) through (3.27) are reduced to

 2 rr  2 zz  2 rz
  2 0 (3.38)
z 2 r 2 rz

 2 1  zz 2  rz
  0 (3.39)
z 2 r r r z
42

 rr   2  2 
r  r 0 (3.40)
r r  r 

 rr 2
r r  r   0 (3.41)
z rz

Since a vertical uniformly distributed circular load is considered, the surface

boundary conditions due to a circular load of radius a (Figure 33) are given by

 zz (r , 0)  q
r  a (3.42)
 rz  0

 zz (r , 0)  0
r  a (3.43)
 rz  0

where q is the external uniformly distributed vertical load.

At the interface layers, the boundary conditions (Figure 33) are given by

ur(i ) (r , hi )  ur( i 1) (r , 0),


u z(i ) (r , hi )  u z( i 1) (r , 0),
(3.44)
 zz(i ) (r , hi )   zz(i 1) (r , 0),
 rz(i ) (r , hi )   rz(i 1) (r , 0)

where hi is the thickness of layer i.


43

Figure 33. Boundary Conditions

In this case wherein all boundary conditions are of matching type, the above

compatibility equations are a necessary condition to guarantee that the displacement field

is unique. We will henceforth consider only tractions, the so-called Neumann boundary

conditions (Cheng, 2005). Substituting equations (3.34) through (3.37) into compatibility

equations (equations (3.38) through (3.41)) and later substituting equations (3.28) and

(3.29) into the result gives

1    rr    rr 1   rr      zz 
2 2

r   0 (3.45)
r r  r  z
2
1  r 2
44

1        
2

r  0 (3.46)
r r  r  z
2

1    zz    zz 1   rr      zz 
2 2

r   0 (3.47)
r r  r  z
2
1  r 2

1    rz    rz 1   rr      zz 
2 2

r   0 (3.48)
r r  r  z
2
1  r z

which are known as the Beltrami-Michell stress compatibility equations (Patnaik, 2007)

E E
where,   and   .
1  1  2  2 1  

By completing the previous substitutions, the current ten field equations with ten

unknowns are thus reduced to four unknowns and four equations.

The four equations above (equations (3.45) through (3.48)) were used to

determine the four components of stress  rr ,   ,  zz and  rz subject to the boundary

conditions. The next step now is to formulate the problem in a dual mathematical space

which will make the problem simpler to solve.

According to Love (Love, 1926), when a solid of revolution is strained

symmetrically by forces applied at its surface, we may express all quantities that occur in

terms of a single function, and reduce the equation of equilibrium of the body to a single

partial differential equation. Based on that, the stresses can be expressed by


45

 2
 rr  2  (3.49)
r 2

1  2
     
2
(3.50)
r r

 2
 zz   2   2  (3.51)
z 2

where  is Poisson’s ratio and  is the primary variable in this dual space.

Substituting equations (3.49) and (3.50) into equation (3.48) allows  rz to be

expressed in terms of a function  such that   / z . Based on that, equations (3.49),

(3.50) and (3.51) become:

 2 2  
 rr     2  (3.52)
z  r 

  2 1  
       (3.53)
z  r r 

 2  
 zz        
2
2  (3.54)
z  z 2 
46

Substituting equations (3.52), (1.53) and (1.54 ) into equation (3.48) gives

 2  
 rz   1    2
   (3.55)
r  z 2 

Plane strain compatibility condition in term of stress can be expressed as

1
2 (  )  f , (3.56)
1 

where  2 (  )    ,

Substituting equations (3.52) through (3.55) into equation (3.56), and considering

no body forces, Beltrami-Michell equation given by

4   0 (3.57)

ise satisfied.

From stress-strain relations the displacements are given by

1   2  
ur     (3.58)
E  rz 

1    2  1  
uz       2
  
E  r 2 r r 
1 2 (3.59)

The formulation in thus are believed by the author to be the most appropriate to

obtain analytic predictions of the stresses, strains, and displacements.

3.3 Damage Analysis

Over the years, numerous empirical correlations have been developed to predict the

number of load cycles necessary to cause failure of pavement. Although these equations

are not entirely physically based, they are widely used in mechanistic-empirical pavement
47

design. The two most common types of failure are fatigue cracking and permanent

deformation (or rutting). KENLAYER, as most mechanistic-empirical designs, also uses

these failure criteria to analyze and predict pavement life.

In the mechanistic-empirical method, each failure criterion should be calculated

separately to account for each failure mode. KENLAYER predicts fatigue cracking and

permanent deformation separately. The weakest result obtained from both predictions

will be considered as the failure criteria and will be used to predict the design life.

In order to explain the failure criteria adopted by KENLAYER, it was decided, by

this author, to use a continuum mechanics approach to mathematically explain how the

number of load repetitions to prevent, fatigue cracking and permanent deformation can be

obtained based on fracture mechanics principles. The formulation described herein is

used to predict the number of load repetitions to prevent both, fatigue cracking and

permanent deformation.

The conservation of energy (first law of thermodynamics) is given by

u   ij ij  qi ,i   r (3.60)

where qi ,i is the heat flux per unit area.

The second law of thermodynamics is given by

 q  r
s   i   0 (3.61)
 T  ,i T
where T is the temperature.

In fracture mechanics, the crack growth is considered the most dominant over the

whole fatigue process. When a crack is running in a elastic body and there is no energy

dissipated from chemical changes, a statement of the second law (equation (3.61)) can be

written as (Quoc Son, 1973),


48

(G  GIC )a  0 (3.62)

where G and GIC are the energy release rate and the critical energy release rate

respectively and a is the crack length.

The energy release rate is the rate at which energy is absorbed by growth of the

crack. Therefore, when

G  GIC  a  0 (3.63)

meaning the crack can, but not necessarily will propagate. Equation (3.63) is known as a

weak statement. Rewriting equation (3.63) as

G  GIC  a  0 (3.64)

(Griffith, 1921), implies that the crack must propagate, making equation (3.64) a strong

statement of the second law.

Based on the Coleman-Noll procedure (Coleman-Noll , 1959), it can be assumed

that

a  a(G) (3.65)

where G  energy release rate.

Using the chain rule, it can be written that

dt da dt
 A(G)n . (3.66)
dN dt dN

where it is assumed that the functional dependences is in the form of a power law.

Thus, from the above development it can be seen that the crack growth rate is given by
49

da
 C (G)n (3.67)
dN

dt
where C  A .
dN

Equation (3.67) is a special case of equation (3.65). N is the number of load

repetitions and, C and n are curve fitting parameters. Equation (3.67), well known as the

Paris Law (Paris, 1961), is frequently used to represent fatigue crack growth and it has

been shown to be reasonable for some materials.

A relationship between the energy release rate and the stress intensity factor can

be given by (Tada, 1973)

(1  2 )K 2
G (3.68)
E

where K is the stress intensity factor, and  and E are Poisson’s ratio and Young

modulus respectively. Equation (3.68) is applicable only to quasi-static linear elastic

materials, under mode I crack growth and plane strain conditions.

For mode I crack growth, the critical stress intensity factor can be given by (Tada,

1973)

KI     a (3.69)

where  is the far-field load.


Integrating equation (3.67) gives

Nf acrit
1
 dN   C  G 
n
da (3.70)
0 a0

Substituting equation (3.69) and (3.68) into (3.70) gives


50

n acrit

1  1      a  
2  n 1
N     .    (3.71)
C  E   n  1 
a 0

Writing  is a function of  and E , and assuming  , C , a and n can be

represented by constants f1 , f 2 and f3 , the number of load repetitions to prevent fatigue

cracking N f can be given by

N f  f1 ( )  f2 ( E )  f3 (3.72)

In KENLAYER, f1 , f 2 and f3 are used to represent constants from laboratory

tests, and  is the tensile strain on the bottom of the asphalt layer.

Assuming permanent deformation only occurs when energy is dissipated and

applying the same concept used to obtain N f , the number of load repetitions to prevent

permanent deformation N d can be given by

Nd  f 4 ( ) f5 (3.73)

where in KENLAYER  is the compressive strain on the top of subgrade, and f 4 and f5

are constants determined from laboratory tests.

KENLAYER, as mentioned before, is based in the elastic layered theory, and

assumes the pavement to be composed of linear elastic layers (which are characterized by

the time-independent constants of proportionality between stress and strains) on a semi-

infinite elastic subgrade. For most paving materials, this assumption is valid only under

dynamic loads, where the time of loading is relatively short. To account for the

viscoelastic time dependency, KENLAYER uses an approximation called Quasi-Elastic


51

Method. In this method, the viscoelastic solution is approximated by an elastic solution

wherein all elastic constants are replaced by the corresponding time-dependent relaxation

moduli (Schapery, 1962).


52

4 MECHANISTIC ROADWAY MODEL

4.1 Mechanistic Roadway Model

The second part of this thesis is to pose the proposed method to predict pavement

service life. While the first was the mechanistic-empirical method, based on the software

KENLAYER, the second one is based on a finite element analysis using an in-house

FEM code. This chapter reviews a complete formulation of the problem together with a

concise description of all the equations and unknowns necessary to construct a well-

posed boundary value problem.

In mechanistic roadway model, the roadway problem is solved by the mechanistic

approach, described as a linear two-dimensional quasi-static initial boundary value

problem (IBVP). The problem is treated as quasi-static since the frequency applied by the

cyclic motion of vehicles on the roadway is lower than the lowest natural frequency that

the pavement can reach.

The first step is to mathematically describe the problem of interest and pose the

boundary value problem by deriving governing field equations and expressing all

variables involved. Some variables are known a priori while others will be determined by

the field equations. The formulation described herein is based on the continuum

mechanics approach where the body is considered as a continuum and neglected the

structure of materials on a scale much smaller than the characteristic scale of the

problem.

Consider a three-dimensional body containing discrete cracks, shown in Figure


34.
53

Figure 34. General Three-Dimensional Body

The body has an interior volume V and a boundary V that is composed of two

parts: internal boundary without cohesive zone VI , and internal boundary with cohesive

zone Vcz . Essential (displacement) and natural (or traction) boundary conditions were

specified throughout the domain V , where V1 is the boundary subjected to known

essential boundary conditions (or displacement) and V2 is the boundary subjected to

known natural boundary conditions (or tractions).

The field variables used to predict this response are represented by the stress

tensor,  ij ( x , t ) , the strain tensor,  ij ( x , t ) , and the displacement vector, ui ( x , t ) .

To solve the IBVP, similar to the mechanistic-empirical approach, the governing

equations are derived from conservation of mass, conservation of momentum, kinematic

constraints and material constitutive equations. Additional constraints are imposed by

initial and boundary value problems.


54

The conservation of linear momentum is given by

dvi
 ji , j   fi   (4.1)
dt

where  is the mass per unit volume, f i is the body force per unit of volume and vi is the

velocity vector.

For a quasi-static problem, the acceleration term is negligible and equation (4.1)

reduces to

 ji , j   fi  0 (4.3)

The conservation of angular momentum is given by

eijk ij   mk  0 (4.4)

where eijkl is the permutation symbol and  mk is the body moment. It is assumed that

there is no body moment, therefore  mk becomes negligible and equation (4.4) reduces

to

 ij   ji (4.5)

Satisfying the conservation of angular momentum and stating a negligible body moment,

results in a symmetric stress tensor.

Kinematics constraints are satisfied by the strain tensor (called strain-

displacement tensor) and is given by

 ij 
1
2
ui, j  u j ,i  uk ,iuk , j  (4.6)
55

Assuming there is no large gradient deformation, uk ,i uk , j becomes negligible and

equation (4.6) reduces to

 ij 
1
2
ui, j  u j ,i  (4.7)

Note that the strain tensor is symmetric.

In addition to the kinematics and kinetics equations, the constitution of a body

must be defined. The material involved in this research, asphalt concrete, exhibits

viscoelastic mechanical behavior. Therefore, the material constitutive equations for a

general orthotropic, linear viscoelastic material are given by:

 kl ( x , )
t
 ij ( x , t )   Cijkl (t   ) d (4.8)
0


or

 kl ( x , )
t
 ij ( x , t )   Dijkl (t   ) d (4.9)
0


where Cijkl is the fourth order tensor of relaxation moduli relating stress to mechanical

strain, and Dijkl is the fourth order tensor creep compliance relating strain to stress.

Cracks are included in the form of cohesive zones. The constitutive behavior of

the cohesive zones can be expressed by a general traction-displacement relationship

given by

Ti cz  x , t   Ti cz u j ,   on Vcz (4.10)
56

where Ti cz is the cohesive traction vector, and α the damage evolution function. The

cohesive zone boundary, Vcz , is typically time-varying. Braces in equation (4.10)

imply history dependence for damage zones, which have a viscoelastic behavior.

Constraints imposed by boundary conditions can be applied on a body as a

displacement (Dirichlet), traction (Neumann) or a mix of both. Displacement boundary

conditions are given by

ui  known on V1 (4.11)

where ui is the described displacement on the surface. Traction boundary conditions are

given by

ti ( x , t )  known on V2 (4.12)

ti ( x, t )   ji n j (4.13)

where ti is the surface traction per unit area and n j is the unit outer normal. Equation

(4.13) is called Cauchy’s formula.

In addition to the constraints imposed by the boundary, initial conditions can be

specified for all state variables and are given by

ui  x , t  t 0 0
 ij  x , t  t 0 0 x in V  V (4.14)
 ij  x , t  t 0 0

Equations (4.11) to (4.14) provide all the equations necessary to solve the IBVP

for all the unknown variables. This constitutes now a well-posed boundary value
57

problem. The objective now is to find uij  x , t  ,  ij  x , t  and  ij  x , t  which can be found

by equations (4.3), (4.7), (4.8), (4.11) ,(4.12) and (4.14).

4.2 Cohesive Zone Model

Cohesive zones, as shown in Figure 35, are modeled by employing interface

elements to model the growth of new boundary surface to the body through the

incorporation of the damage evolution law. A constitutive relation governing mechanical

behavior of the cohesive zone is typically represented either by rate-independent or rate-

dependent models (Kim, 2003).

In 1999, Yoon and Allen developed a damage evolution law for a nonlinear

viscoelastic cohesive zone model (VCZM) which was used to perform this research. The

cohesive zone model for two-dimensional problems is given by

1 ui (t )  t
 ( ) 
Ti  t   1   (t )  if   E c (t   ) d  (i  n or t ) (4.15)
 (t )  i  0
 

where Ti  t  is the cohesive zone area-averaged traction,  (t ) is the Euclidean norm of

the cohesive displacements, ui (t ) is the cohesive zone displacement,  i is the cohesive

zone material length parameter,  (t ) is the internal state variable representing damage

evolution law,  i f is the requisite stress level to initiate damage, and E c (t ) is the linear

relaxation modulus of the cohesive zone.

The Euclidean norm of the damaged zone opening displacements λ, which is used

to couple normal, radial, and tangential behavior, is given by


58

1/2
 u (t ) 2  u (t ) 2  u (t ) 2 
 (t )   n    r    t   (4.16)
  n    r    t  

Figure 35. Cohesive Zone


59

5 FINITE ELEMENT FORMULATION

After presenting a concise statement of the IBVP (Chapter 4), the next step is to

develop a numerical method for solving the aforementioned problem (Zocher, 1995).

5.1 Method of Weighted Residuals

The finite element method is a computationally based procedure for solving

IBVPs. It consists of multiplying the differential equation to be solved by a weighting

function and integrating over the entire body, which attempts to minimize the error.

Neglecting body forces and inertial effects the conservation of linear momentum

(equation (3.1)) can be written as

 ji , j  0 (5.1)

Transforming equation to integral form and multiplying by the weight function

 ui (which can be any arbitrary admissible displacement vector) gives

 (

ji , j ) ui dV  0 (5.2)

where  is any sub domain, herein termed a finite element.

By applying

( ji ui ), j   ji ui   ji ui , j (5.3)

and

 ji , j ui  ( ji ui ), j   ji ui , j
(5.4)
equation can be rewritten as
60

 (

 ui ), j dV    ji ui , j dV  0
ji

(5.5)

Using Gauss divergence theorem, equation can be rewritten as



 ui , j dV    ji ui n j dS
ji

(5.6)

where  is the boundary of  and n j is the unit outer normal to  . Applying

Cauchy’s formula, ti ( x, t )   ji n j , equation becomes



ji  ui , j dV   t  u dS   t  u dS
nc
i i
c
i i (5.7)

where nc denotes the part of the boundary without cohesive zones and  c the part of

the boundary with cohesive zones. Note that

1 
 ji ui , j   ji 
2
  ui , j   u j ,i    ui , j   u j ,i  
1
2 
(5.8)
 ji ui , j   ji  ij  ij 

where ij is the linear rotation tensor represented by the second term in equation . But

 ijij is zero since it is the product of symmetric and skew-symmetric tensors which is

zero, equation becomes

 ji ui , j   ji ij (5.9)

Substituting equation into gives

  tjit dV  t  uit t dS  t  uit t dS


t t t t t t
ji i i (5.10)
 nc c
61

Assuming that stress, strain, and displacement at time t are known, the state

variables at time t  t are evaluated.

But first, define the following

 ji   tjit   tji   tjit   ji   tji


 ji   ijt t   ijt   ijt t   ji   ijt
(5.11)
t t t t
ui  u i u u t
i i  ui  u t
i

vi  vit t   it  vit t  vi   it

Substituting equation into equation gives

  

ji   tji     ij   ijt  dV 

  t  t   u  u  dS    t  t   u  u  dS
(5.12)
t t t t
i i i i i i i i
nc c

As a consequence of uit ,  it and  ijt be known at time t , the variations of them at

time t are zero. Rearranging gives

 

 ij dV    tji ij dV 
ji

 ti  t ui dS 
(5.13)
  tiui dS   titui dS
t
i
nc c c

Substituting  ij  Cijkl  kl   ij into equation gives


R

 C

ijkl  kl   ijR   ij dV    tji ij dV 

  t  t u dS   t u dS   t u dS


t t
(5.14)
i i i i i i i
nc c c

where
62

 kl ( x, )
t
 ijR   Cijkl d (5.15)
0

The recursive traction difference ti between time t and t  t can be expressed
by:

1   (t  t )
ti   E (t ) ui  tiR
i (5.16)
where
1   (t )  p   E j   
tiR 
i

 j 1 
1  exp    
  j  j 
t  (t ) 
  
(5.17)
  p p 
  E   
 Eui (t )    i (t )   1  exp   j t    i (t )      i f
i  j 1 
 j  
 j 1
   

1 p   E j 
E (t )  E   j  1  exp   t   (5.18)
t j 1    j  

 Ej     E 
 j (t )  exp   t   j (t  t )    j 1  exp   j t   (5.19)
 j  t    j 
   

Simplifying equations and gives

ti  kij u j  tiR (5.20)

where

1   (t  t )
kij   E (t ) (5.21)
i

At each time increment, cohesive zone opening produces a reduction of traction

due to damage evolution. It is therefore necessary that equation needs to be implemented

in a finite element model

ti  kij ui  tiR (5.22)

Substituting equation into equation gives


63

 C

ijkl  kl   ijR   ij dV 
c
 k u u dS 
ij j i

  t
nc
i  tit  ui dS    tji ij dV    ijR ij dV 
 
(5.23)

 t ui dS   t u dS
R t
i i i
c c

Applying the strain-displacement operator     B U  and the displacement

shape function u   N U  , equation can be represented in a matrix form

 U   B  C  B  U  dV   U   N  k  N  U  dS 
T T T T

 c

 U   N  t  t  t   dS   U   B   t  dV 


T T T T

nc 
(5.24)
 U   B     dV   U   N  t  dS 
T T R T T t

 c

 U   N   t R  dS
T T

c

Since U  is arbitrary, equation can be represented as


T

 K e  U e    f1e    f 2e    f3e    f 4e    f5e  (5.25)

where

 K e     B  C  B  dV    N  k  N  dS
T T
(5.26)
 c

 f     N   t  t  t  dS
T
e
1 (5.27)
nc

 f     B
T
2
e
 t  dV (5.28)

64

 f     B
T
3
e
  R  dV (5.29)

 f     N 
T
4
e
t t  dS (5.30)
c

 f     N 
T
5
e
 t R  dS (5.31)
c

 K e  is the element stiffness matrix including the effects of surrounding cohesive ones

and  f , f  , f  , f  and  f  are contributions to the element force vector due
1
e
2
e
3
e
4
e
5
e

to surface tractions, stresses at the previous time step, change of stresses during the time

step, cohesive zone tractions during the time step, respectively. The second term in the

element stiffness matrix and force vector terms f 4


e
and f 
5
e
are excluded if the

cohesive zone interface elements are not specified in the body. It can be noted that f 
3
e

and  f  are terms due to viscoelastic characteristics from viscoelastic solid elements
5
e

and viscoelastic cohesive zone interface elements, respectively. It should also be noted

that part of the body can be elastic, but still be implemented in a time-incremental form.

After assembly each element, the global system of equation reduces to

 K G  U G   F G  (5.32)

where  K  is the total global stiffness matrix,


g

U  is the global displacement increment at time t , and


G

F  is the global force vector.


G
65

After presenting a concise statement of the IBVP (Chapter 4), the next step is to

develop a numerical method for solving the aforementioned problem.


66

6 RESULTS

As described previously, the main objective of this research is to compare asphalt

pavements service life predictions subjected to different truck load configurations. A key

point for this comparison was the use of two different approaches: mechanistic-empirical,

that uses KENLAYER and a mechanistic, that uses a FEM code named SADISTIC3 –

Structural Analysis of Damage Induced Stress in Thermo-Inelastic Composites (Allen,

1994). The first one is a commercial by available software package developed by Huang

at the University of Kentucky (Huang, 2004) that performs mechanistic-empirical

analysis specific for asphalt pavements; the second one has been developed at University

of Nebraska Lincoln (Allen et al, 2004; Soares, 2005)

SADISTIC is capable of solving two- and three-dimensional thermo mechanical

continuum problems with growing cracks using the finite element method. Material

models included are thermoelastic, linear thermoelastic (both isotropic and orthotropic),

elastic-plastic, thermoviscoplastic, and linear thermoviscoelastic. The model has

implemented each constant and linear strain triangular elements for two-dimensional

analysis and 8-node curvilinear bricks for three dimensions. In addition, damage can be

modeled by using nonlinear viscoelastic cohesive zones to simulate crack growth.4

This chapter gives a complete description of the problem together with a detailed

explanation of the procedure and results of the analysis using the Mechanistic Roadway

Model. The computational model used in this research is based on a mechanistic analysis,

and it was developed to predict pavement service life using finite element method. It is

3
A more detailed description of the program can be found in (Allen et al., 1994)
4
A more detailed description of the program can be found in (Zocher et al.,1999)
67

focused on a two-dimensional approach to the roadway problem, modeling each

viscoelasticity and cracks to help account for energy dissipation. Viscoelastic properties

(Monismith and Secor, 1962; Elseifi et al., 2006; Park and Kim, 1998; Al-Qadi et al.,

2002, 2005) and crack associated damage (Song et al., 2006; Castell et at., 2000; Kim et

al., 2007) have been used in several studies to predict stresses and strains in pavement

systems.

Environmental conditions, temperature variation and aging should also be

included in the model to help account for energy dissipation. However, a purely

mechanistic model to predict long-term pavement life has rarely been used due to the

significant analytical/computational complexities (Soares, 2008). In this study,

environmental conditions, temperature variation and aging are not taken in consideration

at this time. In addition, the problem was considered homogeneous for the entire analysis.

A typical two-lane asphalt multilayer roadway was chosen, that consists of a total

of four layers which include surface layer, base, sub-base and sub-grade (Figure 36).

Similarly to other studies (Al-Qadi, 2005; Soares, 2005) HMA and base layers were

modeled as viscoelastic materials. Although known that sub-base and sub-grade

materials usually behave as anisotropic and nonlinear elastic material, it was assumed

here that they behave as isotropic linear elastic materials (Al-Qadi et al., 2004; Elseifi

and Al-Qadi, 2006).


68

Figure 36. Typical Two-Lane Asphalt Multilayer Roadway

6.1 Mechanistic Roadway Model Analysis

6.1.1 Roadway Geometry

For the Mechanistic Roadway Model analysis, the geometry adopted is a typical

two-lane asphalt roadway based on the Roadway Design Manual of the Nebraska

Department of Roads, (2005). The top layer is the HMA and is above a base layer

usually made of either bituminous millings or crushed aggregates. The third layer is the

sub-base composed of soils stabilized with lime or fly ash. The sub-grade is in-situ soil.

The road is 12 meters wide with two 3.50 meters traffic lanes and two 2.50 meters

shoulders. For simplicity, the roadway is assumed to be symmetric about the roadway

centerline5.

5
A more detailed description of the Roadway geometry can be found in (Soares, 2005)
69

Figure 37 shows the half space geometry and the boundary conditions

used in this approach. HMA, base, sub-base and sub-grade thicknesses, as can be seen

from Figure 38, are 10cm, 40cm, 30 cm and 1.10m respectively.

Figure 37. Pavement Boundary Conditions

Figure 38. Pavement Cross-Section

Linear viscoelastic materials can be represented by a convolution integral of the

form shown in equation (4.8). The relaxation modulus E (t ) , may be accurately

represented by a Prony series.


70

As shown in Figure 39, the above may be represented by a generalized Maxwell

model as shown in equation .

Figure 39. Generalized Maxwell Model

n t

E (t )  E   Ei e i
(6.1)
i 1

As said before, the sub-base and sub-grade are treated as linear elastic media.

Table 1 shows the Prony series coefficients of relaxation modulus used herein for the

asphaltic layers and Table 2 shows the materials properties for elastic layers.
71

Poisson Ei (kPa ) i (s)


ratio =
0.35
 3.50E+06 -
1 2.03E+08 1.00E-05
2 3.77E+09 1.00E-04
3 3.59E+09 1.00E-03
4 2.52E+09 1.00E-02
5 1.29E+09 1.00E-01
6 4.09E+08 1.00E+00
7 1.31E+08 1.00E+01
8 3.40E+07 1.00E+02
9 1.22E+07 1.00E+03
10 1.61E+06 1.00E+04
11 1.76E+06 1.00E+05
Table 1. Material Properties for Viscoelastic Layers

Sub-base Sub-grade
E 800 MPa 200 MPa
ν 0.35 0.35
Table 2. Material Properties for Elastic Layers

The truck loads used in this analysis are based on the Nebraska Department of

Roadway limits for a tandem axle truck. The gross vehicle weight is 355.86 kN (80,000

lb); the weight per tandem axle and front axle are 151.24 kN (34,000 lb) and 53.38 kN

(12,000 lb), respectively (Figure 40).


72

Figure 40. Tandem axle truck

As many studies describe (Raj, 2002, Soares, 2007 and Dae-Wook, 2005), tire

contact stress distribution, which is significantly affected by tire inflation, tire type and

tire load, has a considerable impact on pavement life. Tire inflation, tire type, axle

distance and distance between tires will be not treated as variables. The tire inflation

used in this study is 0.6895 MPa (100 psi) and the tire type adopted is a conventional

dual tire with overall width of 29.5cm. The distance between tires is 5cm and the axle

length is 1.80m; their dimensions are shown in Figure 41 .


73

Figure 41. Tire and Axle Dimensions

6.1.2 3-D and 2-D

An ideal case would be consider the roadway used in the analysis as 3-D (Figure

42). However for this analysis, a 2-D plain strain configuration was used, where the third

dimension is considered infinitely long, and the loading is applied is a strip load in the

third dimension. The advantage of 2-D over 3-D is the reduction in computer processing

time. By running a 2-D analysis with a strip loading in the third dimension, an

overestimation of the load is induced. The consideration of uniform load is an

approximation, since the tire applies a non-uniform load to the pavement.


74

Figure 42. 3-D View of a Typical Two-Lane Asphalt Multilayer Roadway

As can be seen from Figure 43, the tire pressure distribution is not symmetric. The

stress concentration is higher at the center than at the edges. A good 3D approximation

would be considering the stress distribution as a uniform sinusoidal curve (Figure 44).

However, by doing a 2-D approximation, there is no variation in the third direction which

makes only one dimension behave as a sinusoidal curve. For the 2-D approximation the

tire load distribution adopted is described in Figure 45.

Figure 43. Tire Pressure Distribution


75

Figure 44. Pressure Distribution

4.0

3.5

3.0

2.5
Load(MPa)

2.0

1.5

1.0

0.5

0.0
0.15 0.25 0.35 0.45 0.55 0.65 0.75 0.85
x(m)

Figure 45. Load Distribution


76

6.1.3 Roadway Mesh

Mechanistic Roadway Model is an in-house finite element analysis program. The

final mesh has 1934 elements and 1101 nodes, and the mesh convergence was conducted

in a previous study by Soares (2005). The elements are more refined close to the tire load

and in the upper layers (Figure 46) where higher stresses will occur.

Figure 46. Roadway Mesh

6.1.4 Analysis

The first step in this analysis was to determine where the first cracks would

develop. In order to enable these first cracks to grow, cohesive zones were included in the

appropriate places.

Initially, a monotonic load was imposed on the pavement, with a maximum load

of 18,900 kPa per axle. The choice for a monotonic load Figure 47 instead of a cyclic one

was not to predict when the pavement will fail, but to locate the most susceptible places

for crack development in a convenient manner.


77

3.50
3.00
2.50
Load(kPa)

2.00
1.50
1.00
0.50
0.00
0 10 20 30 40 50
time(s)

Figure 47. Monotonic Load

After the simulation, tangential (sigma xx) (Figure 48), normal (sigma yy) (Figure

49) and shear (sigma xy) (Figure 50) stresses were plotted along the width of the

pavement.

1.30E+04
1.10E+04
9.00E+03
7.00E+03
Sigma xx (kPa)

5.00E+03
3.00E+03
1.00E+03
-1.00E+03
0 0.5 1 1.5 2 2.5 3 3.5
-3.00E+03
-5.00E+03
1st layer 2nd layer
-7.00E+03
x(m)

Figure 48. Stress in xx direction for first and second layers


78

2.00E+03

0.00E+00
0 0.5 1 1.5 2 2.5 3 3.5
-2.00E+03
Sigma yy (kPa)

-4.00E+03

-6.00E+03

-8.00E+03

-1.00E+04
1st layer 2nd layer
-1.20E+04
x(m)

Figure 49. Stress in xx direction for first and second layers

3.00E+03

2.00E+03

1.00E+03
Sigma xy (kPa)

0.00E+00
0 0.5 1 1.5 2 2.5 3 3.5
-1.00E+03

-2.00E+03

-3.00E+03
1st layer 2nd layer
-4.00E+03
x(m)

Figure 50. Stress in xy direction for first and second layers


79

Based on the Mohr Circle (Figure 51), the region where the stresses obtained in

the analysis reach their maximum values were those closest to the tire load, specifically at

the center of the tire (Figure 52). This was expected, since the tire pressure distribution in

a tire (Figure 43) is highly concentrated at these locations.

Figure 51. Mohr Circle


80

2.00E+04

1.00E+04

0.00E+00
0 0.5 1 1.5 2 2.5 3 3.5
Stress (MPa)

-1.00E+04

-2.00E+04

-3.00E+04

-4.00E+04
max sigma x'x' max sigma y'y' max sigma x'y'
-5.00E+04
x(m)

Figure 52. Maximum Stresses

It is believed that the first cracks appear in the region where stresses are

maximums in the plane where the angles reach their maximum values (Figure 53). The

maximum stresses and maximum angles were obtained for this problem and twenty

cohesive zones were included around the region believed to be the most critical, in the 45

degrees plane (Figure 54).


81

60

40

20
Angle (degrees)

0
0 0.5 1 1.5 2 2.5 3 3.5
-20

-40

-60

-80
shear angle principal angle
-100
x(m)

Figure 53. Maximum Angles

Figure 54. Roadway Mesh with Cohesive Zones


82

7 Results

The results for this analysis provide the predicted pavement design life obtained

from the two different models discussed in this thesis. The life of asphalt pavement

subjected to cyclic loads was predicted for different material properties of the cohesive

zone and asphalt layer thicknesses. What makes this analysis unique is the capability to

include damage evolution in the predicted pavement design life; thus resulting in the

ability to determine which combination of design variables gives the best service life.

One limitation of this analysis is that damage has been included only in the region where

it was believed the stresses reach their critical values. In addition, the predicted values for

life obtained from the finite element analyses do not accurately predict the life of

pavements observed in the field.

The failure criterion used in this analysis is not the main thrust of this research. It

was decided to use a simpler criterion in order to demonstrate the technique. A more

accurate failure criterion will be a subject to be incorporated in a future work. All

pavements used have the same base and sub-base thickness, the only possibility of

variation is the HMA layer. However, the overall thickness is held constant in the

analyses. The pavement is said to fail when the difference between the asphalt layer

thickness and the maximum displacement at the surface becomes less than the average

size of the aggregate in the mixture. Consider the roadway in Figure 55 with thickness h.

At t  0 , a force F ( x, t ) is applied to the surface and the displacement is measured at

time t  t1 . Also, when

h    D MAX (7.1)
83

where DMAX is the maximum size of the aggregate in the mixture, which in this

case was considered to be 1.27 cm, the pavement is considered to fail. Note that the

picture is for demonstration purposes only, such that the displacement (  ) is not actually

greater than the thickness ( h ) of the roadway.

Figure 55. Failure criterion

Each computer simulation requires an extraordinary amount of computer

processing time that would make the work practically unachievable to conduct over a

longer period of time. Since the prediction does not include damage everywhere in the

pavement, it is possible to extrapolate the results after a certain number of cycles. This is

done by running the problem up to 15500 seconds and adding a trend line to the data.

Figure 56 shows the graphic with maximum displacement on the surface versus time and

Figure 57 shows the maximum displacement versus time in log scale. As can be seen

from both plots, the displacement is varying in a cyclic way according to the load. The

top part of the plot reflects the pavement being loaded and the lowest part (where the

trend line is positioned) corresponds to the pavement being unloaded. The trend line was

positioned where a permanent deformation was observed.


84

4.00E-04

3.50E-04

3.00E-04
Displacement(m)

2.50E-04

2.00E-04

1.50E-04

1.00E-04

5.00E-05

0.00E+00
0 200 400 600 800 1000 1200 1400 1600 1800 2000
time(s)

Figure 56. Displacement versus time

Figure 57. Displacement versus time in log scale

The analysis was done by comparing life for different thickness of asphalt layer.

Two thicknesses are presented for three different cases. In the first analysis, HMA

thickness was held fixed at 10 cm and the life was predicted for three meshes with: no
85

damage included; including damage; and from the KENLAYER software. In the second

analysis, life was predicted for meshes with no damage included, including damage, and

from the KENLAYER software for HMA thickness of 20 cm.

By looking at Figure 58 and Figure 59, it can be seen that by including damage

the life decreases and gets closer to the life predicted by KENLAYER. It can be also seen

that by increasing the asphalt layer thickness, the predicted life increases for both

methods. If a 20 cm thick layer is used as opposed to a 10 cm, the life goes up by 15% for

no damage included, 11% including damage, and 17% from KENLAYER prediction.

HMA Thickness

123.4

150
86.1
Time (years)

100
53.1
50

0
10cm
KENLAYER with damage no damage

Figure 58. Different design lifes for HMA thickness of 10 cm.


86

HMA Thickness
189.9

150 97.9
Time (years)

90.6
100

50

0
20cm
KENLAYER with damage no damage

Figure 59. Different design lifes for HMA thickness of 20 cm

A 10 cm thickness of asphalt layer gives the shortest life for all parameters and a

20 cm layer gives the best life. Damage 2 gives the greatest life as predicted by the

Mechanistic Roadway Model and the closest result compared to the one given by

KENLAYER. As can be seen, the Mechanistic Roadway Model appears to be a useful

tool for pavement design, since it gives a reasonable prediction of how the pavement will

respond over the years, including the ability to account for geometric and material

properties, several of which KENLAYER is not able to account for. Some of these are as

follows: crack propagation, different loads sets and geometries.


87

8 Conclusion
The objective of this study was to compare two different asphalt pavement design

software. The comparison consisted of investigating the weaknesses and strengths of each

program. The software KENLAYER, developed at the Kentucky University (Huang,

2004) and based on mechanistic-empirical analysis was one of the programs. The other

program used was the Mechanistic Roadway Model which uses an in-house finite

element code (Allen, 1994). The study compared the life predicted from both models for

the same set of load, geometry and material properties.

It was found that the asphalt layer of 20 cm resulted in a better service life for all

different analyses obtained in this research when compared to the 10 cm thickness HMA

layer. Also, it was observed that when damage was included, the service life decreased

considerably, and the results were closer to results predicted by KENLAYER. All

Mechanistic Roadway Model analyses resulted in a higher life when compared to

KENLAYER.

Mechanistic Roadway Model appears to be a useful tool for pavement design,

since it gave a reasonable prediction of how the pavement will respond over the years,

including the ability to account for different sets of geometric and material properties,

several of which KENLAYER is not able to account for. One of the advantage of the

Mechanistic Roadway Model over KENLAYER, it is the ability to predict pavement life

based on a scientific approach. Also, it considers the roadway as a 2-D analysis subjected

to a sinusoidal load. However, the main tool of Mechanistic Roadway Model is the

capability to account for energy dissipation in the form of cohesive zones.


88

The energy dissipation considered in Mechanistic Roadway Model, in the form of

cohesive zones, was included only in the region where stresses reach their higher values.

However, a better approach that was left for future work would be including cohesive

zones not only in the critical region, but everywhere in the asphalt pavement layer.
89

9 References

ALLEN, D. H., & SEARCY, C. R. (2001). A micromechanical model for a viscoelastic


cohesive zone. International Journal of Fracture 107 , 159–176.
Allen, D. H., & Searcy, C. R. (2000). Numerical aspects of a micromechanical model of a
cohesive zone . Reinforcedd plastic and composites 19 .
ALMÁSSY, K. (2002). EXAMINATION OF MECHANICAL PROPERTIES IN
UNBOUND ROAD BASES. PERIODICA POLYTECHNICA SER. CIV. ENG. VOL. 46,
NO. 1 , 53–69 .
AL-QADI, I. L., & ELSEIFI, M. A. (2006). Mechanism and modeling of transverse
cracking development in continuously reinforced concrete pavement. International
Journal of Pavement Engineering, Vol. 7, No. 4 , 341–349.
Al-Qadi, I. L., Elseifi, M., & Yoo, P. J. (n.d.). In-Situ Validation of Mechanistic
Pavement Finite Element Modeling. Virginia Tech Transportation Institute .
Belouchrani, M., & Weichert, D. (1999). An extension of the static shakedown theorem
to inelastic cracked structures . International Journal of Mechanical Science , 163-177.
Burmister, D. M. (1944). The general theory of stresses and displacements in layered soil
systems. II. Journal of applied physics .
Burmister, D. M. (1944). The general theory of stresses and displacements in layered soil
systems. III. Journal of applied physics .
Burmister, D. M. (1944). The general theory of stresses and displacements in layered
systems. I. Jounal of applied physics .
Castell, M. A., Ingraffea, A. R., & Irwin, L. H. (2000). Fatigue crack growth in pavement
. Journal of transportation engineering .
Ceylan, H., Gopalakrishnan, K., & Coree, B. (2006). Development of a Mechanistic-
Empirical Structural Design Program for HMA Overlaid Rubblized PCC Pavements.
Transportation Research Board .
Choi, S. T., & Earmme, Y. Y. (2003). Elastic Singularity Interacting With Various Types
of Interfaces. Journal ofapplied mechanics 70 .
Ciavarellaa, M., & Pugno, N. (2005). A GENERALIZED PARIS’ LAW FOR FATIGUE
CRACK GROWTH. XXXIV CONVEGNO NAZIONALE , 14-17.
Costanzo, F. (2001). An Analysis of 3D Crack Propagation using Cohesive Zone Models
and the Theory of Configurational Forces. Mathematics and Mechanics of Solids , 149-
173.
Donna Harmelink, P., Scott Shuler, P. P., & Tim Aschenbrener, P. (2008). Top-Down
Cracking in Asphalt Pavements: Causes, Effects, and Cures. JOURNAL OF
TRANSPORTATION ENGINEERING .
Elseifi, M. A., Imad L. Al-Qadi, F., & Yoo, P. J. (2006). Viscoelastic Modeling and Field
Validation of Flexible Pavements. JOURNAL OF ENGINEERING MECHANICS .
Foulk, J., Allen, D., & Helms, K. (2000). Formulation of a three-dimensional cohesive
zone model for application to a fnite element algorithm. Computer Methods in Applied
Mechanics 183 , 51-66.
Gedafa, D. S. (2006). COMPARISON OF FLEXIBLE PAVEMENT PERFORMANCE
USING KENLAYER AND HDM-4. Ames: Midwest Transportation Consortium.
90

GHALY, A. M. (2003). A Mechanistic Analysis of Static and Repeated Loading Effects


on Rubberized Asphalt. The International Journal of Pavement Engineering, Vol. 4 ,
209–219.
Griffith, A. A. (1921). The Phenomena of Rupture and Flow in Solids.
Harichandran, R. S. (2001 ). Flexible pavement design in Michigan: transition to
empirical to mechanistic methods. Transportation research record 1778 , 100-106.
Hasegawa, H. (1975). An extension of Love's solution for axisymetric problems of
elasticity . Bulletin of the JSME 18 .
Huang, Y. H. (1972). Strain and curvature as factors for predicting pavement fatigue .
International Conference on the Structural Design of Asphalt Pavements , 622-628.
Huang, Y. H. (1967). Stresses and displacements under circular load. Second
international conference on the structural design of asphalt pavements , 225-244.
Ioannides, A. M., & hKazanovich, L. (1998). General formulation for multilayered
pavement systems. JOURNAL OF TRANSPORTATION ENGINEERING .
Kim, R., Daniel, J. S., & Wen, H. (2002). FATIGUE PERFORMANCE EVALUATION
OF WESTRACK ASPHALT MIXTURES USING VISCOELASTIC CONTINUUM
DAMAGE APPROACH.
Kim, Y.-R., D. N. Little, P. F., & R. L. Lytton, P. F. (2003). Fatigue and Healing
Characterization of Asphalt Mixtures. JOURNAL OF MATERIALS IN CIVIL
ENGINEERING .
Lee, H.-J., Daniel, o. S., & Kim, Y. R. (2000). CONTINUUM DAMAGE MECHANICS-
BASED FATIGUE MODEL OF ASPHALT CONCRETE. JOURNAL OF MATERIALS
IN CIVIL ENGINEERING .
Li, Y., P.E., M., & John B. Metcalf, M. (2002). Crack Initiation Model from Asphalt Slab
Tests. JOURNAL OF MATERIALS IN CIVIL ENGINEERING .
Loulizi, A., Al-Qadi, I. L., & Elseifi, M. (2006). Difference between In Situ Flexible
Pavement Measured and Calculated Stresses and Strains. JOURNAL OF
TRANSPORTATION ENGINEERING .
Love, A. E. (1906). A treatise on the mathematical theory of elasticity. Cambridge:
Cambridge University Press.
Majidzadeh, K. (1971). Application of fracture mechanics in the analysis of pavement
fatigue. Proceedings of the association of asphalt paving technologists , 227-246.
Masad, S., Little, D., & Masad, a. E. (2006). Analysis of Flexible Pavement Response
and Performance Using Isotropic and Anisotropic Material Properties. JOURNAL OF
TRANSPORTATION ENGINEERING .
Officials, A. -A. (1993). Guide for design of Pavementss Structures. Washington:
AASTHO.
Papagiannakis, A. T. (2002). Micromechanical analysis of viscoelastic properties of
asphalt concretes. Transportation research record , 113-120.
Park, D.-W., Amy Epps Martin, A., & Eyad Masad, A. (2005). Effects of Nonuniform
Tire Contact Stresses on Pavement Response. JOURNAL OF TRANSPORTATION
ENGINEERING .
91

Parka, S. W., & Schapery, R. A. (1999). Methods of interconversion between linear


viscoelastic material functions[ Part I-a numerical method based on Prony series.
International Journal of Solids and Structures 25 , 1653-1665.
Patnaik, S. N., & Hopkins, D. A. (2007). Completed Beltrami-Michell formulation in
polar coordinates. International Journal of Physical Sciences Vol. 2 , 128-139.
Program, N. C. (2004). Guide for Mechanistic-Empirical Design of new and rehabilitated
pavement structures. Champaign: Transportation Research Board.
Sangpetngam, B. (2004). A muli-layer boundary element method for the evaluation of
top-down cracking in hot-mi asphalt pavements . Transportation research record .
Schram, S. (2006). Improving prediction accuracy in mechanictic-empirical pavement
design guide. Transportation research record 1947 , 59-68.
SEARCY, C. R. (2004). A MULTISCALE MODEL FOR PREDICTING DAMAGE
EVOLUTION IN HETEROGENEOUS VISCOELASTIC MEDIA.
Soares, J. B. (2003). Considering material heterogeniety in crack modeling of asphaltic
mixtures . Transportation research record , 113-120.
Soares, R. F., Allen, D. H., Kim, Y.-R., Berthelot, C., Soares, J. B., & E.Rentschler, M.
(2007). A Computational Model for Predicting the Effect of Tire Configuration on
Asphaltic Pavement Life. Road Materials and Pavements Design .
Son, N. Q. (1973). Matériaux élasto-visco-plastiques à potentiel généralisé. Comptes
Rendus de l’Académie des Sciences 277 , 915–918.
Song, S. H., Paulino, G. H., & Buttlar, W. G. (2006). Simulation of Crack Propagation in
Asphalt Concrete Using an Intrinsic Cohesive Zone Model. JOURNAL OF
ENGINEERING MECHANICS .
Souza, F. V., & Soares, J. B. (2003). EFEITO DA CONSIDERAÇÃO DO
COMPORTAMENTO VISCOELÁSTICO LINEAR DO REVESTIMENTO NO
CÁLCULO DE TENSÕES E DESLOCAMENTOS EM PAVIMENTOS ASFÁLTICOS.
ANPET .
THEYSE, H. L., BEER, M. D., & RUST, F. C. (n.d.). Overview of South African
Mechanistic Pavement Design Method. TRANSPORTATION RESEARCH RECORD 1539
.
Widyatmoko, C. E., & Read, J. M. (1999). Energy dissipation and the deformation
resistance of bituminous mixtures. Materials and Structures 32 , 218-223.
Xu, Q., & Rahman, M. S. (2008). Finite element analyses of layered visco-elastic system
under vertical circular loading. INTERNATIONAL JOURNAL FOR NUMERICAL AND
ANALYTICAL METHODS IN GEOMECHANICS 32 , :897–913.
Yang, B., Mall, S., & Ravi-Chandar, K. (2001). A cohesive zone model for fatigue crack
growth in quasibrittle materials. International Journal os solids and structures 38 , 3927-
3944.
Yong-Rak Kim, A., Allen, D. H., & D. N. Little, F. (2005). Damage-Induced Modeling
of Asphalt Mixtures through Computational Micromechanics and Cohesive Zone
Fracture. JOURNAL OF MATERIALS IN CIVIL ENGINEERING .
Yong-Rak Kim, A., Allen, D. H., & D. N. Little, F. (2005). Damage-Induced Modeling
of Asphalt Mixtures through Computational Micromechanics and Cohesive Zone
Fracture. JOURNAL OF MATERIALS IN CIVIL ENGINEERING .
92

Yong-Rak Kim, M., Allen, D. H., & D. N. Little, F. (n.d.). Computational Constitutive
Model for Predicting Nonlinear Viscoelastic Damage and Fracture Failure of Asphalt
Concrete Mixtures. INTERNATIONAL JOURNAL OF GEOMECHANICS , 2007.
YOO, P. J., AL-QADI†, I. L., ELSEIFI, M. A., & JANAJREH, I. (2006). Flexible
pavement responses to different loading amplitudes considering layer interface condition
and lateral shear forces. The International Journal of Pavement Engineering, Vol. 7, No.
1 , 73–86.
YOON, C., & ALLEN, D. H. (1999). Damage dependent constitutive behavior and
energy release rate for a cohesive zone in a thermoviscoelastic solid. International
Journal of Fracture 96 , 55–74.
ZOCHER, M. A., & GROVES, S. E. ( 1997). A THREE-DIMENSIONAL FINITE
ELEMENT FORMULATION FOR THERMOVISCOELASTIC ORTHOTROPIC
MEDIA. INTERNATIONAL JOURNAL FOR NUMERICAL METHODS IN
ENGINEERING VOL. 40 , 2267-2288.

You might also like