Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Construction and Building Materials 141 (2017) 2335

Contents lists available at ScienceDirect

Construction and Building Materials


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

Finite element modeling and parametric analysis of viscoelastic and


nonlinear pavement responses under dynamic FWD loading
Maoyun Li, Hao Wang , Guangji Xu, Pengyu Xie
Department of Civil and Environmental Engineering, Rutgers, The State University of New Jersey, United States

h i g h l i g h t s

 Develop finite element models to simulate FWD loading on pavement system.


 Effects of dynamic analysis, temperature, bedrock depth, delamination, and material models.
 Better understanding of using FWD for in-situ pavement condition evaluation.

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

Article history: Falling Weight Deflectometer (FWD) is the common non-destructive testing method for in-situ evalua-
Received 17 May 2016 tion of pavement condition. This paper aims to develop finite element (FE) models that can simulate
Received in revised form 24 January 2017 FWD loading on pavement system and capture the complexity in material properties, layer interface,
Accepted 19 February 2017
and boundary conditions. Parametric analysis was conducted considering the effects of dynamic analysis,
Available online 8 March 2017
temperature gradient, bedrock depth, asphalt layer delamination, viscoelasticity, and unbound material
nonlinearity on pavement surface deflections and critical strain responses. Although the parametric anal-
Keywords:
ysis findings vary depending on the specific pavement response, the study results illustrate the appropri-
Falling Weight Deflectometer
Surface deflection
ate selection of analysis type, constitutive models of pavement material, and layer boundary conditions
Finite element modeling on the accuracy of FE modeling results. In particular, the analysis findings show that delamination in
Strain responses asphalt layers induces the greater strain responses; while neglecting bedrock effect overestimates surface
deflections. The developed FE models can directly benefit the use of FWD testing for in-situ pavement
condition evaluation, such as pavement performance prediction and/or backcalculation of layer moduli.
2017 Elsevier Ltd. All rights reserved.

1. Introduction through iterations. To overcome the limitations pertaining to sim-


ple modeling assumptions, finite element (FE) models were devel-
Falling Weight Deflectometer (FWD) is the common non- oped considering more realistic loading and materials assumptions
destructive testing method for in-situ evaluation of pavement con- in simulation of FWD testing.
dition. The surface deflections measured under FWD testing are Previous researches mainly focused on one or two factors
used to assess structure capacity of pavement system and backcal- among the interaction between environmental, material, and
culate the moduli of pavement layers. Therefore, accurate mecha- structural behavior under impulsive dynamic loading applied in
nistic models are needed to analyze FWD testing results, FWD testing. The dynamic analysis approach was used with con-
considering the interaction between loading, environmental, and sideration of viscoelastic asphalt layer to simulate the impulsive
pavement structure. Analytical and numerical models have been FWD loading and wave propagation in the pavement structure
used for years to simulate FWD testing and predict pavement [1,2]. The simulation results showed better agreements in match-
responses. In the early studies, multi-layer models with assumed ing the field FWD measurements as compared to non-dynamic
linear elastic moduli for all layers were developed for analysis of analysis or elastic layer assumption. The dynamic behavior of
pavement structural behavior and backcalculation of layer moduli FWD can be magnified with the existence of bedrock as a rigid
layer underlying the subgrade [3]. Few researches have considered
Corresponding author. the effect of boundary condition due to bedrock and investigated
E-mail addresses: ml929@scarletmail.rutgers.edu (M. Li), hwang.cee@rutgers. the threshold depth after that there was no negative effect of bed-
edu (H. Wang), ml929@scarletmail.rutgers.edu (G. Xu), px19@scarletmail.rutgers. rock on pavement surface deflections.
edu (P. Xie).

http://dx.doi.org/10.1016/j.conbuildmat.2017.02.096
0950-0618/ 2017 Elsevier Ltd. All rights reserved.
24 M. Li et al. / Construction and Building Materials 141 (2017) 2335

" #
The correlation of asphalt concrete layer modulus to the tem- X
n

perature along the layer depth was investigated. The results indi- Kt K 0 1 K i 1  et=si 2
i1
cated that the mid-depth temperature can be considered
representing the in-depth temperature distribution for the predic- where,
tion of elastic modulus [4]. However, in consideration of viscoelas- G is shear modulus;
tic property of asphalt layer, temperature profile in positive or K is bulk modulus;
negative distributions can cause different stress and strain distri- t is relaxation time;
butions as compared to the constant temperature profile [5]. The G0 and K0 are instantaneous shear and volumetric elastic mod-
FWD testing results were also affected by the behavior of base uli; and
layer and subgrade. It was reported that neglecting nonlinearity Gi, Ki, and si are Prony series parameters.
of unbound material properties produced substantial errors in cal-
culating the responses of flexible pavement [6,7]. The BELLS3 equation, which was validated with measurements
Another significant issue related to the existing pavement con- from field sections in the Long Term Pavement Performance (LTPP)
dition is the degradation of interlayer bonding between asphalt study, was used to predict temperatures within asphalt layer, as
layers. An investigation on several newly constructed asphalt pave- shown in Eq. (3) [12]. Compared to other temperature prediction
ments was carried out to study the effect of interlayer debonding models developed by Huber [13] and Park [14], the BELLS3 equa-
behavior on back-calculation of pavement layer moduli [8]. The tion predicted the distributed temperatures that were closest to
effect of layer debonding could become more serious in the com- the measured ones [15].
posite pavement. It was found that the broken bonds between
asphalt overlay and existing concrete slabs could cause an oscilla- T d 0:95 0:892IR logd  1:250:448IR
tion of the deflection-time history due to the vibrating behavior 0:6211  day 1:83sinhr18  15:5
experienced by the asphalt overlay [9]. Therefore, to better under-
0:042IR sinhr 18  13:5 3
stand pavement behavior under FWD loading, all the affecting fac-
tors in FWD testing need to be considered in the simulation models where,
in order to enhance the accuracy modeling results. Td is pavement temperature at depth d, C;
IR is pavement surface temperature, C;
2. Objectives d is depth at which the temperature is to be predicted, mm;
1-day is average air temperature the day before testing, C;
The objectives of this study have two folds. The first objective is sin is sine function on an 18-h clock system, with 2p radians
to develop finite element (FE) models that can simulate FWD load- equal to a 18-h cycle; and
ing on pavement system and capture the complexity in material hr18 is time of day, in a 24-h clock system, but calculated using
properties, layer interfaces, and boundary conditions. The second an 18-h AC temperature rise-and-fall time cycle.
objective is to analyze the important factors that should be consid-
ered in the FE analysis, including dynamic analysis, temperature 3.2. Nonlinear modulus of unbound material
gradient, depth to bedrock, asphalt layer delamination, viscoelas-
ticity of asphalt layer, and nonlinearity of unbound material. The The unbound layers including aggregate base and subgrade
development of FE models can directly benefit the use of FWD test- were both modeled as the nonlinear materials. A generalized
ing for in-situ pavement condition evaluation, such as pavement model was introduced to express the vertical modulus of unbound
performance prediction and backcalculation of layer moduli. layer material, as shown in Eq. (4) [16]. Moreover, the horizontal
and shear modulus ratios (n and m) were introduced for the aggre-
3. Finite element modeling gate base layer for cross-anisotropic behavior. It was found that
horizontal modulus ratio and shear modulus ratio have relatively
3.1. Viscoelastic modulus of asphalt layer small ranges of variation. Therefore, typical values of 0.15 and
0.34 for n and m in aggregate base layer were used in this study
Constitutive models of each pavement layer are critical for [17]. After consideration of stress-dependency of unbound mate-
mechanistic analysis of pavement responses. The time- or rial, the material modulus can change as the stress state in the
frequency- and temperature-dependent behavior of asphalt layer material changes due to different factors (such as load, layer thick-
was considered with the viscoelastic model in this study. The ness, temperature, etc.)
relaxation modulus of asphalt mixture was modeled as a general-  k2  k3
ized Maxwell solid model in terms of Prony series, as shown in Eqs. h soct
Mvr k1 pa 1 4
(1) and (2). The relaxation modulus can be obtained from pa pa
laboratory-tested creep compliance or dynamic modulus by means q
of an inter-conversion relationship [10]. The constant Poissons with h r1 r2 r3 and soct 13 r1  r2 2 r2  r3 2 r1  r3 2
ratio of 0.35 was assumed in the calculation of shear and bulk where,
modulus from relaxation modulus. The temperature dependency M vr is vertical resilient modulus (kPa);
of AC modulus is characterized by time-temperature superposition
h is bulk stress (kPa);
principle. This behavior introduces the horizontal shifting of the
soct is octahedral shear stress (kPa);
material property to form a single characteristic master curve as
r1, r2, and r3 are maximum, middle, and minimum principal
a function of reduced time (or frequency) at a desired reference
stresses;
temperature. The relationship between the shift factor and the
k1, k2, and k3 are exponent parameters; and
temperature can be approximated by the WLF (Williams-Landell-
pa is atmospheric pressure (100 kPa).
Ferry) function [11].
" #
X
n ABAQUS/Standard uses the iterative Newton-Raphson method
Gt G0 1  Gi 1  et=si 1 to solve nonlinear equations. The applied load in this method is
i1
augmented incrementally, and at each increment the program
M. Li et al. / Construction and Building Materials 141 (2017) 2335 25

solves a system of equations through iterations. The iterations con- where,


tinue on the basis of the previous solutions until it reaches a rea- [M] is mass matrix;
sonable convergence [11]. Because the modulus of the unbound [C] is damping matrix;
material is a function of the total stress states, a modified [K] is stiffness matrix;
Newton-Raphson approach with secant stiffness was used in this is acceleration vector;
fUg
study and implemented in a user material subroutine (UMAT), _ is velocity vector;
fUg
which was developed in the authors previous work [18]. {U} is displacement vector; and
{P} is displacement vector.
3.3. Dynamic analysis

Static, quasi-static, and dynamic analysis have been employed 3.4. Finite element model development
in different studies for pavement analysis. Dynamic transient anal-
ysis allows for the inertia associated with the FWD loading and the The modeled pavement structure was constructed with three
dependency of viscoelastic material properties on loading fre- structure layers, as shown in Fig. 1. It consists of a 150-mm asphalt
quency. For a nonlinear dynamic analysis problem, the direct inte- pavement placed on a 300-mm aggregate base layer over subgrade.
gration method is commonly employed. Therefore, the implicit The master curve of dynamic modulus and temperature shift fac-
dynamic analysis was selected in this study because it provides tors were obtained from the prediction model that was developed
better numerical stability than explicit analysis and is generally from the LTPP study using the asphalt binder type of PG 64-22 and
efficient for structural dynamic problems [19]. 4% air void [7]. The material parameters for base aggregate and soil
The equation of motion of a multi-degree of freedom system subgrade were obtained from the laboratory-developed empirical
with damping is shown in Eq. (5). When using viscoelastic material equations that related the physical properties of unbound material
behavior for an asphalt layer, it is not necessary to introduce addi- to nonlinear modulus parameters [21,22]. Three groups of param-
tional structural or mass damping for that layer. A popular spectral eters were selected to represent the variations of nonlinear modu-
damping scheme used in structure dynamic analysis is Rayleigh lus. Pavement temperature profiles were characterized using the
damping. For the dynamic loading on a pavement structure, the BELLS model and the climate data extracted from the LTPP data-
two frequencies in calculating the Rayleigh coefficients may be base [23].
taken as the lowest natural frequency of the structure and the An axisymmetric 2-D FE model of the thin asphalt pavement
highest loading frequency. The critical damping ratio of soil falls was developed by applying the general-purpose software ABAQUS.
in the range of 2% to 5% depending on the confining pressure and The 2-D FE model is more appropriate for simulating the FWD test-
shear strain level, which can be measured from the resonant col- ing in comparison to the 3-D FE model. It can better evaluate a real
umn test or cyclic triaxial test [20]. Because the subgrade is circular load on the FWD-pavement interface, because the axisym-
considered as elastic material without any other energy metric elements are capable of shaping an absolute circular load
dissipation sources (such as plastic), the maximum damping ratio, other than the elements in 3-D FE model. In this study, four-
5%, is used. node bilinear axisymmetric solid elements were used in the finite
CfUg
_ KfUg fPg domain; while axisymmetric infinite elements were applied to
MfUg 5
reduce a large number of far-field elements without significant loss

Fig. 1. Flexible pavement structure with material properties of different layers.


26 M. Li et al. / Construction and Building Materials 141 (2017) 2335

of accuracy and then to build a silent boundary for the dynamic 2.5  8 m to achieve the balance between computation cost and
analysis. accuracy.
Fig. 2 presents an axisymmetric 2-D FE model that discretizes Contact conditions at layer interfaces are important parameters
the pavement structure. The FWD loading was represented by a that could significantly affect pavement responses under FWD
circular impulse loading applied on the surface of asphalt layer. loading. It is expected that the layers within the pavement struc-
The FE mesh is refined around the circular loading area; instead ture remains in contact with no gap-opening since the contact area
a relatively coarse mesh is applied far away from the loading area. is very large and compressive loading due to gravity and traffic
The length of elements within the loading area is selected at loading. Therefore, it is reasonable to assume that both relative
12.7 mm in the radial direction. The element thickness is selected and absolute motions of contacting surfaces at layer interfaces
to be 8.5 mm for the asphalt layer. The selection for the element are small. In this study, the Coulomb friction model was used with
sizes is based on the mesh convergence analysis conducted in a friction coefficient of 1.0 for the HMA-base layer interface and 0.3
the early study [24]. for the base-subgrade interface [25].
A sensitivity analysis was performed to find the required finite
domain size in terms of radial length and depth, as shown in Fig. 3. 3.5. Model validation
After comparing the maximum central and edge deflections with
different lengths of finite domain, the location of the infinite A validation study for the developed FE model was conducted
boundary from the loading center need to be greater than 2.2 m with field measurement results under FWD loading. A 127-mm
in the radial direction and 7.5 m in the depth (vertical direction) asphalt pavement with a 300-mm aggregate base layer that was
in order to obtain a stable solution (less than 5% changes). The final used in the field study was modeled to simulate the in-situ condi-
domain size of the axisymmetric model was selected to be of tion. The material properties were obtained from laboratory test-

Fig. 2. FE model layout in axisymmetric 2-D domain.


M. Li et al. / Construction and Building Materials 141 (2017) 2335 27

Fig. 3. Sensitivity analysis for (a) central deflection vs. radial length of finite domain; (b) edge deflection vs. radial length of finite domain; (c) central deflection vs. depth of
finite domain; and (d) edge deflection vs. depth of finite domain.

Fig. 4. Validation of FE model with measurement under FWD loading of (a) 53 kN; (b) 40 kN; and (c) 27 kN.

ing and pavement temperature profiles were recorded in the tions under three different loading levels. The results show that
asphalt layer [24,26]. The temperature in the asphalt layer is in with the FE model results have good agreements with the deflec-
the range of 2023 C (6873.4F). These data were directly used tion measurements at different sensor locations.
in the FE model as inputs and the accuracy of prediction results In order to illustrate the importance of considering dynamic
as compared to field measurements was examined for model vali- analysis and appropriate constitutive models in the simulation of
dation. Fig. 4 compares the measured and predicted surface deflec- FWD loading, the calculation results using different model
28 M. Li et al. / Construction and Building Materials 141 (2017) 2335

Table 1
Comparisons of calculated and measured deflections using different model assumptions.

Case No. Dynamic analysis Viscoelasticity Nonlinearity Temperature gradient RMSE between calculation and measurement (mm)
* * * *
1 N Y Y Y 0.15
2 Y* N* Y* Y* 0.28
3 Y* Y* N* Y* 0.16
4 Y* Y* Y* N* 0.08
5 Y* Y* Y* Y* 0.05
*
Y means that the factor is considered in the model; N means that the factor is not considered.

Fig. 5. Comparison of dynamic and quasi-static analysis through (a) deflection-time histories; (b) strain-time histories; (c) D0 hysteresis loops; and (d) D900 hysteresis loops.

assumptions were compared to the field measurements. Table 1 deflections can represent structure behavior of the entire pave-
summarizes the root mean squared error (RMSE) between the ment system and the subgrade, respectively [27]. The critical
measured and predicted surface deflections using different model strains analyzed included tensile strains at the bottom of HMA
assumptions. For each affecting factor, simple model assumptions layer, compressive strains on the top of subgrade. It has been
were used for comparison, such as quasi-static analysis, linear elas- widely accepted that critical tensile strains in HMA layer are
ticity, and average temperature. The results show that the full con- related to bottom-up fatigue cracking distress and compressive
sideration of dynamic analysis, viscoelasticity and nonlinearity, strains in subgrade are attributed to rutting failure. In addition,
and temperature gradient results in the smallest discrepancy with the load-deflection curve was used to indicate the energy dissipa-
field measurements. Instead, simple model assumptions cause the tion of pavement system in terms of the hysteresis loop. The hys-
higher errors, especially when the viscoelasticity of asphalt mix- teresis loop was represented using the FWD loading-deflection
ture was not considered. (vertical displacement) rather than stress-strain loop, because the
deflection can better represent the integrity of the pavement sys-
tem than the critical strains [28].
4. Results and analysis

The effects of primary factors on pavement responses under 4.1. Effect of dynamic analysis
FWD loading were investigated using the developed FE model.
Both the surface deflections and critical strains were considered Fig. 5 shows the comparisons between dynamic analysis and
in the analysis. Two specific deflection indicators D0 (deflection quasi-static analysis through the deflection-time and strain-time
under central load) and D900 (deflection 900 mm away from cen- histories as well as the hysteresis loops of D0 and D900. The load-
tral load) were primarily used in the analysis. The two typical time history is plotted to indicate the time shift between the
M. Li et al. / Construction and Building Materials 141 (2017) 2335 29

Fig. 6. Comparison of viscoelasticity and elasticity through (a) deflection-time histories; (b) strain-time histories; (c) D0 hysteresis loops; and (d) D900 hysteresis loops.

applied load and resulting responses. It is believed that the time lag material damage. Fig. 5(c) shows that the dissipated energy in D0
here is due to the viscoelastic and damping behaviors as well as the using dynamic analysis is higher than the one using quasi-static
stress wave propagation from loading center. As expected, all the analysis. It reveals that more energy is dissipated under the
response-time history curves in Fig. 5 (a) and (b) resemble the dynamic loading because of the damping behavior. The energy dis-
shape of the load-time history. It was found that the time lags to sipation under quasi-static analysis can be attributed to the vis-
the load-time history were obvious in the D0- or D900-time histo- coelasticity of asphalt layer. In the case of Fig. 5(d) shows that
ries under the dynamic analysis as compared to quasi-static analy- the discrepancy of energy dissipation using two analysis methods
sis. Moreover, the magnitudes of D0 or D900 under dynamic is more significant for D900. This is mainly because the dynamic
analysis are larger than the ones under quasi-static analysis. It is analysis incorporated the inertia force and mass damping into
worth mentioning that the D900 has a longer time lag than D0 in the equation of motion (as shown in Eq. (5)). Thus, these findings
the displacement-time history. As mentioned before, the D0 repre- emphasize that the influence of dynamic analysis cannot be
sents the behavior of the entire pavement structure but D900 is ignored for FWD loading.
uniquely attributed to the subgrade behavior. Therefore, the differ-
ent time lags imply that the stress wave propagation plays a dom- 4.2. Effect of HMA viscoelasticity
inating role in addition to the viscoelastic behavior of asphalt layer
and the damping effect of unbound layers. Fig. 6 shows the comparisons between the viscoelastic asphalt
On the other hand, as shown in Fig. 5 (b), no time lags were material and elastic asphalt material through the deflection- and
observed for critical strains under both analysis approaches. How- strain-time histories as well as hysteresis loops of D0 and D900.
ever, the critical strains under dynamic analysis are slightly greater The elastic modulus of asphalt layer was selected according to the
than the ones under quasi-static analysis. However, the discrepan- dynamic modulus at the loading frequency of 30.3 Hz that is calcu-
cies of strains under two analysis methods were not as significant lated from the pulse duration of FWD loading. In Fig. 6(a), the longer
as the deflections. This indicates that the difference of both analy- time lag and greater magnitude of D0 were observed for viscoelastic
sis approaches casts more effect on the back-calculation of the material; while the time lag and magnitude of D900 were similar
layer moduli through FWD deflections rather than on the assess- using both material models. As shown in Fig. 6(b), no time lags were
ment of pavement performance through critical responses. found for critical strains using both material models, although the
Fig. 5(c) and (d) show the hysteresis loops in terms of D0 and magnitudes of critical strains in the elastic asphalt material are
D900 to indicate the energy dissipation resulted from the FWD greater than the ones in the viscoelastic one. It is noted that the
loading. If the loading and unloading paths coincide, it means that assumption of elastic asphalt layer may underestimate or overesti-
all the strain energy caused by the load is recovered after unload- mate pavement responses, depending on the selection of loading
ing. If not, the area between the loading and unloading curves indi- frequency; therefore, it cannot capture the frequency-dependent
cates the dissipated energy due to the viscoelasticity, damping, or response of viscoelastic material.
30 M. Li et al. / Construction and Building Materials 141 (2017) 2335

Fig. 7. Effect of temperature gradient on (a) deflection-time histories at the intermediate temperature; (b) strain-time histories at the intermediate temperature; and (c)
deflection-time histories at the high temperature; and (d) strain-time histories at the high temperature.

Fig. 6(c) and (d) show the hysteresis loops in terms of D0 and 37.5 C (99.5F) and 25.2 C (77.4F) for the high temperature case,
D900. It shows that the dissipated energy of D0 in the viscoelastic respectively.
asphalt material is higher than the ones in the elastic asphalt Fig. 7 illustrates the comparisons of the deflection-time and
material, while the discrepancy between the two material models strain-time histories using two different temperature gradients
in terms of D900 is negligible. This is because that the energy and the corresponding average constant temperatures. In Fig. 7
dissipation in terms of D900 is mainly affected by the damping (a) and (c), the magnitudes of D0 or D900 under different temper-
of subgrade; while the energy dissipation in terms of D0 is ature assumptions were found close to each other at both cases.
affected by both the viscoelastic asphalt layer and the damping Similarly, in Fig. 7(b) and (d), for the compressive strains on the
of unbound materials. The dissipated energy of viscoelastic mate- top of subgrade, no significant discrepancies of the strain ampli-
rial is due to the time lag between stress and strain that is usu- tudes were found with regard to the effect of temperature gradi-
ally described through the phase angle in dynamic modulus ents. However, it was found that tensile strains at the bottom of
testing. It is noted that for elastic pavement system without asphalt layer under the constant average temperature gradient
damping, the induced deflection recovers immediately after the were slightly greater than the ones under the nonlinear tempera-
load pulse reaches zero. This will cause the same load- ture gradient. This is because the temperature at the bottom of
deflection curves between loading and unloading processes, asphalt layer is higher when the average temperature assumption
which means zero energy dissipation. is used. The effect of temperature gradient on tensile strains
became more significant at the high temperature case. This indi-
cates that the assumption of average constant temperature may
be valid for calculating surface deflections but not for predicting
4.3. Effect of temperature gradient
bottom-up fatigue cracking potential of the relatively thin asphalt
pavement.
The assumption of temperature profiles in the asphalt layer can
also affect viscoelastic modulus of asphalt layer and accordingly
pavement responses. Two temperature conditions were considered 4.4. Effect of asphalt layer delamination
to evaluate the effect of in-depth nonlinear temperature gradient
on pavement deflections and critical strains as compared to the Although the asphalt layer is designed to be full-bonded, layer
constant temperature gradient using average temperature. The delamination could happen in the field due to poor construction
pavement surface temperature and average air temperature in quality. Fig. 8 illustrates the effect of layer delamination within
the BELLS3 model (Eq. (5)) were selected as 23.5 C (74.3F) and the asphalt layer on deflection- and strain-time histories as well
14.2 C (57.6F) for the intermediate temperature case, and as hysteresis loops. The debonded behavior was defined in the FE
M. Li et al. / Construction and Building Materials 141 (2017) 2335 31

Fig. 8. Effects of layer delamination on (a) D0-time history; (b) D900-time history; (c) tensile strain-time history; (d) compressive strain-time history; (e) hysteresis loop of
D0; and (f) hysteresis loop of D900.

model by dividing the asphalt layer into two sub-layers (upper and This can be attributed to the reasons that the bottom of upper layer
lower) with frictional interaction at the mid-depth of 75 mm. with the debonded asphalt layer bears the most flexural deforma-
As expected, Fig. 8(a) shows that deflection-time histories of D0 tion; while the bottom of upper layer with the fully-bonded asphalt
with fully-bonded asphalt layer are much smaller than the ones with layer experiences approximately zero strain on the neutral axis of
debonded asphalt layer. However, Fig. 8(b) indicates that no differ- bending. The layer debonding also causes the decrease of tensile
ence can be found in D900 between the two assumptions of bonding. strain at the bottom of lower layer as compared to the fully bonded
It is reasonable because D0 represents the entire pavement behavior case. Fig. 8(d) presents that the compressive strains in the debonded
so that the layer debonding causes the larger magnitude of D0. case are much greater compared to the fully bonded case. Therefore,
Fig. 8(c) shows the tensile strain-time histories under two bond- the delamination behavior can induce premature failure in fatigue
ing assumptions of the asphalt layer. The tensile strains at different cracking initiating at the bottom of asphalt upper layer or the perma-
depths were used to study the delamination effect. The critical ten- nent deformation in the subgrade. These observations signify that
sile strain at the bottom of upper layer with the deboned asphalt the asphalt layer bonding condition should be considered in FE mod-
layer has the maximum magnitude among all scenarios. As eling for prediction of strain responses.
expected, the strains at the bottom of upper layer with the fully- Figs. 8(e) and (f) show the hysteresis loops in terms of D0 and
boned asphalt layer are around zero or have small compressions. D900 in consideration of the asphalt layer delamination. It shows
32 M. Li et al. / Construction and Building Materials 141 (2017) 2335

Fig. 9. Effect of nonlinear model of aggregate base on (a) D0 and D900; (b) critical strains; (c) hysteresis loop of D0; and (d) hysteresis loop of D900.

Fig. 10. Effect of nonlinear model of subgrade on (a) D0 and D900; (b) critical strains; (c) hysteresis loop of D0; and (d) hysteresis loop of D900.
M. Li et al. / Construction and Building Materials 141 (2017) 2335 33

Fig. 11. FWD deflection variations under multiple loadings considering nonlinear unbound material for (a) D0 and (b) D900.

that the energy dissipation of D0 in the full-bonded case is differ- elastic moduli for both base aggregate layer and subgrade were cal-
ent from the one in the debonded case. However, the energy dissi- culated as the average moduli from the modulus distribution
pation of D900 in both cases keeps approximately similar to each throughout the base layer and subgrade, respectively. As shown in
other. In general, the results indicate that the layer debonding Fig. 9(a) and (b), the linear and nonlinear isotropic models for base
could be detected through both the magnitude and energy dissipa- aggregate underestimated the deflections and critical strains in
tion of D0 if in-situ testing data are available. comparison to the nonlinear cross-anisotropic model. This trend is
similar for the effect of nonlinear subgrade model, as shown in
4.5. Effect of unbound material nonlinearity Fig. 10(a) and (b). Therefore, it is concluded that ignoring the nonlin-
earity of the unbound materials can lead to adverse effects on the
Fig. 9 and Fig. 10 show the comparison of different unbound modulus back-calculation and pavement performance assessment.
material models through the deflections and critical strains as well Fig. 9 and Fig. 10(c) and (d) present the hysteresis loops using
as hysteresis loops of D0 and D900. In the reference case, the linear different unbound material models in terms of D0 and D900. The

Fig. 12. Effects of bedrock on FWD deflections for (a) D0-time history; (b) D900-time history; (c) sensitivity of D0 to the depth to bedrock; and (d) sensitivity of D900 to the
depth to bedrock.
34 M. Li et al. / Construction and Building Materials 141 (2017) 2335

dissipated energy of D0 and D900 in the linear model is the lowest the parametric analysis findings vary depending on the response
in general. It was found that the dissipated energy of D0 was of interest, the study results illustrate the appropriate selection
mainly affected by the nonlinear model of aggregate base; while of analysis type, constitutive models of pavement material, and
the dissipated energy of D900 was mainly affected by the nonlinear layer boundary conditions on the accuracy of FE modeling. In par-
model of subgrade. Therefore, the linearity assumption for both ticular, the analysis findings show the delamination in asphalt
unbound materials might result in inaccurate results if energy dis- layer induces the greater strain responses; while neglecting bed-
sipation is used as an analysis indicator. rock effect overestimates surface deflections. These two factors
It is expected that the effect of nonlinearity is affected by the were usually neglected in the previous analysis work.
thickness of asphalt layer or the loading magnitude. As the asphalt The developed FE model successfully captures the distinctive
layer thickness decreases or the load magnitude increases, more constitutive model for each pavement layer and the interaction
stresses are transmitted into the base layer and subgrade. This between different layers and boundary conditions. With the proper
may signify the modulus dependency of unbound material on the development of FE models, further research will be conducted for
stress state. In order to investigate the stress-dependent behavior backcalculation of layer moduli using the combination of synthetic
of unbound material, multiple loading levels were considered in database from the forward FE analysis and artificial intelligence
the FE analysis. The stiffness sensitivity analysis to the loading techniques.
levels was conducted using the ratio of loading to deflection (D0
and D900), as shown in Fig. 11. This ratio would be equal to a con-
Acknowledgement
stant value under different loading levels or with different pave-
ment structures if the linear material properties are used in the
This study was partially funded by the USDOT Region II Univer-
FE model. The positive slope indicates that the nonlinear stress
sity Transportation Research Center.
hardening behavior is dominant, while the negative slope indicates
that the stress softening behavior of subgrade is dominant. The
greater slope would be observed if the stress-dependency of mod- References
ulus is more significant. Fig. 11 indicates that D0 is affected by the
nonlinearity in base layer and subgrade, although with different [1] I.L. Al-Qadi, H. Wang, E. Tutumluer, Dynamic analysis of thin asphalt
pavements by using cross-anisotropic stress-dependent properties for
trends. However, D900 is mainly governed by the nonlinearity in granular layer, Transp. Res. Rec. 2154 (2010) 156163.
subgrade that shows significant stress-softening behavior. [2] Q. Xu, J.A. Prozzi, Static versus viscoelastic wave propagation approach for
simulating loading effects on flexible pavement structure, Constr. Build. Mater.
53 (2013) 584595.
4.6. Effect of bedrock depth [3] M. Broutin, Assessment of Flexible Airfield Pavements using Heavy Weight
Deflectometers (Ph.D. dissertation), LCPC, Paris, 2010.
[4] H.M. Salem, F.M. Bayomy, M.G. Al-Taher, I.H. Genc, Using long-term pavement
Fig. 12 shows the effect of stiff layer (bedrock) underlying sub- performance data to predict seasonal variation in asphalt concrete modulus,
grade on surface deflections through the deflection-time histories Transp. Res. Rec. 2004 (1896) 119128.
and the sensitivity of D0 and D900 to the depth to bedrock. The [5] Y. Lu, P.J. Wright, Y. Zhou, Effect of temperature and temperature gradient on
asphalt pavement response, Road Transport Res. 18 (1) (2009) 1930.
selected stiffness of bedrock is based on typical rock properties [6] C.W. Schwartz, Effect of stress-dependent base layer on the superposition of
with an elastic modulus of 7000 MPa and a Poissons ratio of 0.2. flexible pavement solutions, Int. J. Geomech. 2 (3) (2002) 331352.
As shown in Fig. 12(a) and (b), deflection-time histories of D0 [7] M. Kim, E. Tutumluer, J. Kwon, Nonlinear pavement foundation modeling for
three-dimensional finite-element analysis of flexible pavements, Int. J.
and D900 sharply decreased if the bedrock exists at the depth of
Geomech. 9 (5) (2009) 195208.
3000 mm from pavement surface. It is noted that the variations [8] B.A. Hakim, L.W. Cheung, R.J. Armitage, Use of FWD Data for Prediction of
of strain-time histories for tensile and compressive strains are Bonding between Pavement Layers, Int. J. Pavement Eng. 1 (1) (2000) 4959.
[9] S.N. Shoukry, D.R. Martinelli, O.I. Selezneva, Dynamic performance of
insignificant (difference smaller than 3%); therefore they were
composite pavements under impact, Transp. Res. Rec. 1570 (1997) 163171.
not plotted here due to the reason of brevity. It indicates that the [10] S.W. Park, Y.R. Kim, Interconversion between Relaxation Modulus and Creep
existence of bedrock casts more effect on the back-calculation of Compliance for Viscoelastic Solids, J. Mater. Civil Eng. 11 (1) (1999) 7682.
layer moduli through FWD deflections than on the assessment of [11] ABAQUS, ABAQUS/Standard Users Manual, Version 6.10, Hibbitt, Karlsson &
Sorenson, Inc., Pawtucket, RI, 2010.
pavement performance through critical strains. [12] FHWA, Temperature Predictions and Adjustment Factors for Asphalt
Fig. 12(c) and (d) shows the sensitivity of D0 and D900 to the Pavement, FHWA-RD-98-085, Federal Highway Administration (FHWA),
depth to bedrock. It shows that the D0 is increased gradually as McLean, VA, 2000.
[13] G.A. Huber, Strategic Highway Research Program Report SHRP-A-648A:
the location bedrock becomes deeper; and the variation of the Weather Database for the Superpave Mix Design System, Transportation
D0 becomes stable (variation smaller than 5%) as the depth goes Research Board, Washington, DC, 1994.
beyond 8000 mm from pavement surface. However, the stability [14] D.Y. Park, N. Buch, K. Chatti, Effective Layer Temperature Prediction Model and
Temperature Correction via Falling-Weight Deflectometer Deflections, in:
for the D900 variation was not obtained until the depth reached Transportation Research Record, No. 1764, TRB, National Research Council,
to 10,000 mm. Given that the bedrock is closer to and then has Washington, DC, 2001, pp. 97111.
more effect on the subgrade behavior, this helps explain that the [15] D.S. Gedafa, M. Hossain, S.A. Romanoschi. Prediction of Asphalt Pavement
Temperature. Proceedings Airfield and Highway Pavement, Los Angeles, CA,
D900 is more sensitive than the D0 to the change in the depth to
2013, pp. 373382.
bedrock. The threshold range of 8000 mm and 10,000 mm obtained [16] Guide for Mechanistic-Empirical Design of New and Rehabilitated Pavement
from the sensitivity analysis were consistent with previous find- Structures, NCHRP 1-37A Final Report, TRB, ARA, Inc. ERES Division,
Washington, DC, 2004.
ings using field measured deflections [3].
[17] E. Tutumluer, M.R. Thompson, Anisotropic modeling of granular bases in
flexible pavement, Transp. Res. Rec. 1577 (1997) 1826.
[18] H. Wang, I.L. Al-Qadi, The importance of nonlinear anisotropic modeling of
5. Conclusions granular base for predicting maximum viscoelastic pavement responses under
moving vehicular loading, J. Eng. Mech. 139 (1) (2013) 2938.
[19] K.J. Bathe, Finite Element Procedures in Engineering Analysis, Prentice-Hall, NJ,
In this study, FE models were developed to simulate flexible
1982.
pavement behavior under impulsive FWD loading. Parametric [20] X.Z. Zhong, X. Zeng, J.G. Rose, Shear modulus and damping ratio of rubber-
analysis was conducted to analyze the primary factors affecting modified asphalt mixes and unsaturated subgrade soils, J. Mater. Civil Eng. 14
model results, including dynamic analysis, temperature gradient, (6) (2002) 496502.
[21] K.P. George, Prediction of Resilient Modulus from Soil Index Properties, Final
bedrock depth, asphalt layer delamination, viscoelasticity of Report to Federal Highway Administration, Mississippi Department of
asphalt layer, and nonlinearity of unbound materials. Although Transportation Research Division, Jackson, MS, 2004.
M. Li et al. / Construction and Building Materials 141 (2017) 2335 35

[22] Y.J. Xiao, E. Tutumluer, J. Siekmeier, Mechanistic-empirical evaluation of [26] J. Kwon, Development of a Mechanistic Model for Geogrid Reinforced Flexible
aggregate base and granular subbase quality affecting flexible pavement Pavements (Ph.D. dissertation), University of Illinois at Urbana-Champaign,
performance in Minnesota, Transp. Res. Rec. 2227 (2011) 97106. 2007.
[23] FHWA. LTPP InfoPave. https://infopave.fhwa.dot.gov/, 2016. [27] Y.H. Huang, Pavement Analysis and Design, 1st ed., Prentice Hall, Upper Saddle
[24] H. Wang, I.L. Al-Qadi, Impact quantification of wide-base tire loading on River, NJ, 1993.
secondary road flexible pavements, J. Transport. Eng. 137 (9) (2011) 630639. [28] K.A. Ghuzlan, S.H. Carpenter, Energy-Derived, Damage-Based Failure Criterion
[25] S.A. Romanoschi, J.B. Metcalf, Effects of Interface Condition and Horizontal for Fatigue Testing, in: Transportation Research Record: Journal of the
Wheel Loads on the Life of Flexible Pavement Structures, in: Transportation Transportation Research Board, No. 1723, TRB, National Research Council,
Research Record, No. 1778, TRB, National Research Council, Washington, DC, Washington, DC, 2000, pp. 141149.
2001, pp. 123131.

You might also like