You are on page 1of 12

International Journal of Pavement Engineering Vol. 11, No.

2, April 2010, 95105

Use of creep compliance interconverted from complex modulus for thermal cracking prediction using the M E pavement design guide
Hao Yina*, Ghassan R. Chehaba1, Shelley M. Stoffelsa2, Tanmay Kumarb3 and Laxmikanth Premkumara4
a

Department of Civil and Environmental Engineering, The Pennsylvania State University, University Park, PA, USA; bApplied Research Associates, Inc., Transportation, Sacramento, CA, USA ( Received 15 May 2007; nal version received 12 November 2008 ) The objective of this study is to evaluate the creep compliance (D(t)) of asphalt concrete (AC) mixtures for thermal cracking prediction of exible pavements. Various AC overlay design factors inuencing the thermal cracking resistance of exible pavements, such as mixture properties and pavement structure, were included in the evaluation. Two sources of D(t) data were considered: (1) measurement from indirect tensile test and (2) numerical interconversion of complex modulus E*. Design input levels in the mechanistic empirical design guide software do signicantly impact the predicted thermal cracking distresses. For AC maintenance overlay design purposes, level 1 and 2 analyses yield very similar thermal cracking predictions, whereas level 3 analysis signicantly underpredicts the extent of cracking when compared with to level 1 analysis. Some discrepancy exists in the thermal cracking predicted from measured and interconverted D(t) due to the inherent approximation nature of numerical interconversion methods. However, level 1 and 2 analyses using interconverted D(t) values provide results closer to the predictions from measured D(t) values than does level 3. Three mixtures are used in this study. The various analyses indicate that using a more ductile AC mixture for the surface layer signicantly reduces the amount of thermal cracking at failure. Keywords: MEPDG; thermal cracking; complex modulus; dynamic modulus; indirect tensile test; creep compliance

Introduction Background Thermal cracks appear as regularly spaced cracks propagating transversely to the wheel path along the width of the pavement. Thermal cracking is environmentally induced by sharp and rapid drops in pavement temperature, which cause extreme thermal contraction and subsequent fracture of the surface asphalt concrete (AC) layer, in addition to cyclic freeze thaw action. Thermal cracking is predominant when AC surface mixtures exhibit low strength and ductility at low temperatures. With time, thermal cracks widen and propagate into underlying layers of the pavement leading to water inltration, loss of smoothness (Bae et al. 2007) and ultimately structural failure of the pavement. The thermal cracking model (TCMODEL) adopted in the mechanistic empirical pavement design guide (MEPDG; ARA and ERES 2004) is based on mechanistic empirical principles, wherein the prediction of pavement distresses and performance incorporate fundamental properties of AC materials as well as empirical transfer functions (Timm and Newcomb 2003). The three key material properties are average tensile strength (TS) at 2 108C (148F), creep compliance (D(t)) at 2 20, 2 10 and 08C (2 4, 14 and 328F, respectively) and the thermal coefcient of contraction (CTC) of the AC mixture.

Both creep compliance and strength are obtained from testing conducted in indirect tensile (IDT) mode. The D(t) is a unit response function that characterises the AC material behaviour in the linear viscoelastic (LVE) range, where it is assumed that the material has not been damaged and exhibits linearity in the response; i.e. principles of proportionality and homogeneity apply. Yin et al. (2006a,b) and Freeman et al. (2006) showed that creep compliance displays a signicant sensitivity on thermal cracking prediction. In the TCMODEL, D(t) is used to obtain the relaxation modulus, E(t), using the Laplace transformation. Thermal stresses induced by thermal deformation are then calculated using the following constitutive equation:

s j

j
0

E j 2 j 0

d1 0 dj ; dj 0

where s and 1 are thermal stress and strain, respectively; j is reduced time at reference temperature; j 0 is an integration variable and E(j 2 j 0 ) is the relaxation modulus at reduced time j 2 j 0 . Thermal strain due to temperature change can be calculated as follows: 1 aT j 2 T 0 ; 2

*Corresponding author. Email: hxy153@psu.edu


ISSN 1029-8436 print/ISSN 1477-268X online q 2010 Taylor & Francis DOI: 10.1080/10298430802621531 http://www.informaworld.com

96

H. Yin et al.

where a is linear coefcient of thermal contraction of AC mixtures, T(j) is pavement temperature at reduced time and T0 is initial pavement temperature.

Study objectives Evaluation of hierarchical input levels The MEPDG software provides pavement design engineers exibility in selecting the design inputs and their hierarchical levels of accuracy based on the importance of the project and the resources available for obtaining those inputs (MEPDG 2004). In the course of this study, the hierarchical input levels for thermal cracking prediction are varied, while keeping all other inputs, such as trafc and climate, constant. Level 1 inputs provide the highest level of accuracy, and typically require laboratory or eld testing, such as IDT strength and D(t) for AC mixtures. Level 2 inputs provide an intermediate level of accuracy and are typically used when data or testing equipment are not available for level 1 requirements. An example would be estimating D(t) over a wide temperature range from D(t) measured at one temperature (2 108C (2 148F)). Level 3 inputs provide the lowest level of accuracy and are typically used for low-stake projects. For such analysis, inputs would typically be based on default or national/regional averages. One of the objectives of this paper is to evaluate the relative benets to state highway agencies (SHAs) in making the investment needed for implementing level 1 design for thermal cracking prediction; i.e. performing strength and IDT tests at a variety of temperatures.

Figure 1. Sample interconversion between dynamic modulus and creep compliance.

shown in Figure 1. The resulting discrepancy can then be compared with discrepancy resulting from level 3 analysis. Research scope and methodology Pavement structures and AC mixtures Two AC overlaid pavement structures are considered in this study. The structures are similar to AC overlay sections built on I-79 in Butler County, Pennsylvania. The rst overlay structure consists of a single 37.5 mm (1.5 in.) AC surface layer (W) and the other consists of an additional 50 mm (2 in.) binder layer (WB). The layers are constructed on top of an existing 241 mm (9.5 in.) HMA layer, a 203 mm (8 in.) granular (crushed stone) base, and compacted silty sand subgrade. The three Superpave mixtures are considered for the surface layer. All mixtures are coarse-graded and consist of a PG 76-22-modied binder. Mixtures 1 and 2 have a nominal maximum aggregate size (NMAS) of 9.5 mm and optimum asphalt content of 5.5, while mix 3 is a 12.5-mm NMAS with 10% reclaimed asphalt pavement, and 4.6% optimum asphalt content. Gradation charts of three mixtures are given in Figure 2. Laboratory testing Specimen geometry and fabrication Specimens of 150 mm in diameter and 180 mm in height were compacted using a Superpave gyratory compactor. Specimens used for uniaxial testing were cut and cored to 100 mm in diameter and 150 mm in height. Specimens used for IDT testing were 100 mm in diameter and 38 mm thick. All specimens used for testing met the target air void content of 4 ^ 0.5%.

Evaluation of interconverted D(t) It is widely accepted (Chehab and Kim 2005) that AC response can be characterised as LVE at very low temperatures and fast cooling rates. The LVE behaviour is characterised through unit response functions, mainly creep compliance (D(t)), relaxation modulus (E(t)) and complex modulus (E*). Several interconversion techniques have been developed for obtaining one LVE function from the other (Park and Kim 1999, Schapery and Park 1999) for uniaxial stress modes. Interconversion between uniaxial E*, a level 1 input for rutting and fatigue, and biaxial D(t), a level 1 input for thermal cracking, has a great potential benet. While E* data are becoming relatively easy to obtain through the simple performance test (SPT), obtaining D(t) in IDT mode requires expensive equipment and extensive testing. It thus becomes advantageous to study the inaccuracy resulting in thermal cracking prediction if D(t) in IDT mode is obtained through interconversion from uniaxial E*, especially given that E* is already needed for level 1 rutting and fatigue prediction. An example of interconversion between dynamic modulus and creep compliance is

Testing set-up and instrumentation Mechanical testing was conducted at the Pennsylvania Transportation Institute labs using an MTS closed-loop

International Journal of Pavement Engineering

97

Figure 2.

Gradation chart.

hydraulic universal testing machine, and a 16-bit National Instruments DAQ card. Spring-loaded linear variable differential transformers (LVDTs) were used on uniaxial specimens with a gauge length of 75 mm (3 in.) while loose-core LVDTs with 25 mm (1 in.) gauge length were used on IDT specimens.

The phase angle, f, represents the time lag between the stress and strain cycles. The complex modulus tests were conducted at various temperatures and frequencies, and for three replicates (Table 1). A sigmoidal function was then used to construct the jE*j mastercurves (Figure 3(a)).

Creep compliance and strength tests Complex modulus test The complex modulus test, also referred to as the SPT, was performed according to AASHTO TP62 to obtain the LVE material properties of AC mixtures. The amplitude of the compressive uniaxial haversine load was selected based on the mixture stiffness, temperature and frequency to ensure that the strain response remained within 70 100 microstrain. Two parameters obtained from the test are dynamic modulus, jE*j, and phase angle, f. The dynamic modulus, jE*j, is the magnitude of E* and equivalent to the ratio of the amplitude of a stress cycle (samp) to the amplitude of a strain cycle (1amp) at steady-state conditions: jE* j Creep compliance and TS tests were conducted in IDT following AASHTO T322. The creep compliance D(t) testing was conducted at 2 20, 2 10 and 08C (2 4, 14 and 328F, respectively) for a duration of 100 s. The loads applied ensured that specimen response stayed within the LVE range; i.e. no damage. After the creep test was completed, the TS test was conducted on the same specimen at 2 108C (148F). Tests were conducted on three replicates to ensure repeatability. Three replicates were used for each test. Loads applied in IDT mode create a relatively complex state of stress and strain in the specimen. Determining accurate material properties requires that the analysis address the biaxial state of stress and strain, nonuniform strain distribution, localisation and bulging effects

samp : 1amp

Table 1.

Measured jE*j (MPa) from complex modulus test. Frequency, Hz

Mixture 1

Temperature, 8C (8F) 4 10 20 35 4 10 20 35 4 10 20 35 (39.2) (50.0) (68.0) (95.0) (39.2) (50.0) (68.0) (95.0) (39.2) (50.0) (68.0) (95.0)

0.1 10,098 5277 2169 582 10,781 7682 3818 795 7708 4135 2164 1028

0.5 13,142 7682 3579 1084 13,617 11,078 5922 1678 10,857 6579 3618 1612

1 14,489 8855 4352 1398 14,835 12,239 6958 2209 12,221 7838 4484 1997

5 17,590 11,804 6528 2424 17,586 14,976 9585 3816 15,137 10,998 7048 3338

10 18,884 13,136 7619 3015 18,719 16,221 10,781 4665 16,227 12,359 8343 4148

25 20,530 14,917 9178 3703 20,154 17,874 12,391 5918 17,489 14,068 9701 5450

98

H. Yin et al.

Figure 3.

Dynamic modulus and creep compliance mastercurves at 208C (68.08F).

(Roque and Buttlar 1992, Buttlar and Roque 1994). For more accurate material characterisation, a LVE solution for D(t) in IDT mode has been recently developed for AC mixtures (Kim and Wen 2002), where d Dt 2 cU t eV t; p 4

where d is the thickness of the specimen; p is applied load; c and e are the specimen diameter and gauge length specic coefcients, respectively and U(t) and V(t) are horizontal and vertical displacement, respectively. The displacements are obtained by integrating the strain over the entire gauge length (l) of the horizontal and vertical LVDTs mounted on the specimen: l
2l

Interconversion from E* to D(t) A LVE material is a rheological material that is characterised using unit response functions, mainly D(t) and relaxation modulus, E(t), in the time domain, and complex modulus, E*, in the frequency domain. It is sometimes necessary and/or practical to interconvert between LVE unit response functions. Schapery and Park (1999) introduced the following approximate method (Equations (7)(11)) to convert E* in the frequency domain to D(t) in the time domain for uniaxial mode of loading: (1) Calculate the storage modulus (E 0 ) from dynamic modulus (jE*j) and phase angle (f): E 0 jE* jcos f; 7

U t

1x; tdx;

where the phase angle f is a function of the frequency f and time lag Dt between stress and strain cycles during the complex modulus test:

V t

l
2l

1y; tdx:

f 2pf Dt:
(2) Obtain (E(t)) from E 0 : E0 ; G1 2 ncosnp=2

For the study presented in this paper, 100-mm (4 in.) diameter specimens were tested with a gauge length of 25 mm (1 in.), yielding values of 0.7874 and 2.2783 for c and e, respectively (Momen 2004). The creep compliance D(t) values measured from IDT are tabulated in Table 2.

E t

International Journal of Pavement Engineering


Table 2. Measured D(t) from IDT. Creep compliance (D(t)), 1/MPa Mixture 1. TSa 4.20 MPa (609 psi) Loading time (s) 1 2 5 10 20 50 100 1 2 5 10 20 50 100 1 2 5 10 20 50 100 Low temperature, 2 208C (2 48F) 5.02 5.42 6.12 6.77 7.59 8.99 1.04 6.25 6.90 8.89 9.08 1.16 1.34 1.60 1.30 1.43 1.65 1.87 2.16 2.70 3.28 1026 1026 1026 1026 1026 1026 1025 1025 1025 1025 1025 1024 1024 1024 1025 1025 1025 1025 1025 1025 1025 Intermediate temperature, 2 108C (148F) 9.47 1.10 1.36 1.64 1.99 2.64 3.32 5.53 5.74 6.46 6.95 7.35 7.85 8.74 2.09 2.45 3.12 3.83 4.80 6.69 8.75 1026 1025 1025 1025 1025 1025 1025 1025 1025 1025 1025 1025 1025 1025 1025 1025 1025 1025 1025 1025 1025

99

High temperature, 08C (328F) 3.09 3.92 5.47 7.14 9.37 1.36 1.81 4.27 4.41 4.73 5.04 5.25 5.24 5.63 3.92 4.92 6.85 8.98 1.20 1.77 2.39 1025 1025 1025 1025 1025 1024 1024 1025 1025 1025 1025 1025 1025 1025 1025 1025 1025 1025 1024 1024 1024

2. TS 5.23 MPa (759 psi)

3. TS 3.98 MPa (577 psi)

TS denotes the average tensile strength at 2 108C (148F).

where n is the local slope of E 0 in log log space and G denotes the gamma function. 8 1 x2 1 2 a > < 0 a e da ; x . 0; G x G x 1 > ; x , 0: : x

10

Reduced frequency ( fR) and reduced time (tR) are related through the following relationship: tR 2p/fR. (3) Determine (D(t)) from (E(t)): sinnp ; np

EtDt

11

where t is the time of interest and n is the localised slope of the loglog curve of G at time intervals t 2 Dt to t Dt. Considering the objective of this study, the dynamic moduli, jE*j, measured during the complex modulus tests are used to obtain interconverted creep compliance (Table 3). The (D(t)) mastercurves are plotted for the three mixtures in Figure 3(b) (d). The mastercurves at a reference temperature of 208C (688F) are plotted using measured D(t), interconverted D(t) and default D(t) from level 3 analysis.

MEPDG software runs As discussed earlier, level 2 analysis provides an intermediate accuracy in performance prediction compared with levels 1 and 3. For D(t) level 1 input requires D(t) values at three temperatures, whereas level 2 requires D(t) at only one temperature (2 108C (148F)). Level 3, providing the least accuracy, does not require testing; default D(t) values correlated to user input AC mixture properties such as air voids and volumetric effective binder content are used. For all hierarchical input levels, TCMODEL requires the value of TS at 2 108C (148F). Levels 1 and 2 require actual test data for TS, whereas for level 3, default values are used based on user input AC mixture properties. As shown in Table 4, default values of D(t) are higher than both measured and interconverted. In addition to D(t) and TS, another essential material property required in the TCMODEL is the coefcient of thermal contraction of AC mixture, CTC (Equation (2)). Prior studies (Easa et al. 1996, Yin et al. 2006a) showed that the greatest sensitivity of thermal cracking to CTC is observed for AC mixtures of moderate strength and ductility, in which case, considerable interaction exists among CTC, D(t), and TS. Default values of CTC that lie within the range of typical CTC of AC mixtures are used in this study. These CTC values are kept constant for all analyses conducted with the MEPDG. Another critical factor affecting thermal cracking is the pavement temperature over the entire analysis period,

100
Table 3. Interconverted D(t) from jE*j.

H. Yin et al.

Creep compliance (D(t)), 1/MPa Mixture 1. TSa 4.20 MPa (609 psi) Loading time (s) 1 2 5 10 20 50 100 1 2 5 10 20 50 100 1 2 5 10 20 50 100 Low temperature, 2 208C (2 48F) 4.45 4.61 4.86 5.11 5.41 5.90 6.38 4.49 5.28 6.47 7.01 7.84 8.43 9.37 5.29 5.69 6.37 7.02 7.88 9.35 1.09 1025 1025 1025 1025 1025 1025 1025 1025 1025 1025 1025 1025 1025 1025 1025 1025 1025 1025 1025 1025 1024 Intermediate temperature, 2 108C (148F) 5.82 6.28 7.05 7.82 8.76 1.04 1.21 3.75 4.21 4.84 5.12 5.53 5.82 6.26 7.51 8.48 1.02 1.20 1.42 1.84 2.26 1025 1025 1025 1025 1025 1024 1024 1025 1025 1025 1025 1025 1025 1025 1025 1025 1024 1024 1024 1024 1024 High temperature, 08C (328F) 9.54 1.10 1.34 1.60 1.93 2.51 3.12 3.40 3.65 3.99 4.14 4.35 4.49 4.71 1.25 1.49 1.94 2.39 2.99 4.08 5.16 1025 1024 1024 1024 1024 1024 1024 1025 1025 1025 1025 1025 1025 1025 1024 1024 1024 1024 1024 1024 1024

2. TS 5.23 MPa (759 psi)

3. TS 3.98 MPa (577 psi)

TS denotes the average tensile strength at 2 108C (148F).

preferably at hourly time intervals. Consequently, the thermal stress at any given depth and time within the AC layer can be predicted. The MEPDG recommends
Table 4. Level 3 default D(t).

interpolating climatic data from as many applicable weather stations near the project site as possible. At times, topographic maps may be consulted for intervening features

Creep compliance (D(t)), 1/MPa Mixture 1. TS 4.17 MPa (605 psi)


a

Loading time, s 1 2 5 10 20 50 100 1 2 5 10 20 50 100 1 2 5 10 20 50 100

Low temperature, 2 208C (2 48F) 9.96 1.10 1.24 1.37 1.51 1.71 1.87 6.79 7.39 8.26 9.00 9.79 1.10 1.19 8.05 8.82 9.95 1.09 1.20 1.35 1.48 10 1024 1024 1024 1024 1024 1024 1027 1027 1027 1027 1027 1026 1026 1025 1025 1025 1024 1024 1024 1024
25

Intermediate temperature, 2 108C (148F) 1.29 1.51 1.86 2.18 2.54 3.12 3.65 7.68 8.91 1.08 1.26 1.46 1.77 2.05 1.04 1.21 1.49 1.74 2.02 2.48 2.89 10 1024 1024 1024 1024 1024 1024 1027 1027 1026 1026 1026 1026 1026 1024 1024 1024 1024 1024 1024 1024
24

High temperature, 08C (328F) 1.73 2.22 3.10 3.99 5.15 7.19 9.27 9.61 1.22 1.69 2.15 2.74 3.77 4.80 1.39 1.78 2.48 3.18 4.09 5.69 7.30 1024 1024 1024 1024 1024 1024 1024 1027 1026 1026 1026 1026 1026 1026 1024 1024 1024 1024 1024 1024 1024

2. TS 3.99 MPa (579 psi)

3. TS 4.57 MPa (663 psi)

TS denotes the average tensile strength at 2 108C (148F).

International Journal of Pavement Engineering

101

Figure 4.

Daily air temperatures at Pittsburgh International Airport.

Figure 5.

Crack length vs. pavement age for different pavement structures.

102

H. Yin et al. (0.6 mile) of length. To include all cases depicted in this study (Figure 1), a total of 15 runs are conducted for 90% reliability using MEPDG version 0.91.

to avoid erroneous or microclimate data. In this study, the weather station at Pittsburgh International Airport was selected to generate climate data for the AC overlay section. Measured maximum and minimum daily air temperatures from 2005 to 2006 are shown in Figure 4(a) and (b), respectively. For the MEPDG design runs, a typical 10-year design life for AC overlay is chosen with no scheduled maintenance or rehabilitation activities. A value of 1.578 m/km (100 in./mile) is assigned to the initial International roughness index, and for this study, the performance criterion for thermal cracking is set to 189.4 m (621.4 ft) of transverse cracking per 1 km

Results and analysis Crack lengths predicted from MEPDG runs are plotted in Figures 5 7. Table 5 lists the year at which pavement failure is predicted and the total amount of thermal cracking at the end of the 10-year design life for each case. The amount of thermal cracking at failure is also reported. Detailed evaluations of each experiment design factor, pavement structure, hierarchical input level and

Figure 6.

Crack length vs. pavement age for different AC mixtures.

International Journal of Pavement Engineering method of obtaining D(t), are presented in the following sections.

103

Evaluation of pavement structure In general, as shown in Figure 5, the two pavement structures exhibit similar trends of thermal cracking development and progression, regardless of the level of hierarchical inputs and method of obtaining D(t). With the same type of mixture, two structures display a similar amount of thermal cracking at the end of a 10-year design life. The small reduction (about 70 m/km) in crack length of structure WB may be attributed by the

addition of a binder layer. An evaluation of pavement structure suggests that an inclusion of binder layer in an AC pavement will not signicantly reduce the amount of thermal cracking, but may extend the service life by 1 2 years in this mode. However, it would be necessary to evaluate the sensitivity of thermal cracking prediction to the surface AC layer thickness in subsequent studies.

Evaluation of hierarchical input levels For the same pavement structure and method of obtaining D(t), levels 1 and 2 show comparable levels of thermal cracking, with level 3 consistently underpredicting thermal

Figure 7.

Crack length vs. pavement age for interconverted D(t) and level 3 default.

104
Table 5. Summary of thermal cracking predictions. Mixture 1 Hierarchical input level 1 2 2 3 1 2 3 3 1 2 3 1 2 2 3 1 2 3 3 1 2 3

H. Yin et al.

Pavement structure W

Creep compliance Measured Interconverted Measured Interconverted Default Measured Interconverted Measured Interconverted Default Measured Interconverted Measured Interconverted Default Measured Interconverted Measured Interconverted Default Measured Interconverted Measured Interconverted Default Measured Interconverted Measured Interconverted Default

Fail? Yes Yes Yes Yes No Yes Yes Yes Yes No Yes Yes Yes Yes No Yes Yes Yes Yes No Yes Yes Yes Yes No Yes Yes Yes Yes No

Year@failure 5 7 5 7 7 8 7 9 8 9 8 9 6 8 7 9 7 8 7 10 8 9 8 9

Total thermal cracking (m/km) 1430 1106 1401 1073 858 1287 1051 1261 1019 686 1260 1100 1235 1067 780 1359 1051 1331 1019 815 1223 1030 1198 1009 652 1197 1045 1173 1014 741

WB

cracking. As demonstrated in Figure 5, the use of measured D(t) rather than default level 3 values leads to earlier failures. One possible reason is that level 3 inputs provide a much higher D(t) than measured D(t) (Figure 3). For all cases presented in this study, no early failure is predicted from level 3 analyses. Comparable thermal cracking predictions from levels 1 and 2 indicate that the MEPDG software reasonably constructs the complete creep compliance mastercurve using D(t) measured at only one temperature, 2 108C (148F). Thus, it can be concluded that level 2 can be used for preliminary pavement design and performance prediction when resources or testing equipment are not available for level 1 inputs. Evaluation of interconverted D(t) Predicted crack lengths are plotted for the three mixtures in Figure 6. As expected, signicant divergence in D(t) leads to both divergence of time to pavement failure and extent of thermal cracking experienced throughout the pavement life for each mixture. Those observations, and Equation (1), suggest that using the interconverted D(t) underpredicts the resulting thermal stresses for a given thermal strain and, consequently, underpredicts the extent of thermal cracking.

Compared with mix 3, mixtures 1 and 2 always result in a larger amount of thermal cracking at the end of the 10-year design life, regardless of the level of hierarchical inputs and the method used for obtaining D(t). Again, lower D(t) makes mixtures 1 and 2 less ductile at low temperatures than mix 3. It should be pointed out that TS also plays an important role in predicting thermal cracking. Given a similar creep compliance for mixtures 1 and 2, the variation in predicted thermal cracking is possibly due to the difference in measured values of TS (4.2 MPa (609 psi) of mix 1 vs. 5.23 MPa (759 psi) of mix 2). Another purpose of evaluating interconverted D(t) is to examine the practicality of using D(t) obtained from jE*j to predict thermal cracking, instead of using level 3 default values. Compared with the thermal cracking prediction from interconverted D(t) values (Figure 7), level 3 default D(t) values signicantly underpredict both pavement life and intensity of thermal cracking. Since D(t) can be easily extended to the temperature of interest once the creep compliance mastercurve is known, level 1 interconverted D(t) does not require much more effort than level 2 interconverted D(t). Thus, using level 1 interconverted D(t) for thermal cracking prediction is preferable over using the level 3 default.

International Journal of Pavement Engineering Conclusions and recommendations The objective of this paper is to evaluate the feasibility and benets of using interconversion techniques based on linear viscoelastic principles to obtain D(t) for predicting thermal cracking in the MEPDG software. The thermal cracking predictions using level 1 with interconverted D(t), level 1 with measured D(t) and level 3 with default D(t) are evaluated. Different pavement structures, hierarchical input levels for the MEPDG and methods to obtain creep compliance are considered. Two sources of D(t) are considered: direct measurement from IDT test and interconversion from complex modulus, E*. Because of the approximation nature of numerical interconversion methods, some variation was observed in the thermal cracking predicted from these two D(t) methods. However, interconverted D(t) still offers thermal cracking predictions closer to those from measured values than that of level 3 with default D(t) values. Since the E* test is becoming relatively common and practical to conduct using the simple performance tester, this nding gives SHAs a better alternative to the use of the level 3 default, if conducting IDT for obtaining D(t) is not feasible. The conclusions from this study are limited to the pavement structures and materials tested. Further structures and materials need to be evaluated to verify the ndings of this paper. Notes
1. 2. 3. 4. Email: gchehab@engr.psu.edu Email: sms26@psu.edu Email: tkumar@ara.com Email: lxp213@psu.edu

105

References
ARA Inc. and ERES Consultants Division, 2004. Guide for mechanistic empirical design of new and rehabilitated pavement structures. Final Report, NCHRP 1-37A. Bae, A., et al., 2007. Direct effects of thermal cracks on pavement roughness. Presented at the 2007 AAPT Annual Meeting, San Antonio, TX.

Buttlar, W.G. and Roque, R., 1994. Development and evaluation of the strategic highway research program measurement and analysis system for indirect tensile testing at low temperature. Transportation Research Board , 1454. Washington, DC: Transportation Research Board. Chehab, G. and Kim, R., 2005. Viscoelastoplastic continuum damage model application to thermal cracking of asphalt concrete. Journal of Materials in Civil Engineering, ASCE, 17 (4). Easa, S.M., Shalaby, A., and El Halim, A.O.A., 1996. Reliabilitybased model for predicting pavement thermal cracking. Journal of Transportation Engineering, ASCE, 122 (5). Freeman, T., et al., 2006. Sensitivity analysis of and strategic plan development for the implementation of the M E design guide in TXDOT operations. College Station, TX: Texas Transportation Institute. Kim, Y.R. and Wen, H., 2002. Fracture energy from indirect tension testing. Proceedings of the Journal of the Association of Asphalt Paving Technologists, 71, 816 827. MEPDG, 2004. Available from: http://www.trb.org/mepdg/soft ware.htm Momen, M., 2004. Complex modulus determination of asphalt concrete using indirect tension test. Thesis (MS). North Carolina State University. Park, S.W. and Kim, Y.R., 1999. Interconversion between relaxation modulus and creep compliance for viscoelastic solids. Journal of Materials in Civil Engineering, 11 (1), 76 82. Roque, R. and Buttlar, W.G., 1992. The development of a measurement and analysis system to accurately determine asphalt concrete properties using the indirect tensile mode. Proceedings of the Journal of Association of Asphalt Paving Technologist, 65. Schapery, R. and Park, S., 1999. Methods of Interconversion between linear viscoelastic material functions. Part II an approximate analytical method. International Journal of Solids and Structures, 36. Timm, D.H. and Newcomb, D.E., 2003. Calibration of exible pavement performance equations. Transportation Research Record, 1853. Washington, DC: Transportation Research Board. Yin, H., Chehab, G., and Stoffels, S., 2006a. A case study: assessing the sensitivity of the coefcient of thermal contraction of AC mixtures on thermal crack prediction. ASCE, Geotechnical Special Publication, No. 146. Yin, H., Chehab, G., and Stoffels, S., 2006b. Sensitivity of thermal cracking prediction to AC mixture properties using the ME pavement design guide. Proceedings of the 10th International Conference on Asphalt Pavements. Quebec, Canada.

Copyright of International Journal of Pavement Engineering is the property of Taylor & Francis Ltd and its content may not be copied or emailed to multiple sites or posted to a listserv without the copyright holder's express written permission. However, users may print, download, or email articles for individual use.

You might also like