You are on page 1of 9

Combined Effect of Moving Wheel Loading

and Three-Dimensional Contact Stresses


on Perpetual Pavement Responses
Hao Wang and Imad L. Al-Qadi

Tire–pavement interaction was analyzed with measured three-dimensional of the bending stiffness within the tire structure. The restricted inward
(3-D) tire contact stresses at various load levels (35, 44, and 53 kN) and movement of the tire ribs causes transverse stresses to develop, and the
constant tire pressure (720 kPa). The combined effect of moving wheel longitudinal stresses are primarily controlled by the tire–pavement
load and 3-D contact stresses on flexible pavement response was evaluated friction forces (2). The three-dimensional (3-D) tire contact stresses
with a developed 3-D finite element (FE) model, which incorporated the result in a complex stress state near the pavement surface, which
measured 3-D tire contact stresses, hot-mix asphalt (HMA) viscoelasticity, increases the potential for pavement damage, including top-down
and continuous moving load by using implicit dynamic analysis. In cracking, near-surface cracking, and hot-mix asphalt (HMA) rutting.
FE modeling, a perpetual pavement design with 254-mm HMA placed In addition, stress rotations and loading rates induced by moving
on 305-mm lime-modified subgrade was exposed to dual tire loading. loads accelerate pavement deterioration.
The critical pavement responses under two loading conditions (uniform Many researchers have analyzed the effects of 3-D tire contact
contact stresses and measured 3-D contact stresses) at various load levels stresses on flexible pavement damage. De Beer et al. (3) found that
were calculated and compared. The 3-D tire contact stresses induced pavement responses of thin HMA pavements are sensitive to vertical
greater pavement stresses and strains at the pavement near surface (shear load shape and distribution. Drakos et al. (4) concluded that the 3-D
strains and octahedral shear stresses) and at deeper depths (transverse tire contact stresses increase HMA rutting potential. Romanoschi and
tensile strains and compressive strains) comparable to the uniform contact Metcalf (5) and Al-Qadi and Yoo (6) reported that the effect of surface
stresses; these results suggest that using uniform contact stresses could tangential stresses on pavement response is significant. However,
underestimate pavement damage, especially near-surface cracking poten- limited investigations were conducted on the effect of 3-D contact
tial and shear flow in perpetual pavement. The transverse tangential stresses on perpetual pavement responses.
stresses induce the outward shear flow from the tire center and shear The perpetual pavement concept is to prevent cracks from begin-
strain concentration at the pavement near surface. Increasing the wheel ning at the bottom of the HMA layer through increasing HMA
load mostly increases contact stresses at the tire edge and the correspond- thickness. The design method of perpetual pavement is usually based
ing shear strains and octahedral shear stresses. The difference between on mechanistic–empirical approaches. In the conventional pavement
pavement responses caused by uniform contact stresses and 3-D contact design method, the dynamic moving transient load is considered as
stresses decreases when wheel load increases. a stationary uniform stress distribution on a circular contact area.
When the wheel load increases, the contact stress or the area or both
increase uniformly. This situation is inconsistent with field-based
The pavement–tire contact stress distribution arising from the moving experience. The effect of this assumption on resulting pavement
wheel load carried by each tire depends largely on tire geometry and responses may be minimal when considering the responses far from
loading conditions. Results from tire–pavement contact stress mea- the surface, but the resulting errors could be high near the pavement
surements indicate that different tire contact stress patterns develop surface. To achieve a realistic mechanistic–empirical design method,
when the tire load changes while the tire inflation pressure is constant the effects of a moving wheel load and 3-D tire contact stresses on
(1). For better understanding of the behavior of flexible pavements a pavement system need to be considered.
under real traffic loading, it is necessary to incorporate moving wheel
load and tire–pavement contact stresses into the vehicle–pavement
interaction analysis. OBJECTIVE AND SCOPE
When a tire loading is applied to a pavement surface, three contact
stress components are generated: vertical, transverse, and longitudinal. Tire–pavement interaction was analyzed by using the measured
The vertical contact stresses are not uniformly distributed because 3-D tire contact stresses at various load levels (35, 44, and 53 kN)
and constant tire pressure (720 kPa). The combined effect of a moving
wheel load and 3-D tire contact stress on perpetual pavement response
H. Wang, Department of Civil and Environmental Engineering, and I. L. Al-Qadi,
Illinois Center for Transportation, University of Illinois at Urbana–Champaign, was analyzed with a 3-D finite element (FE) model. This model
205 North Mathews Avenue, MC-250, Urbana, IL 61801. Corresponding author: incorporates the measured 3-D tire contact stresses and continuous
I. L. Al-Qadi, alqadi@uiuc.edu. moving load using implicit dynamic analysis. A perpetual pavement
design having 254-mm HMA placed on 300-mm lime-modified sub-
Transportation Research Record: Journal of the Transportation Research Board,
No. 2095, Transportation Research Board of the National Academies, Washington,
grade was exposed to dual tire loading. HMA linear viscoelasticity
D.C., 2009, pp. 53–61. was characterized by the relaxation modulus from creep compliance
DOI: 10.3141/2095-06 obtained by the indirect tensile test. The subgrade elastic modulus

53
54 Transportation Research Record 2095

was backcalculated from the falling weight deflectometer test. The The transverse distributions of contact stresses under each rib at
critical pavement responses for different failure mechanisms were various loading levels are shown in Figure 2b and c for the vertical
calculated and compared for two loading conditions (uniform contact contact stresses and transverse tangential stresses, respectively. The
stresses and 3-D measured contact stresses) at various loading levels. vertical contact stress is usually greater underneath inner tire ribs
(crown) than outer tire ribs, and the middle point of each rib has the
lowest vertical stress. The maximum vertical stress under the center
TIRE–PAVEMENT INTERACTION rib is about 1.6 times the tire inflation pressure.
The transverse tangential stresses show the distinct asymmetric
Tire–Pavement Contact Area distribution beneath each rib. If averaged over the tire width, the total
transverse stress is near zero. However, the transverse stresses may
Two important factors are at play when considering the tire– be either tension or compression at different positions along each tire
pavement interaction mechanism: contact area and contact stress. rib. In addition, the transverse tangential stresses range from about
Many researchers used the circular or equivalent rectangular contact 24% of vertical stress at the center rib to 35% to 52% of vertical con-
area in the pavement loading analysis (7). However, the contact area tact stress at edge ribs. Thus, the localized surface tangential stresses
of a truck tire is, in reality, closer to a rectangular than to a circular under each rib should be considered for predicting the pavement
shape, especially for a wide-based tire (8). In addition, both the response at the near surface.
circular and equivalent rectangular contact areas overestimate the Comparing Figure 2a, b, and c, the maximum vertical contact
net contact area without considering the tread pattern of the tire or stresses beneath the three center ribs are almost constant as the wheel
the localized stress distribution under each tire rib. load increases, and the vertical contact stresses at the two outside
Figure 1a shows the measured tire imprint for one tire of a dual tire ribs increase from around 600 to 800 kPa, probably due to the
assembly. The rectangular contact area of each rib is clearly indicated. influence of the structural stiffness of the tire sidewalls. The higher
Figure 1b shows the contact areas and contact lengths of the middle the load, the greater is the contribution of the sidewall, which is mainly
rib at three loading levels for 275/80R22.5 dual tires. The patterns transferred to the outside ribs in the contact area. However, the
of increased contact area and contact length with increased tire load increased transverse stress is insignificant as the load increases. The
can be clearly observed. As load increases, the increase in contact general shapes of the longitudinal stresses remain relatively constant
length is more significant for the middle rib than for other ribs. The in spite of the changes in load, with the exception of a slight difference
contact widths are almost constant under various loading levels. in the exit part. The peak longitudinal stress under each rib increases
as the loading increases. The contact stress data at various loading
levels are summarized in Table 1.
3-D Tire–Pavement Contact Stress

The measured longitudinal distributions of contact stresses (vertical, MATERIAL CHARACTERIZATION


transverse, and longitudinal) under the center rib of a 275/80R22.5
dual tire at a tire pressure of 720 kPa and various wheel loads on flat The perpetual pavement design used in the analysis is an existing
pavement surface are shown in Figure 2a. Both the vertical com- test section built as part of a project focused on the extended-life
pression stresses and transverse tangential stresses have a convex flexible pavement study (9), as shown in Figure 3a. The HMA was
shape along the longitudinal contact length. The longitudinal tan- prepared in accordance with the Superpave® volumetric design pro-
gential stresses vary significantly between the entrance and exit parts cedure. The laboratory mix–design criterion is based on 90 gyrations
of a tire imprint, with backward stresses in the front half and forward to achieve 4% air voids. Two asphalt binders were used in the HMA
stresses in the rear half. The location of a point where the longitudi- layers: a PG 64-22 for the standard binder course and a styrene–
nal stress directions change is determined by the rolling condition butadiene–styrene PG 70-22 for the polymer-modified binder courses
of the tire. and dense-graded surface. The asphalt content of the standard and

1000 400
Net Contact Area
Contact Length (mm)

800 Contact Length (Middle Rib)


Contact Area (cm2)

300
600
200
400

100
200

0 0
35 44 53
Load on Dual-Tire Assembly (kN)
(a) (b)

FIGURE 1 The 275/80R22.5 dual tire: (a) measured tire imprint and (b) contact area
and length.
Wang and Al-Qadi 55

1400
35kN
1200 44kN
Vertical contact stress 53kN
1000

800
Contact Stress (kPa)
600
Transverse contact stress
400

200

0
-120 -100 -80 -60 -20-40 0 20 40 60 80 100 120
-200
Longitudinal contact stress
-400
Longitudinal Contact Length (mm)
(a)

1500
Vertical Contact Stress (kPa)

1200

900

600

300
35kN 44kN 53kN

0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16
Data Points in a Rib
(b)

400
Transverse Contact Stress (kPa)

200

0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16

-200

35kN 44kN 53kN


-400
Data Points in a Rib
(c)

FIGURE 2 Distribution of (a) 3-D, (b) vertical, and (c) transverse contact stresses
at various loading levels (275/80R22.5 tire under 720-kPa inflation pressure).
56 Transportation Research Record 2095

TABLE 1 Contact Stresses at Various Loading Levels

Peak Stress (kPa)

Contact Average Vertical Transverse Longitudinal


Load (kN) Area (cm2) Stress (kPa) Stress Stress Stress

35 507 700 1,205 351 148


44 614 725 1,208 363 190
53 686 779 1,205 372 223

polymer-modified binder courses is 4.5%, and the HMA dense-graded FINITE ELEMENT MODEL DEVELOPMENT
asphalt content is 5.4%. No antistripping was used in the HMA. The
aggregate used in all mixes was limestone. The subgrade was 305-mm A 3-D FE model was developed using ABAQUS version 6.7 to predict
lime stabilized to increase the strength of the natural soil. pavement responses under various loading conditions (see Figure 3b).
In this study, an indirect tensile creep compliance test was con- The elements used were eight-node, linear brick elements with reduced
ducted to characterize the three HMA materials used in the perpetual integration (C3D8R). The infinite elements (CIN3D8) were used to
pavement section. The indirect tensile setup was used to allow testing reduce the number of far-field elements without significant loss of
of thin HMA layer cores taken from the field. In addition, this test’s accuracy and to create a “silent” boundary for the dynamic analysis.
tensile stress state is close to stress conditions in the field at the bottom Sensitivity analysis shows that the dimensions of the element thickness
of the HMA layer (10). The relaxation moduli were interconverted can be selected as 9.5 mm for the upper HMA layers, 20 to 50 mm
from the creep compliance data. The bulk (K) and shear (G) relax- for the granular base layers, and 200 to 500 mm for the subgrade (11).
ation moduli were calculated assuming a constant Poisson’s ratio The element horizontal dimensions along the vehicle loading area
and fitted into the Prony series as a generalized Maxwell solid model were dictated by the tire rib and groove geometries: 15 to 18 mm lat-
(see Equations 1 and 2). The elastic modulus of the subgrade was erally and 20 mm in the traffic direction. Full interface bonding was
backcalculated from the falling weight deflectometer test. assumed between HMA layers, and the Coulomb friction model was
used at the HMA and subgrade interfaces.
The loading time actually changes at various pavement depths
⎛ ⎞
G ( t ) = G0 ⎜ 1 − ∑ Gi (1 − e − t τi )⎟
N
(1) and the principal stresses rotate in the pavement under the moving
⎝ i =1 ⎠ wheel loading. A recently developed concept of continuous moving
load was used. In this approach, the tire loading imprint is gradually
⎛ ⎞
K ( t ) = K 0 ⎜ 1 − ∑ K i (1 − e − t τi )⎟
N
shifted over the loading area (step loading) until a single tire pass
(2)
⎝ i =1 ⎠ is completed. More details about the developed FE model and
continuous moving load are presented elsewhere (11, 12).
where
G = shear modulus, DYNAMIC TRANSIENT ANALYSIS
K = bulk modulus,
t = reduced relaxation time, Three different approaches can be used in pavement analysis: static,
G0 and K0 = instantaneous elastic modulus, quasi-static, and dynamic transient analysis. The static approach tra-
Gi, Ki, and τi = Prony series parameters, ditionally has been used in multilayer elastic analysis. The quasi-static
N = number of terms in the equation, and approach is based on the concept of moving the load at later positions
e = base of natural logarithm. along the pavement for each new time step and assuming the load is

Loading
DG Surface 50 mm Infinite area
element
Polymer Binder 115 mm

Standard Binder 89 mm

Lime Modified
Subgrade 305 mm
Infinite
element
(a) (b)

FIGURE 3 The 3-D FE model: (a) perpetual pavement structure and (b) cross
section (DG ⴝ dense-graded).
Wang and Al-Qadi 57

static at each position. No inertia or damping effects are quantified Transverse Tensile Strain (micro)
in quasi-static analysis. -150 -100 -50 0 50 100 150
Two important factors need to be considered in dynamic transient 0
analysis of the pavement, including the inertia associated with the
moving load and the dependency of the material properties on the 50

Depth (mm)
loading frequency. Zafir et al. (13) considered the continuum-based
100
finite layer approach and concluded that the dynamic effects of
moving loads on pavement strain responses are important and should 150
not be ignored. Yoo and Al-Qadi (14) found that the dynamic transient
analysis induces greater strain response and residual stress. 200
The FE method was used in this study for dynamic transient
analysis. This method permits modeling of more complex material 250
properties (such as linear or nonlinear viscoelasticity of HMA and Uniform contact stress
300 stress
Measured 3D contact
nonlinear elasticity of aggregate base) and pavement geometries
(discontinuity such as cracking or interface debonding) than is pos- (a)
sible with the analytical solution method. In ABAQUS, the dynamic
equilibrium equation can be solved by a direct integration method Longtitudinal Tensile Strain (micro)
such as implicit or explicit modes. -150 -100 -50 0 50 100 150
In pavement analysis, the rise time (loading duration) of the tran- 0
sient dynamic loading, which in this study depends on vehicle speed
[e.g., 8 km/h (0.009 s/20-mm element length in traffic direction)], 50
exceeds by a few multiples the time required for a stress wave to speed

Depth (mm)
through the flexible pavement structure: 100 to 600 m/s (15). Hence, 100
this problem may be classified as a structure dynamic problem instead
150
of a wave propagation problem. Using an implicit method is usually
most effective for a structure dynamic problem such as this one. The 200
damping ratio of 5% and the Rayleigh damping scheme were used
for the subgrade in the implicit dynamic analysis. 250
Uniform contact stress
Measured 3D contact stress
EFFECT OF 3-D CONTACT STRESS (b)
ON PAVEMENT RESPONSES
FIGURE 4 Horizontal tensile strains at bottom of
To evaluate the effect of 3-D contact stresses on flexible pavement HMA in (a) transverse direction and (b) longitudinal
responses, two groups of analysis were conducted: one using measured direction.
3-D contact stress and one using uniform contact stress distribution.
The uniform contact stress at each rib was calculated as the total load downward. Many field studies have proven that top-down cracking
divided by the actual net contact area, as shown in Table 1. Both groups is the major cracking mechanism in thick flexible pavement. Several
of analysis were conducted at a vehicle speed of 8 km/h at 25°C. factors have been proposed as the cause of top-down cracking: load-
induced factors (high tensile and shear stress or strain at the edges
of truck tires or in between truck tires), material factors (low-fracture
Tensile Strain
energies, HMA aging, and segregation during construction), and
The horizontal tensile strain at the bottom of HMA is thought to temperature-induced factors (extreme cooling rate) (16).
be related to the bottom-up fatigue cracking of flexible pavement. In this study, the maximum shear strain was found to be greater than
As shown in Figure 4, the longitudinal and transverse tensile strain the tensile strain at the bottom of HMA for both loading conditions,
distributions with depth are similar for the two loading conditions as shown in Figure 5a and b, consistent with the near-surface fatigue
(uniform stress and 3-D measured contact stress): compressive in cracking phenomenon reported in a recent study (11, 12). At inter-
the upper half of the HMA layer and inverted to tensile in the lower mediate to high temperatures, the vertical shear strains at the shallow
part of the layer. The transverse tensile strain at the bottom of HMA depth (up to 100 mm below the surface) are significantly greater
under 3-D contact stress was found to be 17% greater than the than the critical transverse and longitudinal tensile strains at the
transverse tensile strain under a uniform contact stress condition. surface and bottom of HMA layers for thick flexible pavement. This
Similar longitudinal tensile strains were obtained at the bottom of shear strain could induce a new phenomenon of crack development
the HMA layer for the two loading conditions, which suggests that the known as near-surface fatigue cracking. The difference between
pavement response in the transverse direction is more sensitive to maximum tensile strain and shear strain is 47% to 50% for 3-D
stress nonuniformity than the response in the longitudinal direction. contact stress and 25% to 37% for uniform contact stress. Hence,
the uniform contact stress underestimates the near-surface cracking
potential for thick pavement.
Shear Strain Figure 6 (a and b) shows the schematic distribution of the shear
strain within the HMA layer at 35-kN dual tire loading under uniform
Top-down cracking is recognized as longitudinal or transverse crack- contact stresses and 3-D contact stresses, respectively. The maximum
ing, or both, that appears at the pavement surface and propagates shear strain is concentrated at the upper part of the HMA layer at the
58 Transportation Research Record 2095

240 Maximum tensile strain


Maximum shear strain
180
Strain (micro)

120 (a)

60

0
35kN 44kN 53kN
(a)
(b)

FIGURE 7 Schematic distribution of octahedral shear stresses


240 Maximum tensile strain within HMA for (a) uniform contact stresses and (b) 3-D contact
Maximum shear strain stresses.
180
Strain (micro)

Octahedral Shear Stress


120
Two types of deformation exist in HMA layer: volume reduction
60
caused by traffic densification and permanent movement at a constant
volume caused by shear flow. Rutting is caused mainly by shear flow
because volumetric rutting by postcompaction appears only in the
0
initial phase.
35kN 44kN 53kN
Octahedral shear stress is independent of the first stress invariant.
(b)
Thus, it is possible to use octahedral shear stress to evaluate the HMA
FIGURE 5 Maximum tensile strains at bottom of shear flow potential without considering the volumetric deformation.
HMA and maximum shear strains for (a) uniform The crack can also be induced by shear because the shear stress can
contact stresses and (b) 3-D contact stresses. induce dilation due to aggregate particle movement and result in
tensile in the asphalt mastic or the interface between mastic and
tire edge for both loading conditions. The 3-D contact stress causes aggregate (18). Octahedral shear stress is calculated as the equivalent
greater shear strain at the pavement near surface than the uniform shear stress at the octahedral plane and reflects the ability of ductile
contact stress loading condition. The greater shear strain and the material to resist shear failure (see Equation 3) (19).
low confinement at the tire edge indicate that the 3-D tire contact
( σ 1 − σ 2 )2 + ( σ 2 − σ 3 )2 + ( σ 1 − σ 3 )2
stresses could cause significant cracking or shear flow, or both, at 1 2
τ oct = = J2 (3)
the pavement near surface. 3 3
An outward shear flow trend away from the tire center is clearly
observed under the 3-D contact stress loading condition, probably where
due to the effect of transverse surface tangential stresses and the
nonuniform distribution of vertical contact stresses. In addition, high τoct = octahedral shear stress;
shear strain concentration is observed close to the pavement surface σ1, σ2, and σ3 = maximum, middle, and minimum principal
under the tire ribs when the 3-D contact stresses are considered. This stresses; and
finding is consistent with the findings of top-down cracking reported J2 = second deviatoric stress invariant.
by Myers et al. (17 ). The schematic distribution of octahedral shear stresses within the
HMA layer at 35-kN dual tire loading is shown in Figure 7a and b for
uniform stress and 3-D contact stress loading conditions, respectively.
The octahedral shear stress concentrates near the surface because
of the effect of 3-D tire contact stresses. This high concentration
of octahedral shear stress indicates that the 3-D contact stresses can
cause significant HMA shear flow at the upper part of the HMA
(a)
layer and lateral vaults at tire edges compared with the uniform
contact stresses.

EFFECTS OF VARIOUS LOADS


ON PAVEMENT RESPONSES
(b)
As described in the tire–pavement interaction analysis, the vertical
FIGURE 6 Schematic distribution of shear strains within HMA contact stresses under each rib increase in a nonuniform manner as
for (a) uniform contact stresses and (b) 3-D contact stresses. load increases. In this section, the measured 3-D contact stresses under
Wang and Al-Qadi 59

various loading levels were used, and their effects on pavement 240

Maximum Shear Strain (micro)


responses were analyzed. Uniform contact stress
3D contact stress
180

Vertical Shear Strain and Octahedral


Shear Stress 120

Figure 8 (a and b) shows the schematic distribution of vertical shear 60


strain within HMA for 35- and 53-kN dual tire loading, respectively,
using measured 3-D contact stresses. As the load increases, the
maximum shear strain concentrates at the upper part of the HMA 0
35kN 44kN 53kN
layer at the tire edge, while the increase of shear strain concentra-
(a)
tion under tire ribs becomes less significant, possibly because the
increase in wheel load mainly increases the vertical contact stress

Maximum Octahedral Shear Stress


at the tire edge.
300
Figure 8 (c and d) shows the schematic distribution of octahedral
Uniform contact stress
shear stresses within HMA for 35- and 53-kN dual tire loading,
3D contact stress
respectively, using the measured 3-D contact stresses. As the load
200
increases, octahedral shear stresses increase, and a small concentration

(kPa)
of octahedral shear stress is observed at the inner edge of the dual
tire assembly.
100
The maximum shear strains and octahedral shear stresses at various
load levels were compared for the two loading conditions (uniform
and 3-D contact stresses), as shown in Figure 9a and b, respectively. 0
The 3-D contact stress resulted in a 7% to 18% greater maximum shear 35kN 44kN 53kN
strain and a 19% to 25% greater maximum octahedral shear stress than (b)
under uniform stress loading conditions. The strain differences
between the two loading conditions decrease as the load increases, FIGURE 9 Comparisons of (a) maximum shear
consistent with the fact that the nonuniformity of vertical contact strains and (b) maximum octahedral shear
stresses for various loading levels.
stress distribution at each rib decreases as load increases.

Tensile Strain

The transverse and longitudinal tensile strains at the bottom of


HMA at various loading levels were compared for the two loading
conditions (uniform and 3-D contact stresses; see Figure 10a and b).
(a) The 3-D contact stresses resulted in 9% to 17% greater transverse
tensile strains and 1% to 3% greater longitudinal tensile strains than
under the uniform stress loading condition. Similar to the shear strains
and octahedral shear stresses, the tensile strain differences for the
two loading conditions decrease as the load increases.

(b)
Compressive Strain

The compressive strains at the pavement near surface and the top
of the subgrade at various loading levels were compared for the two
loading conditions (uniform and 3-D contact stresses; see Figure 10c
and d). The 3-D contact stresses resulted in 28% to 35% greater
(c)
compressive strains at the pavement’s near surface and 7% greater
compressive strains at the top of the subgrade than under the uni-
form stress loading condition. The strain differences are significant
at the pavement’s near surface and become negligible at a greater
depth. The pavement responses at greater depth are mainly controlled
by total wheel load, suggesting that the 3-D contact stresses may
(d) not be neglected when considering pavement responses at the near
surface. In addition, the compressive strains at the pavement surface
FIGURE 8 Schematic distribution within HMA for (a) shear strain
under 35-kN load, (b) shear strain under 53-kN load, (c) octahedral change insignificantly as the load increases, because the vertical
shear stress under 35-kN load, and (d) octahedral shear stress contact stresses at the center rib are almost constant as the total wheel
under 53-kN load. load increases.
60 Transportation Research Record 2095

Transverse Tensile Strain (micro)


160 160

Longitudinal Tensile Strain


Uniform contact stress Uniform contact stress
120 3D contact stress 3D contact stress
120

(micro)
80 80

40 40

0 0
35kN 44kN 53kN 35kN 44kN 53kN
(a) (b)
Compressive Strain at Near-surface

200

Subgrade Compressive Strain


200
Uniform contact stress
Uniform contact stress
150 3D contact stress 3D contact stress
150
(micro)

(micro)
100 100

50 50

0 0
35kN 44kN 53kN 35kN 44kN 53kN
(c) (d)

FIGURE 10 Comparisons of (a) transverse tensile strains at bottom of HMA, (b) longitudinal tensile strains
at bottom of HMA, (c) compressive strains at pavement near surface, and (d ) compressive strains at top
of subgrade.

CONCLUSION nois Department of Transportation, Division of Highways; and the


U.S. Department of Transportation, FHWA. The authors acknowledge
This study demonstrates that 3-D contact stresses are important when David Lippert’s support and the assistance of the following members
analyzing perpetual pavement responses. The following conclusions of the Technical Review Panel for ICT-R59: Mark Gawedzinski,
can be drawn from this study: Amy Schutzbach, Bruce Peebles, Charles Wienrank, and Rich
Telford. Earlier work by J. Yoo and M. A. Elseifi is also greatly
1. The 3-D tire contact stresses induce greater compressive strain,
appreciated.
shear strain, and octahedral shear stress at the pavement near surface
than under the uniform contact stresses, suggesting that uniform con-
tact stresses might underestimate the near-surface cracking potential REFERENCES
and shear flow in perpetual pavement.
2. Although the pavement responses at greater depths are mainly 1. Yap, P. A Comparative Study of the Effects of Truck Tire Types on Road
controlled by wheel loading, the 3-D tire contact stresses induce greater Contact Pressures. Vehicle Pavement Interaction—Where the Truck Meets
transverse tensile strains at the bottom of HMA and compressive the Road. SP765. Society of Automotive Engineers, Inc., Warrendale, Pa.,
strains at the top of the subgrade. 1988.
2. Tielking, J. T., and F. L. Roberts. Tire Contact Pressure and Its Effect
3. The nonuniform distribution of vertical contact stresses and on Pavement Strain. Journal of Transportation Engineering, Vol. 113,
transverse tangential stresses induce the outward shear flow from the 1987, pp. 56–71.
tire center and shear strain concentration under tire ribs at the pavement 3. De Beer, M., C. Fisher, and F. J. Jooste. Evaluation of Non-uniform Tire
near surface. Contact Stresses on Thin Asphalt Pavements. Proc., 9th International
Conference on Asphalt Pavements (ICAP 2002), Copenhagen, Denmark,
4. Increasing the wheel loading mostly increases the contact August 17–22, 2002.
stresses at the tire shoulder and the corresponding shear strain and 4. Drakos, C., R. Roque, and B. Birgisson. Effects of Measured Tire Contact
octahedral shear stress at the tire edge. Stresses on Near Surface Rutting. In Transportation Research Record:
5. The difference in pavement responses for uniform contact Journal of the Transportation Research Board, No. 1764, TRB, National
stresses and 3-D contact stresses decreases as wheel loading increases. Research Council, Washington, D.C., 2001, pp. 59–69.
5. Romanoschi, S. A., and J. B. Metcalf. Characterization of Asphalt Con-
crete Layer Interfaces. In Transportation Research Record: Journal of
the Transportation Research Board, No. 1778, TRB, National Research
ACKNOWLEDGMENTS Council, Washington D.C., 2001, pp. 132–139.
6. Al-Qadi, I. L., and P. J. Yoo. Surface Tangential Contact Stresses Effect
on Flexible Pavement Response. Journal of the Association of Asphalt
This publication is partially based on the results of ICT-R59, Eval- Paving Technologists, Vol. 76, 2007, pp. 558–582.
uation of Pavement Damage Due to New Tire Designs, conducted 7. Huang, Y. H. Pavement Analysis and Design, 1st ed. Prentice Hall,
in cooperation with the Illinois Center for Transportation; the Illi- Englewood Cliffs, N.J., 1993.
Wang and Al-Qadi 61

8. Al-Qadi, I. L., M. A. Elseifi, and P. J. Yoo. Characterization of Pavement 15. Dynamic Loading of Pavements. Organisation for Economic Co-operation
Damage Due to Different Tire Configurations. Journal of the Association and Development, Paris, 1992.
of Asphalt Paving Technologists, Vol. 84, 2005, pp. 921–962. 16. Baladi, G. Y., M. Schorsch, and T. Svasdisant. Determining the Causes
9. Carpenter, S. H. Extended Life Hot Mix Asphalt Pavement (ELHMAP) of Top-Down Cracking in Bituminous Pavements. Final report for
Test Sections at ATREL. Research Report FHWA-ICT-08-017. Illinois Michigan Department of Transportation, Department of Civil and Envi-
Center for Transportation, Rantoul, 2008. ronmental Engineering, Michigan State University, East Lansing,
10. Buttlar, W. G., and R. Roque. Development and Evaluation of the Strate- 2002.
gic Highway Research Program Measurement and Analysis System 17. Myers, L. A., R. Roque, and B. E. Ruth. Mechanisms of Surface-Initiated
for Indirect Tensile Testing at Low Temperatures. In Transportation Longitudinal Wheel Path Cracks in High-Type Bituminous Pavements.
Research Record 1454, TRB, National Research Council, Washington, Journal of the Association of Asphalt Paving Technologists, Vol. 67,
D.C., 1994, pp. 163–171. 1998, pp. 401–432.
11. Yoo, P. J., and I. L. Al-Qadi. The Truth and Myth of Fatigue Cracking 18. Wang, L. B., L. A. Myers, L. N. Mohammad, and Y. R. Fu. Micro-
Potential in Hot-Mix Asphalt: Numerical Analysis and Validation. mechanics Study on Top-Down Cracking. In Transportation Research
Journal of the Association of Asphalt Paving Technologists, Vol. 77, Record: Journal of the Transportation Research Board, No. 1853, Trans-
2008, pp. 549–590. portation Research Board of the National Academies, Washington, D.C.,
12. Al-Qadi, I. L., H. Wang, P. J. Yoo, and S. H. Dessouky. Dynamic 2003, pp. 121–133.
Analysis and In-Situ Validation of Perpetual Pavement Response to 19. Doweling, N. E. Mechanical Behavior of Materials, 2nd ed. Prentice
Vehicular Loading. In Transportation Research Record: Journal of the Hall, Englewood Cliffs, N.J., 1999.
Transportation Research Board, No. 2087, Transportation Research Board
of the National Academies, Washington, D.C., 2008, pp. 29–39.
13. Zafir, Z., R. Siddharthan, and P. E. Sebaaly. Dynamic Pavement-Strain The contents of this paper reflect the views of the authors, who are responsible
Histories from Moving Traffic Load. Journal of Transportation Engi- for the facts and accuracy of the data. The contents do not necessarily reflect
neering, Vol. 120, No. 5, 1993, pp. 821–843. the official views or policies of the Illinois Center for Transportation, the Illinois
14. Yoo, P. J., and I. L. Al-Qadi. Effect of Transient Dynamic Loading on Department of Transportation, or FHWA. This paper does not constitute a standard,
Flexible Pavements. In Transportation Research Record: Journal of the specification, or regulation.
Transportation Research Board, No. 1990, Transportation Research Board
of the National Academies, Washington, D.C., 2007, pp. 129–140. The Flexible Pavement Design Committee sponsored publication of this paper.

You might also like