Вы находитесь на странице: 1из 25

Construction and Building Materials 235 (2020) 117406

Contents lists available at ScienceDirect

Construction and Building Materials


journal homepage: www.elsevier.com/locate/conbuildmat

Numerical investigation of the mechanical response of semi-rigid base


asphalt pavement under traffic load and nonlinear temperature gradient
effect
Ogoubi Cyriaque Assogba a,⇑, Yiqiu Tan a,b,⇑, Xingye Zhou a,c, Chao Zhang a, Josita Neurly Anato d
a
School of Transportation Science and Engineering, Harbin Institute of Technology, Harbin 150090, People’s Republic of China
b
State Key Laboratory of Urban Water Resource and Environment, Harbin Institute of Technology, Harbin 150090, People’s Republic of China
c
Key Laboratory of Road Structure and Material of Transport Industry, Research Institute of Highway Ministry of Transport, NO. 8 Xi Tu Cheng Road, Beijing 100088, People’s Republic
of China
d
School of Civil Engineering, Harbin Institute of Technology, Harbin 150090, People’s Republic of China

h i g h l i g h t s

 Three pavement structures with typical functional and structural requirements, specifically designed to withstand various distresses of semi-rigid
pavements were presented.
 This study provides evidence that non-linear temperature gradient is a key parameter for accurate prediction of stresses and strains distribution in
pavement system.
 This study provides a reference for the design of more durable semi-rigid base asphalt pavement structures.

a r t i c l e i n f o a b s t r a c t

Article history: The effectiveness of any pavement design based on the mechanistic procedure depends on the accuracy
Received 5 September 2019 of the mechanical parameters used such as stresses and strains. Hence, an effective and realistic predic-
Received in revised form 22 October 2019 tion of these parameters when traffic and environmental data are provided is a step towards the future
Accepted 26 October 2019
design of more sustainable pavement structures. This study investigated the distribution of the mechan-
Available online 14 November 2019
ical parameters in three new semi-rigid pavement structures with typical functional and structural
requirements, specially designed to withstand various distresses of semi-rigid pavements. The response
Keywords:
of these pavements to the combined effect of the nonlinear thermal gradient and moving axle load was
Semi-rigid base asphalt pavement
Thermal modeling
examined by means of an advanced 3D pavement FE modeling. The model integrated mechanical and
Numerical analysis thermophysical properties of each pavement material, studied the transient dynamic effect of wheels
Pavement mechanical response load, and employed transient heat transfer and implicit dynamic analysis. Field measurements of these
Pavement temperature experimental pavement structures at the RIOH-Track test site in Beijing, China, as well as relative climatic
Nonlinear temperature gradient data were used to verify and validate the accuracy of the developed model. Comparison of the numerical
Heat transfer and experimental results indicated that the developed model predicted accurately the characteristic non-
Full-scale pavement test track (RIOH-Track) linear distribution of temperature along with the pavement depth and the pavement response to the
combined effect of the nonlinear temperature gradient and traffic load. The influence of three main fac-
tors affecting the distribution of the mechanical parameters, namely; the bonding conditions between the
pavement’s layers, the vehicle speed and the axle load amplitude were evaluated. In addition, fatigue
analysis of the asphalt layers and the semi-rigid layers was performed to evaluate the theoretical service
life of each pavement structure. The findings of this study underscored that the non-linear temperature
gradient, despite being often neglected in mechanistic modeling, is a key parameter for accurate predic-
tion of the distribution of stresses and strains in pavement system. Moreover, the results indicate that the
lowest traffic speed and the extra weight extension of the truck axle have an adverse effect on mechanical
parameters. Furthermore, the improvement of contact conditions at the interface results in good perfor-
mance. Overall, the conclusions obtained in this study provides a reference for the design of semi-rigid
pavement structures.
Ó 2019 Elsevier Ltd. All rights reserved.

⇑ Corresponding authors.
E-mail addresses: assocyr@hotmail.fr, assocyr@stu.hit.edu.cn (O.C. Assogba), tanyiqiu@hit.edu.cn (Y. Tan).

https://doi.org/10.1016/j.conbuildmat.2019.117406
0950-0618/Ó 2019 Elsevier Ltd. All rights reserved.
2 O.C. Assogba et al. / Construction and Building Materials 235 (2020) 117406

1. Introduction formation of the different forms of distress exhibited by the newly


proposed semi-rigid base asphalt pavements.
1.1. Background
1.2. Objective and scope
The mechanistic-empirical design approach that combines the
mechanistic response of a pavement structure with an estimate The present study aims to examine the distribution of
of the damage that it will sustain over time is widely used around mechanical parameters in three new semi-rigid pavement struc-
the world. An accurate estimate of the mechanical response field tures with typical functional and structural requirements, spe-
(stress, strain and deformation fields) of the pavement is a prereq- cially designed to withstand various distresses of conventional
uisite for an effective and realistic prediction of pavement perfor- semi-rigid pavements. In this regard, this study presents an
mance over the design lifetime. The layered elastic theory (LET), experimental validation of an advanced 3D pavement FE model
originally developed by Burmister et al, is commonly used to com- using measurements performed on these experimental pavement
pute the pavement response to truck loading [1]. Although layered structures located at the RIOH track test site in Beijing, China, as
elastic analysis is relatively simple and faster, it impedes the incor- well as relative climatic data. First, based on the principle of heat
poration of real tire-pavement loading condition and effective transfer and meteorological data, the prediction model of the
characterization of materials in pavement analysis [2]. temperature field of pavements was established. The model was
With advances in computer technology and algorithmic validated and used for preliminary analysis of the distribution
improvements, the use of the finite element method (FEM) has and the influence of temperature on each pavement. Finally,
become widespread in the mechanistic analysis of pavement the corresponding temperature profile was integrated into the
response. This is due to the fact that FEM considers the variables mechanistic analysis model for a realistic simulation of the
omitted in LET such as; dynamic traffic load simulation, non- mechanical response of the pavement subjected to the combined
uniform tire-pavement contact stresses, a realistic interface model effect of the nonlinear thermal gradient and moving axle load.
for the simulation of inter-layer bonding conditions, properties of The influence of three main factors affecting the distribution of
viscoelastic and nonlinear materials, discontinuities in pavements, the mechanistic parameters namely; bonding conditions between
etc. The application of finite element techniques makes it possible the pavement’s layers, speed of the vehicle and amplitude of the
to take these variables into account and to provide more realistic load on the axles were evaluated. Furthermore, fatigue analysis of
results as well as help pavement researchers to understand com- the asphalt layers and the semi-rigid layers of the three new
plex pavement behavior under moving dynamic wheel loading. semi-rigid pavement structures was performed to compare their
A large body of studies have utilised FEM to analyze pavement theoretical service life.
structure. This method can be applied to the mechanistic analysis
of pavement responses, including: modeling the dynamic behavior 2. Numerical modelling
of flexible pavement subjected to dynamic solicitation [3],
investigating the effect of vehicle speed and overload on the 2.1. Thermal analysis
dynamic response of a semi-rigid base asphalt pavement and its
fatigue life [4], characterization of rutting behavior of the asphalt The state of stress and strain of the pavement is strongly influ-
mixes [5,6], understand the influence of non-linear characteriza- enced by the temperature profile. Thus, the distribution of the tem-
tion of unbound materials in pavement response [7,8], assessing perature field in the asphalt pavement has an inherent link with
the impact of wide-base tire and dual tire on pavement perfor- the mechanical behavior and service performance of the latter.
mance [9,10] are just some examples of the main applications of Therefore, the distribution characteristics and the variation of the
FEM analysis on pavements. temperature profile in the pavement system at different points in
In most of the previous FE models that studied the mechanistic time are necessary for a better prediction of the mechanical
response of pavement structures, it was often assumed that the response of semi-rigid base asphalt pavement under traffic load.
temperature field in the pavement structure is constant, thus, gen- The thermal analysis makes it possible to determine the tempera-
erating linear temperature gradients over the entire depth of the ture field in the pavement structure, and to provide both thermal
pavement structure [11–14]. This assumption is rather incompati- strains and actual rheological behavior of the asphaltic layer for
ble with the actual thermal environmental conditions to which subsequent mechanical analysis. The theory of thermal analysis
pavements are exposed. Moreover, studies on temperature field is mainly based on the principle of heat transfer and meteorology.
in road system have elucidated this inconsistency. By showing that Fig. 1 provides a schematic diagram illustrating the process of heat
the thickness of a pavement can be subjected to different nonlinear transfer in an asphalt pavement structure. As shown in Fig. 1, the
thermal gradient depending on the seasons, solar radiation and surface of the pavement is simultaneously subjected to convective
climate history [15–17]. Thus, distribution of the temperature and radiant exchanges with constantly changing ambient temper-
inside the structure also evolves according to instant daily atures and to the incident solar flow, whose evolutions are
measurements. obtained by means of local meteorological records. Theoretical
The flexible layer of semi-rigid pavements is generally made of principles and basic equations governing the main means of heat
asphalt mixture, which is a visco-elastic material whose behavior transfer in an asphalt pavement system are presented in the subse-
is strongly related to temperature. Thus, the mechanical properties quent subsections.
and performance of the asphalt layer are significantly influenced
by pavement temperature. This in return, affects the stiffness of 2.1.1. Transient conduction
the pavement profile and the state of stress and strain in the pave- The problem of conductive heat transfer within the pavement
ment. Therefore, an accurate prediction of the temperature field in system can be solved by combining the first law of thermodynam-
the road system is thus important to assess the mechanical ics, which stipulates that thermal energy is conservative, and the
response of the structure, when subjected to traffic and tempera- Fourier law relative to the heat flux with thermal gradient. Under
ture. On the other hand, an effective and realistic prediction of constant thermal conductivity, the heat transfer for an isotropic
the distribution of mechanistic parameters, such as stress and medium is governed by the following differential equation
strain, will also help to understand in depth the mechanisms of [Eqs. (1) and (2)] [18,19]:
O.C. Assogba et al. / Construction and Building Materials 235 (2020) 117406 3

Fig. 1. Heat transfer process in an asphalt pavement structure.

 
1 @T where:
r2 T ¼  ð1Þ
a @t
qo = maximum solar radiation at midday, w=m2 ;
with:
qd = total daily solar radiation, w=m2 ;
k c = actual effective sunshine hours in a day, h
a¼ ð2Þ
qC w = angular frequency, rad.
where:
The formula above [Eq. (4)] is defined on a sequence of inter-
a = thermal diffusivity; vals, which is neither smooth nor continuous. When calculating
k = thermal conductivity, W=m: C; the temperature field, the discontinuities of jumps will appear.
q = density, kg=m3 ; Therefore, it is necessary to extend Eq. (4) as a series to obtain a
C = Specific heat, j=kg= C; smooth and continuous function expression. According to the cor-
t = time, s. relation principle of the Fourier series, Eq.4 can be extended to the
Fourier series as a cosine function, that is, formula [Eqs. (8)–(10)]
In a cartesian coordinate system, the three-dimensional tran- computes the k-order up to 30, which can meet engineering preci-
sient heat conduction under normal condition is governed by the sion requirements.
following expression [Eq. (3)]
! a0 X 1
kpðt  12Þ
qðtÞ ¼ þ ak cos ð8Þ
@T @2T @2T @2T 2 12
qc ¼ k þ þ ð3Þ k¼1
@t @x2 @y2 @z2
with:
where: x, y, and z are the coordinates of the point, and t is the time.
2qo
a0 ¼ ð9Þ
mp
2.1.2. Direct solar radiation
Direct solar radiation is highly dependent on the time of the 8 h i
>
< qpo ¼ mþk
1
sinðm þ kÞ 2m
p þ p k¼m
day, atmospheric conditions, and the angle of incidence of sun’s 2m
ak ¼ h i
rays on the pavement surface. According to the results of Barber >
: qpo ¼ mþk
1
sinðm þ kÞ 2m
p þ 1 sinðm  kÞ p k–m
mk 2m
et. al and Yan et. al, the daily variation in solar radiation q(t) can
be approximated by the following functions [Eqs. (4)–(7)] [20,21]. ð10Þ
8
>
<0 0  t  12  2c
2.1.3. Pavement surface effective radiation
qðtÞ ¼ q0 cosmwðt  12Þ 12  2c  t  12 þ 2c ð4Þ The effective radiation of pavement is mainly related to many
>
:
0 12  2c  t  24 factors, such as soil temperature, air temperature, air humidity,
and transparency. Most previous studies have estimated the effect
with:
of heat release from the effective radiation on the road surface by
q0 ¼ 0:132mqd ð5Þ appropriately modifying the coefficient of heat exchange on the
road surface to correct the temperature or reduce the magnitude
12 of solar radiation. However, this method has large errors and the
m¼ ð6Þ
c boundary conditions for effective radiation can be directly realized
by the following formula [Eq. (11)].
2p h i
x¼ ð7Þ
qF ¼ er ðT 1 jZ¼0  T Z Þ4  ðT a  T z Þ4 ð11Þ
24
4 O.C. Assogba et al. / Construction and Building Materials 235 (2020) 117406

where: The stress tensor can be defined using the traditional Hooke’s
law [Eq. (17)]
qF = effective radiation of pavement surface, W/m2°C;
e = Pavement emissivity; r ¼ krU  I þ lðrU þ rU T Þ ð17Þ
r = Stefan-Boltzmann constant (blackbody radiation coeffi- where:
cient), W=ðm2  K4 Þ;
T Z jz ¼ 0 = Pavement surface Temperature, °C; I = Identity matrix;
T a = Atmospheric temperature, °C; k,l = lame’s coefficients;
T Z = Absolute zero value, °C; U = displacement vector.

2.1.4. Ambient temperature and convective heat transfer The discretization of the governing equation of the dynamic
Due to the influence of solar radiation, the atmospheric temper- response of a pavement system results in the following form
ature has periodic variation characteristics. The minimum daily [Eq. (18)]
temperature usually appears around dawn, about 4–6 am, and
the daily maximum temperature generally occur about 2 h after € g þ ½Cfug
½Mfu _ þ ½Kfug ¼ F ðtÞ ð18Þ
maximum solar radiation (14 p.m.). Thus, it takes more than 14 h where:
to rise from the lowest temperature to the highest temperature,
while it takes less than 10 h to decrease from the highest temper- ½M = mass matrix;
ature to the lowest temperature; a single sinusoidal function can- ½C = damping matrix;
not accurately simulate the actual temperature change process. ½K = stiffness matrix;
Therefore, the linear combination of two sinusoidal functions _ = velocity vector;
fug
[Eqs. (12)–(14)] was used to simulate the diurnal variation in € g = acceleration related to the nodes;
fu
temperature. _ = displacement vector;
fug

T a ¼ T a þ T m b0:96sinx ðt  t0 Þ þ 0:14sinx ðt  t 0 Þc ð12Þ F ðtÞ = external force vector related to the structure dynamic
system.
with:
 1  Max  In general, implicit or explicit integration method is used to
Ta ¼ T a þ T Min
a ð13Þ solve the dynamic response-related problems in the finite element
2
method. The implicit direct integration dynamic analysis was used
 1  Max  to investigate the mechanical response of the instrumented
Tm ¼ T a  T Min
a ð14Þ semi-rigid base asphalt pavement. This method is based on
2
Hilbert-Hughes-Taylor’s time integration method and the New-
where:
mark integration schema [22]. The use of an implicit method is
 generally more efficient for a structural dynamics problem such
T a = average daily temperature, °C; as pavement dynamic response because its integration is uncondi-
T m = diurnal temperature variation, °C; tionally stable and allows a better numerical convergence than an
T Max
a = maximum daily temperature, °C; explicit method. However, this type of integration leads to an
T Min = minimum daily temperature, °C; impractical time step and, consequently, to computationally
a
t0 = Initial phase, and t = time, h. expensive analysis.

The heat exchange between the surface of the roadway and the 2.2.2. Modelling of material damping
atmosphere is mainly affected by the wind speed, the linear rela- Structural damping is one of the factors that seriously affects
tion between them can be expressed as follows [Eq. (15)]: the dynamic behaviors of the pavement system. Thus, for realistic
hc ¼ 3:7v w þ 9:4 ð15Þ simulation of asphalt pavement, it is essential to adopt realistic
structural damping in the nonlinear dynamic analysis. The source
where: of energy dissipation present in the pavement material can be
modeled using damping ratios. The modal damping method, the
hc = Surface coefficient of heat transfer, W=ðm2  CÞ; stiffness factor method, the stress-energy factor, and the Rayleigh
v w = Average daily wind velocity, m=s. damping matrix are the most commonly used for calculating the
damping ratio. In this analysis, the Rayleigh damping matrix
2.2. Mechanistic analysis expressed below [Eq. (19)] were used to define the damping rate
[Eq. (20)].
2.2.1. Transient dynamic analysis
Using the virtual work principle applied to a domain X and con- C ¼ a½M  þ b½K  ð19Þ
sidering small strains, the characteristic dynamic equations of a
a bw
solid consist of elastic and linear isotropic materials can be written n¼ þ ð20Þ
as [Eq. (16)]: 2w 2
! ! ! where:
div r þ f ¼ q c ð16Þ
where: a = mass damping coefficient;
b = stiffness damping coefficient.
r = Cauchy stress tensor;
!
f = denote the applied force; The mass and stiffness proportional damping coefficient can be
q = density; obtained by the vibration modes and the corresponding specific
!
c = acceleration vector. modal damping ratio, as indicated below [Eq. (21)]
O.C. Assogba et al. / Construction and Building Materials 235 (2020) 117406 5

  " #
a wi wj wj wi  ni  chosen to be 15 m  20 m  25 m, where the x-axis indicates the
¼2 ð21Þ
w2j  w2i  wj transverse direction; the y-axis denotes the depth and the z-axis
1 1
b wj nj
represents the traffic direction.
where: Previous research indicated that the consistency of the mesh
size has some influence on the predicted result by the FE model.
th th Thus, in order to increase the accuracy of the model, relatively fine
wi and wj = vibration frequencies at the i andj modes,
mesh size was adopted along the truck wheel path, where the
respectively;
th th
stresses and displacement are high. Dense size mesh was used near
fi and fj = specific damping ratio at the i andj modes, the truck well path, while a relatively coarse mesh size was
respectively. adopted far away from the loading area. The in-plane Geometry
and the 3-D view of the established pavement FE model are given
In a certain frequency range, the damping rate can be approxi- in Fig. 2[(a) and (b)]. In order to enhance the convergence rate and
mated as constant and the damping coefficients have the same computational efficiency, two distinct mesh element types have
critical damping ratio. Thus, the mass and stiffness proportional been defined for the FE model in this study. An 8-node linear heat
damping can be expressed with the following simplified form transfer brick (DC3D8) was assigned as the mesh element type for
[Eq. (22)]: thermal analysis, while the Continuum 3-D 8-node Reduced inte-
    gration element (C3D8R) was adopted for mechanical analysis.
a 2n wi wj
¼ ð22Þ
b wi þ wj 1
3.1.2. Initial and boundary condition
where: To avoid an adverse effect of the artificial boundary on the accu-
racy of numerical simulation results, it is imperative to adopt
w1 and w2 = natural frequencies defined from modal analysis; appropriate and adequate boundary conditions [4]. Consequently,
f = critical damping ratio for a given i order. during the thermal analysis, the bottom of the 3D FE model was
set as outflow. The four sides of the model were assumed to be adi-
The first natural frequencies of the structure are often used to abatic to allow for sufficiently large horizontal expansion. Note
determine the Rayleigh damping coefficients. Moreover, to avoid that temperature changes are much greater in the vertical direc-
over-damping the system, research suggests using the shape of tion than in the horizontal direction. Thus, the horizontal heat
the 10th node first to determine the natural frequency of the transfer across the four sides can be neglected. The initial temper-
model. The lowest and highest load frequencies are often consid- ature field of the entire model was determined by interpolating the
ered as critical frequencies in pavement structure analysis. Exper- field monitoring data.
imentally, the critical damping rate of asphalt pavement was found As for the mechanical analysis, it was assumed that all four
to range from 2 to 5% [23,24]. A critical damping coefficient of 5% sides of the model are constrained by any perpendicular move-
has been adopted in this numerical simulation; thus, the mass ment to the model side, the displacements of the x-axis and the
damping coefficient and the stiffness damping coefficient were y-axis of the longitudinal and transverse sides, respectively, are
set at 0.342 and 6.90 E3, respectively. constrained to be zero [27–29]. In addition, the bottom of the
model was constrained to zero in the x, y, and z directions so that
there was no vertical or horizontal movement at the bottom of the
3. Pavement structure and material
natural ground [4].
3.1. Modelling of the 3D nonlinear FE
3.1.3. Interface model
3.1.1. Geometry and meshing The bonding condition between pavement layers could exert a
Considering three typical test sections of the full-scale RIOH- considerable influence on the pavement structural response.
Track pavement test [25,26] as a prototype, a 3-D nonlinear FE Therefore, the model used to define the bonding conditions at the
model of the pavement system was developed using the commer- layers interface is another important parameter for efficient pave-
cial FE software, ABAQUS. Table 1 presents the thickness and the ment simulation. As the investigated semi-rigid pavement struc-
materials constituting the layers of each semi-rigid asphalt pave- tures were newly built, the upper, middle and bottom layers of
ment structure that were investigated. Three main issues, namely asphalt concrete layer were assumed to be completely continuous
1) the lateral size of the model, 2) the depth of the model and 3) (perfectly bonded) to each other in the developed model. Likewise,
the size of the mesh; were addressed during the development of a fully bonded interfacial condition has been adopted for the inter-
the pavement FE model. Existing 3D FE models in literature were face between the cement stabilized graded crushed stone base and
considered as reference and sensitivity analysis of the mesh was subbase, as well as the interface between the cement stabilized
performed. Then, the full-size FE model dimension (x, y, z) was subgrade and the natural soil. On the other hand, a simple friction

Table 1
Pavement Structures and material composition investigated in this study.

N° Layer Experiment Structure-A (STR-A) Experiment Structure-B (STR-B) Experiment Structure-C (STR-C)
1 Asphalt Concrete upper layer 4 cm SBS1-AC13 4 cm SBS1-AC13 4 cm SBS3-PAC13
3 Asphalt Concrete middle layer 10 cm A30-AC25 6 cm SBS2-AC20 6 cm SBS2-AC20
5 Asphalt Concrete bottom layer 2 cm SBS1-AC10 8 cm A70-AC25 8 cm A70-AC25
7 Semi-rigid base 20 cm CBG-A 18 cm CBG-A 18 cm CBG-B
8 Semi-riigid Subbase 20 cm CBG-A 20 cm CBG-A 20 cm CBG-B
9 Cement stabilized Subgrade 20 cm CS 20 cm CS 20 cm CS
10 Natural Soil (In place soil) Infinite Infinite Infinite
Total thickness of Asphalt Concrete structure layer 16 cm 18 cm 18 cm

Note: For pavement material abbreviation referred to the subsection ‘‘Test section and materials properties”.
6 O.C. Assogba et al. / Construction and Building Materials 235 (2020) 117406

2.5
3.885

10
2.23
3.885

2.5
2.5 25 2.5
a) In plane geometry

Left side Boundary U(x)=0


U(z)=0
condition UR(y)=UR(z)=0
UR(x)=UR(y)=0
Front Boundary
condition

20 m

U(x)=0
UR(y)=UR(z)=0
Right side Boundary
condition

Notation for
Boundary condition
Displacement
U(z)=0 U(x)
UR(x)=UR(y)=0 Direction
Rear Boundary
condition U(x)=U(y)=U(z)=0 Rotation
UR(x)=UR(y)=UR(y)=0 UR(x)
Bottom Bondary condition Direction

b) 3D View of Finite Element

Fig. 2. Finite element mesh of the pavement structure.

model (coulomb model) with a friction coefficient (l) was used to is a transient temperature field. As a pavement material, the
simulate the interface bonding conditions between the asphalt viscoelasticity of asphalt mixture is very significant, and the prop-
layer and the semi-rigid base, as well as the interface between erties of the material are greatly affected by temperature. There-
the subbase and subgrade. In this model, two friction coefficient fore, for a realistic simulation of the actual mechanical response
values of 0 (full slip) and 1 (frictionless) were used to define inter- of the pavement, the actual temperature field of the pavement
layer conditions. Normal and tangential contact properties were structure must be integrated. Thus, a steady-state heat transfer
incorporated to correctly capture the mechanical behavior of the analysis was firstly conducted to obtain the initial temperature of
interface. Thus, the normal behavior of the load transfer system the pavement structure, then transient heat transfer analysis was
was modeled using a hard contact pressure overlay between the performed to determine the relevant temperature profile. FILM
surfaces. The tangential behavior of the interface between the and FLUX user subroutines were coded via a Fortran computer
two layers was modeled using Coulomb friction. Note that no shear program to predict the temperature distribution in the pavement
stress limit was included and separation between the surface was layers. In this respect, the ambient temperature, solar radiation,
allowed after contact. pavement surface effective radiation and transient conduction
predefined in the previous section ‘‘2-1 Thermal analysis” were
3.1.4. Thermal loading considered. The meteorological parameters were collected from
Under the influence of environmental factors, the actual tem- meteorological stations installed at the RIOH-Track test site. For
perature within the pavement structure changes with the time of more details, refer to sub-sections ‘‘3-2-1 Thermophysical Proper-
the day and the depth of the pavement, and its temperature field ties” and ‘‘4-1 In Situ Measurements”.
O.C. Assogba et al. / Construction and Building Materials 235 (2020) 117406 7

3.1.5. Dynamic traffic loading axles, the measured tire-pavement contact area and the contact
The loads on the wheels of the vehicle in motion applied to the pressure under each load configuration are summarized in
pavement system at a given moving speed are undeniably Tables 3.
dynamic. Thus, the magnitude of the load and the position of the
load on the wheels are time-dependent and can be considered 3.2. Materials mechanical and thermophysical parameters
equal to the sum of the static load and a constantly changing load.
In this regard, a user subroutine DLOAD was coded via a Fortran 3.2.1. Thermophysical properties
computer program to integrate the dynamic load and its contact Previous studies indicated that the thermophysical parameters
stress effect on the pavement into the implicit dynamic analysis. of pavement materials have a significant influence on the temper-
The DLOAD subroutine specifies the variation of the amplitude of ature profile of the pavement structure. The thermophysical
the distributed load (F) as a function of time (TIME *) in a specific parameters of a multilayer asphalt pavement are related to the
coordinate system (COORDS *). Note that the truck wheel load con- composition of materials, moisture, and porosity, which affects
figurations, the tire-pavement surface contact, and the associated the properties and mechanical behavior of materials. Thus, in addi-
contact stress were required parameters for modeling the transient tion to environmental and meteorological factors, the thermal and
dynamic load of the vehicle in motion. Stery-King 6  4 Dump physical properties of the materials must be carefully defined.
Truck which includes a steer axle and a tandem drive axle was Asphalt mixture is an extremely temperature sensitive material,
used as controlled load truck. The detailed parameters of the latter however, it has been found that its thermal properties remain
are summarized in Table 2. The contact area between each tire and stable regardless of temperature variation. Consequently, the ther-
the pavement surface was measured and recorded, as shown in mophysical parameters of the pavement structure were defined as
Fig. 3[(a) and (b)]. For more simplicity, it was assumed that the constant in this numerical simulation. The parameters were
tire-road contact area induced by each wheel load (Table 2) is selected to provide a good correspondence between the field
rectangular with an equivalent area Ac. The length and width of observed and predicted pavement temperatures. The adopted ther-
the converted rectangle were 0.8712 L and 0.6 L, respectively, mophysical parameters presented in Table 4 are based on the Chi-
and can be inferred from the use of the following expressions nese specifications for the design of highway asphalt pavement
[Eqs. (23) and (24)] [30]: (JTG D50-2017).

3.2.2. Mechanical properties


Ac ¼ pð0:3LÞ2 þ ð0:4LÞð0:6LÞ ¼ 0:5227L2 ð23Þ
The mechanical properties used to characterize the materials of
the pavement structure were defined in two ways. In the first
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi stage, where the pavement temperature profiles are generated, it
Ac pffiffiffiffiffi
L¼ ¼ 1:3832 Ac ð24Þ was considered that the asphalt materials exhibited a linear ther-
0:5227
moelastic behavior defined by the stiffness modulus, which
The weighbridge was also used to obtain the load distribution depends on the temperature. The elastic properties assigned for
data on each axle of the truck under different loading conditions. each asphalt mixtures during that stage are listed in Table 7. In
The truck weighing method consisted of weighing the truck’s total the second stage, where both temperature and traffic loading were
load and weighing each axle of the truck. The actual weights of the applied, the asphalt layers were considered as thermo-rheological
simple material with linear viscoelastic behavior. The constitutive
equation of such a material can be expressed as per the following
Table 2
Controlled-load test truck dimensions (Story-King 6  4 Dump Truck).
form [Eqs. (25)–(28)] [31]:
Z Z
n
@ ekl ðxk ; n0 Þ 0 n
@hkl ðxk ; n0 Þ 0
Dimensions (mm) Value
rij ðxk ; nÞ ¼ cijkl ðn  n0 Þ dn  bij ðn  n0 Þ dn
Wheelbase 5175 0 @n0 0 @n0
Distance between the front axle and the rear first axle 3825 ð25Þ
Distance between the rear first axle and the rear second axle 1350
Front-wheel track 2022 with:
Rear-wheel track 1830 Z t
Spacing between rear tires 350 1
n ¼ nðt Þ ds ð26Þ
0 aT

Fig. 3. Tire-pavement contact area: (a) Tire Footprint Measurement and (b) Measured Footprint.
8 O.C. Assogba et al. / Construction and Building Materials 235 (2020) 117406

Table 3
Test truck axles weight, measured contact surface and pressure under various loading condition.

Side Tire contact Surface (cm2) Contact Pressure (MPa) Axle Weight (kN)
Left Right Left Right —
Test truck axles weight, measured contact surface and pressure under Empty truck load
Tire position 1 2 1 2 —
Front axle 276 238 0.98 1.13 54.80
Tire position 1 2 1 2 1 2 1 2 —
Rear first axle 165 171 103 121 0.70 0.67 1.12 0.95 46.83
Rear second axle 274 274 239 267 0.42 0.42 0.48 0.43 46.83
Test truck axles weight, measured contact surface and pressure under full loading condition
Tire position 1 2 1 2 —
Front axle 447 443 0.83 0.86 75.53
Tire position 1 2 1 2 1 2 1 2 —
Rear first axle 321 331 292 310 0.78 0.76 0.86 0.81 101.93
Rear second axle 462 462 334 337 0.54 0.54 0.75 0.75 101.93
Test truck axles weight, measured contact surface and pressure under overloading condition
Tire position 1 2 1 2 —
Front axle 459 375 0.85 1.04 78.61
Tire position 1 2 1 2 1 2 1 2 —
Rear first axle 387 322 267 300 0.86 1.07 1.29 1.15 140.19
Rear second axle 448 448 386 383 0.77 0.77 0.89 0.90 140.19

Table 4
Thermophysical and mechanical parameters of pavement materials for temperature analysis.

Thermal conductivity Specific heat capacity Coefficient of Density Poisson’s


(J/m h °C) (J/kg.°C) Expansion (m3/m) (kg/m3) ratio
Asphalt Mixture
SBS1-AC13 5240 1086 2e-5 2345 0.25
SBS3-PAC13 4672 926 2e-5 2148 0.25
A30-AC25 4325 984 2e-5 2437 0.25
SBS2-AC20 4586 976 2e-5 2430 0.25
SBS1-AC10 4880 1168 2e-5 2350 0.25
A70-AC25 4684 994 2e-5 2433 0.25
Non-asphaltic layer
Base 5680 946 1e-5 2428 0.20
Subbase 5680 946 1e-5 2428 0.20
Subgrade 5562 1040 25e-5 1898 0.25
Natural Soil 4050 1012 55e-5 1780 0.35
Solar absorptance 0.90 (Summer)
0.80 (Spring)
Pavement surface Emissivity 0.85
Absolute zero temperature (Tz/°C) 273
Stefan Boltzmann constant (J/h m2 K4) 2.041  10-4

Z t0
1 [Eqs. (29) and (30)], using the kernel functions of the generalized
n0 ¼ nðt 0 Þ ds ð27Þ Maxwell elements [32].
0 aT
 

aT ¼ aT ½TðsÞ X
nG  n
ð28Þ GðnÞ ¼ G0 þ Gi e
kG
i ð29Þ
aT ¼ aT ½hðsÞ
i¼1

where:  
X
nk  n
kK
aT = time-temperature shift factor; K ðnÞ ¼ K 0 þ Kie i ð30Þ
s = integration variable; i¼1
t and t0 = time moment;
The kernel functions are expressed in terms of the Prony-
rij and ekl = stress and strain tensor, respectively; Dirichlet series which is an efficient analysis algorithm, its formu-
xk = function of position lation is indicated by the following equation [Eqs. (31) and (32)]:
bij = second-order tensor of relaxation moduli relating stress to
"   !#
thermal strain; X
N
st
C ijkl = fourth-order tensor of relaxation moduli relating stress to GðtÞ ¼ G0 1  Gi 1  e i
ð31Þ
mechanical strain. i¼1

"   !#
The viscoelastic characteristic of the asphalt mixture with X
N
st
regards to relaxation used in numerical simulation through the K ðt Þ ¼ K 0 1  Ki 1  e i
ð32Þ
i¼1
finite element method can be expressed in hereditary integral form
O.C. Assogba et al. / Construction and Building Materials 235 (2020) 117406 9

where: soil and the natural soil in-place by way of back-calculation.


The elastic moduli of non-asphaltic layers obtained by back-
N = number of Prony Dirichlet series terms; calculation and laboratory experiments are presented in Table 7.
G0 and K 0 = instantaneous shear modulus and bulk modulus,
respectively;
4. Numerical model validation
Gi , K i and si = Prony Dirichlet series coefficients.

Although efforts were made to approximate the actual pave-


The Williams Landell Ferry function (WLF) presented in Eq. (33)
ment conditions in the developed 3-D FE model based on labora-
was used to model the time-temperature superposition of asphalt
tory results (material characterization), environmental conditions
concrete materials [33].
(temperature profile), and modeling techniques, certain assump-
C 1 ðT  T ref Þ tions were unavoidable. Therefore, this section is dedicated to
logðaT Þ ¼ ð33Þ
C 2 þ ðT  T ref Þ the validation of the model. The first part presents the measure-
ments in situ, including tests section design, material properties,
where:
pavements instrumentation, and full-scale field testing. The second
part assesses the accuracy of the response to the thermal load and
aT = time-temperature shift factor;
the last section investigates the accuracy of the response to both
C 1 , C 2 = regression coefficients;
thermal and dynamic load.
T and T ref = analysis and reference temperature, respectively.

The regression coefficients were obtained from the sigmoid 4.1. In situ measurements
function introduced in the AASHOTO pavement design guide
(AASHTO, 2004), as shown in Eqs. [(34) and (35)] 4.1.1. Test section and materials properties
The instrumented semi-rigid asphalt pavement modeled in this
a
logðEðtr ÞÞ ¼ d þ ð34Þ FE analysis is part of the full-scale RIOH-Track pavement test facil-
1 þ expðb þ clogðtr ÞÞ
ity located in Yi-Zhuang, Beijing, China. In order to generate com-
with: prehensive test data needed to validate and refine current Chinese
design methodology for pavement structures and materials, the
t
aT ¼ ð35Þ test section included 19 asphalt pavements structure, 13 cement
tr
concrete structures, and 6 rutting resistance asphalt pavement
where: structures. However, only three typical structures, representing
three distinct groups of the RIOH-Track semi-rigid pavement struc-
a; b; c; d = regression coefficients; ture were modeled in this study. These pavement structures iden-
t = time before conversion under temperature test; tified as STR-A, STR-B, and STR-C correspond respectively to STR-6,
tr = time after conversion under reference temperature. STR-7, and STR-9 of the RIOH-Track project. For more details on
these structures, see [36]. The first structure (STR-A) is a semi-
To obtain the viscoelastic material parameters, dynamic com- rigid structure with a stress-absorbing layer. The second structure
plex modulus tests were conducted as required by the American (STR-B) differs from conventional semi-rigid structures in the high
Standard Test Method for Dynamic Modulus of Asphalt Mixtures strength of its semi-rigid layer. The last structure (STR-C) was an
(ASTM-D3497). For each asphalt mixture, the test was carried out improvement of the most used traditional semi-rigid structures.
at a temperature of 5, 15, 20, 30, 40 and 50 °C, and at loading fre- The three pavement sections were built on the same cement
quencies of 0.01, 0.10, 0.20, 0.50, 1, 2, 5, 10, 20 and 25 Hz for each stabilized subgrade (CS) with a relatively small longitudinal slope
temperature. The approximate method proposed by Park et al. was and all the natural soil (in-place soil). The subgrade layer contains
used to determine the relaxation modulus from the dynamic com- 2 to –4% cement and the required compressive strength after
plex modulus and the phase angle of the test results [34,35]. The 7 days was 3 MPa. The top of the in-place soil was improved by
latter was used to determine the shear and bulk relaxation compaction without the addition of a stabilizer (hydraulic binder,
modulus. The test result was then fitted in the kernel functions cement, etc.). Two types of base layers, which are widely used in
[Eqs. (31) and (32)] to obtain the Prony Dirichlet series parameters, China, namely a high strength semi-rigid base and a conventional
and considering the thermal-rheological simplicity principle semi-rigid base were used. The first semi-rigid base consists of
[Eqs. (33)–(35)] the WLF constant can be obtained at each test cement stabilized graded crushed stone (CBG-A) containing 6 to
temperature. The Prony Dirichlet series parameter and the instan- 9% cement, with a compressive strength of 6 MPa at 7 days. The
taneous time scale elastic modulus of each asphalt mixture at the second consists of cement stabilized graded crushed stone
aforementioned testing temperature are provided in Tables 5 and (CBG-B) with 4–6% cement content, had a compressive strength
6, respectively. Note that at any desired temperature, asphalt of 4.6 MPa at 7 days. The gradation and the composition of the
material parameters can also be obtained by fitting the above- semi-rigid layer used in this study are listed in Table 8.
mentioned equations [Eqs. (31)–(35)]. The asphalt layers of the pavement test sections presented in
It was assumed that all non-asphaltic layers were isotropic lin- this study consisted of six distinct asphalt mixtures from a
ear elastic. The material parameters of the semi-rigid base and styrene-butadienestyrene modified compound (SBS) [SBS1, SBS2,
sub-base were obtained using a laboratory uniaxial compression and SBS3] or a standard asphalt binder [A30#, and A70#]. Refer
module test method in accordance with Chinese test methods to chinise technical specification for construction of higway
JTG D50-2017 (Specification for Design of Highway Asphalt asphalt pavement [JTG F40-2015]. SBS1-AC13 and SBS3-PAC13
Pavement-Appendix E). Note that the test samples were made as were used as a wear-resistant surface layer to improve the
soon as the construction of the experimental structure section resistance to rutting and durability of the pavement structure.
was completed in order to obtain the module of the uncracked A30-AC25 and SBS2-AC20 were applied in the asphalt middle
semi-rigid base and sub-base. On the other hand, Falling Weight course to enhance resistance to load-associated pavement distress.
Deflectometer tests on the constructed structure were carried out A70-AC20 and the anti-fatigue layer SBS1-AC10 were used as
to determine the elastic modulus of the cement stabilized subgrade asphalt underlying layer. The gradation of the aggregates applied
10 O.C. Assogba et al. / Construction and Building Materials 235 (2020) 117406

Table 5
Prony parameters of each asphalt mixture at the selected test temperature.

s G/K
5 15 20 30 40 50
SBS1-AC13
0.00001 0.02378 0.06063 0.09236 0.18423 0.31288 0.45334
0.0001 0.02372 0.05965 0.08712 0.15131 0.20212 0.21401
0.001 0.04529 0.10726 0.14720 0.21059 0.21377 0.17327
0.01 0.08344 0.17304 0.20840 0.20828 0.14498 0.08664
0.1 0.14207 0.21951 0.21063 0.13324 0.06803 0.03570
1 0.20071 0.18828 0.13729 0.05948 0.02679 0.01373
10 0.21267 0.10574 0.06266 0.02356 0.01078 0.00586
100 0.13701 0.04208 0.02341 0.00887 0.00428 0.00240
1000 0.08435 0.02140 0.01228 0.00512 0.00264 0.00154
A30-AC25
0.00001 0.02421 0.04830 0.06366 0.13876 0.24469 0.34647
0.0001 0.01782 0.03743 0.05071 0.09623 0.14293 0.18534
0.001 0.03094 0.06335 0.08509 0.14323 0.18199 0.19464
0.01 0.05073 0.10175 0.13260 0.18287 0.18039 0.14012
0.1 0.08403 0.14990 0.18036 0.18357 0.13090 0.07273
1 0.12059 0.18535 0.19279 0.13272 0.06819 0.03050
10 0.17558 0.18213 0.15055 0.06977 0.02866 0.01257
100 0.16013 0.11771 0.07775 0.02783 0.01053 0.00499
1000 0.21059 0.07759 0.04294 0.01428 0.00565 0.00316
SBS2-AC20
0.00001 0.04755 0.10633 0.15027 0.27106 0.42854 0.59675
0.0001 0.03680 0.07687 0.10274 0.15560 0.18515 0.17155
0.001 0.06233 0.11994 0.15058 0.19000 0.17640 0.12618
0.01 0.10030 0.16677 0.18700 0.17430 0.11444 0.06033
0.1 0.14846 0.19108 0.17994 0.11452 0.05524 0.02511
1 0.18509 0.16287 0.12338 0.05448 0.02179 0.00938
10 0.18408 0.09930 0.06155 0.02189 0.00858 0.00399
100 0.12036 0.04272 0.02370 0.00800 0.00328 0.00160
1000 0.07927 0.02146 0.01188 0.00442 0.00204 0.00110
A70-AC25
0.00001 0.02854 0.06927 0.08675 0.19733 0.34127 0.51440
0.0001 0.02655 0.06110 0.08049 0.14742 0.20200 0.21245
0.001 0.04908 0.10558 0.13784 0.20514 0.21428 0.16205
0.01 0.08764 0.16616 0.20473 0.21240 0.14543 0.07176
0.1 0.14519 0.21519 0.22587 0.14346 0.06354 0.02477
1 0.20335 0.19795 0.15828 0.06162 0.02084 0.00767
10 0.21679 0.11647 0.06898 0.02038 0.00674 0.00281
100 0.14162 0.04359 0.02194 0.00625 0.00224 0.00102
1000 0.07625 0.01712 0.00890 0.00286 0.00121 0.00063
SBS3-PAC13
0.00001 0.06614 0.15150 0.19864 0.30625 0.43836 0.56186
0.0001 0.04187 0.08031 0.09941 0.13594 0.16075 0.16892
0.001 0.06564 0.11275 0.13288 0.16101 0.15985 0.13422
0.01 0.09720 0.14295 0.15555 0.15440 0.11775 0.07249
0.1 0.13464 0.15904 0.15435 0.11745 0.06759 0.03322
1 0.16125 0.14482 0.12144 0.06831 0.03089 0.01352
10 0.16704 0.10533 0.07501 0.03229 0.01299 0.00594
100 0.11886 0.05536 0.03459 0.01268 0.00502 0.00247
1000 0.09802 0.03283 0.01881 0.00680 0.00302 0.00172
SBS1-AC10
0.00001 0.15963 0.12108 0.13548 0.22530 0.41092 0.44959
0.0001 0.08008 0.08190 0.10926 0.15994 0.18096 0.20483
0.001 0.11061 0.14309 0.16712 0.20658 0.17713 0.17466
0.01 0.13821 0.18756 0.20722 0.19120 0.12041 0.09452
0.1 0.15303 0.19511 0.18335 0.11894 0.06144 0.04034
1 0.14060 0.13998 0.10879 0.05316 0.02563 0.01554
10 0.10557 0.07337 0.04805 0.02116 0.01054 0.00651
100 0.05780 0.02855 0.01794 0.00797 0.00415 0.00265
1000 0.03671 0.01485 0.00960 0.00465 0.00265 0.00173

in each type of mixture is illustrated in Fig. 4, and some basic installed throughout the pavement structure following a compre-
design properties of the mixtures are listed in Table 9. More details hensive instrumentation plan and a procedure of installation with
can be found in [26,36]. a high survival rate. Fig. 5 shows a detailed view of the entire
instrumentation plan and the sensor’s distributions layout in each
4.1.2. Pavement section instrumentation pavement structure. A total of 55 sensors were installed at differ-
State-Of-The-Art field instrumentation and data acquisition ent elevations and positions in each instrumented pavement sec-
program was developed as part of the RIOH-Track field monitoring tion to measure significant parameters such as stress, strains,
program. The field instrumentation consisted of numerous sensors deformations, moisture and temperature inside the pavement lay-
O.C. Assogba et al. / Construction and Building Materials 235 (2020) 117406 11

Table 6
The elastic modulus for viscoelasticity with instantaneous time scale at selected Temperature.

Temperature (°C) Instantaneous Modulus of each asphalt mixtures (MPa)


SBS1-AC13 SBS3-PAC13 A30-AC25 SBS2-AC20 SBS1-AC10 A70-AC25
5 18,364 12,839 31,901 23,299 18,215 25,953
15 12,223 9093 25,724 15,476 11,068 18,076
20 9217 7189 21,888 11,649 8451 13,855
30 4787 3993 13,966 5791 47,361 6690
40 2513 1958 7788 2659 2764 2755
50 1472 930 4056 1276 1707 1130

Table 7
Elastic moduli of non-asphaltic layers.

Layers Base / Subbase Subgrade Natural Soil


Materials CBG-A CBG-B CS In place soil
Elastic Modulus (MPa) 11,200 8000 3000 120

Table 8
Gradation and composition of cement stabilized graded crushed stone (CBG).

Sieve Size (mm) 26.5 19 13.2 9.5 4.75 2.36 1.18 0.6 0.3 0.15 0.705
Gradation (% passing) 100 84 68 57 40 26 17 11 7 5 4
Composition CBG-A CBG-B
Water 5.6 5.29
Cement 5.9 4.5

100 AC10 impervious to moisture absorption producing excellent stability


AC13 for long term monitoring. A total of 8 KM-100 HAS strain transduc-
Percent passing ()

80 PAC13 ers, 4 oriented in a longitudinal direction and 4 oriented trans-


AC20 versely were installed at the bottom of the asphalt layer of each
60 AC25 structure to measure the asphalt strain. Two KM-50F strain trans-
40 ducers were integrated to measure the dynamic vertical strain at
the bottom of the asphalt layer. Fig. 6[(a)–(f)] shows some pictures
20 of the installations of the dynamic sensors at the bottom of the
asphalt layer. The asphalt strain gauge was mainly installed by
0 the following methods. (1) Sensors layout: Before the asphalt layer
0.075 0.15 0.3 0.6 1.18 2.36 4.75 9.5 13.2 16 19 26.5
was paved, the embedding position of strain gauge and the posi-
Sieve size (mm)
tion of the sensor connection lines were placed on the top of the
Fig. 4. Aggregate gradation of different asphalt mixture. waterproof bonding layer with wood and other materials, as
shown in Fig. 6(a). After the asphalt layer was completed, the strain
gauge installation groove and sensors connection lines were cut
ers. The position of the sensors determined according to the survey out at this location [Fig. 6(b)]. (2) Installation of strain gauges: A
results based on the wheel truck was centered on the left side of small amount of asphalt mixture was spread to the bottom of
the wheel path of the lane reserved for the controlled-load test the installation hole, then the asphalt strain gauges were placed
trucks. on top. Rubber band was used to clamp the two sides of the
The KM series strain transducers made by Tokyo Measuring I-shaped strain gauge to exert prestressing force, as shown in
Instrument Lab. (TML) were adopted considering their waterproof Fig. 6(c). After carefully covering the strain gauge with asphalt con-
construction and extremely low modulus which is ideally suited crete [Fig. 6(d)], a roller was used to compact it without damaging
for pavement internal strain measurement. They were also, totally the sensors [Fig. 6(e)]. (3) Sensor signal verification: Sensor signal

Table 9
Design properties of asphalt mixtures.

Asphalt Mixtures SBS1-AC13 SBS3-PAC13 A30-AC25 SBS2-AC20 SBS1-AC10 A70-AC25


Asphalt cement SBS-1 SBS-3 A30# SBS-2 SBS-1 A70#
Gradation AC13 PAC13 AC25 AC20 AC10 AC25
VMA (%) 16.93 24.04 13.28 13.18 14.59 13.42
VV (%) 5.15 15.02 3.81 2.93 2.95 4.06
VFA (%) 70.12 37.83 71.44 76.44 80.66 68.17
VCA (%) 42.42 39.46 35.11 39.61 53.27 35.22
Pa (%) 5.46 4.55 4.12 4.48 5.47 4.07
cmb(g/cm3) 2.4741 2.2452 2.5369 2.5404 2.5025 2.5312
cmd(g/cm3) 2.3455 2.1487 2.4372 2.4307 2.3514 2.4333

Note: VFA = Void filled with asphalt volume; Pa = Asphalt aggregate ratio; VV = Volume of air in the compacted mix; cmb = bulk density of the mixture; VMA = Void mineral
aggregate; cmd = dry density of the mixture; VCA = Voids in coarse aggregate of asphalt mix.
12 O.C. Assogba et al. / Construction and Building Materials 235 (2020) 117406

200
90 10 Vertical Strain Sensor
30
Asphalt Transversal Strain Sensor
AC Upper Layer W1
*
AC Middle Layer W2
* Asphalt Longitudinal Strain Sensor
AC Underlying layer Y1 W3
* D1 Base, Subbase and Subgrade
Base Y2 Transversal Strain Sensor
S1 W4
* D2
Subbase Base, Subbase and Subgrade
Y3 S2 W5
* D3
Longitudinal Strain Sensor
Subgrade Hydraulic pressure cells
Y4 S3 W6
20
* D4
Y5 S4 Humidity sensors
Natural Soil

54
W7 Y6 S5
*
50
W6
* Y7 S6 * Thermocouple

Displacement Transduscer
50 S7
W8
* 30 Unit : Cm (Note in Scale)

(a)

40 20 30 30 80 30 30 60

Road centerline Road centerline


Reference Line

Reference Line
60

60
L1-1 H1-3 H1-1
SX 1-1
L1-4 L1-3 L1-2 H1-4 H1-2
30

30
Traffic Direction
Y1 Y2
S-1
30

30
L1-5 Traffic Direction H1-8 H1-6
L1-8 L1-6
SX 1-2
L1-7 H1-7 H1-5

Unit : Cm (Note in Scale) Unit : Cm (Note in Scale)

SX 1-1 and SX 1-2 = Vertical Asphalt Strain Sensor S-1 = Humidity Sensor
L1-1, L1-3, L1-5 and L1-7 = Horizontal Asphalt Strain Sensor (Longitudinal Direction) H1-1, H1-3, H1-5 and H1-7 = Horizontal Asphalt Strain Sensor (Longitudinal Direction)
L1-2, L1-3, L1-6 and L1-8 Horizontal Asphalt Strain Sensor (Transversal direction) H1-2, H1-4, H1-6 and H1-8 Horizontal Asphalt Strain Sensor (Transversal direction)
Y1=Hydraulic pressure cells Y2=Hydraulic pressure cells
(b) (c)
30 30 60 30 30 60

Road centerline Road centerline


Reference Line
Reference Line

60
60

H2-3 H2-1 H3-3 H3-1

H2-4 H2-2 H3-4 H3-2


30
30

Y3 Y4
S-2 S-3
30
30

Traffic Direction H2-8 H2-6 Traffic Direction H3-8 H3-6

H2-7 H2-5 H3-7 H3-5

Unit : Cm (Note in Scale) Unit : Cm (Note in Scale)

S-2 = Humidity Sensor S-3 = Humidity Sensor


H2-1, H2-3, H2-5 and H2-7 = Horizontal Asphalt Strain Sensor (Longitudinal Direction) H3-1, H3-3, H3-5 and H3-7 = Horizontal Asphalt Strain Sensor (Longitudinal Direction)
H2-2, H2-4, H2-6 and H2-8 Horizontal Asphalt Strain Sensor (Transversal Direction) H3-2, H3-4, H3-6 and H3-8 Horizontal Asphalt Strain Sensor (Transversal Direction)
Y3=Hydraulic pressure cells Y4=Hydraulic pressure cells
(d) (e)
Fig. 5. Typical dynamic instrumentation layout of STR-B pavement structure: (a) Profile view; (b) Detailed plan view at the bottom of asphalt layer; (c) Detailed plan view at
the bottom of the Base; (d) Detailed plan view at the bottom of the subbase and (e) Detailed plan view at the bottom of the subgrade.

readings were performed from time to time before and during the At the bottom of the cement stabilized graded crushed stone
asphalt pavement to assess its signal presence and stability in base and subbase, and at the bottom of the cement stabilized sub-
order to confirm that the sensors were working properly [Fig. 6 grade soil, a total of 24 KM-100A strain transducers were installed
(f)]. In the absence of a signal, the sensors were replaced immedi- in order to capture both horizontal and longitudinal dynamic
ately and even if the paving was completed, it was necessary to strain. At each of the above-mentioned positions in each pavement
remedy the sensor by re-digging. section, there were 8 dynamic concrete sensors, in which 4 were in
O.C. Assogba et al. / Construction and Building Materials 235 (2020) 117406 13

Arms of KM-
100HA S

Thermocoupl e
Rubber band (T-F-0.65)

(a) (b) (c)

(d) (e) (f)


Fig. 6. Installation process of the dynamic sensors at the bottom of the asphalt layer: (a) Layout; (b) Sensors installation point; (c) Positioning of strain transducer; (d)
Coverage by asphalt concrete, (e) Roller compaction and (f) Data reader for on-site reading.

longitudinal direction and the other 4 were transversely oriented. hydraulic pressure cells were embedded at the bottom of the
Fig. 7[(a)–(d)] shows some pictures of the installations of the asphalt layer, base, subbase, subgrade and at a different depth
dynamic sensors at the bottom of the cement stabilized graded within the natural soil. Fig. 6 show some pictures of the installation
crushed stone base. The concrete strain gauge was installed using of the pressure transducer on the site. The measurement range [FS]
the following methods: (1) Prefabrication of the mounting block of the pressure transducer was (0–200 kPa) up to (0–20 MPa) and
of the strain gauges: According to the mix ratio and compactness the sensitivity was 0.002% FS. The operating temperature was from
of the semi-rigid material to be embedded, the specimens were 20 °C to +80 °C. The pressure cells were located in the center of
prefabricated in the laboratory by static pressing, and the strain the outer wheel lane, where the wheel was expected to generate
gauge mounting blocks were prepared as illustrated in Fig. 7(a). the greatest stress on the pavement structure. The installation pro-
(2) Sensors layout: Similar to the installation process of the asphalt cess of pressure gauge was similar to that of asphalt strain gauge.
strain gauge, the embedding position of the strain gauges were In addition to mechanical responses such as stress, strain and
marked. Once the layer of the semi-rigid material was completed, deformation, moisture and temperature inside pavement layers
the sensors installation groove and connection lines were cut out are of great importance. Thus, in order to measure temperature
in this location [Fig. 7(b)]. (3) Embedding of the strain gauges: profile throughout the pavement section, a total of 7 platinum
The prefabricated strain gauge mounting blocks were placed in resistance temperature sensor PT100 of Jinzhou Fine Instrument
the mounting trough, and cement stabilized materials with a smal- Co., Ltd were inserted at different depths of each pavement section.
ler particle size were used to fill the gap, then the semi-rigid mate- A roadside weather station was erected for the accurate measure-
rial is backfilled and compacted, as illustrated in Fig. 7(c) and (d). ment and modeling of the climate. Radiation and air temperature
(4) Sensor signal checking: Sensor signal readings were taken from conditions recorded on two representative days, 04th March
time to time before and during sensor installation to evaluate the 2017 (Spring day) and 2nd July 2017 (Summer day) are depicted
presence of the signal and confirm its correct operation. in Fig. 9[(a) and (b)]. For each pavement section, a total of 7 humid-
Piezo-resistive pressure transducer P252A with the hydraulic ity sensors were installed at different depth within the pavement
pressure cells OL111D05000 manufactured by the Italian company to capture the moisture level. Once the construction and installa-
SISGEO were installed to record the pressure at various depth in tion of the instruments were completed, the survivability assess-
the pavement structure, as shown in Fig. 8[(a)–(c)]. A total of 7 ment of the sensors was carried out by moving a truck on the

Fig. 7. Installation process of the dynamic strain sensors at the bottom of the semi-rigid layer (a) Precast concrete strain sensor mounting block; (b) Sensor installation slot;
(c) Positioning and installation, and (d) Coverage by semi-rigid material.
14 O.C. Assogba et al. / Construction and Building Materials 235 (2020) 117406

Pressure cell

(a) (b) (c)


Fig. 8. Installation of pressure cells in (a) Asphalt layer; (b) Semi-rigid layers, and (c) Roadbed.

40 1.2
07/02/2017 03/04/2017
35 Tmax =36.94qC 07/02/2017

Radiation Intensity (MJ/m2)


1.0
30 Tmin =25.81 qC
Air Temperatue (qC)

25 0.8
20
0.6
15 03/04/2017
Tmax =16.75 qC
10 Tmin =-3.26 qC 0.4
5
0.2
0
-5 0.0
0 2 4 6 8 10 12 14 16 18 20 22 24 0 2 4 6 8 10 12 14 16 18 20 22 24
Time of Day (h) Time of Day (h)
(a) (b)

Fig. 9. Variation of: (a) solar radiation and (b) air temperature in a selected day (03/04/2017 and 07/02/2017).

instruments to assess the presence and the stability of the signal. 4.2. Accuracy of the response to the temperature load
Note that data of mechanical response was continuously gathered
at a rate of 2,000 points per second by a high-frequency data The result of the 3-D transient thermal analysis and the field
acquisition system; dynamic data acquisition system (DDAS) of observed variation of pavement temperature at different depth
the Austrian company Gantner. within STR-A over two representative days, 4th March 2017
(Spring day) and 2nd July 2017 (Summer day) are illustrated in
Fig. 11[(a) and (b)]. It appears that the predicted and observed tem-
4.1.3. Controlled truck load testing perature curves are similar and that the amplitude of the temper-
Controlled truck load testing was performed by driving Stery- ature fluctuations decreases as the depth increases. Likewise, the
King 6  4 Dump Truck loaded to a known weight at a controlled maximum temperature in the pavement structure decreases with
speed over the pavement structure with the objective of quantify- increasing depth, however, there is a delay in the appearance of
ing the pavement response while being exposed to varying daily the maximum temperature. Overall, it can be stated that the two
temperature, different wheel load configuration at target moving results lie close together. The effectiveness of the established 3D
speed of 20, 40 and 60 km/h. Taking into account the climatic con- FE models in simulating the thermal behavior of the pavement
ditions of the test site, located in Beijing, China, measurements structure was evaluated in terms of the prediction error at the
were carried out at two representative days, 04th March 2017 maximum and minimum daily temperature, er [Eq. (36)]:
(Spring day) and 2nd July 2017 (Summer day). The dump truck
 
used to conduct the controlled load full-scale field test had a single Tp  Tm
er ¼  100 ð36Þ
steering axle and a tandem driving axle, as previously indicated in Tm
Table 3. In order to take into account the impact of truck weight
distribution and the overload phenomenon of the heavy-duty where:
trucks on the semi-rigid pavement structures, measurements were
made under three load levels namely empty, fully loaded and T p predicted temperature, °C;
overloading conditions, as presented in Table 3. It is important to T m measured temperature, °C.
mention that the instrumentation area was centered along the left
side of the wheel path of the truck lane, as shown in Fig. 8. The A negative value of er indicates an under-prediction of temper-
location of the sensing zone was indicated on the pavement surface ature from FE simulations, while a positive value of er suggests an
[Fig. 10(a)] to ensure that the truck left wheels passed directly over over-prediction. As indicated by the prediction errors tabulation
the sensors [Fig. 10(b)]. Video cameras were also installed to verify (Table 10), the developed 3-D transient heat FE model generally
the passage of the truck wheels over the instrument area. Each test predicts the temperature observed at different depths of the pave-
was repeated at least three times to ensure that the measured ment structure with a prediction error of less than 10%. Therefore,
characteristic was representative and replicable. it can be concluded that the 3-D transient heat FE model is capable
O.C. Assogba et al. / Construction and Building Materials 235 (2020) 117406 15

1.350 m 3.825 m

Instrumentation Area
1.830m

2.022 m
Rear Drive Front Drive
Steering axle
Tandem Axle Tandem Axle

Front Axle
Rear Axles
(a) (b)
Fig. 10. Identification of sensors positions during the field test (a) expected disposition of wheel load on the instrumentation area during a controlled load test. and (b) sensor
positions on the road surface and controlled load test in progress.

25 25
Actual Air Temperature Observed Temperature Approximate Air Temperature Predicted Temperature
Depth = 0.04m Depth = 0.14m 03/04/2017 Depth = 0.04m Depth = 0.14m 03/04/2017
20 Depth = 0.16m Depth = 0.36m 20 Depth = 0.16m Depth = 0.36m
Depth = 0.56m Depth = 0.76m Depth = 0.56m Depth = 0.76m
Temperature ( qC)

15 Temperature ( qC) 15

10 10

5 5

0 0

-5 -5
0 2 4 6 8 10 12 14 16 18 20 22 24 0 2 4 6 8 10 12 14 16 18 20 22 24
Time of Day (h) (a) Time of Day (h)
45 45 Approximate Air Temperature
Actual Air Temperature Observed Temperature Predicted Temperature
Depth = 0.04m Depth = 0.14m Depth = 0.04m Depth = 0.14m 07/02/2017
07/02/2017
Depth = 0.16m Depth = 0.36m Depth = 0.16m Depth = 0.36m
Depth = 0.56m Depth = 0.76m Depth = 0.56m Depth = 0.76m
40 40
Temperature ( qC)

Temperature ( qC)

35 35

30 30

25 25
0 2 4 6 8 10 12 14 16 18 20 22 24 0 2 4 6 8 10 12 14 16 18 20 22 24
Time of Day (h) Time of Day (h)
(b)
Fig. 11. Comparison between predicted (right) and observed temperature (left) for: (a) 04th March 2017 and (b) 2nd July 2017.

of simulating the time-dependent temperature profile in asphalt predicted response using laboratory-characterized material prop-
pavement structure under daily temperature variation. And, the erty, measured tire contact stress, and the dynamic implicit analy-
thermal properties issued from the Chinese specifications for the sis were compared to the field measurements of 2nd July 2017
design of highway asphalt pavement (JTG D50-2017) adopted in (summer day).
the model yielded a great temperature prediction. Figs. 12 and 13 illustrate the comparison between field-
measured and FEM predicted dynamic strain for pavement struc-
4.3. Accuracy of the response to dynamic moving load ture STR-A under a full load truck (front axle loaded at 75.53 kN
and rear tandem axle loaded at 203.864 kN) operating at 40 km/h
In addition to the accuracy of the response to the thermal load, speed, respectively, at the bottom of the asphalt layer and
the accuracy level of the developed FE model in predicting mecha- semi-rigid layer. As expected, the time histories of horizontal
nistic response of the pavement under vehicular loading and the [Fig. 12(a)] and vertical [Fig. 12(b)] dynamic strain taken at the
temperature gradient must be verified and validated. Thus, FE bottom of the asphalt layer, as well as the horizontal [Fig. 13(a)]
16 O.C. Assogba et al. / Construction and Building Materials 235 (2020) 117406

Table 10
Result of the predictions error analysis.

Depth from pavement surface 0.04 0.14 0.16 0.36 0.56 0.76
07–02-2017 (Summer)
Field Observed Maximum Temperature, °C 41.81 39.47 39.02 37.23 34.67 31.96
Minimum Temperature, °C 33.92 34.84 35.07 35.67 34.26 31.57
FEM Predicted Maximum Temperature, °C 40.80 37.80 37.49 35.10 33.40 31.92
Minimum Temperature, °C 34.44 35.35 35.37 34.66 33.27 31.79
Prediction error, % Maximum Temperature 2.42 4.22 3.92 5.72 3.67 0.14
Minimum Temperature 1.52 1.46 0.84 2.84 2.90 0.69
03–04-2017 (Spring)
Field Observed Maximum Temperature, °C 13.13 10.90 10.60 9.30 7.99 7.09
Minimum Temperature, °C 4.51 5.80 6.16 7.60 7.49 6.84
FEM Predicted Maximum Temperature, °C 13.33 10.85 10.49 8.37 8.37 7.50
Minimum Temperature, °C 4.89 6.24 6.48 7.46 7.46 7.27
Prediction error, % Maximum Temperature 1.54 0.53 1.05 9.96 4.77 5.78
Minimum Temperature 8.48 7.43 5.05 1.76 0.36 6.26

5 25

0 0
Dynamic Strain (PH )

-5 Dynamic Strain (PH ) -25


Stering Axle
-10 -50
Stering Axle Field Measurement
Field Measurement
FEM Prediction FEM Prediction
-15 -75
Anmbient Temperature: 36 qC Rear Tandem Axle Anmbient Temperature: 36 qC
-20 Target Truck Speed : 40 km/h -100 Target Truck Speed : 40 km/h Rear Tandem Axle
Steering Axle: Single ~75.53kN Front Tandem Axle Steering Axle: Single ~75.53kN
Drive Axle: Tandem ~ 203.86kN Front Tandem Axle
-25 -125 Drive Axle: Tandem ~ 203.86kN

2 3 4 5 6 7 8 2 3 4 5 6 7 8
Time (s) Time (s)
(a) (b)

Fig. 12. Comparisons between field-measured and FEM predicted dynamic strain at the bottom of the asphalt layer in the (a) longitudinal direction and (b) vertical direction.

20 15
Field Measurement Field Measurement
FEM Prediction FEM Prediction
Rear Tandem Axle
15 Rear Tandem Axle
Dynamic Strain (PH )

Front Tandem Axle


Dynamic Strain (PH)

Front Tandem Axle


10 Anmbient Temperature: 36 qC
Anmbient Temperature: 36 qC Target Truck Speed : 40 km/h
10 Target Truck Speed : 40 km/h Steering Axle: Single ~75.53kN
Steering Axle: Single ~75.53kN Drive Axle: Tandem ~ 203.86kN
Stering Axle 5
Drive Axle: Tandem ~ 203.86kN
5 Stering Axle

0
0

-5 -5
2 3 4 5 6 7 8 2 3 4 5 6 7 8
Time (s) Time (s)
(a) (b)

Fig. 13. Comparisons between field-measured and FEM predicted dynamic strain at the bottom of the semi-rigid layer in the (a) horizontal direction and (b) longitudinal
direction.

and longitudinal [Fig. 13(b)] obtained at the bottom of the semi- to accurately control truck speed and wheel position. The latter
rigid layer, were in agreement with those recorded during field should ideally be symmetrically applied to the wheel path and just
test. It can be seen that the time histories of both measured and below the center of dual tires. (2) The dynamic responses recorded
calculated strains showed the same trend of change over time by strain gauge are influenced by surrounding material constraints.
and peak values were close. However, there were slight differences The sensors were still under compressive stress whether the
observed in response time and stress peaks. The differences gauges were subjected to compressive stress when approaching
between the measured and calculated responses may be due to the vehicle or after the passage of the vehicle. (3) It was assumed
the following reason: (1) In large-scale field trials, it was difficult that the wheel load was uniformly distributed over a rectangular
O.C. Assogba et al. / Construction and Building Materials 235 (2020) 117406 17

contact area between each tire and pavement surface, which dif- addition, the non-linearity of the thermal gradient causes a varia-
fers from the actual contact and tire pressure induced by the tires tion of stiffness modulus as a function of depth, which directly
in field tests. Only a realistic three-dimensional contact pressure influences the response and performance of the pavement struc-
and non-uniform contact area can simulate with high accuracy ture. Therefore, the assumption of constant temperature for these
level the actual stress distribution on the pavement structure. layers may be misleading for a realistic prediction of mechanical
Overall, the dynamic responses predicted by numerical simula- response, which is essential for the future design of more durable
tion are fundamentally consistent with field data. Therefore, the pavement structures. For this reason, the nonlinear characteristics
developed FE model can be used to predict the pavement response of temperature were incorporated into the mechanistic model.
under the effect of traffic load and temperature gradient. Accordingly, the asphalt layers were subdivided into several sub-
layers and the corresponding mechanical properties were assigned
to take into account the variation with the depth of the dynamic
5. Results and discussion properties of the asphalt mix.

5.1. Temperature analysis


5.2. Combined effect of axle load and temperature gradient on the
Prior studies indicated that the modulus of elasticity of asphalt pavement mechanical response
concrete can depend strongly on the temperature. The temperature
at pavement surface varies with the time during the day, which in As indicated in the previous section the temperature gradient
turn changes the temperature gradient along with the pavement along the pavement structure depth varies with the time during
structure. For instance, Fig. 14[(a) and (b)] shows a typical temper- the day. Also, the temperature gradient along the asphalt layer
ature variation in STR-A pavement sections at different times dur- was generally larger than that along with the semi-rigid layer (base
ing a summer and spring days. As indicated, the temperature of the and subbase) and the subgrade. Thus, in this study, the nonlinear
pavement surface increases with increasing solar radiation, while thermal gradient profiles at 5: 00 and at 16:00 in summer and
the temperature inside the pavement structure decreases with spring day were applied to account for most adverse temperature
increasing depth. When solar radiation reaches its maximum gradient of the day. The temperature gradient in the asphalt layer
value, the temperature on the pavement surface starts to decrease at 5: 00 and at 16:00 on summer days was 10.47 °C/m and
progressively, although the temperature inside the pavement con- +36.56 °C/m, respectively. In spring, the temperature gradient
tinues to increase. This indicates that there is a continuous transfer along with the asphalt layer at 5: 00 and at 16:00 was 34.85 °C/m
of thermal energy within the pavement system while sun radiation and 49.54 °C/m, respectively. In order to allow a better analysis
decreases. After sunset, the heat exchange between the cold air and of the coupling effect of traffic load and temperature on the
the pavement structure leads to a rapid decrease in the tempera- mechanical response of the roadway, specific analysis points were
ture of the AC upper layer. At this stage, the temperature in AC selected (Fig. 15) to investigate the pavement response according
middle layer is higher than that of the AC upper layer, AC bottom to depth. Figs. 16 and 17 illustrate the effect of thermal gradients
layer and layers beneath. It was observed that the temperature on the three directional strain and stress that occur inside the
inside the cement stabilized subgrade remains almost constant in pavement structure STR-B under a fully loaded test truck [Table 3]
spring day and showed a slight variation in summer day (less than moving at a speed of 20 km/h. It can be seen that the amplitude
2 °C). This reveals that the pavement structures studied could and distribution of mechanical responses along with the depth of
reduce the adverse effect of the temperature variation on the foun- the pavement structure changed with change in temperature gra-
dation soil. The temperature variation inside the pavement struc- dient. This indicates that the effects of the temperature gradient
ture, especially in the asphalt layer (AC upper layer to bottom have a significant impact on the distribution of stresses and strains
layer) was found to be nonlinear. The temperature gradient along in the pavement system.
the asphalt layer was generally larger than that along with the The result of the dynamic strain shown in Fig. 16[(a)–(c)] indi-
semi-rigid layer (base and subbase) and the subgrade. Overall, cates that there were greater compressive and tensile strains in the
the temperature field within the pavement structure is strongly asphalt layer. The transverse strain [Fig. 16(a)] at the top of the
affected by the physical properties of the non-asphaltic layers upper asphalt layer is in tension, it decreases with depth and
and by the viscoelastic nature of the asphalt concrete layers. In inverted to compressive approximately at the mid-depth of the

(a) (b)

Fig. 14. Variations of pavement temperature along with pavement depth at a different time periods: (a) on a spring day and (b) on a summer day in STR-A.
18 O.C. Assogba et al. / Construction and Building Materials 235 (2020) 117406

Rear axle dual tire Moving Direction


(Left tire footprint)
1 - Horizontal strain ɛx and stress σx
Wheel path
2 - Vertical strain ɛy and stress σy
- Longitudinal strain ɛz and stress σz
- Maximum shear stress τmax
1 3 - Shear stress τxy
4 - Shear stress τxz
5 - Shear stress τyz
4 2
Wheel path
3 5 Face
Rear axle dual tire Shear stress τyz
(Rigth tire footprint) Direction

Fig. 15. Analysis points where the dynamic response of the pavement structure is taken according to the depth.

Dynamic Strain (H) Dynamic Strain (H) Dynamic Strain (H)


-60μ -40μ -20μ 0 20μ 40μ -250μ -200μ -150μ -100μ -50μ 0 50μ -60μ -40μ -20μ 0 20μ
0.0 0.0 0.0
AC Layer

AC Layer

AC Layer
0.1 0.1 0.1
Depth From Pavement Surface (m)

Depth From Pavement Surface (m)

Depth From Pavement Surface (m)


0.2 0.2 0.2
Base

Base

Base
0.3 0.3 0.3

0.4 0.4 Thermal gradient 0.4 Thermal gradient


Subbase

Subbase

Subbase
Thermal gradient
in a spring day in a spring day in a spring day
+49.54 qC.m-1 +49.54 qC.m-1 +49.54 qC.m-1
0.5 0.5 0.5
-34.84 qC.m-1 -34.84 qC.m-1 -34.84 qC.m-1
Subgrade

Subgrade

Subgrade
0.6 Thermal gradient 0.6 Thermal gradient 0.6 Thermal gradient
in a summer day in a summer day in a summer day

0.7 +36.56 q C.m-1 0.7 +36.56 qC.m-1 0.7 +36.56 qC.m-1


-10.47 qC.m-1 -10.47 qC.m-1 -10.47 qC.m-1
Compression Tension Compression Tension Compression Tension
0.8 0.8 0.8
(a) (b) (c)
Fig. 16. Combined effect of thermal gradient and axle load on the distribution of the dynamic strain along with the depth of the pavement structure STR-B: (a) Transverse (b)
Vertical and (c) Longitudinal direction.

Dynamic Stress (kPa) Dynamic Stress (kPa) Dynamic Stress (kPa)


-375 -250 -125 0 125 -750 -600 -450 -300 -150 0 150 -625 -500 -375 -250 -125 0 125
0.0 0.0 0.0
AC Layer

AC Layer

AC Layer

0.1 0.1
Depth From Pavement Surface (m)

Depth From Pavement Surface (m)

0.1
Depth From Pavement Surface (m)

0.2 0.2 0.2


Base

Base

Base

0.3 0.3 0.3

0.4 Thermal gradient 0.4 Thermal gradient Thermal gradient


0.4
Subbase

Subbase

Subbase

in a spring day in a spring day in a spring day


0.5 +36.56 qC.m-1 0.5 +36.56 qC.m-1 +36.56 qC.m-1
0.5
-10.47 qC.m-1 -10.47 qC.m-1 -10.47 qC.m-1
0.6 Thermal gradient 0.6 Thermal gradient Thermal gradient
Subgrade

Subgrade

0.6
Subgrade

in a summer day in a summer day in a summer day


+49.54 qC.m-1 +49.54 qC.m-1 +49.54 qC.m-1
0.7 0.7 0.7
-34.84 qC.m-1 -34.84 qC.m-1 -34.84 qC.m-1
Compression Tension Compression Tension Compression Tension
0.8 0.8 0.8
(a) (b) (c)
Fig. 17. Combined effect of thermal gradient and axle load on the distribution of the dynamic stress along with the depth of the pavement structure STR-B: (a) Transverse (b)
Vertical and (c) Longitudinal direction.
O.C. Assogba et al. / Construction and Building Materials 235 (2020) 117406 19

surface layer. The compressive strain reaches its highest value tion of a semi-rigid base asphalt pavement. In fact, the quality of
around the mid-depth of the AC middle course, then decreases lin- adhesion, which depends on the type and rate of application of
early along with the depth of the AC bottom course. The strain in the tack coat, as well as the construction technologies and environ-
the upper half part of the semi-rigid base remains compressive ments, greatly influences the field interface bond between these
whereas, that of the lower part of the latter and the subgrade are layers. Thus, the contact condition of the interface could signifi-
tensile. In contrast to the transverse strain, the longitudinal strain cantly affect pavement response to temperature and traffic load.
[Fig. 16(c)] at the top of the upper asphalt layer is in tension. It Therefore, the response under two different bonding conditions,
decreases with depth and changes to compression approximately namely full slip (m = 0) and frictionless (m = 1) were modelled in
in the mid-depth of the AC middle course. The compression strain addition to perfectly bonded condition (m = 1), and the results
reaches its peaks value approximately at the bottom of the middle are demonstrated in Figs. 18[(a)–(c)] and 19[(a) to (c)]. Note that
layer. The semi-rigid layer and the subgrade were in tension and the dynamic response of the pavement was acquired by moving
strain reponses magnitude increases with the depth. Regarding a fully loaded truck [Table 3] at a constant speed of 20 km/h.
the distribution of the vertical strain [Fig. 16(b)], it was observed And, the measurement was taken under the rear axle of the truck
that they are all in compression and decreased with change in according to the depth.
pavement depth, especially when there was a change of layers. It As expected, Figs. 18 and 19 revealed that the distribution of
was also found that the decrease is more pronounced in the asphalt stresses and strain in the pavement structure are strongly affected
layers than in the other layers. by the contact condition of the interfaces. The curvature and the
The dynamic stress output obtained from the coupled thermal trend of the distribution of stresses and strains along the depth
and mechanistic modeling, illustrated in Fig. 17[(a)–(c)], were con- of the pavement are distinctly different for each bonding condition,
sistent with the results of the dynamic strain. In that case, it can be in particular for a perfectly bonded interface condition. It was also
observed that transverse stresses [Fig. 17(a)] at the top of the apparent that the tensile stress and strain at the bottom of the
upper asphalt layer are in tension on summer days, whereas they asphalt layer decreased and change into compression with the
are in compression on spring days. The maximum compressive improvement of the interface bonding conditions, while the verti-
stress appears in the AC middle course. The stresses in the sub- cal compressive stress and strain increased. This reduction of the
base and the subgrade layer are in tension and decrease with the tensile strain and the tensile stress could be attributed to the coop-
increase of the depth. The shape of the distribution of vertical eration of each layer because of the perfect contact between the
stresses [Fig. 17(b)] is very similar to that of strain. However, it interfaces of the layers. For full slip (l = 0) and frictionless (l = 1)
can be noted that, in this case, the stresses decreased more slowly bond conditions, it was obvious that there existed both tensile
with the increase of the pavement depth, and the magnitude of the and compressive stresses and strains at the bottom of the asphalt
stresses tending to zero in the subgrade. The longitudinal stress layer. Longitudinal tensile stress and strain were found to be much
[Fig. 17(c)] indicates that there is a high concentration of compres- greater and could lead to premature bottom-up fatigue cracking.
sive stress at the top of the upper layer, which can easily cause
deformation of the pavement surface. The tensile stress occurs 5.3.2. Influence of vehicle speed and axle load
when moving deeper into the semi-rigid layer and cement stabi- To investigate the impact of loading amplitude on the mechan-
lized subgrade. ical response of a semi-rigid base asphalt pavement, the pavement
system was subjected to three load levels: empty truck load (rear
5.3. Parametric study single axle load of 46.83 kN), full load truck (rear single axle load
of 101.93 kN) and overloading conditions (rear single axle load of
5.3.1. Influence of interfacial conditions 140.19 k0N), as shown in Tables 2–4. For each load, the controlled
The contact conditions of the asphalt layer and the interface of load test truck was assumed to move at a constant speed of
the semi-rigid layer are of significant importance in the construc- 20 km/h. The influence of vehicle speed on the dynamic response

(a) (b) (c)


Fig. 18. Effect of interfacial condition on the distribution of the dynamic strain along with the depth of the pavement structure STR-B: (a) Transverse (b) Vertical and (c)
Longitudinal direction.
20 O.C. Assogba et al. / Construction and Building Materials 235 (2020) 117406

(a) (b) (c)


Fig. 19. Effect of interfacial condition on the distribution of the dynamic stress along with the depth of the pavement structure STR-B: (a) Transverse (b) Vertical and (c)
Longitudinal direction.

of the pavement was also evaluated by moving a full load truck (transverse strain), 1.109 times (vertical strain) and 1.119 times
[Table 3] at three different traffic speeds, namely; 20 km/h, (longitudinal strain). Similarly, the response to dynamic stresses
40 km/h and 60 km/h. Although, for each type of pavement, the in the asphalt layer increases by 1.105 times (transversal stress),
mechanical response was computed under the aforementioned 1.107 times (vertical stress) and 1.042 times (longitudinal stress).
loads and speed, however, only the distribution of the dynamic In this case, it may be appropriate to regulate the lowest traffic
stress and stress along the depth of the STR-C pavement structure speed on the pavement structure, as heavy trucks always drive at
were reported in this section. a relatively lower speed. The effect of the vehicle speed on the
Figs. 20 and 21 presents the influence of truck moving speed on stress distribution extends in the semi-rigid layer, while for the
the distribution of the dynamic strain and the dynamic stress, distribution of the strain, its influence is limited to the asphalt
respectively, along with the depth of the pavement structure. It layer. The influence of vehicle speed on the response to stresses
can be seen that the stress and strain distributions along with and strains decreases with increasing depth.Fig. 22.
the depth of the pavement, at each velocity, are consistent in terms Figs. 22 and 23 show the impact of the truck axle load level on
of curvature and trend, but not in magnitude. This magnitude the distribution of the dynamic strain and the dynamic stress,
increases as the vehicle speed decreases. For instance, with the respectively, along with the depth of the pavement structure.
decrease in travel speed from 60 km/h to 20 km/h, the dynamic These figures show that the distributions of stresses and strains
strain response in the asphalt layer increased by 1.104 times along with the depth of the pavement system, whatever the load

(a) (b) (c)


Fig. 20. Effect of the truck moving speed on the distribution of the dynamic strains along with the depth of the pavement structure STR-C: (a) Transverse (b) Vertical and (c)
Longitudinal direction.
O.C. Assogba et al. / Construction and Building Materials 235 (2020) 117406 21

(a) (b) (c)


Fig. 21. Effect of the truck moving speed on the distribution of the dynamic stresses along with the depth of the pavement structure STR-C: (a) Transverse (b) Vertical and (c)
Longitudinal direction.

(a) (b) (c)


Fig. 22. Influence of truck axle load level on the distribution of the dynamic strain along with the depth of the pavement structure STR-C: (a) Transverse (b) Vertical and (c)
Longitudinal direction.

condition, were consistent in terms of curvature and trend. It also and the amplitude of the mechanical response within the system.
indicates that the stresses and strains within the pavement struc- This could affect its structural behavior and subsequently result
ture increases with increase of overload rate. For example, consid- in adverse effects on its long-term performance. Therefore, it is
ering the mechanical response of the pavement under the weight imperative to regulate the traffic load against the overload and
of the axle under an empty truck as the reference value; when integrate the influence of the latter in the design phase of the pave-
the axle load of the empty truck increases to an overload condition ment structure.
while the truck’s travel speed remains constant, the response of The tensile strain in the asphalt layers and in particular at the
strain in the asphalt layer increased by 2.989 times (transverse bottom of all proposed new semi-rigid pavement structures was
strain), 4.950 times (vertical strain) and 3.107 times (longitudinal below the critical strain of a conventional pavement structure with
strain). Similarly, the response to dynamic stresses in the asphalt an ordinary asphalt (70 me) [37] or with a modified asphalt mixture
layer increases by 2.992 times (transversal stress), 3.008 times (100 me) [38], even under critical conditions such as overload or
(vertical stress) and 3.127 times (longitudinal stress). Overall, all low speed traffic. Therefore, it is expected that the asphalt layer
of these responses demonstrated that the extra weight extension of these typical semi-rigid pavements will be able to withstand
of the truck axle has a significant influence on the distribution premature fatigue failure, especially bottom cracking.
22 O.C. Assogba et al. / Construction and Building Materials 235 (2020) 117406

(a) (b) (c)


Fig. 23. Influence of truck axle load level on the distribution of the dynamic stress along with the depth of the pavement structure STR-C: (a) Transverse (b) Vertical and (c)
Longitudinal direction j.

5.4. Pavement structure damage analysis where:

Semi-rigid asphalt pavement consists of three components: N f = allowed load repetition until failure of the semi-rigid layer;
asphalt layer, semi-rigid layer, and subgrade. Each layer has a rt = tensile stress at the bottom of the semi-rigid layer;
specific failure mode. The failure mode of each layer is related to M R = 28 days modulus of rupture (flexural strength);
the critical response of that layer (e.g., stress, strain and deforma- b1 ; b2 = field calibration factors, the default factor of 1 was used
tion) to a specific location under loading. In this study, fatigue in this study.
cracks, which represent the most serious type of load-associated
road damage on semi-rigid pavements, were investigated and used The aforementioned transfer functions [Eqs. (37) and (38)] were
to estimate the allowable number of load repetitions in order to used to examine the bottom-up fatigue failure progressions for the
avoid failure in each structure. Two approaches were adopted to asphalt concrete layer and the semi-rigid layer of each structure.
model distress due to fatigue cracking: (1) model the number of Since the maximum tensile stress in the asphalt layer was
allowable load repetitions to prevent fatigue failure of the asphalt observed at the intermediate layer, the tensile stress at the bottom
layers, and (2) model the number of allowable load repetitions to of the asphalt intermediate layer was chosen as the critical
avoid fatigue failure of the semi-rigid layers. The asphalt institute’s response. The flexural strength (M R ) of the CBG-A and CBG-B layers
fatigue model presented in Eq. (37) was used to model fatigue of was 2 MPa and 1.5 MPa, respectively.
the asphalt layer [39]. Figs. 24 and 25 illustrate the results of fatigue analysis for each
pavement structure. As indicated in Fig. 24[(a) and (b)], the num-
0:854
Nf ðet Þ ¼ A  0:00432  104:84ðVFA0:69Þ  e3:291
t  jE
j ð37Þ ber of allowable load repetitions to avoid fatigue failure of the
asphalt layer and the semi-rigid layer decreases sharply with
where:
increase of axle load of the truck. For instance, when the axle load
of the empty truck (rear single axle load of 46.83 kN) increases to
N f = allowable load repetitions to prevent fatigue cracking of
an overload condition (rear single axle load of 140.19 kN), the
the asphalt layer;
allowable load repetitions to avoid fatigue failure of the asphalt
A = is the laboratory-to-field adjustment factor, the default fac-
layer decreased by 96.547% (STR-A), 97.285% (STR-B) and
tor of 18.4 was used for this study;
97.286% (STR-C). Likewise, the allowable load repetitions to avoid
VFA = void filled with asphalt;
fatigue failure of the semi-rigid layer decreased by 76.876% (STR-
et = tensile strain at the bottom of asphalt layer;
A), 78.688% (STR-B) and 79.669% (STR-C). As shown in Fig. 25[(a)
jE
j = dynamic modulus.
and (b)], the fatigue life of the asphalt layer and semi-rigid layer
increases gradually with increase in the truck’s moving speed.
In the case of fatigue failure of the semi-rigid layer, the
For instance, increase of moving speed of fully loaded truck from
Mechanistic-Empirical Pavement Design Guide (MEPDG) fatigue
20 km/h to 60 km/h, the fatigue life of the asphalt layer increased
failure transfer function [Eq. (38)] for a layer of chemically
by 1.524 times (STR-A), 1.131 times (STR-B) and 1.515 times
stabilized material was used to model the number of traffic load
(STR-C). Similarly, the fatigue life of the semi-rigid layer increased
to fatigue failure of the semi-rigid layer.
by 1.158 times (STR-A), 1.167 times (STR-B) and 1.173 times (STR-
   C). Under a fixed moving speed and axle load, the fatigue life of the
0:9723b1  MrtR asphalt layer of the three pavements can be arranged from smallest
logðNf Þ ¼ ð38Þ
0:0825b2 to the largest as follows: STR-A < STR-B < STR-C, while that of the
O.C. Assogba et al. / Construction and Building Materials 235 (2020) 117406 23

(a) (b)
Fig. 24. Comparison oft he fatigue life of the asphalt layer of the three pavements structure under different axle load: (a) Asphalt layer and (b) Semi-rigid layer.

(a) (b)
Fig. 25. Comparison of the fatigue life of the semi-rigid layer of the three pavements structure under full loaded truck moving at a different speed: (a) Asphalt layer and (b)
Semi-rigid layer.

semi-rigid layer of the three pavements can be ordered from lar- pavement test was used to verify and validate the level of accuracy
gest to smallest as follows: STR-C < STR-B < STR-A. of the developed 3D-FE model for the prediction of the pavement
temperature field and the mechanical response under thermal
6. Summary and conclusions and dynamic loads. Overall, the main findings of this numerical
investigations led to the following conclusions:
This study investigated the mechanical response of three new
semi-rigid base asphalt pavements under traffic load and nonlinear (1) The result of the 3D transient thermal analysis and the vari-
temperature gradient. Typical semi-rigid base asphalt pavement ation in pavement temperature observed in the field at dif-
structures were produced which was specifically designed to mit- ferent depths of the pavement, indicates that the model is
igate various functional and structural problems that traditional able to simulate with good accuracy, the time-dependent
pavement structures tend to exhibit. For each type of pavement temperature profile of the road system under a daily tem-
structure, the distribution of mechanical responses within the perature variation. This will assist pavement engineers in
pavement system was evaluated using a model of thermal and the design process. In addition, the thermal properties
mechanistic finite elements coupled in 3D. The models integrated derived from the Chinese specifications for highway asphalt
mechanical and thermophysical properties of each pavement’s pavement design adopted in the model yielded an excellent
materials, examined the effect of transient dynamic wheels load, temperature prediction.
and utilized transient heat transfer and implicit dynamic analysis. (2) The dynamic response of pavement to vehicular loading and
The comprehensive data recorded from the full-scale RIOH track non-linear temperature gradient at the bottom of the
24 O.C. Assogba et al. / Construction and Building Materials 235 (2020) 117406

asphalt layer and the semi-rigid layer were predicted and there are some limitations. In order to achieve a more accurate
the results are consistent with those recorded from RIOH- evaluation of the pavement response and its performance, it is rec-
track full-scale test facility. The developed FE model is there- ommended that future research should consider further improving
fore applicable for predicting the pavement response under the mechanistic pavement modeling. Albeit the developed 3D FE
the effect of both traffic load and non-linear temperature model predicts the mechanistic response with great accuracy,
gradient. however, some assumptions were inevitable. For instance: 1) the
(3) The analysis of the temperature of the three pavements asphalt concrete had a linear viscoelastic behavior. This does not
revealed that the surface temperature of the pavement and entirely reflect the actual dynamic properties of the asphalt
the atmospheric temperature were strongly correlated, concrete; 2) the behavior of the non-asphaltic layers was assumed
while the temperature inside the pavement structure linear and elastic in an isotropic system. This may not be represen-
decreased with depth. The temperature variation within tative of the stress-dependent state of these layers at some stage;
the pavement structure, particularly in the asphalt layer, 3) the tire-pavement contact stresses being uniformly distributed
was non-linear. This non-linearity of the thermal gradient on a rectangular equivalent areas may not entirely be indicative
causes variation of stiffness modulus as a function of depth, of realistic loading condition. Thus, the authors recommend afore-
which directly influences the response and performance of mentioned concerns for further investigations in future studies.
the road system. Therefore, the assumption of constant tem-
perature for these layers can be misleading for a realistic 8. Data availability statement
prediction of mechanical response, which are essential for
future design of more durable pavement structures. The numerical model and subroutine code (FILM, DFLUX, and
(4) Analysis of the mechanical response of the three pavements DLOAD) generated in this study are available from the correspond-
under the combined effect of axle load and temperature gra- ing author upon reasonable request. Also, the data supporting the
dient indicates that the non-linear temperature gradient plots and tables in this paper, as well as other findings of this
over the depth of the pavement structure significantly study, are available from the corresponding author upon reason-
affects the magnitude and the distribution of stresses and able request.
strain in the pavement system. The results also show that
the distribution of stresses and strains in the pavement
Funding
structure is strongly affected by the contact condition of
the interfaces. It appears that improvement of the field inter-
This work was funded by the National Key of Research and
face bond between pavement layers led to a considerable
Development Plan under Grant number 2016YFE0202400 and the
reduction in the mechanical responses, which is beneficial
Natural Science Foundation of China under Grant number
for the service life of the pavement structure.
U163320005.
(5) Analysis of the effects of vehicle speed and axle load on the
mechanical response of the pavement revealed that lowest
traffic speed and the extra weight extension of the truck axle Disclosure statement
have a significant influence on the distribution of the
mechanical response within the system. This may affect its No potential conflict of interest was reported by the authors.
structural behavior and, in turn, have adverse effects on its
long-term performance. Therefore, it may be appropriate, Acknowledgments
to regulate the lowest traffic speed and the load of the vehi-
cles against the overload, and to integrate the influence of The authors gratefully acknowledge the technical support
the latter in the design phase of the pavement structure. (equipment, software, etc.) provide by the Key Laboratory of Road
However, the tensile strain at the bottom of all proposed Structure and Materials, Research Institute of Highway Ministry of
new pavement structures are below the critical strain of a Transport, Beijing, China, and Harbin Institute of Technology. The
conventional pavement structure with an ordinary asphalt authors also express their sincere gratitude to all the people
(70 me) or with a mixture of modified asphalt (100 me), even involved in this research project. Finally, the authors would like
under critical conditions such as overloading or low-speed to thank reviewers for useful comments and editors for improving
traffic. Therefore, it is expected that the asphalt layer of the manuscript.
these typical semi-rigid pavements will be able to withstand
premature fatigue, especially bottom cracks References
(6) The analysis of the fatigue life of the three pavements
revealed that the fatigue life of the asphalt layer and the [1] D.M. Burmister et al., The theory of stress and displacements in layered
systems and applications to the design of airport runways, Highway Research
semi-rigid layer is strongly dependent on the speed of the Board Proceedings, 1944.
truck and the magnitude of load on the axles. As a result, [2] I.L. Al-Qadi, H. Wang, Evaluation of pavement damage due to new tire designs,
factors such as low-speed traffic or overloading could have Illinois Center for Transportation (ICT), 2009.
[3] R. Tautou, B. Picoux, C. Petit, Temperature influence in a dynamic viscoelastic
a negative impact on the service life of the pavement modeling of a pavement structure, J. Transp. Eng. Part B Pavements 143 (3)
structure. Under a fixed axle load and moving speed, the (2017) 04017012.
fatigue life of the asphalt layer of the three pavements are [4] O.C. Assogba et al., Effect of vehicle speed and overload on dynamic response of
semi-rigid base asphalt pavement, Road Mater. Pavement Des. (2019) 1–31.
as follows: STR-A < STR-B < STR-C; while that of the semi- [5] R. Imaninasab et al., Rutting performance of rubberized porous asphalt using
rigid layer of the three pavements are as follows: STR-C < Finite Element Method (FEM), Constr. Build. Mater. 106 (2016) 382–391.
STR-B < STR-A. [6] H.K. Shanbara et al., Predicting the rutting behaviour of natural fibre-
reinforced cold mix asphalt using the finite element method, Constr. Build.
Mater. 167 (2018) 907–917.
7. Limitation and further research direction [7] M. Li et al., Finite element modeling and parametric analysis of viscoelastic and
nonlinear pavement responses under dynamic FWD loading, Constr. Build.
This study extended the body of knowledge on the mechanical Mater. 141 (2017) 23–35.
[8] M. Kim, E. Tutumluer, J. Kwon, Nonlinear pavement foundation modeling for
response of semi-rigid pavements subjected to the combined effect three-dimensional finite-element analysis of flexible pavements, Int. J.
of non-linear thermal gradient and moving axle load, however, Geomech. 9 (5) (2009) 195–208.
O.C. Assogba et al. / Construction and Building Materials 235 (2020) 117406 25

[9] I.L. Al-Qadi, et al., Effects of tire configurations on pavement damage. 74(1) [26] X.D. Wang et al., Review of researches of RIOHTRACK in 2017, J. Highway
(2005) 921–961. Transp. Res. Develop. (2018).
[10] O. Gungor, et al., Quantitative assessment of effect of wide base tire on [27] J. Liao, S. Sargand, Viscoelastic FE modeling and verification of a US 30
pavement response using finite element analysis, 1990. perpetual pavement test section, Road Mater. Pavement Des. 11 (4) (2010)
[11] J. Ling et al., Analysis of airfield composite pavement responses using full-scale 993–1008.
accelerated pavement testing and finite element method, Constr. Build. Mater. [28] G. Wang, R. Roque, D. Morian, Evaluation of near-surface stress states in
212 (2019) 596–606. asphalt concrete pavement three-dimensional tire-pavement contact model,
[12] A. Sarkar, Numerical comparison of flexible pavement dynamic response Transp. Res. Rec. J. Transp. Res. Board 2227 (-1) (2011) 119–128.
under different axles, Int. J. Pavement Eng. 17 (5) (2016) 377–387. [29] I.V.L. Chango et al., Dynamic response analysis of geogrid reinforced
[13] I.L. Alqadi, M.A. Elseifi, Viscoelastic modeling and field validation of flexible embankment supported by CFG pile structure during a high-speed train
pavements, J. Eng. Mech. 132 (2) (2006) 172–178. operation, Latin Am. J. Solids Struct. 16 (7) (2019) p. NA-NA.
[14] B. Saad, H. Mitri, H. Poorooshasb, Three-dimensional dynamic analysis of [30] Y.H. Huang, Pavement Analysis and Design:United States Edition. Pearson
flexible conventional pavement foundation, J. Transp. Eng. 131 (6) (2005) 460– Schweiz Ag, 2003.
469. [31] S. Hu et al., Development of a viscoelastic finite element tool for asphalt
[15] D. Wang, J.R. Roesler, D.-Z. Guo, Analytical approach to predicting temperature pavement low temperature cracking analysis, Road Mater. Pavement Des. 10
fields in multilayered pavement systems, J. Eng. Mach. 135 (4) (2009) 334– (4) (2009) 833–858.
344. [32] M.J. Minhoto et al., The temperature effect on the reflective cracking of asphalt
[16] A. Mammeri et al., Temperature modelling in pavements: the effect of long- overlays, Road Mater. Pavement Des. 9 (4) (2008) 615–632.
and short-wave radiation, Int. J. Pavement Eng. 16 (3) (2015) 198–213. [33] M.L. Williams, R.F. Landel, J.D. Ferry, The temperature dependence of
[17] X. Gu, X. Liang, Q. Dong, Numerical Simulation of Long Term Pavement relaxation mechanisms in amorphous polymers and other glass-forming
Temperature Field, GeoShanghai International Conference, Springer, 2018. liquids, J. Am. Chem. Soc. 77 (14) (1955) 3701–3707.
[18] F.P. Incropera, D.P. DeWitt, Fundamentals of heat and mass transfer, 1996. [34] S. Park, R.A. Schapery, Methods of interconversion between linear viscoelastic
[19] M.N. Ozisik, Heat transfer: a basic approach, 1985. material functions. Part I—A numerical method based on Prony series, Int. J.
[20] E.S. Barber, Calculation of maximum pavement temperatures from weather Solids Struct. 36 (11) (1999) 1653–1675.
reports, Highway Research Board Bulletin (168) (1957). [35] R. Schapery, S.W. Park, Methods of interconversion between linear viscoelastic
[21] Y. Zuoren, Analysis of the temperature field in layered pavement system, J. material functions. Part II—an approximate analytical method, Int. J. Solids
Tongji Univ. 3 (1) (1984) 76–85. Struct. 36 (11) (1999) 1677–1699.
[22] A.U. Manual , Abaqus Theory Guide. Version 6.14, Dassault Systemes Simulia [36] W. Xu-dong, Design of pavement structure and material for full-scale test
Corp, USA, 2014. track, J. Highway Transp. Res. Develop. 34 (6) (2017) 30–37.
[23] I. Alqadi et al., Dynamic analysis and in situ validation of perpetual pavement [37] D.E. Newcomb, M. Buncher, I.J. Huddleston, Concepts of perpetual pavements,
response to vehicular loading, Transp. Res. Rec. J. Transp. Res. Board 2087 Transp. Res. Circ. 503 (2001) 4–11.
(2087) (2008) 29–39. [38] V. Quintus, L. Harold, Application of the endurance limit premise in
[24] X.G. Zhong, X. Zeng, J.G. Rose, Shear modulus and damping ratio of rubber- mechanistic: empirical based pavement design procedures, International
modified asphalt mixes and unsaturated subgrade soils, J. Mater. Civ. Eng. 14 Conference on Perpetual Pavements, Applied Research Associates Press,
(6) (2002) 496–502. Ohio, 2006.
[25] X.D. Wang, Design of pavement structure and material for full-scale test track, [39] A. Institute, Research and Development of the Asphalt Institute’s Thickness
J. Highway Transp. Res. Develop. 34 (6) (2017) 30–37. Design Manual, Asphalt Institute, 1982.

Вам также может понравиться