Download as pdf or txt
Download as pdf or txt
You are on page 1of 139

THREE-DIMENSIONAL FINITE ELEMENT ANALYSIS

OF FLEXIBLE PAVEMENTS

BY
Jia Wang
B.S. North Industry University, China, 1996
A THESIS
Submitted in Partial Fulfillment of the
Requirements for the Degree of
Master of Science
(in Civil Engineering)

The Graduate School


The University of Maine
May, 2001

Advisory Committee:
William G. Davids, Assistant Professor of Civil Engineering, Advisor
Eric N. Landis, Associate Professor of Civil Engineering
Vincent Caccese, Associate Professor of Mechanical Engineering

LIBRARY RIGHTS STATEMENT


In presenting this thesis in partial fblfillment of the requirements for an advanced
degree at the University of Maine, I agree that the Library shall make it freely available
for inspection. I further agree that permission for "fair use" copying of this thesis for
scholarly purposes may be granted by the Librarian. It is understood that any copying or
publication of this thesis for financial gain shall not be allowed without my written
permission.
Jan. 22.2001

THREE-DIMENSIONAL FINITE ELEMENT ANALYSIS


OF FLEXIBLE PAVEMENTS

By Jia Wang
Thesis Advisor: Dr. William G. Davids
An Abstract of the Thesis Presented
in Partial Fulfillment of the Requirements for the
Degree of Master of Science
(in Civil Engineering)
May, 2001

Flexible pavements or roads surfaced with asphalt have been in use for the past 100
years. Currently, the design of flexible pavements is largely based on empirical methods.
However, there is currently a shift underway towards more mechanistic design
techniques. While layered elastic analysis and two-dimensional finite element (FE)
methods have been generally been used to deternline stresses, strains and displacements
in flexible pavements, they suffer several severe limitations. To overcome these
difficulties, three-dimensional (3D) FE analysis must be used to analyze pavement
structqres. This study focuses on exploring the use of 3D finite-element methods to
examine the response of flexible pavements.
For this study, an efficient 3D FE meshing tool was developed. This meshing tool
allows us to develop models of layered system, inter-layer debonding and slip, various
wheel and axle loadings.
The 3D FE models were tested by comparing predicted results with experimentally
measured field data. Critical finite-element model dimensions were determined, and

material properties were back-calculated to give good comparisons with field-measured


data.
Since the stress-strain response of granular materials is non-linear, the use of the
stress-dependent K-8 model was investigated. Two implementation methods were
considered. It was shown that the analysis results are very different depending on the
method of implementation.
When applying wheel loads, it is quite typical that a constant tirefpavement pressure
distribution is assumed over a rectangular or circular area. However, prior investigations
have shown that spatially varying tirelpavement contact pressures can affect the response
of pavements significantly. In this study, a model was developed to simulate spatially
varying tire contact pressures based on previous published data. Paranletric studies were
performed to examine the effect of spatially varying tire pressures on pavement response.
There studies showed that using spatially varying tirefpavement pressures yields stresses
at the bottom of asphalt that are up to 30% larger than those predicted when uniform
tirefpavement pressures are assumed. The largest differential occurs in thin flexible
pavement structures with a sub-base having a low stiffness. Further, the studies show
that for thick pavement sections, predicted tensile stresses in the top of the asphalt are
much larger when spatially varying tirefpavement contact pressures are considered.

ACKNOWLEDGEMENTS
I would first like to acknowledge the sponsors of this project. Financial support was
provided by the Washington State Department of Transportation and the University of
Maine.
Special thanks to my advisor, Dr. William Davids, for accepting me into the master's
program and providing direction for this project. His time and effort with the teaching,
thesis corrections, and many meetings despite his increasingly busy schedule are greatly
appreciated. Dr. Davids was very helpful in offering guidance, encouragement, and
recommendations for my scholastic and professional interests as well.
Special thanks go to members of my advisory committee: Prof. Eric Landis, Prof.
Vincent Caccese. They were always available to answer my questions and comment on
my work. I also received much guidance on this project from Prof. Dana Hamphrey.
Finally, I would like to my family and friends for their encouragement and
understanding throughout my graduate career.

TABLE OF CONTENTS
..
ACKNOWLEDGEMENTS...................................................................11
LIST OF TABLES...............................................................................vi
LIST OF FIGURES ............................................................................. vii..
CHAPTER 1. INTRODUCTION............................................................ 1
1.1 Background............................................................................. 1
1.2 Objectives of Study .....................................................................3
1.3 Organization of the Thesis............................................................. 4
CHAPTER 2 . LITERATURE REVIEW............................................................................6
2.1 Introduction...................................................................................................................... 6
2.2 Layered Elastic Analysis ................................................................................................. 7
2.2.1 Two-layered System..........................................................................................8
2.2.2 Examples of Layered Analysis Programs........................................................9
2.2.3 Back-calculation Using Layered Elastic Analysis..................................... 10
2.2.4 Advantages and Disadvantages ..................................................
11
2.3 Two-dimensional Finite Element Methods......................................... 12
2.3.1 Axis-symmetric vs . Plane Strain ..................................................
13
2.3.2 Currently Available 2D FE Models for Flexible Pavement Analysis .......14
2.3.2.1 ILLI-PAVE Model ...................................................... 14
2.3.2.2 MICH-PAVE Model ...................................................
14
2.3.2.3 Commercially available FE programs ...............................15
2.3.3 Advantages and Disadvantages of 2D FE Analysis ............................15
2.4 Three-dimensional (3D) Finite Element Methods................................. 16
2.4.1 Cho et a1 (1 996).................................................................... 17
2.4.2 Hsien H . Chen, Hurt M . Marshek, and Chhote L . Saraf (1990) ........ 17
2.4.3 K. D . Hjelmstad, J . Kim and Q. H . Zuo (1997)..............................................18
2.4.4 Sam Helwany, John Dyer, and Joe Leidy (1998) ...............................18
2.4.5 Samir N . Shoukry and Gergis W . William (1999) ..............................19
2.4.6 A Bensalem, A . J . Brown, M.E. Nunn, D . B. Merrill,
and Wyn G. Lloyd (2000)................................................
20
2.4.7 Ronald Blab, and John T. Harvey (2000) ........................................ 20
2.5 Summary................................................................................ 21

CHAPTER 3 . MODELING AND MESHJNG STRATEGIES.................... 23


3.1 Introduction ............................................................................ 23
3.2 Overview of FE Model ............................................................... -24

3.2.1 General Layout of Flexible Pavement System..................................24


3.2.2 Local Refinement Region ..........................................................25
3.2.3 Finite Boundaries of Model ...................................................25
3.2.4 Interface Slip and Constrains...................................................26

3.3 Element Types .........................................................................26


3.3.1 20-noded Brick Element........................................................26
3.3.2 16-noded Interface Element .................................................. 27
3.4 Meshing an Individual Block of Element ......................................... 28
3.5 Merging and Forming 3D Model .................................................... 29
3.5.1 A NaYve Method........................................................................................... 31
3.5.2 Grid Sort Method .....................................................................
31
3.5.3 Numerical Comparison of Two Methods.......................................35
3.6 Implementation....................................................................... -36
3.6.1 Typical Procedure for Generating a Layered Pavement System..............36
3.6.2 Control Parameters for Meshing Code ........................................... 37
3.7 Summary ....................................................................................... 39

CHAPTER 4 . MODEL DEVELOPMENT AND VERIFICATION............41


4.1 Introduction ........................................................................... -41
4.2 Effect of Multiple Wheel Loads..................................................... 41
4.2.1 Single Dual-wheel, Single Axle and Dual-wheel Tandem Axle ..................41

4.3 Determination of Critical Model Dimensions ..................................... 46


4.3.1 Model description.............................................................. 47
4.3.2 Results of study on critical model dimensions.................................50
4.4 Comparison with experimental results............................................. 51
4.4.1 Model Description ............................................................. 51
4.4.2 Experimentally Measured Data ................................................53
4.4.2.1 Benkelman Beam ......................................................53
4.4.2.2 Testing Procedure ......................................................54
4.4.3 Back-calculation for North Yarmouth Pavement Section.....................57
4.4.4 Back-calculation for TWP3 1-MD pavement section .........................61
4.5 Summary ............................................................................. -65

CHAPTER 5. EFFECT OF NON-LINEAR SOIL RESPONSE.................66


5.1 Introduction ...................................................................................66
5.2 Background on K-8 Model ........................................................... 66
5.3 Implementation of K-8 Model .......................................................67
5.4 Convergence of K-8 Model ......................................................... 69
5.5 Comparison with Measured Displacement Basins ................................. 76

5.5.1 Model Description ............................................................. 76


5.5.2 Comparison Using Published Parameters...................................................... 77

5.6 Summary and Conclusions.............................................................


83

CHAPTER 6. EFFECT OF SPATIALLY VARYING TIRE


PRESSURE................................................................................................84

6.1 Introduction ....................................................................................


84
6.2 Modeling the Spatially Varying Tire Pressures........................................................ 85
6.2.1 Background .........................................................................85
6.2.2 Representation of Spatially Varying Tire Pressures .........................89
6.2.3 Adjustment of Total Vertical Load ..............................................94
6.2.4 Different Combinations.............................................................96
6.3 Convergence Study........................................................................................................98
6.4 Parametric Study ........................................................................................................ 101
6.4.1 Model Description ........................................................... 101
6.5 Results and Discussions....................................................................
103
6.5.1 Stresses at Bottom of Asphalt ................................................. 103
6.5.2 Stresses at Top of Asphalt .................................................... 109
6.6 Summary ...................................................................................................................... 114

CHAPTER 7 . SUMMARY. CONCLUSIONS AND


RECOMMENDATIONS...................................................................................................... 116
7.1 Summary .....................................................................................
116
7.2 Conclusions..................................................................................................................117
7.3 Recommendations for Future Work ....................................................119

BIBLIOGRAPHY .............................................................................. 120


BIOGRAPHY OF THE AUTHOR...................................................... 124

LIST OF TABLES
Table 4.1

Material Properties................................................................ -44

Table 5.1

Summary of Max . Tensile Stresses and Displacements for two


Methods .............................................................................-74

Table 5.2

Material Properties and Vertical Model Dimensions ..........................77

Table 5.3

Load Parameters used in MICHPAVE .......................................... 78

Table 5.4

Material Properties used in MICHPAVE....................................... 78

Table 6.1

Vertical Stresses at edge and center of tire ...................................... 98

Table 6.2

Material Properties .................................................................. 99

LIST OF FIGURES
Figure 1.1

A Typical Cross Section of Flexible Pavement System ....................... 1

Figure 2.1

A Typical Cross Section of Flexible Pavement System ....................... 6

Figure 2.2

A 3-layered Flexible Pavement System Subjected to a Wheel Load ........8

Figure 3.1

A Typical Plan View of a Meshed Pavement System ...........................24

Figure 3.2

Elevation View ..........................................................................


24

Figure 3.3

A Typical 20-noded Brick Element .............................................27

Figure 3.4

16-noded Quadratic Isoparametric Interface Element ..........................28

Figure 3.5

Brick Element Mesh ............................................................... 29

Figure 3.6

Merging the Meshes .................................................................... 30

Figure 3.7

32
Building Grid Sort Element Block .................................................

Figure 3.8

Merging Meshes ........................................................................ -33

Figure 3.9

35
A 3D View of Fully Merged Mesh .................................................

Figure 3.10

Computational Time Comparison of Two Methods ............................36

Figure 3.1 1

38
Plan View of Merged Meshes .......................................................

Figure 4.1

Typical Dual-wheel Tandem Axle ..................................................43

Figure 4.2

Model Description (Multiple-wheel Loads) .......................................


43

Figure 4.3

3D View of Fine Mesh (3840 elements) ........................................ 44

Figure 4.4

Maximum Principal Stress at the bottom of AC (single dual wheel.


loading case I) ...........................................................................
-45

Figure 4.5

Maximum Principal Stress at the bottom of AC (single axle. loading


case I+II)...................................................................................
45

Figure 4.6

Maximum Principal Stress at the bottom of AC (dual axle. loading case


I.IV) .........................................................................................46
vii

Figure 4.7

Layered System to be Modeled ................................................... 47

Figure 4.8

Elevation View of Pavement System............................................ 48

Figure 4.9

Plan View of System (quarter symmetric) with B.C ........................... 49

Figure 4.10 3D View of Meshed Model (quarter symmetric).............................. 49


Figure 4.1 1 Stress Variation vs. Plan Dimensions............................................ 50
Figure 4.12 Control Sections...........................................................................53
Figure 4.13 Plan View of Location of Benkelman Beam Probe between Dual Truck
Tires ................................................................................. 55
Figure 4.14 The Modified Benkleman Beam Used to Measure Pavement
Deflection .......................................................................... -56
Figure 4.15 The Best Pair of E2 and E3. for z = 740 mm ......................................-58
58
Figure 4.16 Displacement Basin Comparison. z = 740 mm ....................................
Figure 4.17

Principal Stress Contour at the bottom of the AC with E2 = 67 MPa.


E3 = 10 MPa for North Yarrnouth Control Section. z = 740 mm ............59

Figure 4.18 The Best Pair of E2 and E3. for z = 1940 mm .....................................60
60
Figure 4.19 Displacement Basin Comparison. z = 1940mm...................................
Figure 4.20 Principal Stress Contour at the bottom of the AC with E2 = 46 MPa.
E3 = 28 MPa for North Yarrnouth Control Section. z = 1940mm............61
Figure 4.21

The Best Pair of E2 and E3. for z = 1040 mm ......................................62

Figure 4.22 Displacement Basin Comparison. z = 1040 mm ...................................62


Figure 4.23 Principal Stress Contour at the bottom of the AC with E2 = 68 MPa.
E3 = 30 MPa for TWP-31MD Control Section. z = 1040 mm ................ 63
Figure 4.24

The Best Pair of E2 and E3. for z = 2240 mm ......................................63

Figure 4.25 Principal Stress Contour at the Bottom of the AC with E2 = 49 MPa.
E3 = 76 MPa for TWP-3 1MD Control Section. z = 2240mm .................64

...

Vlll

Figure 5.1

Maximum Vertical Deflections vs . Number of Material Iterations.


Fully Bonded. Integration Point Method..................................................... 70

Figure 5.2

Max. Principal Tensile Stresses in Asphalt vs . Number of Material


Iterations. Fully Bonded Integration Point Method.......................................71

Figure 5.3

Maximum Vertical Deflections vs.Number of Material iterations.


Fully Bonded. Mid-layer Method................................................................71

Figure 5.4

Max. Tensile Stresses vs . Number of Material iterations. Fully


Bonded. Mid-layer Method.......................................................................... 72

Figure 5.5

Maximum Vertical Deflections vs . Number of Material iterations. No


Tension at Asphalt / Sub-base Interface. Integration Point Method ............72

Figure 5.6

Max. Principal Tensile Stresses vs . Number of Material iterations.


Loss Contact. Integration Point Method .........................................................73

Figure 5.7

Maximum Vertical Deflections vs . Number of Material iterations. UnBonded. Mid-layer Method.......................................................................... 73

Figure 5.8

Max. Tensile Stresses vs . Number of Material iterations. Un.Bonded.


Mid-layer Method ....................................................................................... -74

Figure 5.9

Displacement Basin Comparison. Integration Point Method.......................78

79
Figure 5.10 Displacement Basin Comparison. Mid-layer Method ...................................
Figure 5.1 1 The Best Pair of KI and K2 in the Least Square Sense. Integration
Point Method............................................................................................... 80
Figure 5.12 Displacement Basin Comparison. K I = 10 MPa and K2 = 0.78, Integration
Point Method................................................................................................ 81
Figure 5.13 Stresses Contour at the Bottom of the AC with KI = 1OMPa. K2 = 0.78
for North Yarmouth Control Section. Integral Points Method .....................81
Figure 5.14 Maximum Principal Stresses at the Bottom of the AC with K1 =
30 MPa. K2 = 0.53 for North Yarmouth Control Section. Mid-layer
Method................................................................................................................
82
Figure 6.1

Types of Tire Contact Area..............................................................................


84

Figure 6.2

Load. Inflation Pressure Relationship on the Vertical Contact Stress


Distribution (de Beer et al. 1997)................................................................
87

Figure 6.3

Maximum Vertical Contact Stresses (de Beer et al. 1997)...........................88

Figure 6.4

Configurations of Vertical Stress in 2D ......................................................90

Figure 6.5

Configuration of Transverse Stress in 2D................................................... 91

Figure 6.6

Integration Domain...................................................................................... 91

Figure 6.7

Typical Vertical Contact Pressure Distribution (tire load. L = 40kN.


Inflation pressure. P = 620KPa).......................................................................
93

Figure 6.8

Typical Transverse Contact Pressure Distribution (tire load. L = 40


kN. inflation pressure. P = 620 kPa)................................................................... 93

Figure 6.9

Refinement of Contact Area .............................................................................


94

Figure 6.10 A Typical Triangulation.................................................................................


95
Figure 6.1 1 Vertical Stress Distribution (L = 20 kN. P = 420 kPa).................................96
Figure 6.12 Vertical Stress Distribution (L = 20 kN. P = 720 kPa).................................97
Figure 6.13 Vertical Stress Distribution (L = 50 kN. P = 420 kPa).................................97
Figure 6.14 Vertical Stress Distribution (L = 50 kN. P = 720 kPa).................................98

. .

Figure 6.15 Model Description ...........................................................................................


99
Figure 6.16 3D View of Meshed Model (entire pavement system. 3024 solid
elements)........................................................................................................
100
Figure 6.17 Convergence Study on Max . Stresses and Deflections..............................100

. .

102
Figure 6.18 Model Descnption. Plan View......................................................................
Figure 6.19 Model Description. Elevation View............................................................103
Figure. 6.20 Max . Tensile Stresses at bottom of AC (thickness = 75mm. E2 =
35MPa).....................................................................................104

Figure. 6.21

Max. Tensile Stresses at bottom of AC (thickness = 75mm, E2 =


67 MPa). .................................................................................105

Figure. 6.22

Max. Tensile Stresses at bottom of AC (thickness = 75mm, E2 =


135MPa).................................................................................lo5

Figure. 6.23

Max. Tensile Stresses at bottom of AC (thickness = 125mm, E2 =


35MPa)................................................................................... 106

Figure. 6.24

Max. Tensile Stresses at bottom of AC (thickness = 125mm, E2 =


67MPa)................................................................................... 106

Figure. 6.25

Max. Tensile Stresses at bottom of AC (thickness = 125mm, E2 =


135MPa). ...............................................................................
.lo7

Figure. 6.26

Max. Tensile Stresses at bottom of AC (thickness = 200mm, E2 =


35MPa). ..................................................................................
107

Figure. 6.27

Max. Tensile Stresses at bottom of AC (thickness = 200mm, E2 =


67MPa). .................................................................................. 108

Figure. 6.28

Max. Tensile Stresses at bottom of AC (thickness = 200mm, Ez =


135MPa)................................................................................
.lo8

Figure 6.29

Max. Tensile Stresses at top of AC (thickness = 75mm,


E2 = 35MPa). ............................................................................
109

Figure 6.30

Max. Tensile Stresses at top of AC (thickness = 75mm,


E2 = 67MPa).............................................................................110

Figure 6.3 1

Max. Tensile Stresses at top of AC (thickness = 75mm,


E2 = 135MPa)...........................................................................
110

Figure 6.32

Max. Tensile Stresses at top of AC (thickness = 125mm,


E2 = 35MPa).............................................................................
111

Figure 6.33

Max. Tensile Stresses at top of AC (thickness = 125mm,


E2 = 67MPa)...........................................................................
..I11

Figure 6.34

Max. Tensile Stresses at top of AC (thickness = 125mm,


E2 = 135MPa)........................................................................
1

Figure 6.35

Max. Tensile Stresses at top of AC (thickness = 200mrn,


E2 = 35MPa).............................................................................
112

Figure 6.36

Max . Tensile Stresses at top of AC (thickness = 200mm.


E2 = 67MPa)............................................................................. 113

Figure 6.37

Max . Tensile Stresses at top of AC (thickness = 200mm.


E2 = 135MPa)...........................................................................
113

xii

CHAPTER 1. INTRODUCTION

1.1 Background
The first asphalt road was constructed in the US about 100 years ago in New Jersey.
There are currently about 2.2 million miles of roadway surfaced by asphalt concrete
pavements (Huang, 1993). Flexible pavements are made up of bituminous and granular
materials. A typical flexible pavement section can be idealized as a multi-layered system
consisting of asphalt layers resting on soil layers having different material properties as
shown in Figure 1.1.

Asphalt Concrete
Sub-base (often gravel)
Sub-grade (existing soil)

Figure 1.1 A Typical Cross Section of Flexible Pavement System


Methods of designing flexible pavements can be classified into several categories:
empirical method with or without a soil test, limiting shear failure, and the mechanisticempirical method (Huang, 1993). Currently, the design of flexible pavements is largely
empirical (Helwany et al, 1998; Huang, 1993). However, mechanistic design is
becoming more prevalent, which requires the accurate evaluation of stresses and strains
in pavements due to wheel and axle loads.

In general, there are three approaches that can be used to compute the stresses and
strains in pavement structures: layered elastic methods, two-dimensional (2D) finite
element modeling, and three-dimensional (3D) finite element modeling.
The layered elastic approach is the most popular and easily understood procedure. In
this method, the system is divided into an arbitrary number of horizontal layers (Vokas et
al. 1985). The thickness of each individual layer and material properties may vary from
one layer to the next, but in any one layer the material is assumed to be homogeneous and
linearly elastic. Although the layered elastic method is more easily implemented than
finite element methods, it still has severe limitations: materials must be homogenous and
linearly elastic within each layer, and the wheel loads applied on the surface must be
axi-symmetric. Those shortcomings make it difficult to simulate realistic scenarios. For
example, it is very hard to rationally accommodate material non-linearity and incorporate
spatially varying tire contact pressures, which can significantly affect the behavior of the
pavement systems (de Beer et al. 1997; Bensalem et al, 2000).
For 2D finite element analysis, plane strain or axis-symmetric conditions are
generally assumed. Compared to the layered elastic method, the practical applications of
this method are greater, as it can rigorously handle material anisotropy, material nonlinearity, and a variety of boundary conditions (Zienkiewicz and Taylor, 1988).
Unfortunately, 2D models can not accurately capture non-uniform tire contact pressure
and multiple wheel loads.
To overcome the limitations inherent in 2D modeling approaches, 3D finite element
models are becoming more widespread. With 3D FE analysis, we can study the response
of flexible pavements under spatially varying tirefpavement contact pressures, capture the

effect of non-linear materials or the effect of combination of loads, including unsymmetric or different loading types. However 3D FE models can be very difficult and
time-consuming to use. In particular, problems that often arise when using 3D FE
analysis include:

Mesh generation. Mesh generation must be automated to allow a variety of


geometries to be considered and to minimize the time spent developing and checking
meshes.

Problem size. To solve a large 3D FE problem, tremendous computer resources


are required to achieve a solution.

Determination of critical model dimensions. In reality, the flexible pavement


could extend infinitely along the vertical direction and traffic directions. A critical
question raised here is how large the model extent should be with the fixed boundary
conditions.

1.2 Objectives of Study


This thesis focuses on the structural response of flexible pavement structures under
single or multiple wheel loads. The objectives of this thesis are:

Develop eflective meshing tool for flexible pavements.


Determination of critical pavement dimensions for a 3 0 FE model. Practically
speaking, the vertical dimension of sub-grade layer can be essentially infinite.
What are the minimum vertical, longitudinal and transverse dimensions that must be
modeled to accurately capture response?

Examine the eflect of sub-base non-linearity on pavement response. The well-

known stress-dependent K-8 model (Huang, 1993) will be incorporated in the 3D


finite element models of this study, which is widely used to analyze non-linear
granular materials. Unfortunately, this model was developed in 2D, and its
application in 3D is still questionable. This study will examine its usefulness and
efficiency in 3D.
Develop a method of modeling spatially varying wheel tire contact pressures.
Previous researchers (Yue and Svec, 1995; Bensalem et a1 2000, etc) have shown that
modeling tire contact pressures as uniform load tends to overestimate the response
(stresses, strains) of pavements and that the spatial variation in tire-pavement contact
pressures significantly affects the results. Further, recent research has shown that topdown cracking is a concern in pavements (Bensalem et al, 2000; Uhlmeyer et al,
2000). The ability to study the effect of spatially varying tire pressures may provide
insight into the cause of this problem.
An adaptation of the 3D finite element code underlying EverFE (Davids et al,
1998), which employs highly efficient iterative solvers and allows the modeling of
general displacement constraints will be used for the finite element models of this
study.

1.3 Organization of the Thesis


This thesis is organized into seven chapters. In Chapter 2, the literature related to
this thesis is reviewed. This includes topics on background of pavement structures and
the finite element modeling of flexible pavements.
Chapter 3 details the development of the 3D finite element meshing strategies.

Choice of element type, how to form a individual mesh, how to merge smaller meshes
together to form a complete, locally refined mesh, including constraints, and other topics
will be explored.
Chapter 4 covers general issues regarding modeling of the flexible pavements.
Critical dimensions in three directions will be determined from parametric studies and
the model will be verified through conlparison with experimentally measured data.
Chapter 5 discusses the effect of material non-linearity in the sub-base. First,
background on and implementation of the K-0 material model will be reviewed. Next,
issues regarding convergence will be studied. Finally, the best combination of non-linear
model parameters will be back-calculated by comparing with the same experimentally
measured data used in Chapter 4.
Chapter 6 focuses on the spatially varying tirelpavement contact prssures and their
effects on the response of flexible pavement structures. A method for capturing tirepavement contact pressure variations will be developed based on published data (de Beer
et al, 1997). Parametric studies will be performed to examine the effects of load, tire
pressure, asphalt thickness, and sub-base stiffness on pavement response.
Chapter 7 summarizes this study and presents conclusions and recommendations for
future research.

CHAPTER 2. LITERATURE REVIEW

2.1 Introduction
In United States, the first asphalt roadway constructed of bituminous and granular
materials was built in the late 1800's. From the very beginning of flexible pavement
construction history to early 1900's, experience dominated pavement design and
construction. Through this experience gained over the years, many design methods were
developed for detenning some critical features, like thickness of the asphalt surface. As
of 1990, there were millions of miles of paved road in US, 94% of which are topped by
asphalt (Huang, 1993). A typical flexible pavement cross section is shown in Figure 2.1.

r Wheel Load
Asphalt Concrete
Base
Sub-grade

Figure 2.1 A Typical Cross Section of Flexible Pavement System

Three types of flexible pavement constructions have been used: conventional flexible
pavement, full-depth asphalt pavement, and contained rock asphalt mat (CRAM). As
knowledge increased and other technologies developed, a composite pavement made up
of hot mix asphalt concrete (HMA) and portland cement concrete (PCC, beneath the

HMA) came into being with the most desirable characteristics. However, the CRAM
construction is still relatively rare and composite pavement is very expensive, and hence
seldom used in practice (Huang, 1993).
As mentioned in Chapter 1, various empirical methods have been developed for
analyzing flexible pavement structures. Due to limitations of analytical tools developed
in the 1960's and 19707s,the design of flexible pavements is still largely empiricallybased. The empirical method limits itself to a certain set of environmental and material
conditions (Huang, 1993). If the conditions change, the design is no loner valid. The
mechanistic-empirical method relates some input, such as wheel loads to some output,
such as stress or strain. The mechanistic method is more reliable for extrapolation from
measured data than empirical methods. However, the effectiveness of any mechanistic
design method relies on the accuracy of the predicted stresses and strains. Due to their
flexibility and power, three-dimensional (3D) finite element methods are increasingly
being used to analyze flexible pavements.
This chapter will review in detail the background and use of three major approaches
to the structural modeling of flexible pavements: layered elastic analysis, and twodimensional (2D) finite element and 3D finite element methods.

2.2 Layered Elastic Analysis


Flexible pavements can be viewed as layered systems as shown in Figure 2.1. In
layered elastic analysis, the system is divided into an arbitrary number of horizontal
layers (Vokas et al, 1985). The thickness of the individual layer and material properties
may vary from one layer to the next, but in any one layer the material is assumed to be

homogeneous and linearly elastic with an elastic modulus, E and a Poisson's ratio, v.
Figure 2.2 shows a 3-layered system.

Figure 2.2 A 3-layered Flexible Pavement System Subjected to a Wheel Load

To apply the layered elastic method, assumptions need to be made, such as:
1. Each layer is homogeneous and linearly elastic with a finite thickness hi.

2. The self-weight will not be considered.


3. A circular uniform pressure is applied on the surface.
4. Continuity conditions at the interface are h l l y satisfied. In another words, the
interface between two adjacent layers have the same vertical stress, deflections, shear
stress and radial displacement.

2.2.1 Two-layered System


The solutions for two-layered system were first developed by Burmister (1943) and
then extended them to a three-layered system and then multiple-layered system
(Burmister, 1945). With the appearance of computers, the theory can be applied to a

system with any arbitrary number of layers. The stresses in a two-layered system depend
on the modulus ratio EI/E2, and thickness-radius ratio hl/a (a is the radius of loading
area). Vertical surface deflections for two layered systems and the critical tensile strain
can be computed via analysis charts.
The case of a two-layered system is the full-depth construction in which a thick layer
of hot mixed asphalt (HMA) is placed on the sub-grade. This type of construction is
quite popular in areas where good local sub-base materials are not available. The threelayered or multiple-layered system employs one or more sub-base layers. The basic
assumptions and construction of three-layered system are very similar to the two-layered
system. Three-dimensional strains can be determined from analysis charts, such as those
found in (Huang, 1993). Using superposition, the analysis of multiple layered systems
can incorporate various loading cases in terms of different number of wheel loads, such
as single wheel, dual wheels, and dual-tandem wheels.

2.2.2 Examples of Layered Analysis Programs


A number of computer programs based on Burmister's layered elastic theory have
been developed. The program CHEV was developed by the Chevron Research Company
(Warren and Dieckmann, 1963). However CHEV can only be applied to linear elastic
materials. Hwang and Witczak (1981) modified the program to account for material nonlinearity and named it DAMA. De Jong et a1 (1973) introduced another program named
BISAR, which incorporates both vertical loads and horizontal loads. ELSYMS was
developed by Koppemxm et al. (1986).
The KENLAYER computer program developed at University of Kentucky in 1985,
incorporates the solution for an elastic multiple-layered system under a circular load.

KENLAYER can be applied to layered systems under single, dual, dual-tandem wheel
loads with each layer's material properties being linearly elastic, non-linearly elastic or
visco-elastic. To deal with material non-linearity, KENLAYER divides the granular base
layer into a number of sub-layers and the stresses at the mid-height are used to compute
the modulus of each layer. Based on the computed stresses, a new elastic modulus is
determined; this iterative process is repeated until the modulus converges.
The DAMA computer program can be used to analyze a multiple-layered elastic
pavement structure under a single- or dual-wheel load. The number of layers can not
exceed five. In DAMA, the sub-grade and the asphalt layers are considered to be
linearly elastic and the untreated sub-base to be non-linear. Instead of using iterative
method to determine the modulus of granular layer, the effect of stress dependency
(Hwang and Witczak, 1981) is included by an effective elastic modulus computed
according to in Eq. 2.1 :
-0 139 0287
E2 = 10.447hl-0.471 h2-0041
. El . E3. K ; . ~ ~ ~

(2.1)

where, El, E2, E3, are the modulus of asphalt layer, granular base, and sub-grade,
respectively; hl, h2 are the thickness of the asphalt layer and granular base. Kl, and Kz
are parameters for K-8 model with K2 = 0.5.
2.2.3 Back-calculation Using Layered Elastic Analysis
The inverse problem of determining material properties of a flexible pavement
structures from its response to surface loading (often called back-calculation) is a
common application for layered analysis method. In back-calculation, iterative schemes
are employed where the material properties are varied until the theoretical solutions

match the measured deflections. Since Poisson's ratio of pavement layers do not affect
results by much, generally only layer moduli are varied (Maestas, et a1 1994).
The PC-compatible computer programs, including ELSDEF (Bush, 1980),
MODCOMP2 (Irwin 1983), MODULUS4 (Scullion et al, 1990) and ISSEM4
(Stubstad 1988) were used to perform the back-calculation study by Maestas et al.
(1994). Deflections measured from falling weight deflectometer (FWD) field were used
to approximate layer moduli of all pavement sections. A total of 29 flexible pavement
sections were chose from different traffic loading and sub-grade soil types in Arizona.
ELSDEF uses a linear elastic analysis technique to perform the back-calculation. It
updates the layer moduli iteratively until the computed deflections match the measured
ones within an error tolerance. Similarly, MODCOMP2, MODULUS4 and ISSEM4
(Maestas, et a1 1994) all use iterative strategies when performing the back-calculation.
The results from various back-calculation programs are not significantly different except
for the sub-grade. Maestas et al. (1994) state that back-calculation is not reliable for the
use in routine design. Further research is needed to refine the process to reduce the
effects of extraneous errors.

2.2.4 Advantages and Disadvantages of Layered Elastic Analysis


With the advent of high-performance computers, the implementation of the layered
elastic method can be extended to multiple-layer system with any number of layers.
Layered elastic models are widely accepted and easily implemented. Although layered
elastic methods have such advantages, they still can not be employed to accurately
approximate the response of the flexible pavement systems. First of all, this method
assumes that each layer is homogenous with the same material properties throughout the

entire layer. This assumption makes it difficult to analyze layered systems consisting of
non-linear granular materials, such as untreated sub-bases and sub-grade. This difficulty
can be overcome by using the finite element method. Secondly, all wheel loads applied
on the top of the asphalt concrete have to be axi-symmetric, which is not true for actual
wheel loads. Limitations like these do not allow layered elastic analysis to capture the
response of pavement systems and account for all aspects of behavior.

2.3 Two-dimensional (2D) Finite Element Methods


The formal presentation of the finite element method is attributed to Turner et a1
(1956). The term "finite element" was introduced by Clough in 1960. Since its
beginning, the literature on finite element analysis has grown exponentially and there are
many of journals and researchers that are devoted to the theory and application of the
method (Zienkiewicz and Taylor, 1988; Reddy, 1993).
Two-Dimensional (2D) models were the first successful examples of the application
of the finite element method and have been in use since the 1960's (Zienkiewicz and
Taylor, 1988). The finite element method relies discretizing a domain into a number of
smaller elements, each of which is responsible for capturing variations in displacements,
strains, and stresses over its area or volume. Equation 2.2 gives the relationship between
element nodal displacements and strains:
E = SU

(2.2)

where, E is the strain vector; S is a suitable linear operator and U is nodal displacement
vector. The element stiffness matrices are computed using:

where B is a matrix of linear operators (derivatives of shape functions) and D is the


constitutive matrix. The element stiffness matrices are assembled into a system stiffness
matrix, and the resulting system of linear equations in is solved for unknown nodal
displacements. Strains and stresses can then be recovered from the nodal displacements.

2.3.1 Axis-symmetric Model vs. Plane Strain


A flexible pavement can be viewed as a semi-infinite multi-layered system with
different material properties in each layer. Basically, there are two types of 2D FE models
that can be developed for such a system: 2D axi-symmetric and plane strain.
The axi-symmetric modeling approach assumes that the pavement system has
constant material and geometry properties in horizontal planes, and the traffic loading has
bi-axial symmetry (e.g. a circular load acting at the center of the pavement structure).
This type of model is unable to consider interface shear between layers or account for
shoulder conditions or discontinuities in the pavement structures, such as joints or cracks.
It is also unfortunate that it can only be used to capture the response of pavement system
if the loading placement is not close to the shoulder or crack.
In plane strain, the stress in a direction perpendicular to the xy plane of the model is
not zero, but the strain in that direction is zero and no contribution to internal work is
made by that stress, which can be explicitly computed from the three main stress
components. The assumptions of plane strain are realistic for long bodies with constant
cross-section area subjected to loads that act in only the x andlor y directions and do not
vary in the z direction. Load will act uniformly along the outer length of body.

2.3.2 Currently Available 2D FE Models for Flexible Pavement Analysis


2.3.2.1 ILLI-PAVE Model
The ILL1-PAVE computer program (Raad and Figueroa, 1980), developed at
university of Illinois at Urbana-Champaign treats the pavement system as an axisymmetric solid domain. The resilient modulus is stress-dependent and failure criteria
for granular materials and fine-grained soils are incorporated in ILLI-PAVE. The
principle stresses in the sub-base and sub-grade layers are updated iteratively. The MohrCoulomb theory is employed as a criteria to ensure the principle stresses not to exceed
the strength of the materials. When the base or sub-grade layer is divided into several
layers and points used to evaluate the moduli are located at the mid-height of each sublayer, the minor stresses in the upper layers may be very small or become tensile in the
lower layers. Therefore, the replacement of the small or negative stress by a larger
positive stress results in a higher modulus.

2.3.2.2 MICH-PAVE Model


The MICH-PAVE (Harichandran et al, 1990) computer program is very similar to
ILL1-PAVE. It uses the same methods to model granular materials and soils and the
same Morh-Coulomb failure criteria. MICH-PAVE uses a flexible boundary at a limited
depth beneath the surface of sub-grade instead of a rigid boundary at a large depth below
the surface. MICH-PAVE is capable of performing linear and non-linear finite element
analysis of flexible pavements. It assumes axi-symmetric loading conditions and
computes an equivalent resilient modulus for each pavement layer that is obtained as the
average of the moduli of the finite elements in the layer that lie within an assumed 2: 1
load distribution zone. For non-linear materials, MICH-PAVE employs the stress-

dependent K-8 model to characterize the resilient modulus of soils through a stressdependent n~odulusand constant Poisson's ratio.
The comparison of results from MICH-PAVE and those from ILLI-PAVE illustrates
that MICH-PAVE yields more reasonable outcome than the latter with measured data
(Chen et al, 1995).

2.3.2.3 Commercially Available FE Programs


ABAQUS, a commercially available 3D FE program (Hibbitt et al, 1992), has been
used in the structural analysis of pavement system. It has the ability to accommodate
both 2D FE analysis and 3D FE analysis and use reduced integration elements (3D) to
reduce the total computational time (Cho et al, 1996). To investigate whether 3D FE
analyses are necessary or even benefit in the routine design, differences between the
results yield by 3D analyses and those from 2D FE by presented by Chen et a1 (1995).
They showed that ABAQUS tends to generate the lowest tensile strain in all cases and
maximum deflections compared with those obtained from the 2D axis-symmetric
program MICH-PAVE; ILLI-PAVE gave the lowest maximum surface deflection due to
the fixed boundary at a certain depth.

2.3.3 Advantages and Disadvantages of 2D FE Analysis


The two-dimensional FE method has some advantages such as requiring relatively
little computational time and memory and easy of pre- and post-processing (compared to
3D FE methods). Further, the two-dimensional methods can overcome some
shortcomings of layered elastic analysis (see section 2.2). However, the 2D FE methods
are still not sufficient enough to capture detailed response. For example, the 2D plane-

strain model is limited to the use of line loads and the axis-symmetric modeling approach
models wheel truck loads as circular. Many investigations (Weissman, 1997; Cunagin
et al, 1991) have illustrated that the contact area between the pavements and tire is
essentially rectangular, and that titelpavement contact stresses can vary significantly over
the tirelpavement contact area (Tielking et al, 1995).

2.4 Three-dimensional (3D) Finite Element Methods


In order to overcome the limitations of layered elastic analysis and 2D FE methods,

3D finite element methods are increasingly being used to model the response of flexible
pavements. 3D finite element modeling is increasingly viewed as the best approach to
understand pavement performance. A pavement system is typically modeled as a multilayered structure with different material properties in each layer. Interface elements or
springs can be used to transfer the shear between layers, and well-controlled boundary
conditions are critical to analyze behavior of entire pavement system (Blab et al, 2000).
While 3D modeling method can generate more realistic results than 2D modeling, it
generally requires more intensive pre-processing procedures. Further, the number of
elements increases, the total computational time and the amount of memory will increase
tremendously. Fortunately, as computer industry develops exponentially and appearance
of high-performance algorithms, more elements can be used and accordingly the results
will be more and more accurate.
Further, a number of researchers have shown that spatially varying contact pressures
between the tire and pavement can significantly affect response (Tielking et al, 1995;
Weissman 1997). 2D models cannot capture this effect.

To overcome these shortcomings, a number of researchers have moved to 3D finite


element models.
2.4.1 Cho et a1 (1996)
Cho et a1 (1996) compared FEM (2D, axi-symmetric and 3D) solutions with elastic
theory. A reasonable approximation was yielded by 3D model (model size and
I

boundary conditions were well-controlled). Quadratic elements in the 3D model resulted


in stresses that matched layered elastic theory very well (Cho et al, 1996).
2.4.2 Hsien H. Chen, Hurt M. Marshek, and Chhote L. Saraf (1990)
This study documents the effect of high inflation pressure and heavy axle load on
AC pavement performance by using 3D FE model. A computer program TEXGAP-3D
(developed with ABAQUS), was selected to predict the performance of flexible
pavements. The pavement structure is assumed to be linear elastic and homogenous.
Results obtained from TEXGAP-3D were also compared to those of the layer program
ELSYMS (Kopperman et al., 1986) for a uniform circular pressure. There is a close
correspondence between the results from the two models. It was found that the uniforn~
pressure model will predict a higher percentage increase in tensile strain than will the
non-uniform pressure model for either underinflated or overinflated tires; inflation
pressure will have less than 2% effect on the compressive strains at the top of the subgrade for two types of models; axle load has a significant effect on sub-grade
compressive strain.

2.4.3 K. D. Hjelmstad, J. Kim and Q. H. Zuo (1996)


The investigation details the essential aspects of modeling pavement structures with
3D FE analysis. The research group examined some important impacts on overall
problem size such as: mesh refinement, domain extent and element size transitions.
The problems of using 3D FE analytical tools are generation of 3D FE meshes, the
large amount of computer memory and computation time. Due to highly localized wheel
load, the stress gradient are greatest in the neighborhood of the load, and therefore, the
finite element mesh must be finest in that region. The domain size affects the total
displacement response of the pavement system. Infinite elements were used to mesh the
far field region, which allow the displacement and stress fields to decay to zero and
reduce memory and CPU time requirements. The element aspect ratio and smooth
transitions from one element size to another affect the accuracy of the solution and it is
essential to make transition to larger elements in the region of small stress gradients.
Achievement of good aspect ratio also results in more accurate results. Reduction of
computational time to solve large 3D FE pavement problems is important, especially
when including material non-linearity and interface conditions.

2.4.4 Sam Helwany, John Dyer, and Joe Leidy (1998)


This study focused on the usefulness of the finite-element method in the analysis of
three-layered pavement system subjected to different types of loadings. Various factors
such as axle type (single axle dual tire, tandem dual group and tridem axle group), axle
load, tire pressure, vehicle speed, and pavement types were examined and different
material constitutive models were also considered, including linear elastic, non-linear

elastic and visco-elastic. The 3D FE computer program NIKE3D was used for this
analysis. The results indicate that measured tensile stains at the bottom of the AC
layer are very sensitive to vehicle speed and only longitudinal strains at the top and the
bottom of the AC layer were sensitive to tire pressure. It was found that FE modeling of
modeling of flexible pavements, if validated, can be very useful, because it can be used to
predict primary response parameters without resorting to field experiments, which may
be costly. Further, 3D FE analysis has some advantages such as: it may substitute for
full-scale testing; the analysis may be used to form the basis for generalized mechanistic
design procedures; and the analysis may used to verify results from simpler 2D analyses.

2.4.5 Samir N. Shoukry and Gergis W. William (1999)


In this report, the research team used a 3D finite element model to back-calculate the
layer moduli of flexible pavement structures and compared the results with predictions of
back-calculation computer programs: MODCOMP, MODULUS, and EVERCALC
(Deusen 1996). It was found that solution of pavement response problems using 3D
FEM does not require ?he assumptions usually made in layered elastic theory about
the geometry of pavement structure, materials, or interfaces between member layers.
The back-calculation procedure was developed using LS-DYNA3D. The
experimentally measured deflection basin were obtained via Falling Weight
Deflectometer (FWD) test. Then the deflection basin is used with back-calculation
program to evaluate a set of layer moduli. An eight-noded solid brick element mesh was
constructed for the 3-layered pavement structure (AC, base and subgrade). Materials
were considered to behave linearly.

As tabulated results shown in the paper, the mechanistic 3D FEM approach in


backcalculating layer moduli produced almost the same result. In other words, the FE
evaluated moduli do not require correction and can be used as a reference for assessing
the performance of back-calculation.
2.4.6 A Bensalem, et a1 (2000)
In this investigation, the research team detailed the response of un-cracked and
cracked flexible pavement subjected to wheel load using 3D finite element model. The
selected pavement structures have five asphalt thicknesses ranging from 200 rnrn to 400

mm and are considered to be made up of homogenous materials of uniform thickness and


horizontally infinite. A 3D FE model was developed to simulate the pavement system,
which was regarded as extending horizontally to infinity and superimposed on a half
space of infinite depth. The model assumed that each layer has a constant stiffness
throughout the depth and layers were fully bonded to each other (no slip and separation at
interfaces). The effect of single and dual wheel loads were investigated. It was found
that with the presence of shear at the crack tip is likely to cause any crack to propagate;
shear strain plays a important role in crack initiating and propagating and rises to a peak
value in the top 50- 100 mm of asphalt and then reduces with depth. Further, surface
cracking may be more serious in thinner pavements than in thicker pavements. Tension
at the pavement surface in the neighboring region of the wheel edge in thicker asphalt
was higher for sections with thicker asphalt.
2.4.7 Ronald Blab and John T. Harvey (2000)
In this study, having developed an effective method to measure the tirefpavement
contact stresses using Vehicle-Road Surface Pressure Transducer Array (VRSPTA), the

authors tried to predict the probability that failures will occur. It was found that the most
common circular load model with uniform vertical stress equal to the tire inflation
pressure is not sufficient to account for distresses occurring close to the surface. The
VRSPTA directly measures and quantifies the contact stresses between the passing tire an
pavement surface, therefore, it is more realistic. The described FE model validation was
performed using linear elastic layer theory. The graphic results illustrate that the FE
model gives very reasonable results. However$ requires that complex constitutive
relationships be used to accurately predict the rutting test results.

2.5 Summary
Significant progress in both modeling strategies has been made in past years. In this
Chapter, theoretical background, advantages and limitations of layered elastic model, 2D
FE methods and 3D FE methods were reviewed. Layered elastic models can not
accurately capture non-uniform tire contact pressure and multiple-wheel loads and have
difficulty incorporating material non-linearity. 2D finite element models do not allow the
realistic modeling of applied loads. To overcome these problems, 3D FE models have
been developed.
Although a nunlber of studies employing 3D finite element analysis have been
performed, there are some import topics that need further research:
The automatic and efficient generation of locally refined meshes for 3D FE
analysis of flexible pavement systems.
Verify that the models can adequately predict measured response. This must
include studies on the interaction between adjacent wheel loads and an examination
of required model extents.

Investigate the implementation of common non-linear constitutive models for the


base layers in 3D models.
Examine the effects of spatially varying tire pressures on flexible pavement
systems with varying thickness in AC layer and different material properties in the
sub-base. Prior 3D FE-based researchers (i.e. Blab et al, 2000) has shown these
effects to be significant. However, a comprehensive study on the effect of spatially
varying contact pressures on response has not been conducted to date.
The following four chapters directly address these issues the order listed above.

CHAPTER 3. MODELING and MESHING STRATEGIES

3.1 Introduction
As mentioned in chapter 2, both two-dimensional (2D) and three- dimensional (3D)
methods can be employed to capture the structural response of flexible pavements. Many
researchers have investigated the response of flexible pavements and 2D programs have
been developed specifically for flexible pavements such as MICHPAVE and
ILLI-PAVE. However, 2D analysis cannot accommodate non-uniform tire contact
pressure and multiple-wheel loads; 3D finite element models must be used to properly
capture these effects.
However, in order to solve the relatively large 3D problems on the desktop
computers, the meshing code has to be efficient and automatic. This chapter focuses on
the development of effective meshing tool. The meshing code was developed using the
object-oriented programming language, C++.
First of all, an overview of the finite element models will be presented, where
issues such as localized mesh refinement, model size, and boundary conditions will be
discussed. A brief discussion of element types and implementation will be presented.
Secondly, the overall procedure for generating the meshes will be covered. This will
be followed by a discussion of the specific steps taken to efficiently generate and
combine individual portions of the meshes. Finally, a brief summary concludes the
chapter.

3.2 Overview of FE Model


3.2.1 General Layout of Flexible Pavement System
Figures 3.1 and 3.2 show a representative 3D finite element mesh of a flexible
pavement system in plan and elevation.

I-

Wheel Load
X

Figure 3.1 A Typical Plan View of a Meshed Pavement System

Surface
nterface

terface

Sub-grade

Figure 3.2 Elevation View

The finite element pavement structure used in this study is made up of three layers,
including asphalt concrete (AC) layer, sub-base and sub-grade layers. Each layer of the
model has different material properties, and is of uniform thickness.

3.2.2 Local Refinement Region


The dimensions of each element are not uniform as shown in Figure 3.1. More
elements are used around the wheel loads where stresses and displacement gradients are
higher. According to Huang (1993), the maximum tensile stresses and strains at the
bottom of AC layer or maximum deflections are generally used as criterion to design the
flexible pavements. Prior research (Cunagin, et al, 1991) illustrates that critical stresses
and deflection (peak values) tend to occur around the wheel loads and those should
decrease in the far field. In general, with other conditions (traffic load, boundary

'

conditions, model size, material properties and so on) the same, the case with a relatively
large number of FE elements will yield more accurate results than the one with less FE
elements. Yet, the more FE elements used, the more computer resources required.
Therefore, it is effective to refine the region around the wheel loads and use fewer, larger
FE elements away fiom the load.

3.2.3 Finite Boundaries of Model


Various types of boundaries have been examined in different models, including rigid
boundaries and flexible boundaries. The boundary conditions used in this study are
shown in Figure 3.2, where the nodes on each face except the surface are constrained
from moving along the direction perpendicular to that face and are free to rotate.

3.2.4 Interface Slip and Constraints

Because the layers of pavement systems consist of different materials that may be unbonded, we would like to model slip between layers. To accommodate this need, two
interface layers (one between AC and sub-base, the other between sub-base and subgrade) were introduced to capture slippage. Basically, two possibilities, fully bonded and
un-bonded will be considered. For the fully bonded case, nodes on the surface of base
layer are perfectly bonded to nodes at the bottom of AC layer (i.e. the matched nodes
have same displacement in all three (x, y, z ) directions). For the un-bonded case, the
matched nodes will be allowed to separate under tension. Therefore, the interface layer
will be tensionless, and have stiffness constants Kx and Ky (x and y direction stiffness per
unit area, respectively) to capture shear transfer between layers.

3.3 Element Types


Two types of FE elements, 20-noded brick element (Zienkiewicz and Taylor, 1988)
and 16-noded quadratic interface element (Davids et al, 1998) are used to generate the
meshes for both member layers and interface layers.
3.3.1 20-noded Brick Element

There are several other types of brick elements that could be used, including 8nodded linear elements, 27-noded quadratic elements, 32-noded cubic elements and so
on. In general, the more nodes per element, the more accurate the element, but the more
difficult mesh generation and application becomes. The 20-noded brick element achieves
a balance between the difficulty of meshing and accuracy of results. Each model layer is
meshed with 20-noded brick elements; a single 20-noded brick element is shown in

Figure 3.3.

Figure 3.3 A Typical 20-noded Brick Element


Nodes 1-8 are at the corners of the brick and all other nodes are at the middle of
the edges. Note that node numbers 1-20 are the local connectivity, and node number
associates with the local coordinates of that node. Global connectivity and coordinates
will be computed in global coordinates (details will be discussed later in this chapter).

3.3.2 ldnoded Interface Element


To model the interface between the layers, 16-noded interface element will
be used as shown in Figure 3.4.

Figure 3.4 16-noded Quadratic Isoparametric Interface Element


The 16-noded quadratic isoparametric interface elements (Davids et al, 1998) allows
shear transfer between layers in the pavement system. The local connectivity and
coordinates are defined as shown above. The thickness of the interface element, 6, is
very small. The interface element has two important parameters, K, and K , stiffness
constants per unit area in x and y directions. These two parameters allow us to consider
the effect of shear transfer between the surface layer and base layer, or between the base
and sub-grade layers. When Kx and K, are very small, there is no shear transfer; when K,
and K, are very large, there is perfect horizontal bond.

3.4 Meshing an Individual Block of Elements


Each regular portion of the mesh within an individual layer will be meshed
separately. Once the coordinates of 8 comer nodes of a 6-face mesh are specified, three
other control parameters N,, N, and N, can be used to evenly divide the mesh in the 3-

axis directions corresponding to global coordinates system. Figure 3.5 illustrates how
coordinates and local connectivity will be computed. Note that while curved edges or
faces are not considered, the 6-sided region does not need to be a rectangular prism.

NY
Figure 3.5 Brick Element Mesh
The nodes are numbered first along the z axis and then they axis and eventually the x
direction. The interface elements can be meshed in the same way (first go along the zaxis, y-axis and x-axis in that order). The order of brick element mesh in plan will
determine how the interface elements are meshed.

3.5 Merging and Forming 3D Model


After meshing all individual, regular regions of brick elements and the interface
layers, the mesh for entire system is formed. This requires that we have the ability to

merge adjacent meshes. For example, two meshes can be merged together to form a
bigger mesh, then a 3rdblock of elements can be merged to form an even larger mesh,
etc., until all individual meshes are merged together. If distance between any two nodes
is effectively zero, these two nodes take the same number. For simplicity, a 2D example
can be used here to illustrate the situation as shown in Figure 3.6. The connectivity of
the remaining, unmatched nodes on the second mesh are added to new whole mesh
sequentially.

After Merge
Figure 3.6 Merging the Meshes

3.5.1 A Naive Method


As presented above, it seems that all nodes on the two meshes need to looped over to
find each pair of nodes that match at the interface. For instance, if there are N 1 nodes in
mesh I and N2 nodes in mesh 11, then the total work of merging is proportional to N1 X
N2 (for each node in mesh I potentially, all nodes in mesh I1 will have to be checked to
find a match). While this approach can solve the problem of merging, it is expensive in
terms of run-time for large problems. A way to fix this is using a grid-sort method
detailed as follows.

3.5.2 Grid Sort Method


It is true that run-time of meshing and merging depends on how big the model is. In
general, there may be thousands of 20-noded brick elements involved and
correspondingly, there will be tens of thousand of nodes and the naive approach is
expensive in terms of computational time. Another alternative (called the grid-sort
method) was developed to deal with this situation. This basic concept has been used by
others for similar situations (see Davids and Turkiyyah, 1998, for example). Figure 3.7
illustrates basic concept of the grid-sort method in 2D.

mesh

Figure 3.7 Building Grid Sort Element Block


In Figure 3.7, all bold numbers represent the grid sort element numbers and the
remainder are the nodes in the finite element mesh. The grid sort block is constructed in
the same manner we built the brick elements: 0-based element number first increases
along y and then along x direction in that order. Because the dimensions of each grid sort
element are identical, it is very easy to figure out the boundary coordinates of four corner
nodes of each element.
Next we can loop over all nodes in the mesh and compute how many nodes there are
in each individual grid sort element and what their numbers are. For instance, element 0
contains nodes 2 and 3; grid sort element 1 has nodes1 and 6; grid sort element 9 has
nodes 12 and 13, and so on. Therefore, each individual grid sort element will have a few

nodes instead of thousand of them (of course, the number of nodes in each grid-sort
element depends on total number grid sort elements).
When another mesh (called mesh 11) needs to be merged to the existing one (mesh I),
all nodes on the mesh I1 will be looped over. For each node in mesh 11, it is easy to figure
out which grid sort element it belongs to by checking the coordinates nodal coordinates
against those of the grid sort cells. Once we have done that, we can compute the distance
between this node (on mesh 11) to all nodes only in that grid sort element which belong to
mesh I (see Figure 3.8).

Mesh I

2D Grid Sort Element Block


Figure 3.8 Merging Meshes

For example, it is obvious that node 1 on mesh I1 is in the grid element 6 and mesh I
only contributes nodes 14, 17 and 18 to grid element 6. Therefore, we can calculate the

distance between node 1 and nodes 14, 17, and 18 as opposed to looping sequentially
over all of the nodes in mesh I. After repeating the process of checking by corresponding
grid cell and tracking the matching nodes, it is simple to re-number the nodes in mesh I1
and re-define the element connectivities. The small example shown here does not seem
to save much time, but if the number of total nodes are about several thousand, the run
time will be decreased tremendously as shown in the next section. The algorithm of grid
sorting procedure is shown as follows:
Xmin,Y m i Zmin
, = minimum x, y, z coordinates of mesh - 1 respectively
X,, Y,, Z , = maximum x, y, z coordinates of mesh * 1.05 respectively
Ndr, N+ Nd, = number of divisions of grid in x, y, z directions
1 get [ X m , Ymm,
x m i n , ymin, Zmin];
2 6~ = (Xmm- Xmjn)1 Ndx;
3 6y = (Y- - Ymjn) I N4;
4 6~ = (Zmm - Zmjn) 1 Ndr
5 build-sort ( )
6 loop over all grid sort elements
7 for i = 1 to number of elements (Nm* N&,* Nm )
store 8 connectivities and coordinates of each element in order;
8
9 loop over all nodes in mesh I
10 for j = 1 to number of nodes
11
store node number in list of each grid-sort element;
12 End
13 End
14 Loop over all solid elements
15 for i = 1 to number of elements
16 loop over all nodes in mesh 11
17
for j = 1 to number of nodes
determine which grid-sort element the node lies in;
18
loop over all nodes in that grid-sort element;
19
20
for k = 1 to number;
21
find a match, two nodes become one;
22
End
23
End
24 End
25 re-number all other nodes in mesh 11;
26 END

3.5.3 Numerical Comparison of Two Methods


Figure 3.9 illustrates how efficient the grid sort algorithm is with results obtained
fiom real computational cases. Case I (number of grid sort elements l x l x l ) represents
the naYve method, and Case 11, I11 and IV represent grid sorting procedure with different
number of grid sort elements. The model used for testing the efficiency of the grid-sort
algorithm has 2000 solid elements is shown in Figure 3.9:

,
$

I
I I
I :
i i
I :

/
...'

.',..'"
...-.--

,.-'

I:
I

1 0 0 2 0 0 3 0 0 4 0 0 5 0 0 6 0 0 7 0 0 8 0 0 9 0 0
Tim t o m e model (seaxrls)

Figure 3.10 Computational Time Comparison of Two Methods


As showing Figure 3.10, when the number of total solid elements increases beyond a
few hundred, use of the grid sort algorithm drastically speeds the process of merging
meshes. These runs were conducted on a Pentium I1 350MHz PC.

3.6 Implementation
3.6.1 Typical Procedure for Generating a Layered Pavement System
All model layers have exactly the same configuration or plane dimensions. In
general, each portion of an individual layer will be meshed first and then merged together
to fornl the entire layer. Once all model layers and interface layers are meshed, they will

be merged together again to form whole pavement system. The algorithm is shown as
following:
Loop over all member layers
for i = 1 to number of layers
mesh portion (Inear
) the wheel loads;
mesh portions of the s a l e layer in far field and then merge to portion 1;
End
Loop over 2 interface layers
for i = 1 to 2;
mesh portion (I-inter) near the wheel loads;
mesh portions of the same layer in far field and then merge to portion I-inter;
End
Merge the interface layers to pavement member layers;
END

3.6.2 Control Parameters for Meshing Code


A computer program, Flexible-mesh coded in C++, was developed to generate the
mesh and form the model system. As stated previously, some control parameters are
listed as below:
unitflag indicates the unit system will be used; 0 implies English units (inches, kips and
etc), 1 implies metric units (mrn, Newton, and so on);
numberlayers represents how many member layers the pavement structure has (3 or 4
layers are most common, yet it can have any number of layers in theory);
bond-12 and bond-23 indicate bonded control parameters, bond-12 is in charge of
interface layer between AC and base; bond-23 is in charge of interface layer between
base and sub-grade (0 implies perfectly bonded, 1 implies vertical separation and slip
permitted).

N , and ti, represent number of sub-layers and thickness of the ithmember layer;
The plan view is shown in Figure 3.1 1

Plan View

Figure 3.1 1 Plan View of Merged Meshes

Nix,Ni, number of divisions in x and y directions;


dix,di, inner dimensions in x, y directions;
Di, Di, outer dimensions in x, y directions;

Nd, number of elements on the diagonal

KI2,shear stiffness at interface between AC and base


K2& shear stiffness at interface between base and sub-grade
Material -type, 1 implies linearly elastic and 3 implies material non-linearity (k-8 model);

E,, modulus of elasticity of the ithlayer;


y, Poisson's ratio of the ih layer;

Coeg-thermal coefficient of thermal expansion;

pi,density of the i' layer;

Loading-type, (1-7) indicates different loading types;


Num-load, number of loads;

Xi,Yi, Zi, x, y, z coordinates to define the location of rectangular load (i = 1 to 4);


Mag-load

total vertical load

Hz,, Hz2, Hz3, unit hz vector


Mag-load-hz, total hz load
nodal-displ.dat and nodal-stress.dat names of files to store displacement and stresses
joint -stress.dat, contact.dat, LTE.dat, and dowel-contact. dat other make-up files
required by EverFE program
coarse.mes name of fe-input file
end of mesh control file

All control parameters are basically the same for all possibilities except the loading
part, which deals with different traffic loads, for example, to handle circular patch loads,
the solver needs know the diameter of this circular area, etc.

3.7 Summary
In this chapter, an overview of modeling strategies and element types necessary for
the effective 3D finite element modeling of pavement systems were presented. Some
topics related to mesh generation were covered as well, including local refinement of the
region around wheel loads, modeling interface slippage, and incorporation of finite
boundaries.
Basically, each model layer forms a block of the mesh. Individual layer and interface
layer meshes are merged to other meshes of their own type to generate a bigger mesh.

These meshes of brick elements or interface elements are then merged together to form
the entire pavement system. The interface layers can allow slip between layers with
horizontal shear transfer; by specifying different shear stiffness constants for the interface
elements, we can simulate different degrees of slip.
Two types of nodal constraints are considered at layer interface. For the bonded
case, the matched nodes are fully bonded together so that they have the same
displacement in three directions; for the un-bonded case, the matched nodes separate
under tension in the vertical direction, and can slip in the horizontal plane.
To assist in generating the mesh efficiently, a grid sorting algorithm was developed,
and shown to be very effective using a numerical example.

Chapter 4. MODEL DEVELOPMENT AND VERIFICATION


4.1 Introduction
Accurately predicting the maximum stresses and deflections in flexible pavement
systems under realistic loading conditions requires a three-dimensional (3D) finite
element modeling approach as discussed in Chapter 3. However, development of a 3D
finite element (FE) model is not a trivial task. One important problem is how large of a
model should be generated. In reality, flexible pavements can extend infinitely in the
longitudinal and vertical directions. While the infinite extents can be approximated with
infinite elements (Chen et al, 1995; Cho et a1 1996), we would like to model a finite
domain (as presented in Chapter 3) and still get accurate predictions for deflections and
stresses. Further, it is desirable that the model be able to replicate field-measured data.
In this chapter, both of these issues will be examined. First of all, the combined
effect of adjacent wheels on model response is considered. Second, numerical studies are
performed to determine the model extents required to ensure adequate accuracy. Third,
experimentally measured deflections from two different flexible pavement structures in
Maine (Nickels, 1995) will be compared with the finite-element model predictions. Both
pavement structures will be treated as 3-layered elastic systems, and layer moduli will be
back-calculated to give optimized comparisons with experimental data. Finally, a brief
summary finishes the chapter.

4.2 Effect of Multiple Wheel Loads


4.2.1 Single Dual - Wheel, Single Axle and Dual-Wheel Tandem Axle

In general, the wheels of a truck are spaced at a distance of about 1800 mm


transversely and adjacent axles can be as close as 1200 mm in the longitudinal direction
as shown in Figure 4.1. To simulate the actual conditions, one should model the entire
system with multiple-wheel loads. But for the reason of simplicity and to minimize
computational resources, one wants to build a full size model with only one dual wheel
load on it. The critical question is whether the adjacent dual wheel load or adjacent axles
affect the tensile stresses in the asphalt concrete beneath the critical dual wheel load
significantly. To examine this, we will subject the same finite element model (see Figure
4.2) to three different loading cases. The first one will be single dual wheel (I), the
second one will be single axle (four wheels, I + 111), and last one will be two dual-wheel
axles (eight wheels, including I-IV) as shown in Figure 4.1. Each axle will be 80 kN, the
usual magnitude assumed in design. Each wheel will be assumed to have a 200 mm
width by 145 mm long contact area, which gives an average pressure of 690 kPa. All
material properties are assumed to be linearly elastic and are given in Table 4.1.

Figure 4.1 Typical Dual-wheel Tandem Axle

740mnl

Sub-grade

Plan View

Elevation View

Figure 4.2 Model Description (Multiple-wheel Loads)

Table 4.1 Material Properties

Layer
As~haltConcrete
Base
Sub-grade

Dimensions(mm)
125
635
740

E (MPa), v
1400.0.35
50,0.29
20,0.29

Figures 4.4 - 4.6 show the maximum principal stresses at the bottom of the asphalt
concrete under a single dual wheel load (loading case I), single axle load (loading case I

+ loading caseIII), and dual axle load (loading case I-IV) respectively.

Figure 4.4 Maximum Principal Stress at the bottom of AC (single dual wheel,
loading case I)

Figure 4.5 Maximum Principal Stress at the bottom of AC (single axle, loading case
I+II)

45

Figure 4.6 Maximum Principal Stress at the bottom of AC (dual axle, loading case IIV)

The maximum principal tensile stresses at the bottom of AC due to all single dualwheel, single axle and all 8 wheel loads are 0.8865 MPa, 0.8569 MPa and 0.8536 MPa,
respectively. The largest difference among them is 3.85%. Clearly, it is reasonable and
accurate enough that we study the response of pavement system under single dual-wheel
loads.

4.3 Determination of Critical Model Dimensions


Based on the results of the loading study in the previous section, the 3D finite element
models used from this point forward will incorporate only single dual wheel loads. The
focus of this section is the detem~inationof the appropriate effective model dimensions.

This will be determined by subjecting a typical pavement system to a known dual


wheel load. The plan dimensions and total model depth will be varied, and the effect of
the model dimensions on the maximum tensile stress in the asphalt concrete will be
examined. Ideally, the change in tensile stress with increasing model size will become
negligible for a certain set of plan and vertical dimensions.

4.3.1 Model Description


As shown in Figure 4.7, the model to be studied is taken from the North Yarrnouth
control section as reported in (Nickels, 1995). For the purpose of this study, E2 = 50
MPa, E3 = 20 MPa were assumed moduli of elasticity in sub-base and sub-grade
respectively with a typical wheel load of 40 kN. Note that these values are fairly realistic
as will be shown in section 4.4. Further, 2D analysis performed on the same pavement
section by (Nickels, 1995) employed E2 = 50 MPa, E3 = 20 MPa as well.
Wheel Load = 40 kN
-

Asphalt Concrete

E2 = SOMPa, v = 0.29

Base

Sub-grade

Figure 4.7 Layered System to be Modeled


Nodes in each face will not be allowed to move in the direction which is
perpendicular to that face; nodes on the edges formed by the meeting of two or three

faces will be restricted ftom moving in those directions which are perpendicular to the
faces. For simplicity, the model will be square.
Due to the bi-axial symmetry of the loading and boundary conditions, we can use
a quarter-symmetric mesh. Figures 4.8,4.9 and 4.10 show the finite element model and
mesh used here. Each layer will be assumed perfectly bonded for the purpose of
determining critical dimensions.

,-- Dual-Wheel load = 40 kN

Layer 3 Subgrade

Lz restraint

\
x restraint

Figure 4.8 Elevation View of Pavement System


Note that there are two interface layers (gray in color) between layer 1 and layer 2,
and also between layer 2 and layer 3 to capture slippage occurring between member

layers. The loading area is assumed to be a rectangular region and no moving loads are
taken into consideration.

y restraint

restraint

4 v'

restraint

Figure 4.9 Plan View of System (quarter symmetric) with B.C

Figure 4.10 3D View of Meshed Model (quarter symmetric)

4.3.2 Results of Study on Critical Model Dimensions

Critical dimensions in plan are extremely in~portantcan affect such results as


stresses, strains and maximum deflections significantly. Therefore we will first fix the
total depth of the model at 1500 mm, which corresponds to sub-grade vertical dimension,
z = 740 mm of the model and examine the effect of plan dimensions on response. Once

the critical plan dimensions have been determined, the vertical dimension will be varied
to determine its effect on stresses and displacements. The pavement will be modeled as a
square region and the longitudinal (x) and transverse (y) dimensions will be varied
simultaneously. Figure 4.1 1 shows the effect of plan dimensions on the maximum tensile
stresses in the asphalt concrete for different values of the sub-grade depth, z.

Plan Dimensions(mm)

Figure 4.11 Stress Variation vs. Plan Dimensions

Clearly, Figure 4.1 1 illustrates that the stresses converge to a constant value and that
the critical dimensions in plan are about 2000 mm X 2000 mm irrespective of z. The
change in tensile stress in asphalt concrete layer for plan extents beyond 2000 mm X
2000 mm is negligible. Interestingly, stresses peaked for z = 740 mm and then went
down slightly when the sub-grade vertical dimension was increased to 1640 mm.
However, the effect of z on the tensile stress is quite small. Based on the results, it seems
reasonable to fix the quarter-symmetric model dimensions at 2000 mm X 2000 mm in
plan and keep z larger than or equal to 740 mm. Equivalently, if a quarter symmetric
model is not used, we could employ a model with plan dimensions 4000 mm X 4000 mm.

4.4 Comparison with Experimental Results


It is important that any model be able to replicate experimentally observed data.
Achieving this with flexible pavement systems is difficult, since material properties of
the sub-grade soils are often unknown. In practice, this difficulty is often overcome by
back-calculating effective layer moduli or thickness so that the model accurately predicts
the experimentally measured displacements. In this section, we will back-calculate pairs
of El and E2 to give good predictions of field measured displacements. Two different
sub-grade thickness, z will be considered, resulting in two different El and E2 pairs.
Predicted asphalt stresses for each distinct combination of El, E2 and z will be examined,
since asphalt stresses are critical design parameters.

4.4.1 Model Description


Two control sections will be examined as shown in Figure 4.12 and for each of them,
two models with different depths (the same plan dimensions, 2000 mm X 2000 mm) will

be selected. The only variable here remains unknown is the depth of sub-grade, z. For
different values of z, different pairs of Ez, E3 (nlodulus of elasticity of sub-base, subgrade, respectively) can be computed to give a good fit to the experimentally measured
deflections. In this section, we will examine the effect of varying z, and back-calculating
different E2, E3 combinations to give good agreement. From previous studies, z =
740 mm is the critical vertical dimension, and then z will be varied from 1040mm, 1940
mm to 2040 mm. In general, z can be any value larger than or equal to 740 mm. For the
North Yarmouth control section, we pick up z = 740 mm (total depth = 1500 mm, comes
from previous work done in section 4.2) and z = 1940 mm (total depth = 2700 mm). For
the TWP3 1-MD control section, models with z = 1040 mm (total depth = 1800 mm) and z
= 2240 mm

(total depth = 3000 rnrn) are employed, since the thicker the asphalt concrete

layer, the thicker the base and sub-grade layer or bigger difference between the two
models are expected.
The boundary conditions and loading cases will be the same as in section 4.2. A
quarter-symmetric mesh will be used. All corresponding nodes between layers will be
fully bonded (no slip) and the soil will be treated as linearly elastic. No self-weight of any
layer is taken into consideration.

Wheel load

Wheel load

* A

North Yarmouth

Figure 4.12 Control Sections

4.4.2 Experimentally Measured Data


Measured data for the North Yarmouth and TWP3 1-MD was gathered with a
Benkelman beam (Nickels, 1995). The experimental set-up used is shown in Figure 4.13.
The single 80 kN rear axle of a dump truck carrying was positioned on the desired
location and the wheel probe inserted between the dual rear tires.

4.4.2.1 Benkelman Beam


The standard Benkelman Beam (BB, 8 ft) is used to measure pavement deflections
under wheel loads. For this research, it was modified to the standard BB (16 ft) in order
to measure the large diameter deflection basin.
The BB works by inserting the probe between the dual tires of loaded truck as shown
in Figures 4.13-4.14. When the truck drives away, then the pavement, and hence, the
probe, will rebound. The probe which is connected to the beam is pivoted at a fulcrum

point attached to a reference arm resting in back of the influence of the load. The
maximum deflection underneath the dual wheels is taken by the wheel probe mounted 16

fl in front of the fulcrum point.


Movement is measured by a single dial gage mounted 4 fl back of the fulcrum point.
Thus, deflection under the centerline of the wheel load is equal to four times the dial gage
reading and is termed centerline deflection.

4.4.2.2 Testing Procedure


The following procedure was used for measuring pavement deflection at the North
Yarmouth and TWP3 1-MD projects.
1.

A grid system encompassing eight points per 100 fl test section was laid out
for each project.

2.

A single rear axle dump truck loaded to an 18 kip rear axle load was
positioned on the desired location and the wheel probe inserted between the
dual rear tires.

3.

Initial readings were taken for the gages (for our use only).

4.

The truck was driven away and the pavement was permitted to rebound for
about 1 minute.

5.

Final dial gage readings were taken and the BB was moved to the next point.

According to Figure 4.14, the rebound of the pavement directly under the wheel load,
A , is computed using:
A = +dial

- dial,)

where dial, dialo are the final and initial dial gage readings respectively and A is the
corrected deflection value.

(4-2)

-1

Rear of

DY/k

Figure 4.13 Plan View of Location of Benkelman Beam Probe between Dual Truck
Tires

Height
A d justmen t

Note:

The modified Benkleman beam used to measure pavement deflections

N o t t o scole

4.4.3 Back-calculation for North Yarmouth Pavement Section


Here, two different vertical dimensions of sub-grade, z = 740 mm and z = 1940 mm,
are chosen and corresponding values of resilient moduli (E2 ,E3) are computed that give
the best comparison with measured data in a least-squares sense. The least-squares error,
0, is computed using:

where Dm and D, are measured and predicted displacements respectively.

A formal optimization algorithm was not used for this procedure; more simply, 0 was
computed for different combinations of E2 and E3,and the pair that gave the best 0 was
selected. Figure 4.15 shows selected results from the simulations. The best fit to the
experimental data was seen with E2 = 67 MPa and E3 = 10 MPa. Figure 4.16 shows good
agreement between predicted and measured displacen~ents.Figure 4.1 7 shows the
maximum principal stresses at the bottom of the asphalt concrete layer. The largest
maximum principlestress is 0.836 MPa.

'f,

15

E2 =67 MPa
E2 = 100 MPa

20
25
30
35
40
45
Sub-grade Modulus of Elasticity in soil, E3 (MPa)

Figure 4.15 The Best Pair of E2 and E3, for z = 740 mm

Figure 4.16 Displacement Basin Comparison, z = 740 mm

Figure 4.17 Principal Stress Contour at the bottom of the AC with E2 = 67 MPa, E3
= 10 MPa for North Yarmouth Control Section, z = 740 mm
Figures 4.1 8 and 4.19 present the results of the back-calculations of resilient moduli
that give the best comparison with the experimentally measured data in the least square
sense with z = 1940 rnrn. The best combination is E2 = 46 MPa and E3 = 28 MPa; and the
corresponding stress contour at the bottom of asphalt concrete is given in Figure 4.18.
The nmximum principal stress is 0.936 MPa, which is about 10.7% larger than the
0.836 MPa determined with z = 740 mm.

Figure 4.18 The Best Pair of E2 and E3, for z = 1940 mm

Figure 4.19 Displacement Basin Comparison, z = 1940mm

Figure 4.20 Principal Stress Contour at the bottom of the AC with E2 = 46 MPa, E3
= 28 MPa for North Yarmouth Control Section, z = 1940mm
4.4.4 Back-calculation for TWP31MD Pavement Control Section

As shown in Figure 4.12, the control section of TWP3 1-MD has a thicker asphalt
concrete layer. Loads, boundary conditions, and the meshes remain unchanged. The
same procedure will be repeated as done for the control section of North Yarrnouth.
Here, two models with different vertical dimensions z = 1040 mm and z = 2240 rnm
(total depth 1800 mm, 3000 rnm respectively) will be examined. The best pair of
Young's nlodulus, E2 and Ej ,and displacement basins and stress contours for the two
models are illustrated in Figures 4.21-4.23 and Figures 4.24-4.25.

I*

1
:
.......*.......

- ".....
0.8 -

.-

0.6

0.5

E2 = 10 MPa
E2=70MPa
E2=100Mf'a

0.9

0.7

k...
.....

"*-

............
.....'%.................

*-.- *.-- *
................

*.

............. .....................

...................

.*....................

Modulus of Elasticity in soil, E3 (NIPa)

Figure 4.21 The Best Pair of E2 and EJ,for z = 1040 mm

600

I /

&measured
data
%E2 = 68MPa, E3 = 30MPa

1000

1400

1800

Distance(mm)

Figure 4.22 Displacement Basin Ccr=pzrisc=,z = 1040 mm

Figure 4.23 Principal Stress Contour at the bottom of the AC with E2= 68 MPa, E3
= 30 MPa for TWP31MD Control Section, z = 1040 mm

Modulus of Elasticity in soil, Eg(MPa)

Figure 4.24 The Best Pair of Ez and E3, for z = 2240 mm

Figure 4.25 Principal Stress Contour at the Bottom of the AC with E2 = 49 MPa, E3
= 76 MPa for TWP31MD Control Section, z = 2240mm
Two models with different sub-grade depth, z, were taken into considerations for
TWP-3 1MD. The best sub-base and sub-grade moduli were back-calculated to give the
best comparison with the experimentally measured deflections. The largest maximum
tensile stress at the bottom of asphalt concrete is 0.3968 MPa the model with z = 1040
and 0.372 MPa for the model with z = 2240 mm. The error between them is about 7.5%
and illustrates that with all other conditions the same, the finite element model can give
good comparisons with experinlentid results. In particular, as long as a reasonable value
of z (sub-grade depth) is employed, soil properties can be back-calculated to fit the
displacements. Varying z does not significantly affect the stresses.

4.5 Summary
First, the effect of single dual-wheel load, single axle and dual wheel tandem axle
loads was discussed. The results presented in Section 4.1 illustrate that it is reasonable to
build a model around a single dual-wheel load. Secondly, the effective model size (2000
rnrn X 2000 mm X 1500 mm for a quarter symmetric model) was determined on the basis

of the maximum predicted tensile stresses.


Finally, for the two control sections, North Yannouth and TWP3 1-MD, models with
four different depths were selected (two for each). All materials were assumed to be
linearly elastic. The best values of E2 and E3 (the elastic moduli of the sub-base and subgrade respectively) were back-calculated to give the best comparison to the
experimentally measured deflections in a least-squares sense. Using different values of
the sub-grade depth, z, resulted in different combinations of E2, E3 to give a good fit to
the experimental data. However, the stresses computed using different optimal
combinations of E2 and E3, and z varied by only 10.7% for the model with a 125 mm
thick asphalt concrete layer, and 7.5% for the model with a 230 mm thick asphalt
concrete layer. These results indicate that fixing z at a value greater than or equal to 740
mm is reasonable when building a 3-layered flexible pavement system with fixed
boundaries, and that while back-calculated moduli vary with z, different optimal
combinations of E2, E3 and z give similar predicted ashpalt stresses.

Chapter 5. EFFECT OF NON-LINEAR SOIL RESPONSE


5.1 Introduction
In reality, soil (sub-base and sub-grade) is often made up of granular and gravel
material and is not linearly elastic. While assuming linearly elastic material properties
can be reasonable in many cases (see Chapter 4), we would like to capture materially
non-linear soil response. Evaluating the effect of base non-linearity is especially
important here, and relatively few 3D finite element models have considered it.
The goal of this chapter is to examine the feasibility of using the a K-8 model for
granular materials to model base layer response. In addition, the effect of de-bonding
between the asphalt and the sub-base will also be examined. Following a discussion of
possible implementation methods, the study will focus on two particular techniques.
After that, material parameters for the K-8 model will be back-calculated to give good
comparison with field-measured deflections for the North Yarmouth pavement section
considered in Chapter 4. The significance of base non-linearity on pavement response
will then be examined.

5.2 Background on K-8 Model


A nonlinear K-8 model can simply express the relationship between elastic
modulus and the first stress invariant as follows (Huang, 1993):

E=~ , 0 ~ '
where K, and K2 are experimentally derived constants and 8 is the first stress
invariant. Note that 8 is positive if compressive.

6 = o ,+02
+03
= o x+oy+or

(5.2)

Because the sub-base (the 2ndlayer) affects the results more than the sub-grade (the
3rdlayer) does, the sub-base will be treated by using a K-8 model and sub-grade will
be assumed to behave linearly for the models employed in this chapter.

5.3 Implementation of K-8 Model


Generally, when performing a materially non-linear finite element analyses, stresses
and constitutive properties are updated at the integration points within each individual
finite element (Zienkiewicz and Taylor, 1988). When implementing a K-8 model using
this approach, at each iteration we must re-calculate the element stiffness matrix, K' ,
governed by
K' = ~ B ' D B ~ V

(5.3)

In Eq. 5.3, B is a matrix of linear operators (derivatives of shape functions) and D is the
elasticity matrix defined as follows:

We see that D varies throughout the element since E varies; v is generally


considered a constant. In general, the elastic nlodulus will vary over the thickness and

the plan dimensions of the soil layer. This approach will be considered and will be
referred to the "integration point method", which can be regarded as the rigorous way
to implement the K-8 model.
There is a potential difficulty with the integration point method, however. In
particular, we may get negative or zero values for 8 at some locations in the soil layer,
leading to small or zero values for E at those integration points. This can cause
difficulties for iterative solvers like those used in this study, which are sensitive to poor
conditioning of the system stiffness matrix, K. To avoid this, a minimum value, Emin
needs to be set for E. This is an important parameter which will likely affect results.
Another possibility for implementing a K-8 model would be to sanlple the stresses at
the center of the soil layer, computing E based on the value of 8 at this point, and taking
this E as applying through the entire layer thickness. This approach will likely avoid
some of the problems with the integration point method. This second approach will also
be considered, and will be termed the "mid-layer method" for the remainder of this
chapter.
Both the integration point and mid-layer methods will be implemented using a direct
iteration approach, which can be expressed as follows
Until convergence
Update stresses
Re-computed E at each location
Form K
Solve KU = P
end
In this algorithm, K is the secant stiffness matrix based on the element moduli, U is the
total displacement vector, and P is the vector of nodal forces. Convergence will be

defined as the point at which stresses and displacements do not change significantly with
successive iterations.
It should be noted that the K-8 model has been implemented in 2D layered analysis
and 2D FE analysis as well (Huang, 1993). Two methods have been developed to
incorporate the material non-linearity. In first one, the non-linear granular layer is
subdivided into a arbitrary number of layers and the stresses at the mid-depth of each
layer are used to determine the modulus; in the second method, the granular materials are
considered as a single layer and the modulus will be computed at a single point of the
layer. As it is well known that most granular materials do not take tension, the horizontal
stress will be set to zero if it is negative or in tension. Although in method 2, the point
selected is usually in upper half of the layer and tensile stresses occur very rarely, an
arbitrary or minimum modulus Eminmay still be needed to ensure convergence.

5.4 Convergence of K-8 Model


In this section, both the integration point method and mid-layer method will be
considered with respect to convergence. The model selected will be the same as used in
Chapter 4 with a sub-grade depth, z, of 740 mm (see Figures 4.8 and 4.10 for elevation
and plan views) under a single 40 kN wheel load. The material properties of AC and subgrade will be the same as shown in Figure 4.8, but the sub-base will be treated using K-8
model with KI = 30.9 MPa and Kz = 0.53 (Nickels, 1995). The value of Eminand number
of material iterations have to be determined to perform the finite element analysis. As
presented in previous section, the two methods are very different.
For the integration method, modulus of elasticity is updated in each individual finite
element. The E at the top of the sub-base is very different from that at the bottom of sub-

base. For the mid-layer method, n~odulusof elasticity is picked at the mid-height of subbase and used through the entire thickness of sub-base. Therefore, elasticity matrix D
formed from the two methods will be significantly different. As a result, nodal
displacements and stresses will be different as well. In addition to examining the two
different implementations, the influence of de-bonding will also be considered. One
model assumes that the asphalt and sub-base are fully bonded. The second model does
not permit tension at the asphalt sub-base interface, and allows free horizontal sliding
between the asphalt and sub-base.
Maximum deflections and tensile stresses for the fully bonded and un-bonded
condition are shown in Figures 5.1 through Figure 5.6. Table 5.1 summarizes the
maximum tensile stress at the bottom of the asphalt and the largest vertical displacement.

Number of iterations

Figure 5.1 Maximum Vertical Deflections vs. Number of Material Iterations, Fully
Bonded, Integration Point Method

8
10
12
14
Number of Material Iterations

16

18

Figure 5.2 Max. Principal Tensile Stresses in Asphalt vs. Number of Material
iterations, Fully Bonded Integration Point Method

0.825

8
10
12
14
Number of material iterations

16

18

Figure 5.3 Maximum Vertical Deflections vs. Number of Material iterations, Fully
Bonded, Mid-layer Method

8
10
12
14
Number of material iterations

16

18

20

Figure 5.4 Max. Tensile Stresses vs. Number of Material iterations, Fully Bonded,
Mid-layer Method

Number of material iterations

Figure 5.5 Maximum Vertical Deflections vs. Number of Material iterations, Unbonded at Asphalt / Sub-base Interface, Integration Point Method

1I
2

8
10
12
14
Number of material iterations

16

18

20

Figure 5.6 Max. Principal Tensile Stresses vs. Number of Material iterations, Unbonded, Integration Point Method
0.9424
0.9423
Emin = 1OMPa
0.9422
0.9421
h

E
E

-0.942
Q
V)

0.9419
0.9418
0.9417
0.9416
0.9415
Number of material iterations

Figure 5.7 Maximum Vertical Deflections vs. Number of Material iterations, UnBonded, Mid-layer Method

Number of material iterations

Figure 5.8 Max. Tensile Stresses vs. Number of Material iterations, Un-Bonded,
Mid-layer Method
Table 5.1 Summary of Max. Tensile Stresses and Displacements for two Methods

Emin=
2MPa
Emin=
5MPa
Emin =
1OMPa

Integration Point Method


(10 iterations)
Fully
Unbonded
Bonded
Stress (MPa)
1.714
1.7913
Displ. (mm)
3.685 1
2.152
1.6731
1.55
Stress (MPa)
2.076
3.3618
Displ. (nun)
1.4746
1.372
Stress (MPa)
1.945
2.9898
Displ. (rnrn)
--

Mid-layer Method
(5 iterations)
Fully
Unbonded
Bonded
0.8221
0.9663
0.9423 0.8588
0.8221
0.9663
0.8588
0.9423-0.8221
0.9663
0.8588
0.9423

Results shown in Figure 5.2 illustrate that the maximum tensile stresses in the asphalt
become nearly constant after about 10 material iterations for the integration point method

irrespective of the value of Emin.Figure 5.1 shows that while displacements are still
increasing with further iterations, the rate of increase has slowed significantly. The
magnitude of stresses and displacements are, as expected, dependent on the value of Emin,
For example, decreasing Eminfrom 10 MPa to 5MPa increases the maximum tensile
stress from 1.37 MPa to 1.55 MPa, a change of 13%. However, deflections increase by
only 6.7% as Eminis decreased from 10 MPa to 5MPa. Also, it appears that convergence
properties are similar when interlayer slip and debonding is considered. Interestingly,
debonding does not affect stresses, as much as it affects displacements.
Based on the results shown in Figure 5.5 and 5.6, integration point implementation
of the k-8 model for the remainder of this Chapter will rely on a value of Emin= 5 MPa
and use 10 material iterations. All models will assume full bond between the asphalt and
the base.
As expected, there are large differences between the mid-layer method and the
integration point method. First, the mid-layer method converges much more quickly than
the integration point method, reaching nearly constant stresses and strains after only 5
material iterations. Further, the results are insensitive to the assumed value of Emin,
indicating that the value of 8 at the middle of the sub-base does not become negative.
Perhaps most striking, however, is the large difference in predicted stresses and
displacements, with the mid-layer method giving predictions that are approximately half
of the values predicted when using the integration point method. Based on these results,
the mid-layer implementation of K-8 model used for the remainder of the studies of this
chapter will be based on results after 10 material iterations and Eminwill be fixed at 5
MPa as for the integration point method.

The largest difference in results for the integration point method and the mid-layer
implementation does raise an interesting question: can different Kl and K2be backcalculated using each method to give a good fit to field-measured deflections? Further,
what are the asphalt stresses predicted using each these back-calculated values of Kl and
K2? These issues will be explored in the remainder of this chapter.

5.5 Comparison with Measured Displacement Basins


5.5.1 Model Description
It is important that any model be verified by comparing with experimentally
measured data. This was done in Chapter 4 using linearly elastic materials for the
subbase and subgrade; here, we will perform the same back-calculation using a K-8
model for the sub-base. The best combination Kl and K2will be determined through
comparing with the experimental data by using both mid-layer method and the integration
point method. The model used here is exactly the same as the one for the convergence
study. Material properties and vertical model dimensions are shown in Table 5.2.
Note the dead load (self weight) of each layer is applied to the pavements all the time.
The displacement due to the dead load accounts for 30-50% of the total displacement
(Dead Load + Live Load). However, an accurate comparison with experimentally
measured data requires the case in absence of the dead load. Because of material nonlinearity and the effect of dead load on response, when comparing with the
experimentally measured deflections, two analyses have to be done, one for the dead load
plus live load; the other for dead load only. The reported displacements and stresses are
based on the results of dead load plus live load minus the stresses and deflections due to
the dead load alone.

Table 5.2 Material Properties and Vertical Model Dimensions

Material

Layer #

E(MPa)
ness(mm)

Iteration

1
2

Asphalt
Sub-base

1 Sub-grade

1400
K-8

10

740

5.5.2 Comparison using Published Parameters

A previous study conducted at U. Maine (Nickels, 1995) back-calculated the nonlinear parameters Kl and Kr for the North Yarmouth control section considered here by
using MICH-PAVE, a 2D computer finite element program (discussed in Chapter 2).
Load parameters and material properties used in this prior study are presented in Tables
5.3 and 5.4 respectively. Nickels (1995) showed that Kl

= 30.9 MPa and Kr = 0.53

gave

a good fit to field-measured data, and these values will be taken as a starting point here.
Comparison of measured and predicted displacements using integration point method and
the mid-layer method are shown in Figures 5.9 and 5.10.

Table 5.3 Load Parameters used in MICHPAVE

Property
Static Load (kips)
Tire Pressure (psi)
Radius of Load area (in.)

Value
5.35

Table 5.4 Material Properties used in MICHPAVE

&

Properties
Behavior

Asphalt Concrete
Linear Elastic

Modulus (psi)
Poisson's Ratio
Unit Weight (pcf)

E = 200000
0.35
150

+
- -- -- - - - -

Material
Subbase
Nonlinear
Cohesionless
KI=4480,
0.29
130

Subgrade
Nonlinear
Cohesionless
K1=3000, ~ ~ = 0 . 5 7 ~
0.29
125

-.-,-.

measured data
K1 = 30.9MPa, K2 = 0.53
/'

Figure 5.9 Displacment Basin Comparison, Integration Point Method

For the integration point method, Kl and K2 obtained from the 2D approach do not
give displacements that fit the experimental data very well. That is not surprising, since
the 2D MICH-PAVE model bases updates to E for an entire layer on the average value
computed across elements (Harichandran et al., 1993). In contrast, the integration-point
method approach, which uses values for E at each integration point, captures the variation
in E over layer thickness and horizontal dimensions affected by the load. For the midlayer method, results yielded by using published data give a good comparison with the
experimentally measured data as shown in Figure 5.10.
0.9

0.8 -

measured data
K1 = 30.9Wa, K2 = 0.53

Figure 5.10 Displacment Basin Comparison, Mid-layer Method

Figure 5.11 The Best Pair of KI and K2in the Least Square Sense, Integration Point
Method

A back-calculation will now be performed which gives (Kl,,Kz) providing the best
comparison with measured data in the least square sense. Figures 5.1 1 and 5.12 clearly
show that, when Kl

10 MPa and Kz = 0.78, the back-calculated deflection compares

well with the measured data best for the integration-point method. Figure 5.13 shows
the maximum principle stresses at the bottom of the asphalt concrete.

'

0.9

+
0.8 -

measud data
K1= I-, K2 = 0.8

0.7 -

Figure 5.12 Displacment Basin Comparison, Kl = lOMPa and K2 = 0.78, Integration


Point Method

Figure 5.13 Stresses Contour at the Bottom of the AC with KI = lOMPa, K2 = 0.78
for North Yarmouth Control Section, Integration Point Method

It is also necessary to determine what combination of KI and K2 give the best


comparison with the experimental data using mid-layer method. Analysis showed that
accuracy was not significantly improved if values other than Kl

= 30.9

MPa and K2 =

0.53 were used. Figure 5.13 shows the stresses at the bottom of the asphalt larger
predicted using the mid-layer method and Kl

= 30.9

MPa and K2 = 0.53.

Figure 5.14 Maximum Principle Stresses at the Bottom of the AC with Kl = 30MPa,
K2 = 0.53 for North Yarmouth Control Section, Mid-layer Method
It was found that the best combination of Kl and K2 from the mid-layer method was
very similar to the published data (Nickels, 1995). However, the maximum principal
stresses at the bottom of asphalt is about 30% less than results obtained from the
integration point method. The direction of maximum principal stresses are mainly in y
direction. These results indicate that there are still open questions regarding the use of K8 models in 3D finite element analysis.

5.6 Summary and Conclusions


In this Chapter, the stress-dependent K-0 model, which is widely used to model the
non-linear stress-strain response of granular soils, was implemented in a 3D finite
element analysis program for analyzing flexible pavements.
Two alternative methods for implementing the K-0 model within the finite element
solver were examined: the integration-point method, where the material properties are
updated at each integration point within each finite element, and the mid-layer method,
where the material properties are updated only at the mid-depth of the granular layer.
The convergence properties of each method were studied, and appropriate parameters for
their implementation were deternlined.
To study the effect of each implementation on model response, optimal values of the
material parameters K , and Kz were determined for both implementations via comparison
with experimental data. The results of the analyses were striking: while good fits to fieldmeasured displacements were achieved for both implementations, the predicted
maximum principle stresses in the asphalt layer are quite different when two different
implementations are used.
Based on these results, there are clearly still issues that must be resolved when
implementing a K-0 model in a 3D finite element analysis program. In particular, the
locations at which stress updates are perfonned must be resolved. Further, if stress
updates are performed at integration points, the possible mesh-dependency of the results
should be considered.

CHAPTER 6. EFFECT OF SPATIALLY VARYING TIRE


PRESSURE
6.1 Introduction
The most common tire contact areas selected for the analysis of flexible pavements
are circular or rectangular shown in Figure 6.1. A constant pressure is usually assumed
over the entire contact area, often the tire inflation pressure (Huang, 1993).

Figure 6.1 Types of Tire Contact Area

Real tire pavement contact stresses are not constant and experimental data illustrates
that the contact area between the tire and pavement will vary under different wheel loads
and inflation pressures (de Beer et al., 1997; Tielking et al, 1995). Analytical studies have
shown that spatially varying contact pressures can affect the response of a flexible
pavement system significantly (Bensalem et al., 2000). This chapter will examine the
effect of spatially varying tire contact pressure on asphalt concrete displacement and
stresses. For simplicity, we will consider only static loading, although in general the
truck suspension system and vehicle speed do affect stresses in real pavements
(Lee, et al 1993; Cunagin et al, 1993).

First, this chapter will cover a strategy for modeling spatially varying pressures.
Contact patterns will be based on laboratory-measured values taken from de Beer et a1
(1997). A convergence study will be performed to determine the required level of
discretization of the contact area. Next, parametric studies will be performed using the
3D finite element modeling strategies developed in this study to examine the effect of
spatially varying tire contact pressures on pavement response. A variety of loading
conditions and model material and geometric properties will be considered. Finally, a
brief summary finishes this chapter.

6.2 Modeling the Spatially Varying Tire Pressures


6.2.1 Background

When determining flexible pavement response, it is inlportant to know the contact


area between tire and pavement. According to de Beer et a1 (1997), tire contact areas are
close to rectangular. de Beer et a1 (1997) measured vertical contact stress, transverse
(lateral) contact stress and longitudinal contact stress under slowly moving tires on
asphalt pavements. For the smooth tire, Figures 6.2 and 6.3 show typical measured
vertical contact stresses for inflation pressures varying from 420 kPa to 720 kPa and the
load varying from 20 kN to 50 kN. Generally, de Beer et a1 (1997) observed that the ratio
of maximum stresses in the vertical, transverse, and longitudinal directions is 10:3.6: 1.4.
Figures 6.2 and 6.3 also illustrate that tire inflation pressure controlled the vertical
contact stresses at the inner portion of the tire width. Based on the experimentally
measured results, de Beer et a1 (1997) proposed the following equation for the vertical
contact stress at the tire center:

q,, = 0.86P

+ 175

(6.1)

where 460 is the average tire pavement contact stress at the tire center [@a], and P is tire
inflation pressure in kPa (de Beer et al, 1997). The maximum vertical stress at the tire
edge is controlled by loads, and the relationship is given in Eq. 6.2:

q, = - 0 . 5 3 ~ ~57.46L - 534.05

(6.2)

where q, is the vertical tire pavement contact stress at the tire edge, and L is tire load in

kN, ranging between 20 and 50 kN.

Figure 6.2 Load, Inflation Pressure Relationship on the Vertical Contact Stress
Distribution (de Beer et al, 1997)

Z 1eJns!d
VldStlA SSOWV MQWnN Nld
OZ6l~L.LlQlSLPLELZLLIOL6 0 L 9 S b E 2 L

. . . . . . .

...
11 @JnBkJ
V U S W SSOWV MQlrYflNNld

6.2.2 Representation of Spatially Varying Tire Pressure


To efficiently describe the contact stresses acting over the entire contact area, we
need to choose a mathematical description of the contact stress profile. As noted before,
the effect of longitudinal contact stress is relatively small, and it will be accurate enough
to model the tire pavement contact pressures in absence of longitudinal stress. Figures
6.4 and 6.5 show the vertical and transverse pressure distributions assumed in this study.
The vertical contact pressures were assumed to increase linearly from 0 to a maximum
stress, qe,over a distance a or d from each tire edge as shown in Figure 6.4.
Transversely, over the distance w-2d, the pressure q is assumed to vary parabolically.
Longitudinally, over the distance 1-2a, it is constant. This distribution was chosen for
ease of mathematical description and the reasonable good fits it provides to the
experimentally measured data shown in Figures 6.2 and 6.3. The transverse contact
stress is assumed to vary sinusoidally across the tire width and vary parabolically along
the entire length.

v
W

Plan View

1-2a
Section A-A

w-2d

n
Section B-B

Figure 6.4 Configurations of Vertical Stress in 2D

Transverse Direction

Figure 6.5 Configuration of Transverse Stress in 2D

Figure 6.6 Integration Domain

The value of q60, the stress at tire center obtained from Eq. 6.2 is an average
pressure over the inner 60 per cent of the tire width (see Figure 6.6). Hence, q,, the
vertical contact pressure at the tire center needs to be determined so that 460 is truly the
average pressure on interior.
To ensure the average holds, the following integral needs to be satisfied:

where,

q(x') = 4x1' (4, - 4,)


(W- 2d)2

qc

and then, for any given q 6 qe,


~ we can solve for q, to satisfy Eq. 6.3:

To ensure that the average pressure, 960, exists over the entire 60% of the contact area,
the interior area (Ainkrior)is defined as follows:

We can solve for a:

This gives the value of a for any w, I, and d to ensure that AinIerior= 0 . 6 ~ 1
Figures 6.7 and 6.8 show typical vertical and transverse stress distribution for L = 40 kN
(total vertical load), P = 620 kPa (inflation pressure).

1800 '1800

Transverse Direction

Traffic Direction

Figure 6.7 Typical Vertical Contact Pressure Distribution (tire load, L = 40W,
inflation pressure, P = 620KPa)

..
. . . . . ..

'

2200

Transverse Direction

Traffic Direction

Figure 6.8 Typical Transverse Contact Pressure Distribution (tire load, L = 40 kN,
inflation pressure, P = 620 kPa)

6.2.3 Adjustment of Total Vertical Load

Generally, total vertical load yielded fiom integration of the vertical pressures defined
in the previous section is lower than the actual value of L used to determine q, and q,. As
a result, the length of the contact area, I, must be adjusted so that the total applied load
due to the contact pressures is L. Mathematically, this can be expressed as:

where p is the vertical contact pressure which varies over the contact area, A. To
numerically integrate pressures over the contact area, the contact area is divided into a
grid as shown in Figure 6.9. the dimensions of refined regions have been chosen as 2d
and 2a in x, y directions, respectively. For simplicity, the number of divisions in refined
regions as well as in coarse regions is identical and will be called n for the purpose of
discussion.

Fine (2a)

Coarse

w4Number of Divisions

Fine (2a)&
=n

Figure 6.9 Refinement of Contact Area

To do the integration over the entire contact area, we have chosen to generate a
Delaunay Triangulation of the grid shown in Figure 6.9, which is easily implemented.
The mesh of triangles will be generated using the computer program, "triangle"
(Shewchuk, 1999) and a typical mesh is shown below in Figure 6.10:

I
0

50

100

150

200

250

300

Figure 6.10 a Typical Triangulation


The integration can be obtained from Eq. 6.7:

where p is the pressure acting on the contact area, n is the number of triangles, pi is the
average pressure acting on an individual triangle element and Ai is the area of an
individual triangle. Using Delaunay triangulation and Eq. 6.7 to compute the total force
due to the contact pressures, Eq. 6.6 can be solved for the particular length of the contact
area, I, that yields vertical equilibrium. This was accomplished using the bisection
method, and a general computer program was written to generate vertical pressures and
contact areas for a variety of wheel loads and tire inflation pressures.
The error in the numerical solution to Eq. 6.6 is defined as follows:
error =

cP A
L

5 tolerance

where tolerance is a small number (0.001,0.0001, etc).

6.2.4 Different Combinations


Based on the approach developed here, pressure distributions due to different
combinations of loads and inflation pressures are shown in Figures 6.1 1-6.14.
Comparison of the predicted distribution with the measured data of Figures 6.2
and 6.3 is generally good. Lower inflation pressures tend to accentuate the differential
between the peak and average stresses and lower loads give higher predicted center
stresses than edge stresses. The length of contact area decreases when the inflation
increases with the same load and increases when load increases with the same inflation
pressure. Vertical contact stresses yielded from model are also given in Table 6.1.

Transverse Direction

1800 - -1800

Traftic Direction

Figure 6.11 Vertical Stress Distribution (L = 20 kN,P = 420 kPa)

Transverse Direction

1800 '1900

Traftic Direction

Figure 6.12 Vertical Stress Distribution (L = 20 kN, P = 720 kPa)

Transverse Direction

1800 - '1600

Traftic Direction

Figure 6.13 Vertical Stress Distribution (L = 50 kN, P = 420 kPa)

Transverse Direction

1800 '1800

Traffic Direction

Figure 6.14 Vertical Stress Distribution (L = 50 kN,P = 720 kPa)


Table 6.1 Vertical Stresses at edge and center of tire

6.3 Convergence Study


Note that at higher loads, the vertical stresses tend to be largest near the edges and
decrease rapidly towards the center of tire. To capture the steep gradient, pressures are
sampled more frequently over regions near the tire edge as shown in Figure 6.9.
An issue is how many grid points are required to ensure that stresses and
displacements are computed with a sufficient degree of accuracy by the finite element
model, which employ a triangulation scheme to compute equivalent nodal loads that is
similar to that used to solve for the contact length, 1. To determine the required value of

n (see Figure 6.9), a finite element model was constructed with the following properties
as shown in Figures 6.15 and Table 6.2:

Plan View

Sub-base

1740nm

Elevation View
Figure 6.15 Model Description

Table 6.2 Material Properties

Layer
Asphalt Concrete
Base
Sub-grade

Thickness(mm)
125
635
740

E (MPa), v
1400,0.35
67,0.29
10,0.29

The dual wheel load was modeled with spatially varying pressures acting at the
center of the pavement with L = 40kN and P = 620KPa. The finite element mesh used is
shown in Figure 6.16. The results are shown in Figure 6.17:

Figure 6.16 3D View of Meshed Model (entire pavement system, 3024 solid
elements)
h

1.16

1.15 -

2 1.15 5
A

I
I

10
15
Number of Divisions, n

20

1.495 U)

%
E
U)

.-P )

1.49

U)

8'
x

10
15
Number of Divisions, n

20

Figure 6.17 Convergence Study on Max. Stresses and Deflections

It is clear that when number of divisions equals 15, the stresses and displacements
tend to converge. This study will use n = 10, however, since the difference between
predicted stresses and displacements is negligible for larger values of n. For example,
increasing n fiom 10 to 15 increases stresses by only 0.208% and displacements by only
0.198%.

6.4 Parametric Study


It is important to know how spatially varying tire contact pressures affect the
response of the flexible pavement systems. The criteria employed here to quantify these
effects are the maximum principal stresses at the top and bottom of the asphalt concrete.
The maximum principal stresses under the spatially varying tirelpavement contact
pressures will be compared to two different loading cases, including stresses under a
uniform rectangular patch load with the same contact area as spatially varying pressures
having the same value of L; and stresses under a uniform rectangular patch load with the
sanle inflation pressure as the spatially varying pressures and the same vertical load, but a
different contact area. To perform the parametric study, several parameters will be varied:
asphalt thickness, sub-base modulus, tire inflation pressure, and wheel load, L. This will
allow the effect of non-uniform tire contact pressure to be studied under a variety of
conditions.
6.4.1 Model Description

The thicknesses of asphalt concrete that will be examined is 75 mm, 125 mm and 200
mm and Young's modulus of the sub-base is 67MPa (determined via back-calculation, in
Chapter 4), 37 Mpa, or 135 MPa. The sub-grade's modulus of elasticity will be fixed at

10MPa. In each combination, we will compare three different loading cases: I means
spatially varying tirelpavement contact pressures as detailed previously; I1 means the
uniform rectangular patch load with the same contact area as spatially varying pressures \
for the same value of L; and III means the uniform rectangular patch load with a vertical
contact stress equal to the inflation pressure, and a different contact area

Tire loads and

inflation pressures have been considered varying from 20 kN to 50 kN and 420 Wa to


720 H a , respectively. The tire width is fixed at 220 mm. Model dimensions are shown
in Figures 6.18 and 6. 19. The finite element mesh used is shown in Figure 6.16.

-Wheel

Figure 6.18 Model Description, Plan View

Loads

7
Wheel loads
Asphalt Concrete

+ X

f~,=75,125,2OOnun
635mm, Ez

Sub-base
Sub-grade

P0nun

Figure 6.19 Model Description, Elevation View

6.5 Results and Discussion


6.5.1 Stresses at Bottom of Asphalt

The stresses are the maximum principal stresses taken at the bottom of the asphalt.
The location where the maximum principal stresses occur varies, but remains within the
two longitudinal center lines of the two contact areas. The directions of those maximum
stresses are mainly in they (transverse direction) with small x (traffic direction)
components.
According to Figures 6.20-6.28, the spatially varying loads tend to generate larger
tensile stresses at the bottom of asphalt concrete than the uniform pressure loads, with a
maximum difference of about 30%, which implies that assuming a constant pressure over
the entire tire contact area tends to underestimate the response. Generally, the uniform
loads with the same contact area as the spatially varying loads (case 11) yield smaller
results than the loads with the same inflation pressure (case 11).
Under the same loads, the thinner the asphalt concrete, the more pronounced effect of
the spatially varying tirelpavement contact pressures. If we consider the same thickness

of asphalt concrete, subjected to the same vertical load, lower inflation pressures tend to
increase differences in tensile stresses between I and (I1 and 111). Similarly, decreasing
the stiffness of the sub-base (decreasing EZ)tends to somewhat increase the differential
in tensile stresses between case I and (I1 and 111).

Inflation pressure, P = 420KPa

Inflation pressure, P = 520KPa

Vertical Load (kN


Inflation pressure. P = 6 4 O K ~ a

Vertical Load (kN


lnflation pressure. P = 740KPa

Vertical Load (kN)

Vertical Load (kN)

Figure. 6.20 Max. Tensile Stresses at bottom of AC (thickness = 75mm, E2 = 35MPa)

lnflation pressure, P = 420KPa

1
20

r"

30
40
Vertical Load (kN
lnflation pressure. P = 7 4 0 K ~ a

Vertical Load (kN


Inflation pressure, P = 6 OKPa

';;;'

lnflation pressure, P = 520KPa

rn

Vertical Load (kN)

Vertical Load (kN)

Figure. 6.21 Max. Tensile Stresses at bottom of AC (thickness = 75mm, E2 = 67MPa)


Inflation pressure, P = 420KPa

1 1.2
2

tj

lnflation pressure, P = 520KPa


1.6
1.4.

Q,

1 1.2 .

f3 111

ti
.-Q,
g

1.

'r"
x

20

30
40
Vertical Load (kN)

50

0.6

20

30
40
50
Vertical Load (kN
lnflation pressure. P = 7 4 0 ~ ~ a

1.6

Q,
.-g

0.8 .

'r"

--__---__

0.8 :---------I::

f3 111 .

P!

0.6 .
x
20

+.+=J

30
40
Vertical Load (kN)

I
11
111

4'

.
50

Figure 6.22 Max. Tensile Stresses at bottom of AC (thickness = 75mm, E2 =


135MPa)

lnflation pressure, P = 420KPa

lnflation pressure. P = 520KPa

Vertical Load (kN


Inflation pressure, P = 6 OKPa

-$

1
20

30
40
Vertical Load (kN)

Vertical Load (kN


Inflation pressure, P = 7 OKPa

50

r"

1
20

30
40
Vertical Load (kN)

50

Figure 6.23 Max. Tensile Stresses at bottom of AC(thickness = 125mm, E2= 35MPa)
.3

lnflation pressure, P = 520KPa

lnflation pressure, P = 420KPa

Vertical Load (kN


Inflation ~ressure.P = 6 OKPa

0.8
20

30
40
50
Vertical Load (kN
Inflation pressure, P = 7 OKPa

1
30
40
50
I 20

0.8
20

Vertical Load (kN)

30
40
Vertical Load (kN)

50

Figure 6.24 Max. Tensile Stresses at bottom of AC(thickness = 125mm, E2= 67MPa)

lnflation pressure, P = 420KPa

lnflation Dressure. P = 520KPa

1.2

3
V)

%
V)

I-

al
*
L

= 111

Vertical Load (kN


Inflation pressure. P = 6 3 O ~ ~ a

20

30
40
Vertical Load (kN)

50

Vertical Load (kN


Inflation pressure. P = 730KPa

20

30
40
Vertical Load (kN)

50

Figure 6.25 Max. Tensile Stresses at bottom of AC(thickness = 125mm, E2=


135~~a)
A

Inflation pressure, P = 420KPa

Vertical Load (kN


lnflation pressure, P = 6 OKPa

lnflation pressure, P = 520KPa

Vertical Load (kN


Inflation pressure, P = 7 3 0 K ~ a

Vertical Load (kN)

Vertical Load (kN)

Figure 6.26 Max. Tensile Stresses at bottom of AC(thickness = 200mm, E2= 35MPa)

lnflation pressure. P = 420KPa

lnflation pressure, P = 520KPa

Vertical Load (kN


lnflation pressure, P = 6 OKPa

Vertical Load (kN


lnflation pressure, P = 7 OKPa

3 1.4
n

E
u,

2u,

1.2

1
fj

al
.-

0.8

:
8

Z
Vertical Load (kN)

0.6
20

30
40
Vertical Load (kN)

50

Figure 6.27 Max. Tensile Stresses at bottom of AC(thickness = 200mm, E2= 67MPa)
lnflation pressure, P = 420KPa

lnflation pressure, P = 520KPa

Vertical Load (kN


Inflation pressure, P = 6 OKPa

Vertical Load (kN


lnflation pressure. P = 73OKPa

Vertical Load (kN)

Vertical Load (kN)

Figure 6.28 Max. Tensile Stresses at bottom of AC (thickness = 200mm, E2=


135MPa)

6.5.2 Stresses at Top of Asphalt


Engineers have observed that longitudinal and fatigue cracks in thicker asphalt
concrete (AC) pavements appear to crack fiom the top of the wearing course downward
for years (Uhlmeyer et al., 2000), and the top-down cracks are typically longitudinal
appearing in or near the wheelpaths. Stresses at the top of asphalt can affect the topdown cracking significantly and have not been examined in detail. In this section,
parametric study will be performed based on maximum principal stresses at the top of
asphalt and some conclusions will be drawn from results later
Figures 6.29-6.37 show the maximum principal stresses at the top of the AC layer.
The location where the maximum principal stresses occur varies, but remains within the
(longitudinal center lines of the tire contact areas). The directions of those maximum
stresses are mainly in y (transverse direction) with small x (traffic direction) components.
lnflation pressure, P = 520KPa

lnflation pressure, P = 420KPa

0.2 1
20

20

Vertical Load (kN


Inflation pressure. P = ~ O K P ~

20

30
40
Vertical Load (kN)

50

30
40
50
Vertical Load (kN
lnflation pressure. P = 7 1 0 ~ ~ a

30
40
Vertical Load (kN)

50

Figure 6.29 Max. Tensile Stresses at top of AC (thickness = 75mm, Ez= 35MPa)
109

Inflation pressure, P = 420KPa

Inflation pressure, P = 520KPa

0.3 .

Vertical Load (kN


Inflation pressure, P = 6 OKPa

Vertical Load (kN


Inflation pressure. P = 7 3 0 ~ P a

Vertical Load (kN)

Vertical Load (kN)

Figure 6.30 Max. Tensile Stresses at top of AC (thickness = 75mm, E2= 67MPa)
a

Inflation pressure. P = 420KPa

Inflation pressure, P = 520KPa

Vertical Load (kN


lnflation pressure, P = 6 3 0 ~ ~ a

Vertical Load (kN


Inflation pressure, P = 7 OKPa

Vertical Load (kN)

Vertical Load (kN)

Figure 6.31 Max. Tensile Stresses at top of AC (thickness = 75mm, E2= 135MPa)

w
6'

Max. Tensile Stresses (MPa)

Max. Tensile Stresses (MPa)

Max. Tensile Stresses (MPa)

Max. Tensile Stresses (MPa)

Max. Tensile Stresses (MPa)

Max. Tensile Stresses (MPa)

Max. Tensile Stresses (MPa)

Max. Tensile Stresses (MPa)

I
.

Max. Tensile Stresses (MPa)

Max. Tensile Stresses (MPa)

Max. Tensile Stresses (MPa)

Max. Tensile Stresses (MPa)

Max. Tensile Stresses (MPa)

Max. Tensile Stresses (MPa)

Max. Tensile Stresses (MPa)

Max. Tensile Stresses (MPa)

Inflation pressure, P = 420KPa

Inflation pressure, P = 520KPa

Vertical Load (kN


Inflation pressure, P = 6 OKPa

0.05
20

30
40
Vertical Load (kN)

Vertical Load (kN


Inflation pressure, P = 7 OKPa

50
Vertical Load (kN)

Figure 6.36 Max. Tensile Stresses at top of AC (thickness = 200mm, E2= 67MPa)

lnflation pressure, P = 420KPa

lnflation pressure, P = 520KPa

Vertical Load (kN


Inflation pressure, P = 6 OKPa

20

30
40
Vertical Load (kN)

Vertical Load (kN


Inflation pressure, P = 7 OKPa

50
Vertical Load (kN)

Figure 6.37 Max. Tensile Stresses at top of AC (thickness = 200mm, E2= 135MPa)

As shown in Figures 6.29-6.31, for the smallest pavement thickness, the principal
stresses are largest. Increasing E2 tends to decrease the stresses, and the differences
between the non-uniform and uniform pressures are relatively small. The largest
differential occurs for E2 = 135 MPa. Increasing inflation pressures for ZI = 75 mnl does
tend to increase the stresses slightly.
For ZI = 125 mm, the differences between uniform and non-uniform contact stresses
increases significantly with increasing E2, However, the value of E2 does not seem to
have a very large effect on the maximum principal stress predicted for lonading I, and
inflation pressure has a small effect.
For the thickest pavement section, the differential between I and (I1 and 111) grows
with increasing vertical load for all values of E2. Increasing inflation pressure tends to
slightly decrease the differential between I and (I1 and 111) in general. Again, increasing

E2 does not significantly affect the maximum principal tensile stresses at the top of the
asphalt.
Overall, with the exception of the thinnest pavement section with the softest sub-bse,
layers (ZI = 75 rnrn and E2 = 35 MPa), the maximum predicted tensile stresses in the top
of the asphalt tend to remain fairly constant with increasing pavement thickness. This is
important -it implies that irrespective of pavement thickness, the contribution of tensile
stresses to top-down cracking remains fairly constant.

6.6 Summary
In this chapter, a model was developed to simulate the spatially varying tire contact
pressures based on the study by de Beer et al, (1 997). The kernel of the model is

developing a mathematical expression for tire contact pressures as a function of tire load
and inflation pressure.
Parametric studies were perfomled to examine effecs of tire load, inflation pressure,
asphalt thickness and sub-base modulus on pavement response.
It was found that the regular rectangular patch loads tend to underestimate tensile
stresses in the bottom of the asphalt by 10%-30%, especially in thin flexible pavement
structures with low elastic n~odulusof sub-base material; the thicker the asphalt concrete
layer, the less sensitive the pavement in terms of maximum principal stresses at the
bottom of asphalt. Tehnsile stresses at the top of he asphalt are largest for the 75 mnlthick asphalt sections, but remained fairly constant between for a given wheel load
irrespective of inflation pressures and sub-grade stiffness for the thicker pavement
sections. The tensile stresses in the top of the asphalt predicted using spatially varying
tire contact pressures are significantly larger than the stresses predicted assuming a
uniform load.

CHAPTER 7. SUMMARY, CONCLUSIONS AND


RECOMMENDATIONS
7.1 Summary
In the United States, millions of miles of roadway surfaced with asphalt have been
constructing over the past hundred years. Yet the design procedure of flexible pavements
is still largely based on mechanistic-empirical method. This is due part to the limitations
of current analytical tools.
The objective of this thesis was to investigate the response of flexible pavement
structures with a variety of materials, model dimensions and different tire and axle
loadiGs using 3D finite elenlent (FE) analysis. The following tasks were accomplished:
1. A literature review was conducted on the FE modeling of flexible pavements.
2. An effective meshing tool for 3D FE analysis of flexible pavements was

developed. With this tool, we can efficiently build a relatively large 3D FE model of
a pavement system that incorporates multiple layers, inter-layer debonding and slip,
and a variety of load types.

3. The 3D FE models were verified via comparison with field-measured data. To


give the best fit to the experimentally measured data (Nickels, 1995), the best pair of
moduli of elasticity of the sub-base and the sub-grade, respectively were determined
via back-calculation assuming all materials were linearly elastic. It was shown that
sub-grade thicknesses larger than 740 rnrn and model plan dimensions of 4000 rnm X
4000 nun are sufficient to give consistent results.

3. The effect of sub-base material non-linearity on pavement response was studied.


The stress-dependent k-8 model was incorporated in the 3D FE models, which is

widely used to analyze non-linear response of granular materials. Two options for
incorporating the K-0 model were considered in detail: a rigorous integration-point
method and a technique where material properties are updated based on mid-layer
stresses. The convergence properties of these two approaches were studied in detail.
The best combination of non-linear parameters Kl, K2,was back-calculated to give
good comparison to the experimental data.
4. The effect of spatially varying tirelpavement contact pressures on pavement
response was studied. Previous investigations (Yue and Seve, 1995; Bensalem et a1
2000, etc) illustrate that assuming a constant stress distribution over the entire
tirelpavement contact area may not be accurate, and that spatially varying pressures
can affect the pavement stresses significantly. A mathematical model has been
developed to approximate the stress distribution variation in space (assuming a
smooth tire), and then comparison with uniform rectangular patch loads was
presented for a variety of asphalt thicknesses, sub-base stiffnesses, wheel loads and
tire inflation pressures.

7.2 Conclusions
The following conclusions were drawn fiom this study as listed below:
1. When including a three-layer flexible pavement system, the critical model size in
three directions is 4000 mm X 4000 mm sub-grade depth of at least 740 mm.
Modeling only a single dual wheel load is accurate enough to simulate real truck
loads.

2. No matter what the size of the model and material properties of the pavement
structures, the resilient moduli of member layers can generally be back-calculated to
give a good comparison to the corresponding experimentally measured
displacements. For the two control sections in this study (one with an asphalt
thickness of 125 mrn and one with an asphalt thickness of 230 mm), different pairs
of sub-base and sub-grade moduli were computed for different total sub-grade depths
to give good fits to experimentally measured data. The maximum tensile stresses
predicted varied by less than 11%.
3. The back-calculation of material parameters of the K-8 model was performed to
give a good agreement with the experimental displacements using two methods, the
integration point method and the mid-layer method. The K-8 model did not give a
better fit to the experimental data when compared to models using linearly elastic
materials. The best combinations of Kj and K2 are different from two methods and
difference of the corresponding maximum principal stresses at the bottom of asphalt
is about 50%, but the computer predicted displacement basins can fit the field
measured displacement very well (see Figures 5.10 and 5.12). This is not surprising,
due to the limitations of K-8 model and the way we implement the two methods as
discussed in Chapter 5. Further, it takes much longer to achieve solutions when using
K-8 model than when using linearly elastic materials.

4. Spatially varying tirelpavement pressures affect the response of flexible


pavements significantly. The usual way that wheel loads are applied through a
uniform pressure acting on a rectangular or circular area tends to underestimate the
maximum tensile stresses at the bottom of asphalt concrete. For example, when

compared to the rectangular patch loads with the same contact area, the spatially
varying tirelpavement pressures yield stresses at the bottom of asphalt that are 10%30% larger. The largest differential occurs in thin flexible pavement structures with a
sub-base having low elastic modulus.

7.3 Recommendations for Future Work


The procedure and results of this study indicated certain areas where future
investigations are suggested:
1. Consideration of moving loads. Pavements are generally subjected to moving truck
loads instead of static loads and material response is viscoelastic. Further study on the
effect of moving loads and asphalt viscoelasticity is warranted.

2. Implementation of better material models. In this thesis, a K-8 model was employed
to incorporate material non-linearity. Unfortunately, the K-8 model has several
deficiencies. Better constitutive models for the sub-base and sub-grade are required.
3. Extension of current spatially varying pressure proqam. The computer program
developed in this study assumed that the pavements were subjected to static wheel loads
with smooth tires. However, actual tires with tread grooves should be taken in
consideration in future studies. Note that the process of triangulating the pressures and
adjusting total contact areas to presented here would be directly appliciable to any
distribution of contact pressures.

BIBLIOGRAPHY
Bensalem, A, A. J. Brown, M.E. Nunn, D. B. Merrill, and Wyn G. Lloyd, (2000), "Finite
Element Modeling of Fully Flexible Pavements: Surface Cracking and Wheel
Interaction", Preproceedings of the Second International Symposium on 3D Finite
Element for Pavement Analysis, Design, and Research, October 11- 13,2000 pp 103-122.
Blab, Ronald, John T. Harvey, (2000), "Modeling Measured 3D Tire Contact Stresses in
a Viscoelastic FE Pavement Model", Preproceedings of the Second International
Symposium on 3D Finite Element for Pavement Analysis, Design, and Research, October
11-13,2000, pp123-148.
Burmister, D. M. (1943). "The Theory of Stresses and Displacements in Layered Soil
Systems and applications to the Design of Airport Runways," Proceedings, Highway
Research Board, Vol. 23, pp. 126-144.
Burmister, D. M. (1945). "The General Theory of Stresses and Displacements in Layered
Soil Systems," Journal of Applied Physics, Vol. 16, pp. 84-94, 126-127,296-302.
Bush, A. J. 111. Nondestructive Testing for light Aircraft Pavements, Phase 11,
Development of Nondestructive Evaluation Methodology. Final Report FAA-RD-80-9-11.
FAA, Washington, D.C., 1986.
Chen, Dar-Hao, Musharraf Zaman, Joakim Laguros, and Alan Soltani (1995),
"Assessment of Computer Programs for Analysis of Flexible Pavement Structure",
Transportation Research Record, No. 1482, pp. 123-133.
Chen, Hsien H., Kurt M. Marshek, and Chhote L. Saraf, (1990), "Effects of Truck Tire
Contact Pressure Distribution on the Design of Flexible Pavements: A ThreeDimensional Finite Element Approach", Transportation Research Record, No. 1095 pp.
72-78.
Cho, Yoon-Ho, B. Frank McCullough, and Jose Weissmann, (1996), "Considerations on
Finite Element Method Application in Pavement Structural Analysis", Transportation
Research Record, No. 1539, pp. 96-101.
Cunagin, Wiley D. and Albert B. Grubbs, (1991), "Automated Acquisition of Truck Tire
Pressure Data", Transportation Research Record, No. 1322, pp. 112-121.
Davids, William G., George M. Turkiyyah, and Joe P. Mahoney (1998), "Modeling of
Rigid Pavements: Joint Shear Transfer Mechanisms and Finite Element Solution
Strategies", Report No. SGEM 98-6, Department of Civil Engineering, University of
Washington, Seattle, Washington.

Davids, William G., George M. Turkiyyah, and Joe P. Mahoney (1998), "EverFE Rigid
Pavement Three-Dimensional Finite Element Analysis Tool", Transportation Research
Record, No. 1629 pp. 41-49.
de Beer, M., C. Fisher and Fritz J. Jooste, (1997), "Determination of Pneumatic
Tirepavement Interface Contact Stresses Under Moving Loads and Some Effects on
Pavements with Thin Asphalt Surfacing Layers", Proceedings,Volume I, Eighth
International Conference on Asphalt Pavements, Seattle,Washington, August 10-14,
1997, pp 179-227.
De Jong, D. L., M. G. F. Peatz, and A. R. Korswagen, (1973), "Computer Progranl
System Under Normal and Tangential Load", Koin Klijke Shell-Laboratorium,
Amsterdam, External Report AMSR.0006.73.
Deusen D. V. Selection of Flexible Back-calculation Software for the Minnesota Road
Research Project. Report MN/PR-96/29. Minnesota Department of Transportation, St.
Paul, 1996.
Harichandran, Ronald S., Ming-Shan Yeh, and Gilbert Y. Baladi, (1990), "MICH-PAVE:
A Nonlinear Finite Element Program for Analysis of Flexible Pavements",
Transportation Research Record, No. 1286 pp. 123-131.
Helwany, Sam, John Dyer, and Joe Leidy, (1998), "Finite-Element Analysis of Flexible
Pavements", Journals of Transportation Engineering, Vol. 124 No. 5, pp. 491-498.
Hibbitt, Karlsson, and Sorensen, (1992), ABAQUS, version 5.2. Pawtucket, N.Y.
Hjelmstad, K. D., J. Kim and Q. H. Zuo, (1996), "Finite Element Procedures for ThreeDimensional Pavement Analysis", Transportation Research Record, No. 1539 pp. 66-71.
Huang, Yang H., (1993), "Pavement Analysis and Design", Prince-Hall, Inc.
Hwang, D., and M. W. Witczak, (1981), "Program DAMA (Chevron), User's Manual",
Department of Civil Engineering, University of Maryland.
Irwin, L. H. (1983), User's to MODCOMP2. Cornell Local Roads Program Report 83-8.
Cornell University Local Roads Program, Ithaca, N.Y., Nov. 1983.
Kopperman, S., G. Tiller, and M. Tseng, (1986), "ELSYM5, Interactive Microcomputer
Version, User's Manual", Report No. FHWA-TS-87-206, Federal Highway
Admimistration.
Lee Ying-Haur, Alaeddin Mohseni, and Michael I. Darter, (1 993), "Simplified Pavement
Performance Models", Transportation Research Record, No. 1397, pp7- 14.

Maestas, Joseph M. and Michael S. Mamalouk (1994), "Comparison of Pavement


Deflection Analysis Methods Using Overlay Design", Transportation Research Record,
No. 1377, pp. 17-25.
Nickels William L. Jr., (1995), " The Effect of Tire Chips as Subgrade Fill on Paved
Roads", M.S. thesis. Department of Civil Engineering, University of Maine, Orono,
Maine.
Raad, L. and J. L. Figueroa, 1980, "Load Response of Transportation Support Systems",
Transportation Engineering Journal, ASCE, Vol. 106, No. TE 1, pp. 111-128.
Reddy, J. N., (1993), "An Introduction to the Finite Element Method", the second edition,
McGraw-Hill, Inc.
Scullion, T. and C. Michalak, (1990), MODULUS 4.0: User's Manual Research Report
1123-4F. Texas Transportation Institute, Texas A&M University, College Station, Tex.,
Jan. 1990.
Shoukry, Samir N. and Gergis W. William, (1999), "Performance Evaluation of
Backcalculation Algorithm Through Three-Dimensional Finite-Element Modeling of
Pavement Structures", Transportation Research Record, No. 16% pp. 152-160.
Stubstad, R. (1988) Description of and User's Guide for the Dynatest ISSEM4 computer
program. Dynatest Consulting, Inc., Ojai, Calif..
Tielking, John T. and Moises A. Abraham, (1995), "Measurement of Truck Tire
Footprint Pressures", Transportation Research Record, No. 1435 pp. 92-99.
Turner, M., R., W. Clough, H. H. Martin, and L. Topp: "Stiffness and Deflection
Analysis of Complex Structures," Journal of Aeronautical Science, Vol. 23, pp. 805-823,
1956.
Uhlmeyer, Jeff S., Linda M. Pierce, Kim Willoughby and Joe P. Mahoney, (2000) "Topcracking in Washington State Asphalt Concrete Wearing Courses", Paper No. 00-0405,
proceeding of the 2000 Annual Meeting of the Transportation Research Board.
Vokas, Constantine A. and Robert D. Stoll, (1985), "Reinforced Elastic Layered
System",Transportation Research Record, No. 1153, pp. 1-7.
Warren, H. and W. L. Dieckrnann, (1963), "Numerical Computation of Stresses and
Strains in a Multiple-layer Asphalt Pavement System", Internal Report, Cheron Research
Corporation, Richmond, CA.
Weissman, Shmuel L., (1997), "The Influence of Tire-Pavement Contact Stress
Distribution on the Development of Distress Mechanisms in Pavements", Paper No.

005 10, proceeding of the 7gthAnnual Meeting of the Transportation Research Board,
Washington, D.C., Jan. 10-14, 1999.
Yue ZhongQi and Svec, Otto J., (1995), "Effect of tire-pavement contact pressure
distribution on the response of asphalt concrete pavements", Canadian Journal of Civil
Engineering, v22 n5 Oct. 1995, pp 849-860 03 15-1468.
Zienkiewicz, 0. C. and R. L. Taylor, (1988), "The Finite Element Method", Fourth
Edition, Volume I McGraw-Hill International (UK) Limited.

BIOGRAPHY OF THE AUTHOR


Jia Wang was born in Beijing, China, on August 16, 1973. He was raised in
Beijing by his parents, BingGuo, Zhou and YinJu, Wang, and graduated fiom
North Industry University with B.S. in Civil Engineering Department in 1996. He
enrolled in China Aviation Construction and Development Cooperation as an
engineer. He then enrolled in graduate study in Civil Engineering with a
concentration in Structural Engineering as a research assistant under Dr. William
G. Davids. Jia Wang is a candidate for the Master of Science degree in Civil
Engineering fiom The University of Maine in May, 200 1.

You might also like