You are on page 1of 32

Road Materials and Pavement Design

ISSN: 1468-0629 (Print) 2164-7402 (Online) Journal homepage: https://www.tandfonline.com/loi/trmp20

Effect of vehicle speed and overload on dynamic


response of semi-rigid base asphalt pavement

Ogoubi Cyriaque Assogba, Yiqiu Tan, Zhiqi Sun, Nonde Lushinga & Zheng Bin

To cite this article: Ogoubi Cyriaque Assogba, Yiqiu Tan, Zhiqi Sun, Nonde Lushinga & Zheng
Bin (2019): Effect of vehicle speed and overload on dynamic response of semi-rigid base asphalt
pavement, Road Materials and Pavement Design

To link to this article: https://doi.org/10.1080/14680629.2019.1614970

Published online: 22 May 2019.

Submit your article to this journal

Article views: 65

View Crossmark data

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=trmp20
Road Materials and Pavement Design, 2019
https://doi.org/10.1080/14680629.2019.1614970

Effect of vehicle speed and overload on dynamic response of semi-rigid


base asphalt pavement
a
Ogoubi Cyriaque Assogba , Yiqiu Tanb∗ , Zhiqi Suna , Nonde Lushinga a
and Zheng Bina
a School of Transportation Science and Engineering, Harbin Institute of Technology, Harbin 150090,
People’s Republic of China; b School of Transportation Science and Engineering, State Key Laboratory
of Urban Water Resource and Environment, Harbin Institute of Technology, Harbin 150090, People’s
Republic of China

(Received 1 August 2018; accepted 28 April 2019 )

Premature deterioration in semi-rigid base asphalt pavement is still considered as one of the
main distress related to Chinese highway pavements. There are several reasons for this pre-
mature damage and these include poor design and construction, a sudden increase in traffic
loading and the dynamic loading of overloaded heavy truck. This paper investigated the effect
of vehicle speed and traffic overload on the dynamic response as well as on service life of
a semi-rigid base asphalt pavement with typical functional requirements. The study involved
simulating the critical stress–strain response at the bottom of asphalt layer under different
load weight level and vehicle speed. A large scale three-dimensional viscoelastic finite ele-
ment model of transient dynamic analysis of the pavement was developed. The Sinotruck
HOWO-A7 6 × 4 and HOWO-A7 8 × 4 were selected as a representative heavy truck and
the associated contact stress effect of each isolated wheel load on the pavement structure
was incorporated in the implicit dynamic analysis by a developed DLOAD subroutine. Field
dynamic strain measurement from Fiber Bragg Grating strain sensors at the Rizhan-LanKao
G1511 highway test section was successfully used to validate the developed model. The results
of this study indicate that under heavy moving truck load the response time of the longitudi-
nal, vertical and shearing stress–strain at the bottom of the asphalt layer has both a tensile
and compressive component, while the response time history of the transversal stress–strain
has only a tensile component. The result also indicated that a decrease in the vehicle speed
not only induce a significant increase in the load duration on the structure but also amplified
the shock effect induced by the wheel load. The damage caused by the overload is greatly
influenced by the axle weight extra extent. The fatigue analysis results revealed that, factors
such as overloading or low-speed traffic could result in adverse effect on the service life of
pavement. However, as expected the tensile strain at the bottom of the asphalt layer of the
proposed semi-rigid base asphalt pavement structure is less than the critical strain of a con-
ventional pavement structure with an ordinary asphalt mix even under critical conditions such
as overloading or low-speed traffic.
Keywords: semi-rigid asphalt pavement; dynamic response; viscoelastic material; finite
elements simulation; vehicle speed; overloads

1. Introduction
In China, more than 90% of highway pavements are constructed with asphalt concrete which
is placed on a semi-rigid base. Although asphalt pavement is usually designed to last about 15

*Corresponding author. Email: tanyiqiu@hit.edu.cn

© 2019 Informa UK Limited, trading as Taylor & Francis Group


2 O.C. Assogba et al.

years, they often experience premature distress and failure much before the end of their design
life. The most common distresses that occur in asphalt pavement are permanent deformation (rut-
ting), fatigue cracking and low-temperature cracking (Gopalakrishnan & Thompson, 2006). The
manifestation of such failure depends on a number of factors including the material characteris-
tics of each layer, pavement structure, traffic loading and environmental conditions. Traffic load
consists mainly of vehicle axle load configurations, axle loads, tire pressure and vehicle speed.
Vehicular loading induces dynamic stress that is transmitted to the pavement structure through
the complex areas between tire and pavement surface. The vehicle speed affects the amplitude of
stress and strains in the pavement due to the dynamic properties of the asphalt mixtures.
However, with the growth of the Chinese economy and the accelerated development of road
transport infrastructure, the volume of traffic continues to increase drastically. As a result, the
problems of vehicle overload in road transportation, low traffic speed and increasing num-
ber of heavy trucks are becoming more and more serious, thereby, accelerating the premature
appearance of functional and structural failure (Zhao, Tan, & Zhou, 2012).
Furthermore, the current design code and mechanistic analysis procedures for asphalt pave-
ment in China are based on laminar elastic theory under static vehicle load effect. This approach
was mostly based on low traffic volume, without considering the actual overload phenomenon,
the vehicle running speed, the viscoelastic behaviour of asphalt concrete layer and the dynamic
characteristic of the load induced by vehicle tires. Therefore, a big gap exists between this design
theory and the realistic loading condition of pavement under moving multiple axle loads.
Several research studies in recent years have investigated the effect of vehicle speed and over-
loading on the structural behaviour of asphalt pavement. Some researchers have used diverse
full-scale field measurement to perform a quantitative and qualitative analysis on the influence of
vehicle speed and overloading on asphalt pavement response, as well as on the performance and
service life of the asphalt pavement structures. For instance, Sebbaly et al. examined the influence
of Vehicle Speed on Dynamic Loads and Pavement Response using Weigh-in-motion systems
through a full-scale field experiment (Sebaaly & Tabatabaee, 1993). To evaluate asphalt pave-
ment dynamic responses accurately under truck loading, Zhang et al. used a full-scale aggregate
base asphalt pavement accelerated loading facility. Results of this study indicated that vehicle
speed has a significant effect on both the measured dynamic loads and the actual response of the
pavement system.
To investigate the actual strain response of asphalt pavement under a variety of loading and
environmental conditions in order to elaborate relevant prediction models, Ai et al. carried out
experimental measurement with various axle configuration, axle load, speed and testing temper-
ature (Ai, Rahman, Xiao, Yang, & Qiu, 2017). Based on Full-scale field testing on a highway
composite pavement, Chen et al. investigated the dynamic responses of a composite pavement
subjected to heavy traffic load (Chen, Zhang, & Wang, 2015). Their finding showed that the tire
contact pressure beneficially influences the dynamic strains and the dynamic stresses of pave-
ment. The combined effect of high temperature and heavy load on dynamic strain response is not
just a linear combination, but rather their coupling effect would exacerbate the fatigue damage of
asphalt layer. Through experimental and analytical methods on traffic-induced road vibration, Lu
et al. investigated the subgrade vibration induced by heavy-duty truck with various axle weights
and speeds (Lu, Hu, Yao, & Liu, 2016). The result indicates that heavy vehicles can induce
significant subgrade vibrations and thus exert much more damages on road structures than other
vehicles. The vehicle speed slightly affects the vertical stress of the subgrade. However, the effect
of vehicle speed on the acceleration is significant.
Some advanced 3D finite element models for simulating the structural behaviour of pave-
ment have been established. Al-Qadi et al. developed a 3D FE model incorporating measured
three-dimensional tire–pavement Contact Stress, continuous moving dynamic wheel load and the
Road Materials and Pavement Design 3

viscoelastic behaviour of the hot mix asphalt to predict the dynamic behaviour of the pavement
under vehicular loading (I. Al-Qadi, Wang, Yoo, & Dessouky, 2008). The established model was
also validated by the in-situ response of pavement to an accelerated pavement load. To assess the
effect of transient dynamic loading on flexible pavement, Yoo and Al-Qadi developed a 3D FE
model that incorporated the transient dynamic load model (Yoo & Al-Qadi, 2007). Results of this
study indicated an increase in stress and strains of up to 39% due to dynamic effect compared to a
correspondent static response. This was supported by Becky et al. who developed a linear elastic
3D FE model based on ANSYS to study the dynamic and static behaviour of a flexible pavement
response to moving vehicular load (Beskou, Hatzigeorgiou, & Theodorakopoulos, 2016). Sakar
numerically analysed the dynamic responses induced by single, tandem and tridem axles on flex-
ible pavements with two HMA layer thicknesses at different moving speeds (Sarkar, 2016). The
result revealed that the changing trend of critical responses versus speed is affected by axle con-
figuration and the surface layer thickness. Wang et al. analysed the dynamic stress response in
an asphalt pavement with subgrade based on cross-anisotropy under overload (B. S. Wang, Liu,
& Pang, 2013). The results show not only that overloading has a great influence on the stress
within the pavement but also indicates that the horizontal load induced by the deceleration has a
remarkable effect on longitudinal stress and horizontal shearing stress.
The aforementioned studies highlighted that heavy traffic and overload transportation system
have a significant impact on pavement service life and premature failure. The moving speed
affects the dynamic effect of the truck induced load on pavement structure through tire-pavement
interaction, consequently influencing the mechanical response of the pavement. The dynamic
load induced by heavy moving truck accelerates the premature deterioration of the pavement.
Thus, with the rising of vehicle running speed and axle load weight level, the current static design
approach may not perform properly with a more dynamic component highway response. Hence,
effort toward more advanced modelling of asphalt pavement and the need to deepen investigate
depth the influence of vehicle speed and the impact of traffic overload on the mechanical response
of highway pavement is still present.
As part of this effort, the current work developed a large-scale viscoelastic 3D Finite Element
model, which take into account all axle loads under the truck and the influence of each local
dynamic load induced by each isolated wheel load. To investigate the effect of vehicle speed
and overloading on the dynamic response and service life of a semi-rigid base asphalt pave-
ment subjected to heavy moving vehicular load. The pavement structure is new and meets to
typical functional requirements and specific materials, so that it can withstand reflective cracks
and bottom-up fatigue cracking, contrary to those in the literature. In addition, embedded Fiber
Bragg Grating (FBG) sensors which is a relatively novel method for pavement structural health
monitoring were used for the instrumentation of road test sections. Compared to sensors or trans-
ducers often found in the literature, the FBG sensors has a number of advantages including high
accuracy, multiplexing, electromagnetic interference resistance, good repeatability and ability to
measure low micro strain.

2. Objectives and scopes


The main objective of this study was to investigate the effect of vehicle speed and overloading
on the dynamic response as well as on the service life of a semi-rigid base asphalt pavement
typical functional requirements subjected to heavy moving vehicular load. To achieve this object,
a full-scale three-dimensional finite element model that integrates the viscoelasticity behaviour
of the asphaltic layers and consider the influence of each local dynamic load induced by each
isolated wheel under the moving truck was established, in order to accurately simulate the critical
4 O.C. Assogba et al.

stress–strain responses at the bottom of asphalt layer of an instrumented semi-rigid base asphalt
pavement subjected to heavy moving truck load. The predicted three directional dynamic strain
response was compared with the pavement dynamic response measured at the Rizhan-LanKao
G1511 highway test section instrumented with Fiber Bragg grating strain sensors and located in
Shandong province, P.R China.

3. Pavement dynamic analysis by FEM


3.1. Dynamic analysis
The wheel loads of moving vehicles that are applied to the pavement structure at different speeds
along the road system is indubitably dynamic. The oscillatory movement of the vehicle due to
the suspension system causes the dynamic wheel loads to vary around their mean amplitudes
when the vehicle is in motion (Cooperation & Group, 1992). For this reason, the contact stress
and position of the wheel loads are time dependent and considered to be equal to the sum of the
static load with a continuously changing dynamic load. Divers factors can affect the constantly
changing dynamic load such as the pavement surface irregularity, the vehicle speed, mass and
suspension system. The characteristic dynamic equations in the case of materials considered
isotropic and linear elastic can be presented according to the virtual work principle applied to a
domain , in the following form (Tautou, Picoux, & Petit, 2017):
→ → →
div σ + f = ρ γ (1)
=

→ →
where σ and f are Cauchy stress tensor and applied force, respectively; ρ is density and γ is
=
acceleration vector. The classic Hooke’s law can be employed to obtain Cauchy stress factor as
follows:
σ = λ∇U · I + μ(∇U + ∇UT ) (2)
where U and I are respectively the displacement vector and the identity matrix; λ and μ lame’s
coefficients. According to the standard FE Procedure (Bathe, 1982; Lee, Kim, & Jung, 1999),
the governing equation of the dynamic response of a nonconservative (dissipative) system with
damping can be written as follows:
•• •
[M ]{u} +[C]{u} + [K]{u} = F(t) (3)
• ••
where [M ] = mass matrix, [C] = damping matrix and [K] = the stiffness matrix {u} and { u } are
the velocity vector and the acceleration related to the nodes, respectively. {u} = displacement
vector and F(t) = external force vector related to the structure dynamic system. The dynamic
equilibrium equation (Equation (3)) can be solved by a direct integration method such as implicit
or explicit mode provide in Abaqus code. Using an implicit method is usually most effective for
a structure dynamics problem such as asphalt pavement structure (I. L. Al-Qadi, Elseifi, & Yoo,
2006; Bathe, 1982; Yoo & Al-Qadi, 2007).

3.2. Structural damping


The material dynamic response is affected by many factors, among which the damping mech-
anism which is related to the structure itself, the viscosity of the surrounding environment, the
energy dissipation of the soil foundation, etc. It is very hard to determine the damping matrix in
Road Materials and Pavement Design 5

FE analysis due to its complexity. The methods of damping calculation often use are stiffness fac-
tor method, stress-energy factor method, modal damping method and Rayleigh damping method.
The Rayleigh damping method which assumes that the damping matrix is a linear combination of
the mass matrix and stuffiness-proportional damping was used to calculate the dynamic response
in this analysis. The equation is expressed as follows:

C = α[M ] + β[K] (4)

where α is the mass damping coefficient and β is the stiffness damping coefficient. Using the
orthogonality condition, the coefficients α and β can be calculated through the vibration’s modes
and the corresponding modal damping ratio. Their equations are expressed in the following form:

⎪ 2wi wk (ξi wk − ξk wi )

⎨α =
wk 2 − wi 2
(5)

⎪ 2(ξk wk − ξi wi )
⎩β =
wk 2 − wi 2
where wi and wk are vibrations frequencies of the order i and k mode. The corresponding damping
ratios ξi and ξk of the order i and order k mode can be obtained by the resonant column test or
cyclic triaxial test (Zhong, Zeng, & Rose, 2002). The damping ratio can be approximated as a
constant within a certain frequency range by assuming that the damping coefficients α and β
have the same fraction of critical damping ratio ξ . Therefore, Equation (5) can be simplified as
follows (Bathe, 1982):

⎪ 2wi wk

⎨α = ξ
wk + wi
(6)

⎪ 2
⎩β = ξ
wk + wi
For a given order i, the critical damping ratio ξ can be expressed in terms of mass proportional
damping α and stiffness proportional damping β as follows.
α βw
ξ= + (7)
2w 2
The highest loading frequency and smallest natural frequency of the road system is considered
as critical frequencies. The fraction of critical damping of the road structure experimentally is
between 0.02 and 0.05 (I. Al-Qadi et al., 2008; Wu & Shen, 1996; Zhong et al., 2002). Since the
materials from the base to natural soil layer were characterised as elastic with no other source of
energy dissipation, the extreme critical damping rate (5%) was used in this 3-D FE analysis.

4. Finite element modelling of pavement structure


4.1. Model geometry
A full-scale 3-D FE models were developed using Dassault Simulia Abaqus 2017 (Hibbit, Karls-
son, & Sorensen, 2007) to simulate the Rizhan-LanKao G1511 highway test section (mile stake:
K313 + 600 ∼ K314 + 500) located in Shandong province in China. The real road system was
implemented in Abaqus by eight-node linear brick reduced integration elements and extends
horizontally and vertically to infinity. The full-size FE model dimension (x, y, z) was selected to
be 80 m × 20 m × 20 m, where x-axis denote transverse direction; y-axis denote vertical direc-
tion (depth) and the z-axis is the longitudinal direction (traffic direction). This selection has been
6 O.C. Assogba et al.

Figure 1. Controlled-load test truck configuration (not to scale).

done based on the result from in situ measurement of instrumented road test section, and after try-
ing various combinations of length, width and depth of the finite domain for the analysis while
considering the existing models in the literatures as reference (Beskou et al., 2016; Lu et al.,
2016; Mulungye, Owende, & Mellon, 2007; Saad, Mitri, & Poorooshasb, 2005). The in-plane
loading area of each dual tire wheel path is about 0.508 m × 60 m (width × length) with the fine
mesh element to balance the cost of calculation and accuracy. The established full-scale asphalt
pavement structural model consists of three asphaltic layers (Asphalt Concrete surface, middle
and bottom Course); two cement treated bound layers (base and subbase); one cement stabilised
soil (subgrade) all supported by the natural soil (infinite soil). Their respective thicknesses are
summarised in Table 4.

4.2. Modelling of the moving wheel load


The dynamic load and its contact stress effect on the pavement structure were incorporated into
the implicit dynamic analysis by a developed visual studio code for Abaqus user subroutine
DLOAD. The controlled-load test truck axles wheel loads moving was performed as transient
local dynamic load. The user subroutine DLOAD included in the FE programs Abaqus allows
users to specify a nonuniform distributed loads. This by specifying the variation of the distributed
load magnitude as a function of time (TIME*) in a specific coordinate (COORDS*) system
associated with domain element number and applied load integration point number (Hibbit et al.,
2007).

4.3. Wheel load configurations contact area and associated stress


Two types of controlled-load truck were used in order to analyse the dynamic response of
the instrumented Rizhan-LanKao G1511 highway test section subjected to vehicular moving
dynamic load under diverse wheel load configurations and various moving speeds. As can be
seen in Figure 1; One was a SINOTRUK HOWO-A7 6X4 dump truck with single steering axle
and tandem driving axle (Truck-1); The other was a SINOTRUK HOWO-A7 8X4 dump truck
with single unit of two steering axle and tandem driving axle (Truck-2). The dimension of the
wheels and axles of each controlled-load test truck are listed in Table 1.
The shape of each tire footprint measured was elliptical and could be modelled as a combina-
tion of two semicircles and a rectangle having a length L and a width 0.6L (Association, 1966;
Yoder & Witczak, 1975). The contact areas between tire and pavement surface considered in
Road Materials and Pavement Design 7

Table 1. Controlled-load test truck dimensions.

Dimensions (mm) Truck-1 Truck-2

Wheel base wb 5175 7350


Distance between the front first axle and the front second dfs 3825 1800
axle
Distance between the front second axle and the first driving dst 1350 4200
axle (third axle)
Distance between the third axle and the second driving axle dtf – 1350
(Fourth axle)
Front wheel track fwt 2022 2022
Rear wheel track rwt 1830 1830
Spacing between rear tires Srt 350 350

Table 2. Measured tire contact area.

Axle weight (kN) Tire contact surface (cm2)


First axle Second axle Third axle Fourth axle
Front axles Rear axles wheel wheel wheel wheel

Truck-1 71 187 556.618. 545.920 534.695


Truck-2 135 175 595.986 687.116 576.153 576.266

the FE model were generated from actual measurement (Table 2) and were for 11R22.5-11, and
11R22.5-13 radial tire for the test truck 1 and the test truck 2 respectively. The contact area Ac
and the length L were determined by using Equations (8) and (9) (Association, 1984; Hadi &
Bodhinayake, 2003; Huang, 2003).

Ac = π (0.3L)2 + (0.4L)(0.6L) = 0.5227L2 (8)



Ac 
L= = 1.3832 Ac (9)
0.5227
In the present study, the tire-pavement contact area induced by each wheel load was assumed
to be rectangular with an equivalent area. The length and the width of the converted rectangle
were 0.8712L and 0.6L respectively. It can be seen from Table 2 that the contact area may
vary significantly from one tire to another due to the load distribution. However, for numerical
modelling purposes and simplification of the FORTRAN subroutine source code, it was assumed
that all truck tires have the same Contact area. The length and the width of each tire footprint
were taken as 29 and 20 cm respectively.
Pavement structure damage such as top-down cracking, rutting in the asphaltic layer, and
cracking near the surface are due to the complex stress state near the pavement surface caused by
tire-pavement contact pressure (Drakos, Roque, & Birgisson, 2001; H. Wang & Al-Qadi, 2009).
Many field test measurements were performed to measure the tire-pavement contact pressure. It
was found that this contact pressure, is three-dimensional (vertical, horizontal and longitudinal
component) nonuniformly distributed over the contact area. The tire-pavement contact pressure
varies with the vehicle weight and speed, the tire type and its inflation pressure (I. Al-Qadi et al.,
2008; Beer, 1997; Drakos et al., 2001; Huang, 2003).
In FE analysis, since all wheel loads under the test truck axles have been considered and
the complexity of expressing the actual contact pressure in mathematical form considered, it is
difficult to implement this real stress contact in the Fortran subroutine. Hence, only the vertical
8 O.C. Assogba et al.

Figure 2. Finite element mesh of the pavement structure.

contact pressure has been considered to simulate the tire-pavement contact forces generated by
the moving controlled-load test truck. The vertical contact pressure of each tire is assumed to
be uniformly distributed over the contact area and can be easily determined by the wheel load
divided by the tire-pavement equivalent contact area.

4.4. FE mesh and boundary conditions


Several researchers have investigated the adequacy of FE mesh size refinement in a dynamic
model. Al Qadi et al. suggested that in the loading area, 20 mm can be adopted as the mesh length
in the traffic direction and between 15 and18 mm in the lateral direction (Hadi & Bodhinayake,
2003; Yoo & Al-Qadi, 2008). Since in FE analysis the consistency of the mesh size has a certain
influence on the result of the analysis, the mesh must be made in order to provide a best and most
accurate result. In this study, the FE model meshing was performed to increases the accuracy
of the model. The fine mesh around the loading area was used and along the wheel path in
which stresses and displacement are high. A dense size mesh was adopted near the loading area
and a relatively coarse mesh size was used far away from the area influenced by the loading as
shown in Figure 2. In order to enhance the convergence rate, the Continuum 3-D 8-node Reduced
integration element was used for the FE (Hernandez, Gamez, & Al-Qadi, 2016; Reynaud et al.,
2016; Saad et al., 2005; Sukumaran et al., 2004; Xia & Yang, 2012).
The boundary conditions have some effect on the FE analysis result accuracy, thus is essential
to choose correct and adequate boundary conditions. It is clear that when the trucks are moving
Road Materials and Pavement Design 9

on the pavement structure there is no response recorded at infinity. In addition, considering the
principle of Saint Venant (Breuer & Roseman, 1977) the following boundary conditions have
been adopted:

(1) The four sides of the FE model was restrained from any perpendicular movement to the
side of model, thus the x-axis displacements of two longitudinal side are constrained to
be zero and the y-axis displacement of two transverse side are constrained to be zero
(Liao & Sargand, 2010; Saad et al., 2005; Wang, Roque, & Morian, 2011);
(2) It assumed that at the bottom of the natural soil, there is no vertical and horizontal move-
ment, thus the bottom of the FE model was constrained to zero along the x, y and z
direction.

4.5. Model assumptions


The use of FE method to establish a full-scale size layered pavement system requires some
assumption. In this study, the presumptions made in order to simplify the actual road system and
focus on the key factors are enumerated below (Al-Qadi et al., 2008; Liao & Sargand, 2010; Yoo
& Al-Qadi, 2007):

(1) The base, subbase, subgrade and the natural soil (infinite soil) were assumed to respond
linearly and elastically to the applied dynamic loads while the asphaltic layer are mod-
elled as a linear viscoelastic, thus each pavement structure layer materials were assumed
to be homogeneous and isotropic;
(2) Since the road test section is newly built; the contact conditions between two adjacent
layers were assumed to be completely continuous;
(3) The stress and strain generated as well as the displacement of the pavement structure at
an infinite distance from the vehicular loading are zero when the vehicle is in motion;
(4) Tire and pavement are considered as an uncoupled and therefore establish as one-solid
model with well know contact stress (field measured contact stress);
(5) The surface of pavement model is a horizontal plane while the slope and longitudi-
nal unevenness of the road surface are neglected. The pavement structure initial stress
condition was neglected.

5. Pavement material characterisation


Material characteristics are needed as input parameters for pavement system modelling in FE
analysis. The constitutive models adopted for each layer are essential for the mechanistic analysis
of pavement dynamic responses (Li, Wang, Xu, & Xie, 2017). It is therefore necessary to adopt
an ideal constitutive model that will be used to characterise the material properties of each layer
of the structure. This section includes a brief description of the materials of each layer of the
pavement structure and presents their respective property characterisation.

5.1. Bonding layers and natural soil material properties characterisation


The composition of the cement treated base in the mixture is shown in Table 3 and more details
can be found elsewhere (Sun et al., 2018). The subgrade layer consists of 1.5% cement-stabilised
soil (CS) while the semi-rigid base and subbase layers consist of cement-treated bound materials
(CTB).
10 O.C. Assogba et al.

Table 3. Composition of CTB in the mix design.

Composition 20–30 mm 10–20 mm 5–10 mm Sand Cement Water

Weight ratio (%) 28 26 18 28 4 24

Table 4. Pavement layers thickness and Mechanical Materials data.


Elastic Poisson’s Damping
Parameters Thickness Unit weight modulus ratio ratio

Units of measurement cm kg/m3 MPa – –

N° Layer Material
1 AC surface Course SMA-13 3.5 2400 10,000* 0.30 0.05
2 AC middle Course SBS-AC20 13 2400 15,000* 0.30 0.05
3 AC bottom Course HE-AC5 2 2300 6500* 0.35 0.05
4 Base CTB 18 2100 14,000 0.25 0.05
5 Subbase CTB 17 2100 14,000 0.25 0.05
6 Subgrade CS 18 1900 6000 0.35 0.05
7 Natural Soil – Infinite 1800 120 0.40 0.05
*Elastic modulus for viscoelasticity (Moduli time scale instantaneous).

The materials from the base layer to the natural soil were assumed to be isotropic linear elastic.
Falling Weight Deflectometer (FWD) tests on constructed road test section was used to obtain the
subgrade and the natural soil elastic modulus by way of back-calculation as shown in Table 4.
The CTB material parameters were obtained from laboratory conducted uniaxial compression
modulus test method in accordance with the Chinese test methods JTG D50-2017 (Specifica-
tion for Design of Highway Asphalt Pavement-Appendix E). The CTB sample was core drilled
from the field and cut to a size of 150 × 150 mm (diameter × height). The sample after being
immersed in water for 24 h, was uniformly loaded with universal testing machine (UTM) at the
loading speed of 1 mm/min until failure. The load-strain data was recorded to determine a reliable
and accurate mechanical modulus of CTB material by the following relationship:
1.2 Fr
E= (10)
π D2 ε3
where E, Fr , D and ε3 are elastic modulus, maximum applied load, sample diameter and strain,
respectively. The calculated elastic modulus and other material parameters are presented in
Table 4. It should be noted that the cored samples were carried out as soon as the road test
section construction was completed and it was assumed that the CTB layer had not yet cracked.
It has been proven that CTB material in the field is susceptible to cracking and the modulus of
a cracked CTB material are different from the modulus obtained by laboratory test (Freeme, De
Beer, & Viljoen, 1987), therefore this study assumed modulus of uncracked CTB.

5.2. Asphaltic layers’ material property characterisation


From literature review on asphaltic concrete material, it is known that the basic performance of
the asphalt cement used in the mixture, the aggregate gradation and the design properties of the
mixture, have a great influence on the asphalt mixture mechanical properties. These mechanical
properties are the main cause of its viscoelastic behaviour. It is therefore appropriate to make a
brief description of the latter.
Road Materials and Pavement Design 11

Table 5. Basic performance of the binder used in each mixture.

Asphalt mixtures SMA-13 SBS-AC20 HE-AC5


Asphalt cement Compound High elastic
Performance types Asphalt binder modified with SBS asphalt binder

Penetration (25°C, 100G, 5S, 0.1 mm) 70 89 89


Softening point (°C) 77 86 86
Ductility (5 cm/min,5°C, cm) 42 62 62
Viscosity (135°C, Pa. s) 1.048 1.622 1.622

Table 6. Aggregate gradation of each mixture.

Passing percentage for the gradation of mixture (% passing)

Sieve Size (mm) 26.5 19 16 13.2 9.5 4.75 2.36 1.18 0.6 0.3 0.15 0.705
HE-AC5 100 96.2 64.3 46.5 26.8 15.0 11.1 9.3
SMA-13 100 94.1 45.7 27.8 19.7 16.1 14.2 13.4 11.1
SBS-AC20 100 96.1 84.4 75.1 54.8 36.2 25.3 17.9 11.6 7.5 6.1 5.4

Table 7. Volumetric design properties of mixtures.

% Air Voids % VFA % VMA Asphalt content

SMA-13 4.1 76.8 17.8 6.1


SBS-AC20 4.173 69.9 13.67 3.92
HE-AC5 1.84 91.71 21.68 8.73

The asphaltic layer of the road test section consists of three different types of asphalt mixtures
from three distinct asphalt cements. The first asphaltic layer applied stone mastic asphalt (SMA)
with standard asphalt binder, the second layer applied SBS-modified asphalt (SBS-AC) with
Compound modified with SBS and the last applied asphalt concrete with high elastic asphalt
binder (HE-AC). The SMA, SBS-AC and HE-AC mixtures consisted of aggregates of maximum
nominal size 19.5, 19 and 4.75 mm, respectively. Tables 5 and 6 present respectively the basic
performance of each asphalt cement used in each mixture and the Aggregate gradation of each
mixture. The volumetric design properties of each mixture such as the air voids content, the void
filled with asphalt volume (VFA), the void mineral aggregate (VMA) and the optimum asphalt
cement contents are provided in Table 7. More details about design properties of each mixture,
the aggregates gradation and asphalt cement properties can be found here (Sun et al., 2018; Tan
et al., 2017).
The analysis of the asphalt material response under an applied load at various temperature,
revealed that the behaviour of the asphalt concrete is not only time or frequency dependent but
also temperature dependent (de Araújo, Soares, de Holanda, Parente, & Evangelista, 2010; Li,
Guo, & Yang, 2015). Therefore, the different asphalt layers of the pavement structure are sup-
posed to be isotropic linear viscoelastic. The isotropic viscoelastic can be characterised by several
constitutive models such as the Kelvin model, the Maxwell model and Generalised Maxwell solid
model, and so on.
The generalised Maxwell solid model consists of a spring and several Maxwell elements
assembled in parallel. This model was used in this study to characterise the linear viscoelas-
tic behaviour of the asphalt layers, and it can be expressed by the following equations (Blab &
12 O.C. Assogba et al.

Figure 3. Relaxation modulus and master curve for different asphalt mixtures (log-log scale) and
comparison of temperature dependency of asphalt mixtures (log scale).

Harvey, 2002; Hu, Faruk, Zhang, Souliman, & Walubita, 2017):


t
de
s= 2G(t − τ ) dτ (11)

−∞

t
d(tr[ε])
p= 2K(t − τ ) dτ (12)

−∞

where s and e are deviatoric stress and strain, respectively; while p denote the volumetric stress
and tr[ε] denote the volumetric strain; G denote the shear relaxation modulus, K denote the
bulk relaxation modulus of the asphalt mixture in the time domain and t represent the relaxation
time. The shear relaxation modulus and the bulk relaxation modulus can be calculated from the
relaxation modulus which were obtained from laboratory dynamic complex modulus tests (Al-
Qadi et al., 2006; Park & Kim, 1999; Yoo & Al-Qadi, 2008). The dynamic complex modulus tests
were conducted in accordance with the American Standard Test Method for Dynamic Modulus
of Asphalt Mixtures (ASTM-D3497). Figure 3 displays the typical relaxation moduli obtained
from the laboratory and the corresponding master curve for each asphalt mixture at a reference
temperature of 21.1◦ C with their distinctive temperature shift factors function.

5.2.1. Time dependency and Prony–Dirichlet series


The time dependency of the asphalt mixture at each layer can be represented by Prony–Dirichlet
series in the form of shear and bulk modulus in the FE analysis software Abaqus, expressed in
Road Materials and Pavement Design 13

Equations (13) and (14) (Hibbit et al., 2007):



N
(−t/τi )
G(t) = G0 1 − Gi (1 − e ) (13)
i=1

N
(−t/τi )
K(t) = K0 1 − Ki (1 − e ) (14)
i=1

where N is the numbers of Prony–Dirichlet series terms, G0 and K0 denote the instantaneous
shear and elastic modulus, respectively; Gi , Ki and τi are Prony–Dirichlet series coefficients.
The constant poison’s ratio μ of 0.30 was assumed for the AC surface and AC middle course;
while for the AC bottom layer 0.35 were assumed. Then the shear moduli G(t) and bulk moduli
K(t) of each asphalt layer were determined from the relaxation modulus E(t) by the following
well-known basic relationship:
E(t)
G(t) = (15)
2(1 + μ)
E(t)
K(t) = (16)
3(1 − 2μ)
For more simplicity, only the shear modulus G(t) was taken to account for subsequent discussion.
By normalising the shear modulus G(t) by the instantaneous shear modulus G0 , Equation (15)
can be simplified into a ratio and expressed as follows:
G(t)
g(t) = (17)
G0

N
g(t) = 1 − Gi (1 − e(−t/τi ) ) (18)
i=1

where N is the number of Prony–Dirichlet series term; Gi and τi are the material constant and
the retardation time, respectively. In this study, the nonlinear curve-fitting utility included in
the FE analysis software Abaqus was employed to determine the Prony series parameters. The
root means square error (RMSE) function shown in Equation (19) was used to fit the measured
shear modulus g (t) to the calculated shear modulus g(t) for different number of Prony series
terms (Hibbit et al., 2007). Note that the calculated shear modulus denoted the one determined
from Equation (5) while the measured shear modulus is the one obtained by measured dynamic
complex modulus data in laboratory. The fitted Prony series parameter for asphalt concrete at
each layer (AC surface, middle and bottom course) are summarised in Table 8.


1
N
RMSE =  [gi (t) −
gi (t)]2 (19)
N i=1

5.2.2. Temperature dependency and WLF equation


The viscoelastic behaviour of the asphalt layers in the pavement system is highly influenced by
temperature. Therefore, in addition to the need to consider the time dependence, the temperature
dependence must be defined. The time-temperature shift factor aT changes the initial time (or
14 O.C. Assogba et al.

Table 8. Inputs parameters for defining viscoelastic behaviour of each asphaltic layer at 10°C.

Prony–Dirichlet series for generalised Maxwell model

SMA13 SBS-AC20 HE-AC5


N°* τ* G* τ G τ G

1 1.947E − 05 1.593E − 01 1.003E − 04 1.519E − 01 2.921E − 05 2.019E − 01


2 7.661E − 04 1.681E − 01 3.440E − 03 1.326E − 01 9.227E − 04 2.113E − 01
3 3.968E − 02 2.089E − 01 1.651E − 01 1.884E − 01 3.689E − 02 1.775E − 01
4 1.004E + 00 2.022E − 01 2.472E + 00 2.603E − 01 1.248E + 00 1.797E − 01
5 2.267E + 01 1.776E − 01 4.009E + 01 1.789E − 01 6.323E + 01 1.499E − 01
6 1.279E + 03 8.091E − 02 2.857E + 03 6.813E − 02 3.399E + 03 7.165E − 02
Williams–Landell–Ferry equations constants
SMA13 SBS-AC20 HE-AC5
C1 23.74 27.061 25.44
C2 177.80 222.90 198.80
RMSE 0.11 0.02 0.07
N°*: number of Prony Series; τ *: relaxation time; G*: dimensionless relaxation modulus.

frequency) to a reduced time (or frequency), that performs the translation of the master-curve
from the analysis temperature T to the reference temperature Tref . There are different relationships
that have been proposed to express this time temperature shift factor (Chailleux, Ramond, Such,
& de La Roche, 2006; Medani & Huurman, 2003; Pellinen, Witczak, & Bonaquist, 2004). In this
paper Williams Landed Ferry (WLF) was used (Hibbit et al., 2007; Williams, Landel, & Ferry,
1955) and its formulation is shown in Equation (20).
−C1 (T − Tref )
log(aT) = (20)
C2 + (T − Tref )
where aT is the time-temperature shift factor; C1 and C2 are two regression coefficients depen-
dent of the studied material; T the analysis temperature and Tref reference temperature. In order
to obtain accurate regression coefficients, the relaxation modulus data were first fitted using
the sigmoidal function in AASHOTO 2002 pavement Design Guide. The sigmoidal function
is expressed in the following form:
α
log(E(ξ )) = δ + (21)
1 + exp(β + γ log(ξ ))
where α, β, γ , δ are fitting coefficients and ξ reduced time. The regressions parameters obtain by
fitting the WLF equation to the shift factors for each asphaltic mixture are presented in Table 8.
Note that, the analysis temperature was taken equal to the temperature of the asphalt layers during
the field measurement which was approximately around 10°C.

6. FE model validations by full scale field testing


6.1. Design of test section
The semi-rigid base pavement test road section was constructed on Rizhan-LanKao G1511 high-
way, located in Shandong province, China. The pavement asphaltic layer includes a 35 mm
thick upper layer used by stone mastic asphalt (SMA-13) with standard asphalt binder and a
130 mm thick sub-layer layer used by SBS-modified asphalt concrete (SBS-AC20) with com-
pound modified with SBS, for bottom a 20 mm thick asphalt concrete with high elastic asphalt
Road Materials and Pavement Design 15

Table 9. Functional requirements of each asphalt layers in the actual pavement structure.

Layer Main stress characteristics Functional requirements Type of mixture adopted.

AC surface course Three-dimensional • Resistance to permanent Ultra-thin wearing


compression deformation, course(SMA-13)
• Water damage resistance,
• Anti-aging,
• Good surface
performance,
• Low temperature
performance.
AC middle course Vertical shear • Resistance to permanent High modulus asphalt
deformation, Concrete (SBS-AC20)
AC bottom course Horizontal tension and • Anti-fatigue, Stress absorbing layer
compression • Water damage resistance, (HE-AC5)
• Anti-reflective cracks.

binder (HE-AC5) were applied. The basic mechanical parameters of each layer and the volumet-
ric design properties, the gradation of the aggregate, the performance of the binder used in each
mixture were presented in previous sections “Pavement material characterisation” (Tables 3–8).
The usage of SMA-13 is to enhance tire wear, rutting resistance and durability. The sub-layer
(SBS-AC20) was also designed to resist permanent deformation and a stress absorbing bottom
layer designed to resist fatigue cracking, knows as fatigue resistant layer (FEL). Thus the aim
of designing such pavement was to eliminating reflective cracks and bottom-up fatigue cracking
that are very common in the asphalt pavement with semi-rigid base due to the incompatibil-
ity between asphalt materials and the semi-rigid base, more details can be found elsewhere
(Sun et al., 2018; Tan et al., 2017). The main stress characteristics and the functional require-
ments of each asphaltic layers in the actual asphalt pavement structure are summarised below
(Table 9).

6.2. Pavement section instrumentation and field testing


In order to assess pavement responses while being exposed to various environmental conditions
and vehicle induced dynamic loading, Fiber Bragg grating strain sensors and temperature sensors
were embedded in asphalt pavement structure: at the bottom of SMA13; at the mid-depth of SBS-
AC20 and at the top of the HE-AC5 following a comprehensive instrumentation layout plan as
shown in Figure 4. In order to quantify the vertical, transversal and longitudinal dynamic strains
at each location, the strains data collected will subsequently be used to verify the rationality of
the numerical simulation of the pavement dynamic response under moving vehicular loading.
Three test sections were instrumented on Rizhan-LanKao G1511 highway pavement test section
(mile stake: K313 + 600 ∼ K314 + 500). In this study, only the instrumentation data collected in
section K314 + 400 were used to compare the pavement dynamic response obtain from the field
to that obtain by FE analysis in order to validate the developed viscoelastic full-scale 3D-FE
model.

6.3. Field measured dynamic strain response


As the test truck-1 axles load passes over the instrument pavement the output signal from the
embedded sensors is recorded as shown in Figure 5 (left), it can be observed that original signal is
affected by noise. This could be related to the fact that the pavement structural dynamic response
16 O.C. Assogba et al.

Figure 4. Pavement instrumentation and data acquisition in field (Shandong Expressway).

depends on the ratio of the external loading frequency to the road system natural frequency. The
natural frequency of asphalt pavement structure varies between 6 and 12 Hz (Darestani, Thambi-
ratnam, Baweja, & Nataatmadja, 2006; Uddin, 2003) while the truck loading frequency is about
6.5 Hz at 82 km/h and 4.6Hz at 58 km/h (Gillespie, 1993). Therefore, the true amplitudes of the
sensor signal may fluctuate greatly and require the reduction of the signals noise by smooth-
ing before analysis. Digital filtering is usually used to improve the signal-to-noise ratio (SNR).
However, when the original field measured strain signal is filtered, it is unavoidable to lose its
peak. The phenomenon known as peak clipping phenomenon is depicted Figure 5 (right). The
peak value of the signal is reduced to less than the original signal peak after filtering by FFT fil-
ter smoothing offers available in the Scientific Data Analysis and Graphing Software Origin-Pro
2017.
The time-dependent transversal strain recorded by the embedded FBG strain sensors before
and after the passes of each wheel load under the truck are depicted in Figure 6. It can be seen
that, as the controlled load test truck wheel approach, the measurement points (FBG sensors
locations), the longitudinal dynamic strain responses are in compression and change to tension,
then it gradually increases to its maximum tensile strain value reached when the controlled load
test truck wheel load is right below the measurement points. As the truck wheel load moves
far away from the measurement points, the tensile stress decreases progressively and finally re-
changes to compression. Therefore, as the truck approaches and moves away from the measuring
points, the longitudinal dynamic stresses response are subjected to compression while the stresses
are hardly tensile under the wheel load.
Road Materials and Pavement Design 17

Figure 5. Unfiltered (left) and filtered (Right) three-directional strains times histories.

6.4. Comparative analysis between FE simulation result and field data measuring
The 3-D FE analysis result and the full-scale field testing measured three directional dynamic
strain taken at the bottom of AC surface course (Surface), mid-depth of AC middle course (Mid-
dle) and at the top of the AC bottom course (Bottom) under the controlled-load test Truck-1 at a
moving speed 20 km/h are depicted in Figure 7.
As it can be seen from the above comparisons, the numerical simulation and the field mea-
surement curves are in a good agreement with the same trend of change over time and peak
values that are very close to each other, but there is a certain difference in the residual strain and
response time.
Explicitly, in the case of transverse deformation, it is observed that in terms of curves shape the
field measured loading duration is shorter than the calculated. The differences between the max-
imum transverse strain values measured at the surface, middle and at the bottom are respectively
about 2%, 7% and 4%. Regarding the vertical strain: the field measurement response is slower
18 O.C. Assogba et al.

Figure 6. FBG recorded longitudinal dynamic strain at different truck location during the field testing.

to recover, and the load duration is longer. The numerical simulation predicted the vertical strain
with an error range less than 10% at the middle and the bottom of SBS-AC20. For the longitudi-
nal deformation, it may be noted that the peak values of the longitudinal strain measured on the
site are slightly lower than the numerical simulation whereas the compression strain is relatively
high. The difference between the maximum longitudinal strain values measured at the surface, at
the middle and at the bottom on site and the numerical simulation values are approximate of 5%,
4% and 3%, respectively.
The comparative study between the FE simulation result and the field data measurement
revealed that there are minor differences. The reasons for the above differences are that:

(1) It was difficult to accurately control the speed of the test truck and the loading position
during the full-scale field-testing which ideally should be applied symmetrically since
the instrumentation axis is right below the centre of dual tires.
(2) When the trucks approach, the sensors are subjected to compression strain and even after
the passage of the truck, the sensors remain subject to compression strain. The response
of the sensor is therefore affected by the constraints of surrounding materials.
(3) The controlled load test truck wheel loads were assumed to be uniformly distributed
over a rectangular contact area between each tire and pavement which is different to the
real contact stresses and pressure induced by the tires during field testing. Only accurate
three-dimensional contact pressure and non-uniform contact area can simulate with a
very good accuracy the real conditions of stress distribution in the asphalt pavement
structure.
Road Materials and Pavement Design 19

Figure 7. Comparison of three-directional dynamic strains between FE simulation (left) and field
measurement (right).

In summary, the developed viscoelastic 3D-FEM numerical simulation result is basically con-
sistent with the field measured three-directional strain and can be used to accurately predict the
pavement dynamic responses under moving truck loads.

7. Pavement response analysis and discussion


7.1. Effect of vehicle speed on dynamic response of pavement
Few documented literatures have focused and discussed the influence of traffic speed on the
dynamic response of semi-rigid base asphalt pavement. No perfect relationship between speed
20 O.C. Assogba et al.

Figure 8. Strain time histories at different vehicle speed.

and the pavement dynamic response has been established. Nevertheless, studies revealed that
the speed of the vehicle is directly interrelated not only to the role of time but also to the
stress–strain response. Therefore, the vehicle moving speed may have a significant effect on the
pavement dynamic response. In this study, four different traffic speed levels that are 20, 40, 60
and 80 km/h were assumed to analyse the influence of truck speed on the three directional strain,
stress and shearing stress, strain. The axle loads were set at the maximum permitted loading
capacity.
Abaqus can compute any dynamic response at any analysis point. However, only the strain
and stress at the bottom of asphalt layer under the centre of the dual tires were reported in this
paper. This is due to the fact that one of the most important performance indicators required
in pavement performance prediction is the tensile stress at the bottom of the asphalt layer. The
typical time history of the dynamic strain and the dynamic stress at the indicated point under the
controlled load test truck moving at the selected speeds are respectively displayed in Figures 8
and 9. It can be seen that the curves have the same curvature and the same trends, but not the same
response time, nor the same peak value. The response time history of the longitudinal, vertical
and shearing dynamic stress–strain at the bottom of the asphalt layer has both a tensile and
compressive component while the response time history of the transversal dynamic stress–strain
has only a tensile component.
The stress–strain peak response that corresponds to the presence of front axles and rear axles
versus the vehicle speed, and its potential approximation are display in Figures 10 and 11.
The relationship between the dynamic response of semi-rigid base asphalt pavement and the
speed of the controlled load truck were approximated with the greatest precision by the poten-
tial exponential functions below (Equation (22)). Where parameters y0 , a and t are regression
Road Materials and Pavement Design 21

Figure 9. Stress time histories at different vehicle speed.

constant.
y = y0 + ae(− t )
x
(22)
It is highlighted that the peak value of transverse strain and transverse dynamic stress at the bot-
tom of asphalt layer significantly increases as the speed of the controlled load test truck decrease,
as well as the peak value of vertical, longitudinal and shearing stress, strain. For instance, when
the controlled load test truck 1 moving speed decrease from 80 to 20 km/h, rear axle peak value
of the transverse tensile stress increase from 50.908 kPa to about 60.394 kPa (an increase of
18.63%), the maximum vertical tensile stress increase from 37.875 kPa to about 48.734 kPa
(an increase of 28.67%), the maximum longitudinal tensile stress increase from 47.744 kPa to
about 59.618 kPa (an increase of 24.87%) and the maximum tensile shear stress increase from
10.932 kPa to about 14.589 kPa (an increase of 33.45%). Similarly, rear axle peak value of the
transverse tensile strain increases from 2.91 με to about 3.22 με (an increase of 10.65%), the
maximum vertical tensile strain increases from 5.52 με to about 9.21 με (an increase of 66.85%),
the maximum longitudinal tensile strain increases from 2.25 με to about 2.91 με (an increase
of 29.33%) and the maximum tensile shear strain increase from 3.53 με to about 5.12 με. (an
increase of 45.04%).
It can be expected that an increase in traffic speed is beneficial for the road system fatigue
life. However, the decrease in the vehicle velocity not only induce an increase in the wheel load
duration on the structure but also amplified the shock effect of the tire load on the pavement
structure. Although it is not practical at all to apply the lowest traffic speed just based on the
pavement structure’s requirement, it will be appropriate to regulate the pavement structure’s
lowest traffic speed. The influence of traffic speed is not negligible and should be included in the
pavement mechanical analysis.
22 O.C. Assogba et al.

Figure 10. Effect of vehicle moving speeds on three directional strains.

7.2. Effect of load weight level on dynamic response of pavement


Tire pressure induced by the traffic loading is one of the key factors taken into consideration
when designing the pavement structure. In China, the design criterion imposes 0.70 MPa for the
tire pressure, but the pavement structure is subjected to a more vehicular tire pressure during its
service period because of the overload phenomenon. Due to the nature and position of the load,
a vehicle loaded within its gross weight limit can still exceed the weight limit allowed for each
axle over the pavement structure. It is therefore crucial to analyse the influence of the overload on
pavement structure responses. The maximum permitted axle weight and a series of axle weight
overload condition of each controlled load test truck were included in this investigation. The
distribution of each vehicle axle load under the truck load capacity (maximum permitted) and
under the different overload conditions considered in this study are summarised in Table 10.
Figures 12 and 13 depict the time history of the strain and the stress at the bottom of the asphalt
layer, under the centre of the dual tires of the controlled-load test truck 2 moving at the speeds
of 20 km/h, all under different loading conditions. It can be seen that the stress and strain time
histories under different overload condition have the same curve shape and the same trend of
change over the time, but their peak values vary from one overload condition to another. The
response time history of the transversal stress–strain at the bottom of the asphalt layer for each
overload condition has a tensile component, while the response time history of the longitudinal,
vertical and shearing stress–strain has both a tensile and compressive component.
Road Materials and Pavement Design 23

Figure 11. Effect of vehicle speeds on three directional stress.

Table 10. Distribution of axle load under different overload condition.

Axle weight under different overload condition (kN)


Axle loading
capacity (kN) 20% Overload 40% Overload 60% Overload
Front Rear Front Rear Front Rear Front Rear
wheel wheel wheel wheel wheel wheel wheel wheel

Truck-1 70 180 71.25 228.75 73.85 276.15 76.50 323.50


Truck-2 135 175 147.42 234.36 166.47 274.53 171.10 332.90

Figures 14 and 15 exhibit the influence of the overload on pavement dynamic response. It
can be seen that the pavement dynamic peak stress and peak strain take at the bottom of asphalt
layer increases promptly as the overload rate increase. For instance with the rise of axle weight
from axle loading capacity to 60% overload condition, the transverse tensile strain induced by
the rear axle of the controlled load test truck 2 increases from 3.21 με to about 5.92 με (an
increase of 84.42%), the maximum vertical tensile strain increase from 0.97 με to about 1.78 με
(an increase of 78.53%), the maximum longitudinal tensile strain increase from 2.70 με to about
5.47 με (an increase of 102.59%), and the maximum tensile shear strain increase from 5.23 με
to about 10.10 με. (an increase of 93.12%). Similarly, the transverse tensile stress increase from
24 O.C. Assogba et al.

Figure 12. Strain time histories under different overload condition.

Figure 13. Stress time histories under different overload conditions.


Road Materials and Pavement Design 25

Figure 14. Influence of load weight level on three directional strain.

59.575 kPa to about 111.580 kPa (an increase of 87.29%), the maximum vertical tensile stress
increase from 47.924 kPa to about 90.943 kPa (an increase of 89.76%), the maximum longitu-
dinal tensile stress increase from 57.429 kPa to about 111.250 kPa (an increase of 93.72%) and
the maximum tensile shear stress increase from 15.188 kPa to about 28.962 kPa (an increase of
93.72%).
Obviously with the increase of the load rate up to 60% of its maximum allowed load, the
dynamic response of the pavement structure at the bottom of the asphalt layer has increased
drastically by more than 78%. This indicates that the damage caused by the overload is greatly
aggravated with the axle weight extra extent which can greatly shorten the service life of the
pavement structure. It is therefore essential on one hand, to consider the impact of traffic overload
in the design phase of the pavement structure; while on the other hand, to regulate the traffic load
and especially the overloading. Otherwise the pavement structure could be destroyed before the
end of its design life, due to the damage accumulation.

7.3. Effect of vehicle speed and overload on pavement damage


To account for the most severe types of load-associated pavement damage, two damage param-
eters are often considered: fatigue cracking and permanent deformation. This study focuses on
fatigue cracking, which is a type of pavement distress due to localised progressive degrada-
tion caused by fluctuating stress and strain in the material as well as by the accumulation of
irrecoverable strain.
The asphalt institute’s fatigue cracking transfer function presented in Equation (23) was
employed to predict the number of repetitions required for fatigue cracking to occur
26 O.C. Assogba et al.

Figure 15. Influence of load weight level on three directional strain.

(Institute, 1982).

Nf (εt ) = A × 0.00432 × 104.84(VFA−0.69) × εt −3.291 × |E ∗ |−0.854 (23)

Where Nf = allowable load repetitions to prevent fatigue cracking; A = 18.4 is the laboratory-
to-field adjustment factor; VFA = void filled with asphalt; εt = tensile strain at the bottom of
asphalt layer and |E ∗ | = dynamic modulus.
The theory of the fatigue endurance limit for asphalt pavement advocates the existence of a
stress threshold in the asphalt mixture. When the tensile strain at the bottom of the asphalt layer
is greater than the threshold, the asphalt layer will generate fatigue damage. The corresponding
fatigue life of this tensile strain point is the fatigue limit. The research results of many countries
in the world in recent years support the existence of a fatigue limit in asphalt mixtures, and it is
generally agreed that a strain critical point (known as critical strain) of an ordinary asphalt mix-
ture is about 70 με (Newcomb, Buncher, & Huddleston, 2001) and that of the modified asphalt
mixture is about 100 με (Quintus & Harold, 2006).
In the following section, only the dynamic response under the controlled-load test Truck-1
was considered to examine the effect of vehicle speed and overload on pavement damage. The
maximum tensile strain at the bottom of the fatigue resistant layer in axle’s moving direction
and in lateral direction was selected as pavement response parameter. The predicted allowable
number of repetitions to prevent fatigue cracking under different vehicle speed and load weight
level for both strain direction is presented in Table 11.
Road Materials and Pavement Design 27

Table 11. Fatigue life analysis under different overload condition.


Axle weight Axle loading 20% 40% 60%
Strain direction (kN) capacity Overload Overload Overload

In axle’s moving Maximum strain for 2.940 3.830 4.670 5.490


direction fatigue distress (*E − 6)
Fatigue life Nf 1 (*E + 12) 12.647 5.297 2.758 1.620
In lateral Maximum strain for 3.226 4.035 4.754 5.604
direction fatigue distress (*E − 6)
Fatigue life Nf 2 (*E + 12) 9.314 4.462 2.602 1.514
Minimum Nf 1,2 (*E + 12) 9.314 4.462 2.602 1.514
Under different moving speed

Strain direction Moving speed 20 km/h 40 km/h 60 km/h 80 km/h

In axle’s moving Maximum strain for 2.940 2.710 2.460 2.280


direction fatigue distress (*E − 6)
Fatigue life Nf 1 (*E + 12) 12.647 16.536 22.739 29.199
In lateral Maximum strain for 3.226 3.290 3.100 2.900
direction fatigue distress (*E − 6)
Fatigue life Nf 2 (*E + 12) 9.314 8.735 10.623 13.231
Minimum Nf 1,2 (*E + 12) 9.314 8.735 10.623 13.231

The table indicates that the transverse strain at the bottom of asphalt layer is greater than the
longitudinal strain. For instance, the strain at the bottom of asphalt layer in lateral direction is
1.10 times the strain in axle’s moving direction when the truck is loading for a maximum load
capacity allowed and for a speed of 20 km/m. As the truck weight level increases, the static load
and the dynamic tire force distributed on the pavement structure also increases, and the allowable
number of load repetition to prevent fatigue cracking decreases sharply. The minimum fatigue
life for axle loading capacity is 16.25% of that for 60% overload. The analysis of the effect of
vehicle moving speed on the asphalt layer fatigue life as expected shown that the fatigue life
increase as the moving speed increases. For example, with the rise of the moving speed from
20 km/h to 80 km/h, the minimum fatigue life increases by 42.05%. However, the increase in
vehicle moving speed will worsen vehicle vibration which is disadvantageous to driving safety.
The effect of vehicle speed must therefore be considered in all its aspects. All in all, it can
be concluded that overloading or low-speed traffic will negatively affect the asphalt pavement
service life. These factors also can be the main reasons for premature fatigue cracking in asphalt
pavements. Nevertheless, it should be emphasised that: as expected, the horizontal strain at the
bottom of the asphalt layer of the proposed new semi-rigid base asphalt pavement structure is
less than the critical strain of a conventional pavement structure with an ordinary asphalt mix
(70 με) even under critical conditions such as overloading or low-speed traffic.

8. Summary and conclusions


The effect of vehicle speed and overload on the dynamic response of semi-rigid base asphalt
pavement with a typical functional requirement subjected to heavy moving vehicle load was
investigated in this study. A full-scale 3-D FE model of transient dynamic analysis of the pave-
ment comprising of multiple axle loads was established to compute the critical stress–strain
responses at the bottom of asphalt layer. The adopted 3-D FE model incorporated the viscoelas-
tic behaviour of the asphaltic layers and the effect of the transient local dynamic wheels load. The
28 O.C. Assogba et al.

dynamic stress–strain responses at the bottom of the asphalt layer were analysed under different
load weight level and at different vehicle speed. In order to examine in detail the effect of vehi-
cle moving speed and the impact of traffic overload on the pavement dynamic response, the
pavement dynamic response measured at the Rizhan-LanKao G1511 highway pavement test
section was used to verify the validity of the developed viscoelastic 3D FE model. Some of the
preliminary findings from this study are summarised as follows:

• The pavement dynamic responses in term of strain (vertical dynamic strain, transversal
dynamic strain and longitudinal dynamic strain) at various depth (surface, middle and
bottom of SBS-AC20) were calculated and the result are consistent with the full-scale
fields measured responses.
• Under a moving heavy vehicle load the response time history of the longitudinal, vertical
and shearing stress–strain at the bottom of the asphalt layer has both a tensile and com-
pressive component, while the response time history of the transversal stress–strain has
only a tensile component.
• The analyses of vehicle speed effect on the pavement dynamic response revealed that the
decrease in vehicle velocity not only induce a significant increase in the wheel load dura-
tion on the structure but also amplified the shock effect of the tire load on the pavement. It
is therefore, appropriate to regulate the pavement structure lowest traffic speed and include
the effect of traffic speed in the pavement mechanical analysis.
• The analyses of load weight level effect on the pavement dynamic response indicates that
the damage caused by the overload is greatly aggravated by the addition of extra axle
weight which can greatly shorten the service life of the pavement structure. It is therefore
essential on the one hand, to consider the impact of traffic overload in the design phase of
the pavement structure; while on the other hand, regulate the traffic load and especially the
overloading.
• The analyses of the effect of vehicle speed and overload on the pavement damage indi-
cated that the fatigue life of the asphalt concrete layer is strongly dependent on the load
amplitude and the vehicle moving speed. Consequently, factors such as overloading or
low-speed traffic could result in adverse effect on the service life of pavement. However,
as expected the tensile strain at the bottom of the asphalt layer of the proposed semi-rigid
base asphalt pavement structure is less than the critical strain of a conventional pavement
structure with an ordinary asphalt mix even under critical conditions such as overloading
or low-speed traffic.

In spite of the above study results, more improvement in pavement dynamic analysis is
needed in future research for example the assumption made in these 3D viscoelastic FE
model, such as tire-pavement contact stress being uniformly distributed on rectangular equiv-
alent areas is not indicative of realistic loading condition and may be improved in future
study.

Acknowledgements
The equipment and software were supported by Harbin Institute of Technology. The authors express their
sincere gratitude to all the people involved in this research project.

Disclosure statement
No potential conflict of interest was reported by the authors.
Road Materials and Pavement Design 29

Funding
This work was funded by the Chinese Ministry of Science and Technology under [grant number
2014BAC07B00] and the Natural Science Foundation of China under [grant number 51678207]. The
equipment and software were supported by Harbin Institute of Technology.

ORCID
Ogoubi Cyriaque Assogba http://orcid.org/0000-0002-1581-5187
Nonde Lushinga http://orcid.org/0000-0003-2522-8143

References
Ai, C., Rahman, A., Xiao, C., Yang, E., & Qiu, Y. (2017). Analysis of measured strain response of asphalt
pavements and relevant prediction models. International Journal of Pavement Engineering, 18(12),
1089–1097.
Al-Qadi, I. L., Elseifi, M. A., & Yoo, P. J. (2006). Viscoelastic modeling and field validation of flexible
pavements. Journal of Engineering Mechanics, 132(2), 172–178.
Al-Qadi, I., Wang, H., Yoo, P., & Dessouky, S. (2008). Dynamic analysis and in situ validation of
perpetual pavement response to vehicular loading. Transportation Research Record: Journal of the
Transportation Research Board, 2087(2087), 29–39.
Association, P. C. (1966). Thickness design for concrete pavements. Skokie, IL: Portland Cement
Association.
Association, P. C. (1984). Thickness design for concrete highway and street pavements. Skokie, IL: Portland
Cement Association.
Bathe, K.-J. (1982). Finite element procedures in engineering analysis. Engle wood Cliffs, NJ: Prentice-
Hall.
Beer, M. D. (1997). Measurement of tyre/pavement interface stresses under moving wheel loads. Paper pre-
sented at the (paper to) vehicle-road and vehicle-bridge interaction conference, 1994, Noordwijkerhout,
the Netherlands.
Beskou, N. D., Hatzigeorgiou, G. D., & Theodorakopoulos, D. D. (2016). Dynamic inelastic analysis of 3-
D flexible pavements under moving vehicles: A unified FEM treatment. Soil Dynamics & Earthquake
Engineering, 90, 420–431.
Blab, R., & Harvey, J. T. (2002). Modeling measured 3D tire contact stresses in a viscoelastic FE pavement
model. International Journal of Geomechanics, 2(3), 271–290.
Breuer, S., & Roseman, J. J. (1977). On Saint-Venant’s principle in three-dimensional nonlinear elasticity.
Archive for Rational Mechanics & Analysis, 63(2), 191–203.
Chailleux, E., Ramond, G., Such, C., & de La Roche, C. (2006). A mathematical-based master-curve con-
struction method applied to complex modulus of bituminous materials. Road Materials and Pavement
Design, 7(sup1), 75–92.
Chen, X., Zhang, J., & Wang, X. (2015). Full-scale field testing on a highway composite pavement dynamic
responses. Transportation Geotechnics, 4, 13–27.
Cooperation, O. f. E., & Group, D. S. E. (1992). Dynamic loading of pavements. Mathematical Models.
Darestani, M. Y., Thambiratnam, D. P., Baweja, D., & Nataatmadja, A. (2006). Dynamic response of
concrete pavements under vehicular loads. Paper presented at the IABSE Symposium Report.
de Araújo, P. C., Soares, J. B., de Holanda, ÁS, Parente, E., & Evangelista, F. (2010). Dynamic viscoelastic
analysis of asphalt pavements using a finite element formulation. Road Materials & Pavement Design,
11(2), 409–433.
Drakos, C., Roque, R., & Birgisson, B. (2001). Effects of measured tire contact stresses on near-surface
rutting. Transportation Research Record: Journal of the Transportation Research Board, 1764(1),
59–69.
Freeme, C. R., De Beer, M., & Viljoen, A. W. (1987). The behaviour and mechanistic design of asphalt
pavements. Sixth international conference, structural design of asphalt pavements, volume I, proceed-
ings, University of Michigan, July 13–17, 1987, Ann Arbor. Publication of Michigan University Ann
Arbor.
Gillespie, T. D. (1993). Effects of heavy-vehicle characteristics on pavement response and performance.
Washington, DC: Transportation Research Board.
30 O.C. Assogba et al.

Gopalakrishnan, K., & Thompson, M. R. (2006). Effect of dynamic aircraft gear loads on asphalt concrete
strain responses. Journal of Astm International, 3(8), 1–16.
Hadi, M. N., & Bodhinayake, B. (2003). Non-linear finite element analysis of flexible pavements. Advances
in Engineering Software, 34(11–12), 657–662.
Hernandez, J. A., Gamez, A., & Al-Qadi, I. L. (2016). Effect of wide-base tires on nationwide flexible pave-
ment systems: Numerical modeling. Transportation Research Record: Journal of the Transportation
Research Board, 2590(2590), 104–112.
Hibbit, D., Karlsson, B., & Sorensen, P. (2007). ABAQUS/standard analysis user’s manual. Providence,
RI: Hibbit, Karlsson, Sorensen Inc.
Hu, X., Faruk, A. N. M., Zhang, J., Souliman, M. I., & Walubita, L. F. (2017). Effects of tire inclination
(turning traffic) and dynamic loading on the pavement stress–strain responses using 3-D finite element
modeling. International Journal of Pavement Research & Technology, 10(4).
Huang, Y. H. (2003). Pavement analysis and design: United States edition. Upper Saddle River, NJ: Pearson
Schweiz Ag.
Institute, A. (1982). Research and development of the Asphalt Institute’s thickness design manual. Asphalt
Institute.
Lee, I.-W., Kim, D.-O., & Jung, G.-H. (1999). Natural frequency and mode shape sensitivities of damped
systems: Part I, distinct natural frequencies. Journal of Sound and Vibration, 223(3), 399–412.
Li, S., Guo, Z., & Yang, Y. (2015). Dynamic viscoelastic response of an instrumented asphalt pavement
under various axles with non-uniform stress distribution. Road Materials & Pavement Design, 17(2),
446–465.
Li, M., Wang, H., Xu, G., & Xie, P. (2017). Finite element modeling and parametric analysis of viscoelastic
and nonlinear pavement responses under dynamic FWD loading. Construction & Building Materials,
141, 23–35.
Liao, J., & Sargand, S. (2010). Viscoelastic FE modeling and verification of a US 30 perpetual pavement
test section. Road Materials and Pavement Design, 11(4), 993–1008.
Lu, Z., Hu, Z., Yao, H., & Liu, J. (2016). Field evaluation and analysis of road subgrade dynamic responses
under heavy duty vehicle. International Journal of Pavement Engineering, 1–10.
Medani, T., & Huurman, M. (2003). Constructing the stiffness master curves for asphaltic mixes. Delft, The
Netherlands: Delft University and Technology. Report, (7-01), 127–123.
Mulungye, R. M., Owende, P. M. O., & Mellon, K. (2007). Finite element modelling of flexible pavements
on soft soil subgrades. Materials & Design, 28(3), 739–756.
Newcomb, D. E., Buncher, M., & Huddleston, I. J. (2001). Concepts of perpetual pavements. Transporta-
tion Research Circular, 503, 4–11.
Park, S., & Kim, Y. (1999). Interconversion between relaxation modulus and creep compliance for
viscoelastic solids. Journal of Materials in Civil Engineering, 11(1), 76–82.
Pellinen, T. K., Witczak, M. W., & Bonaquist, R. F. (2004). Asphalt mix master curve construction using
sigmoidal fitting function with non-linear least squares optimization. In Recent advances in materials
characterization and modeling of pavement systems (pp. 83–101).
Quintus, V., & Harold, L. (2006). Application of the endurance limit premise in mechanistic: Empiri-
cal based pavement design procedures. Paper presented at the international conference on perpetual
pavements, Ohio, Applied Research Associates Press.
Reynaud, P., Nasr, S. B., Allou, F., Chaise, T., Nelias, D., & Petit, C. (2016). 3D modelling of tyre-pavement
contact pressure. Revue Française De Génie Civil, 21(6), 712–729.
Saad, B., Mitri, H., & Poorooshasb, H. (2005). Three-dimensional dynamic analysis of flexible conventional
pavement foundation. Journal of Transportation Engineering, 131(6), 460–469.
Sarkar, A. (2016). Numerical comparison of flexible pavement dynamic response under different axles.
International Journal of Pavement Engineering, 17(5), 377–387.
Sebaaly, P., & Tabatabaee, N. (1993). Influence of vehicle speed on dynamic loads and pavement response.
Transportation Research Record Journal of the Transportation Research Board, 1410, 107–114.
Sukumaran, B., Chamala, N., Willis, M., Davis, J., Jurewicz, S., & Kyatham, V. (2004). Three dimensional
finite element modeling of flexible pavements. Paper presented at the advances in pavement engineering.
Sun, Z., Xu, Y., Tan, Y., Zhang, L., Xu, H., & Meng, A. (2018). Investigation of sand mixture interlayer
reducing the thermal constraint strain in asphalt concrete overlay. Construction & Building Materials,
171, 357–366.
Tan, Y., Sun, Z., Gong, X., Xu, H., Zhang, L., & Bi, Y. (2017). Design parameter of low-temperature
performance for asphalt mixtures in cold regions. Construction & Building Materials, 155, 1179–1187.
Road Materials and Pavement Design 31

Tautou, R., Picoux, B., & Petit, C. (2017). Temperature influence in a dynamic viscoelastic modeling of a
pavement structure. Journal of Transportation Engineering, Part B: Pavements, 143(3), 04017012.
Uddin, W. (2003). Effects of FWD load-time history on dynamic response analysis of asphalt pavement.
Paper presented at the pavement performance data analysis forum in conjunction with MAIREPAV03,
Guimaraes, Portugal.
Wang, H., & Al-Qadi, I. (2009). Combined effect of moving wheel loading and three-dimensional con-
tact stresses on perpetual pavement responses. Transportation Research Record: Journal of the
Transportation Research Board, 2095(2095), 53–61.
Wang, B. S., Liu, J. Q., & Pang, P. F. (2013). Three-dimensional finite element analysis of dynamic response
of the stress in pavement structure under overload. Paper presented at the advanced materials research.
Wang, G., Roque, R., & Morian, D. (2011). Evaluation of near-surface stress States in asphalt concrete
pavement three-dimensional tire-pavement contact model. Transportation Research Record: Journal
of the Transportation Research Board, 2227(1), 119–128.
Williams, M. L., Landel, R. F., & Ferry, J. D. (1955). The temperature dependence of relaxation mechanisms
in amorphous polymers and other glass-forming liquids. Journal of the American Chemical Society,
77(14), 3701–3707.
Wu, C. P., & Shen, P. A. (1996). Dynamic analysis of concrete pavements subjected to moving loads.
Journal of Transportation Engineering, 122(5), 367–373.
Xia, K., & Yang, Y. (2012). Three-dimensional finite element modeling of tire/ground interaction.
International Journal for Numerical & Analytical Methods in Geomechanics, 36(4), 498–516.
Yoder, E. J., & Witczak, M. W. (1975). Principles of pavement design. Hoboken, NJ: Wiley.
Yoo, P. J., & Al-Qadi, I. L. (2007). Effect of transient dynamic loading on flexible pavements. Transporta-
tion Research Record: Journal of the Transportation Research Board, 1990(1), 129–140.
Yoo, P. J., & Al-Qadi, I. L. (2008). The truth and myth of fatigue cracking potential in hot-mix
asphalt: Numerical analysis and validation. Asphalt Paving Technology: Association of Asphalt Paving
Technologists-Proceedings of the Technical Sessions, 77, 549–590.
Zhao, Y., Tan, Y., & Zhou, C. (2012). Determination of axle load spectra based on percentage of overloaded
trucks for mechanistic-empirical pavement design. Road Materials & Pavement Design, 13(4), 850–
863.
Zhong, X. G., Zeng, X., & Rose, J. G. (2002). Shear modulus and damping ratio of rubber-modified asphalt
mixes and unsaturated subgrade soils. Journal of Materials in Civil Engineering, 14(6), 496–502.

You might also like