Sunteți pe pagina 1din 14

Construction of Dynamic Modulus Master

Curves with Resilient Modulus


and Creep Test Data
Hyung Suk Lee, Sungho Kim, Bouzid Choubane, and Patrick Upshaw

For the past few decades, the stiffness of materials used for roadway design guide uses the dynamic modulus as a primary material input
design and construction has been commonly characterized by the resilient for both new construction and rehabilitation projects.
modulus, defined as the ratio of the applied stress to the recoverable strain. Although the resilient modulus was not adopted for use in the new
However, the resilient modulus is not a fundamental material property design guide, a considerable amount of resilient modulus data have
of a viscoelastic material. Therefore, the concept of resilient modulus been gathered in the past to evaluate asphalt mixtures. However,
has been subsequently diminished in the latest Mechanistic–Empirical because the dynamic modulus test was not needed for calculation of
Pavement Design Guide. Although that design guide could not endorse energy ratio, it has not been conducted during most of Florida DOT’s
the use of the resilient modulus test protocol as the primary means previous research studies. Therefore, a need exists for a reliable and
of characterizing the asphalt concrete modulus in the design of flexible accurate methodology that allows for converting the existing resilient
pavements, that protocol has been a primary mixture test, and much modulus data to dynamic modulus, to make use of the existing resilient
laboratory testing has been completed to date. Analysis methodologies are modulus data along with the new design guide.
introduced for backcalculating the dynamic modulus from the resilient The problem was also recognized by other researchers, and a
modulus test data. To assess the usefulness of the proposed algorithm, research study was conducted to predict the complex modulus from
laboratory experiments in both the uniaxial compression and indirect the resilient modulus data (5, 6). The study, however, employed the
tensile test modes were carried out on asphalt specimens compacted artificial neural network as a link between the resilient and com-
with the Superpave® gyratory compactor. The backcalculated dynamic plex modulus data, indicating that additional effort is necessary in
modulus was used to generate the master curve, and the creep test data building a database of resilient modulus predicted from the complex
were used to enhance the accuracy of the master curve. The advantage modulus data.
of such a methodology is that the existing resilient modulus and creep In a recent study, Lee and Kim proposed a more direct methodol-
test data can be leveraged for estimating the dynamic modulus. The
ogy for backcalculating the fundamental property of a viscoelastic
approach would significantly save time and effort in reevaluating the
material from an arbitrary loading function by means of the method
of least squares (7). The authors showed that if the historical resilient
dynamic modulus of an asphalt mixture when the resilient modulus and
modulus test data are available for a given asphalt mixture, then it is
creep test data are available.
possible to backcalculate the dynamic modulus from the test data
by treating the resilient modulus loading as the arbitrary loading.
For the past few decades, the resilient modulus, defined as the ratio The methodology showed potential, for it was able to backcalculate
of the applied stress to the recoverable strain (1), has been commonly the dynamic modulus of both laboratory-fabricated and field-cored
used by highway agencies for characterizing the stiffness of asphalt asphalt specimens from the resilient modulus test data at multiple
concrete mixtures. As for the Florida Department of Transportation temperatures (0°C, 10°C, and 20°C) in the uniaxial compression mode
(DOT), the resilient modulus test has been typically conducted along as well as in the IDT mode.
with creep and strength tests in the indirect tensile (IDT) mode for
obtaining the energy ratio, a single parameter that depicts the relative
Objective and Scope of Work
cracking performance of a mixture based on the theory of fracture
mechanics (2, 3). Unfortunately, the resilient modulus is not a funda- The objective of this study was to construct the master curve of an
mental property of a viscoelastic material, so the concept of resilient asphalt mixture from the dynamic modulus values backcalculated
modulus has been subsequently diminished in the latest Mechanistic– from the resilient modulus test data. In addition, the creep test data
Empirical Pavement Design Guide (MEPDG) (4). Instead, the new that Florida DOT has typically been collecting with the resilient
modulus data is introduced to enhance the accuracy of the master
H. S. Lee, B. Choubane, and P. Upshaw, Florida Department of Transportation, curve constructed from the backcalculated modulus.
5007 Northeast 39th Avenue, Gainesville, FL 32609. S. Kim, University of North
Florida, 1 UNF Drive, Jacksonville, FL 32224. Current affiliation for S. Kim:
Florida Department of Transportation, 5007 Northeast 39th Avenue, Gainesville, Backcalculation Methodology
FL 32609. Corresponding author: H. S. Lee, hyung.lee@dot.myflorida.com. for Uniaxial Testing Conditions
Transportation Research Record: Journal of the Transportation Research Board,
No. 2296, Transportation Research Board of the National Academies, Washington,
In this section, the mathematical formulation of the backcalculation
D.C., 2012, pp. 1–14. algorithm will be reviewed briefly. A more detailed derivation is
DOI: 10.3141/2296-01 available in Lee and Kim (7).

1
2 Transportation Research Record 2296

Numerical Evaluation of Convolution Integral the least squares approach, the objective function is set up as
follows:
On the basis of the theory of linear viscoelasticity, the relationship
between time-dependent uniaxial stress and strain can be expressed N N
F = ∑ {ε ( tn ) − ε m,n } = ∑ { D0 σ n + D1ξ n − ε m,n }
2 2
as a convolution integral of the following form: (6)
n =1 n =1


∂σ ( τ ) where F is the function to be minimized, εm,n is the measured strain
ε (t ) = ∫ D (t − τ) dτ (1)
−∞
∂τ at time t = tn, and

{(t }
n −1
where ε(t) = strain and D(t) = creep compliance. ξ n = ∑  σ i − ti ) − ( tn − ti +1 )
m m
i
n (7)
i =1
The convolution equation can also be written as the following, known
as the Riemann–Stieltjes integral (8):
The objective function shown in Equation 6 leads to a nonlinear
t
least squares problem owing to the parameter m in Equation 7. To
∂ overcome this difficulty, the m-value has been assumed for each
ε (t ) = D (0) i σ (t ) − ∫ σ (τ) D ( t − τ ) dτ (2)
0
∂τ iteration to simplify the problem into a linear least squares in which
the only remaining unknowns become D0 and D1. Then for a given
m-value, the function F is minimized when its partial derivative with
where σ(t) = time-dependent stress.
respect to D0 and D1 are equal to zero. This formulation results in a set
Following Lee and Kim (7), the foregoing equation can be evaluated of linear equations that can be expressed in the following matrix form:
numerically for each given time step, tn, as follows:
−1
N 2 N
 N 
n −1 ∑ σn
 D0   n =1
∑ σ nξn  ∑ ε m,n σn 
 
ε ( tn ) = D ( 0 ) i σ n + ∑ [ σ i {D (t − ti ) − D ( tn − ti +1 )}] 
n =1 n =1
i
n  =  N  N  (8)
 D1  N
 
 ∑ σn ξ n ∑ ∑ ε m,n ξ n 
i =1
ξ n2 
for tn = t1 , . . . , t N (3)  n =1 n =1   n =1 

where σi = σ(t) for the discrete time interval ti < t < ti+1 and n is the
number of discrete data points in time. The foregoing equation can be Backcalculation Methodology for
applied to any arbitrary functions D(t) and σ(t) that are numerically Indirect Tension Testing Conditions
available.
Elastic Stress Analysis for IDT

The illustration of the IDT test and the plane-stress distribution within
Backcalculation of Creep Compliance an IDT sample are shown schematically in Figure 1. For an IDT
by Least Squares specimen with radius R and thickness d subjected to a strip load
of width a = 2R • sin(α) and magnitude P, the distributions of the
To backcalculate the creep compliance by the least squares approach, tensile stress along the horizontal axis (x) and the compressive stress
it is necessary to identify a function that represents the creep com- along the vertical axis (y) were given by Hondros (14) under the
pliance of viscoelastic materials. For this purpose, the power func- assumption of plane-stress condition as follows:
tion, which has been well accepted as an analytical representation
 1 − ( x R ) sin 2α 
2
of the creep compliance of viscoelastic materials (9–13), has been  4 
2 P  1 + 2 ( x R ) cos 2α + ( x R )  2 P
2
selected owing to its simple form. The function is expressed as
follows: σ x ( x, 0) =  = m(x) (9a)
πad   πad

 − tan −1 
1 − ( x R ) 2

tan α  
D ( t ) = D0 + D1 i t m  1 + ( x R)
2
(4)   

 1 − ( y R ) sin 2α 
2
where D0, D1, and m are the power function parameters. The following  4 
2 P  1 − 2 ( y R ) cos 2α + ( y R ) 
2
equation is obtained by substituting Equation 4 into Equation 3: 2P
σ y ( 0, y ) = −  =− n ( y)
πad  π
−1 1 + ( y R )
 2
  ad
∑ σ i {( t }
n −1
ε ( tn ) = D0 i σ n + D1 − ti ) − ( tn − ti +1 )
m m  + tan  tan α  
 1 − ( y R)
i 2
i =1
i n
  
for tn = t1 , . . . , t N (5) (9b)

This equation can be used to calculate the time-dependent strain of from which the average stresses over the gauge length, l, are calculated
a viscoelastic material with the D(t) shown in Equation 4 subjected as follows:
to any arbitrary loading σ(t).
1
To determine the power function parameters D 0, D1, and m 2P 1 2
1 m ( x ) dx
πad l ∫− 2
σ x,avg = i (10 a)
from the strain output resulting from an arbitrary stress input using
Lee, Kim, Choubane, and Upshaw 3

y ε x = D i  σ x − µσy  (12a)

εy = D i  σ y − µσx  (12b)
a
where D and µ are the elastic compliance and Poisson’s ratio for the
given material, respectively. According to the elastic-viscoelastic cor-
respondence principle, the constitutive equations for a viscoelastic
material in the Laplace domain are obtained by replacing the elastic
2α x material properties with their viscoelastic counterparts multiplied by
the Laplace variable (8, 9, 17). Therefore, the plane-stress constitutive
equations for a viscoelastic material are given as follows:
R
εˆ x ( s ) = sDˆ ( s ) i  σˆ x ( s ) − sµˆ ( s ) σˆ y ( s )  (13a)

εˆ y ( s ) = sDˆ ( s ) i  σˆ y ( s ) − s µˆ ( s ) σˆ x ( s )  (13b)
P
where
(a)

L[ f(t)] = fˆ(s) = ∫ 0 e−st f(t)dt = Laplace transform of the
y
function f(t),
s = Laplace variable, and
D̂(s) and µ̂(s) = Laplace transformed creep compliance and Pois-
σx(x,0) son’s ratio, respectively.
σy(0,y)
Rearranging the terms after eliminating µ̂(s) from Equations 13a and
13b and taking the inverse Laplace transform yield this:
x
t

{ε ( t ) + βε ( t )} = ∫ D ( t − τ ) ∂∂τ {σ ( τ ) + βσ ( τ )} dτ
x y x y (14)
0−

where β was defined in Equation 11. The foregoing equation is


essentially identical to Equation 1 if the following substitutions are
made for the stress, σ(t), and strain, ε(t), terms in Equation 1:

(b) ε ( t ) = {εx ( t ) + βεy ( t )} (15)

FIGURE 1   Schematic illustration of (a) IDT test and σ ( t ) = {σx ( t ) + βσy ( t )} (16)
(b) plane stress distributions.
Then, with the foregoing replacement of variables and the power
2P 1 2
1 function shown in Equation 4 for creep compliance, the same
1 n ( y ) dy
πad l ∫− 2
σ y,avg =− i (10 b) derivation and numerical scheme outlined for the uniaxial case
(Equations 1 through 8) can be followed to backcalculate the creep
Because of the nature of the IDT specimen geometry, another compliance for the biaxial loading conditions. By using the direct
constant can be derived. This constant, β, is defined as the ratio of search method for the m-value, the remaining power function
the horizontal and vertical stress magnitudes and is used to simplify parameters D0 and D1 can be determined from the following:
the equations for the creep compliance formulation in IDT (7, 15, 16): −1
 N N

 ∑ ( σ x,n + βσ y,n ) ∑ (σ x,n + βσ y,n ) ξ n 
2
1

σ y,avg ∫ n ( y ) dx

2
1
 D0   n =1
 =  N
n =1


β=− = 2
(11))  D1  
N
σ x,avg 1

m ( x ) dy  ∑
( σ x,n + βσ y,n ) ξ n ∑ ξ n2 
∫ −
2
1
2
n =1 n =1 
N 
For simplicity, the average horizontal and vertical stresses will ∑ ( ε x,n + βε y,n ) ( σ x,n + βσ y,n ) 
 n =1 
be designated respectively as σx and σy in subsequent sections of N 
the paper.  
(
∑ x,nn ε + βε y,n ) ξ n 
n =1 (17)
Determining Creep Compliance
where
from Random Loading in IDT Test

{(t }
n−1
The plane-stress constitutive equations for a linear, elastic, and ξ n = ∑ ( σ x,i + βσ y,i ) i − ti ) − ( tn − ti +1 )
m m
n (18)
isotropic material are given as follows: i =1
4 Transportation Research Record 2296

Conversion of Creep Compliance TABLE 1   Volumetric Properties of


to Dynamic Modulus Laboratory-Fabricated Mixture

Although the power function has been successfully used for repre- Passing and
Sieve Size and Mixture Property Property Value (%)
senting the creep behavior of viscoelastic materials, its mathematical
deficiency does not provide enough flexibility for analytical inter- 19 mm 100
conversion of the fundamental properties of viscoelastic media. 12.5 mm 99
Instead, the Prony series has been widely used for interconversion 9.5 mm 90
because of its mathematical efficiency. For creep compliance in the
No. 4 64
time domain, the Prony series representation takes the following form:
No. 8 42
M No. 16 30
D ( t ) = D0p + ∑ Dip (1 − e − t τi ) (19) No. 30 22
i =1
No. 50 13
p p No. 100 6
where D0 , and Di , are the Prony series parameters, and τi is the
No. 200 4
retardation times. In fitting creep compliances using the Prony series,
Specific gravity for aggregates (Gsb) 2.647
the retardation times, τi, are usually specified, and then the unknown
coefficients are determined by solving the linear system of equations Specific gravity for mixture (Gmm) 2.494
(13, 18). In this study, the Prony series shown in Equation 19 was Total asphalt content (Pb) 5.2
fitted to the backcalculated creep compliance, expressed in the power RAP asphalt content 0.9
function, with M = 7 and retardation times of one decade interval, Virgin asphalt content 4.3
τi = 10(i−4) (i = 1, . . . , 7).
Note: RAP = reclaimed asphalt pavement.
Once the Prony series parameters have been determined, the real
(D′) and imaginary (D″) parts of the complex compliance can be
obtained by the following (7, 17, 19):
final size of 150 mm tall by 100 mm in diameter in accordance with
AASHTO TP 62. For IDT testing, test specimens 185 mm tall by
M p
D
D′ ( ω ) = D0p + ∑ i
(20) 150 mm in diameter were prepared with the same Superpave gyratory
i =1 ω τ +1
2 2
i compactor. From each specimen, three replicates of 38-mm-thick
IDT samples were carefully cut from the middle of the specimen.
All the uniaxial and IDT samples met the 7% target air void with a
M
ωτ i Dip
D′′ ( ω ) = ∑ (21) tolerance of ± 0.5% in accordance with AASHTO T 322.
i =1 ω 2 τ i2 + 1

where ω is the angular frequency. Then the dynamic modulus is Uniaxial and IDT Testing on Mixtures
obtained as the reciprocal of the complex compliance, as shown in
The uniaxial compressive testing, the primary testing mode in the
the following equation:
MEPDG, is normally conducted in the intermediate to high temper-
ature range (approximately 0°C to 60°C), while the IDT, which has
1 1
E (ω ) = = (22) been used for evaluating the resilient modulus, creep, and tensile
D (ω )

D′ ( ω ) + D′′ ( ω )
2 2
strength (10, 20, 21) of asphalt mixtures, is conducted in the low
to intermediate temperature range (approximately −30°C to 30°C).
Therefore, it was decided to conduct the uniaxial testing at 0°C, 10°C,
Experimental Study with 20°C, and 40°C; and IDT testing at 0°C, 10°C, 20°C, and 30°C.
Laboratory-Compacted Mixtures A picture of the uniaxial test setup with specimen dimensions
is shown in Figure 2. For the uniaxial test, resilient modulus tests
To evaluate the usefulness and effectiveness of the backcalculation followed by dynamic modulus and by creep tests were all performed
algorithm, laboratory experiments have been conducted in both the in the compression mode. The resilient modulus test was conducted
uniaxial compressive mode and IDT mode. The material selected for for five cycles in which each cycle consisted of a 0.1-s haversine load
testing was a dense graded asphalt mixture meeting all Superpave® followed by a 0.9-s rest period. Data were measured according to
requirements. Material was sampled from a local plant and compacted the method documented in a previous study (10). Then the dynamic
in the laboratory. The mixture included 15% reclaimed asphalt pave- modulus tests were conducted at frequencies of 0.1 Hz, 0.5 Hz, 1.0 Hz,
ment, a 12.5-mm nominal aggregate size, a PG76-22 modified binder, 5.0 Hz, and 10.0 Hz, with rest periods of at least 5 min between the
and a total binder content of 5.2%. The volumetric properties of the individual tests. Finally, the creep test was conducted for a duration
mixture are summarized in Table 1. of 1,000 s. All of the aforementioned tests conducted in the uniaxial
mode were also repeated in the IDT mode. A picture of the IDT
specimen and the testing setup is shown in Figure 3. The system for
Preparation of Specimen measuring the stress and strain responses was in accordance with
AASHTO T 322. The maximum strain in the resilient modulus test
For uniaxial compression testing, three replicate specimens 170 mm performed under the uniaxial testing mode and the maximum hori-
tall by 150 mm in diameter were prepared with a Superpave gyratory zontal strain in the same test under the IDT testing mode were kept
compactor. These gyratory specimens were cored and trimmed to a below 100 µε.
Lee, Kim, Choubane, and Upshaw 5

(a)
(a)

25.4 mm

150 mm 100 mm

38.0 mm 150 mm

38.0 mm
100 mm
Gauge Points
(b)
Gauge Points
FIGURE 2   Photograph of (a) uniaxial test setup and (b)
( b) specimen dimensions.
FIGURE 3   Photograph of (a) IDT test setup and
( b) specimen dimensions.
Comparison of Dynamic Modulus Master Curves
from Indirect Tensile and Uniaxial Test Modes
obtaining the dynamic modulus in the IDT mode is briefly described.
The uniaxial compressive testing has been chosen by the MEPDG Detailed mathematical derivations can be found in Lee and Kim (7).
as the standard testing mode for evaluation of the asphalt dynamic From the plane-stress constitutive equation for a linear elastic
modulus and construction of the master curve. However, obtaining material shown in Equation 12, the constitutive equation for a visco-
a field core meeting the specimen size requirement for this testing elastic material in the frequency domain can be obtained by replacing
mode could pose some challenges. As an example, a recent research D and µ in Equation 12 by their viscoelastic counterparts (9, 17):
funded by Florida DOT indicated that out of 25 in-service pave-
ment sections studied as part of that agency’s efforts to implement ε*x = D ∗ i  σ 0x − µ∗σ 0y  (23a)
the MEPDG, only six test sections had a total asphalt thickness in
excess of 150 mm, suitable for carrying out the dynamic modulus
tests in the uniaxial compression mode (22). Also, the asphalt layer ε*y = D ∗ i  σ 0y − µ∗σ 0x  (23b)
in Florida is typically constructed with multiple lifts of thickness
ranging from 40 mm to 50 mm. For this reason, the uniaxial testing where
conducted on field asphalt cores can represent the behavior of only
the composite asphalt mat, which may be composed of layers with ε*x and ε* = 
y complex amplitudes of the horizontal and vertical
different age, gradation, binder content, and so on. For these reasons, strains;
the majority of Florida DOT’s previous tests conducted on field cores σ x0 and σ y0 = horizontal and vertical stress amplitudes, respec-
had been carried out in the IDT mode rather than in the uniaxial mode. tively; and
Therefore, a quick comparison of results from the uniaxial and IDT D* and µ* = viscoelastic, complex compliance and Poisson’s ratio,
test modes is provided in this section of the paper, to evaluate the respectively.
feasibility of the IDT dynamic modulus as input into the new design Combining Equations 23a, 23b, and 11 by eliminating µ* yields a
guide. The comparisons are limited to those from the dynamic mod- single equation for D*:
ulus testing. Backcalculation results will be presented in subsequent
sections of the paper.
ε*x + βε*y
The dynamic modulus from the uniaxial test was obtained accord- D∗( ω ) = (24)
ing to the procedures outlined in AASHTO TP 62. The procedure for σ 0x + βσ 0y
6 Transportation Research Record 2296

3,500

Indirect Tensile Dynamic Modulus (ksi)


3,000
y = 1.02x + 84.57
R2 = 0.98
2,500

2,000

1,500

1,000 Line of Equality

500

0
0 500 1,000 1,500 2,000 2,500 3,000 3,500
Uniaxial Dynamic Modulus (ksi)

0°C 10°C 20°C

FIGURE 4   Comparison of measured uniaxial and IDT dynamic modulus.

As mentioned, taking the reciprocal of the dynamic compliance, conducted (0.1 Hz, 0.5 Hz, 1.0 Hz, 5.0 Hz, and 10.0 Hz). Figure 5
|D*|, obtained from the foregoing equation, results in the dynamic shows a comparison between the measured and the backcalculated
modulus, |E*|, in the IDT testing mode. For simplicity, the dynamic dynamic modulus for both the uniaxial and IDT mode. The average
modulus obtained from the dynamic modulus tests will be referred absolute percent errors between the measured and the backcalculated
to as the measured dynamic modulus in the remaining sections of dynamic moduli were calculated as 9.9% and 10.7% for the uniaxial
the paper. and IDT test modes, respectively. Since these errors were less than the
Also as mentioned earlier, both the uniaxial and the IDT dynamic estimated limit of accuracy, the master curves were constructed with
modulus tests were conducted at four distinct temperatures, among both the measured and backcalculated dynamic modulus values. All
which three temperatures (0°C, 10°C, and 20°C) were common of the master curves presented in this paper were generated for a
in both the uniaxial and IDT mode. Figure 4 shows a comparison reference temperature of 10°C.
of the dynamic modulus values measured in the uniaxial and IDT Figures 6 and 7 show the master curves constructed with mea-
mode at the three temperatures mentioned and at five frequencies sured and backcalculated dynamic moduli, for the uniaxial and
(0.1 Hz, 0.5 Hz, 1.0 Hz, 5.0 Hz, and 10.0 Hz). The figure shows that IDT testing mode, respectively. For clarity, the master curves are
the measured dynamic moduli from the uniaxial and those from the shown in both arithmetic and logarithmic scales. The figures show
IDT modes are in good agreement, with an R2-value of .98. The figure that regardless of the testing mode, the master curve from the back­
also shows that the IDT dynamic modulus values are slightly greater calculated dynamic modulus has reasonable agreement with the
than those from the uniaxial test. Although this was not expected, master curve from the measured modulus at higher frequencies (or
the average absolute percent error between the two test modes was equivalently lower temperatures), but they tend to deviate from
8.7%, which is less than the estimated accuracy limit of the test each other at lower frequencies (or higher temperatures). It was also
method (approximately 12% to 13%) reported in AASHTO TP 62. found that this was more pronounced in the uniaxial mode than in
Therefore, in a way similar to a previous study conducted by Kim the IDT mode. This is believed to be mainly caused by the combined
et al. (23), it can be concluded that the dynamic modulus measured effect of the following: (a) the fast rate of loading (0.1 s for load-
in the IDT mode is a reasonable estimate of the dynamic modulus ing and unloading with a rest period of 0.9 s) and the short duration
measured in the uniaxial mode. (5 s) of the resilient modulus test not being able to disclose the
prominent viscoelastic behavior of the material at elevated tem-
peratures and low frequencies that are equivalent to the long time
Generation of Master Curve with Backcalculated range in the physical time domain; (b) backcalculated frequency
Dynamic Modulus range (0.1 Hz to 10.0 Hz) not being wide enough to provide suffi-
cient over­lap of the dynamic modulus between successive tempera-
With the use of the methodology presented earlier in the paper, the tures; and (c) inevitable errors associated with the backcalculation.
power function parameters for the creep compliance were backcal- As explained by Kim et al., sufficient overlap of modulus values in
culated from the resilient modulus test data. Then the Prony series successive testing temperatures is one of the most important require-
was fitted to the creep compliance and was used to convert the creep ments in construction of the master curve (23). A closer look at Fig-
compliance to the dynamic modulus values corresponding to the ure 5a indicates no significant overlap of modulus between different
same frequencies at which the actual dynamic modulus tests were test temperatures.
Lee, Kim, Choubane, and Upshaw 7

Backcalculated Uniaxial Dynamic Modulus (ksi)


3,500

3,000

2,500

2,000
y = 0.99x – 45.37
R2 = 0.99
1,500

Line of Equality
1,000

500

0
0 500 1,000 1,500 2,000 2,500 3,000 3,500
Measured Uniaxial Dynamic Modulus (ksi)

0°C 10°C 20°C 40°C


(a)
Backcalculated IDT Dynamic Modulus (ksi)

3,500

3,000

2,500

2,000
y = 1.00x – 66.84
R2 = 0.98
1,500

1,000 Line of Equality

500

0
0 500 1,000 1,500 2,000 2,500 3,000 3,500
Measured IDT Dynamic Modulus (ksi)

0°C 10°C 20°C 40°C


(b)
FIGURE 5   Measured versus backcalculated (from resilient modulus) dynamic
modulus in (a) uniaxial test mode and (b) IDT test mode.

According to the reasoning as given, it was hypothesized that Enhancing Backcalculated Dynamic Modulus
the backcalculated creep compliance (and thus the dynamic modu- with Creep Test Data
lus) at higher temperatures and lower frequencies (or longer time
range) could be enhanced by using the creep test data for the same The idea behind the methodology presented here was first proposed by
mixture. According to AASHTO T 322, Florida DOT typically Kim et al. (13), who showed that the creep compliance obtained from
conducts the creep test for a duration of 1,000 s; thus, it is believed dynamic modulus testing at a single temperature and limited frequen-
to give a better estimate of the viscoelastic properties for a longer cies (0.1 Hz to 25.0 Hz) was a better representa­tion of the material
time range than the backcalculated properties from the 5-s resilient behavior at short loading times (or equivalently higher frequencies),
modulus test. In addition, because the creep test is conducted for whereas the creep compliance from the actual creep test represented
a long duration (compared with the resilient modulus test), it was the material behavior at long loading times (or equivalently lower
hypothesized that incorporation of the creep test data may allow frequencies). The creep compliances obtained from both the creep and
for estimating the dynamic modulus at a wider range of frequency. dynamic modulus tests were combined through the use of Prony series
8 Transportation Research Record 2296

3,500

3,000
Dynamic Modulus (ksi)

2,500

2,000

1,500

1,000

500

0
0.00001 0.0001 0.001 0.01 0.1 1 10 100 1,000
Reduced Frequency (Hz)

Measured Uniaxial Dynamic Modulus Backcalculated Uniaxial Dynamic Modulus

(a)

10,000
Dynamic Modulus (ksi)

1,000

100

10

1
0.00001 0.0001 0.001 0.01 0.1 1 10 100 1,000
Reduced Frequency (Hz)

Measured Uniaxial Dynamic Modulus Backcalculated Uniaxial Dynamic Modulus

(b)

FIGURE 6   Master curves at 10°C from measured versus backcalculated


(from resilient modulus test) uniaxial dynamic moduli in (a) arithmetic scale
and (b) logarithmic scale.
Lee, Kim, Choubane, and Upshaw 9

3,500

3,000
Dynamic Modulus (ksi)

2,500

2,000

1,500

1,000

500

0
0.0001 0.001 0.01 0.1 1 10 100 1,000
Reduced Frequency (Hz)

Measured IDT Dynamic Modulus Backcalculated IDT Dynamic Modulus

(a)

10,000

1,000
Dynamic Modulus (ksi)

100

10

1
0.0001 0.001 0.01 0.1 1 10 100 1,000
Reduced Frequency (Hz)

Measured IDT Dynamic Modulus Backcalculated IDT Dynamic Modulus

(b)

FIGURE 7   Master curves at 10°C from measured versus backcalculated


(from resilient modulus test) IDT dynamic moduli in (a) arithmetic scale
and (b) logarithmic scale.
10 Transportation Research Record 2296

to obtain an enhanced creep compliance curve that is accurate in both Figure 8 shows the comparison between the measured dynamic
the short and long time ranges. This was accomplished by replacing modulus and the backcalculated dynamic modulus enhanced with
the last two terms of the Prony series (shown in Equation 19) calculated creep data, for the range of frequency used in the dynamic modulus
from the dynamic modulus test data with the ones obtained from the testing (0.1 Hz to 10.0 Hz). When comparing Figure 8 with Figure 5,
creep test. The enhanced creep compliance was also verified through one can observe that the slopes of the linear regression line had neg-
an experimental study where the sample was subject to a cyclic load- ligible change, and the magnitudes of the y-intercept in both figures
ing of more than 300 cycles. The same methodology as described was had slightly decreased, indicating that the regression line moved
used in this study to enhance the creep compliance and the dynamic closer to the line of equality after the enhancement. After enhance-
modulus estimated at each temperature. The only obvious difference ment, the average absolute percent errors between the measured and
is that the creep data were used to enhance the backcalculated dynamic backcalculated dynamic moduli were determined to be 10.1% and
modulus (rather than the measured dynamic modulus), especially at 10.2% for the uniaxial and IDT test modes, respectively, and they
lower frequencies. were both below the estimated limit of accuracy. The percent error for
Backcalculated Uniaxial Dynamic Modulus (ksi)

3,500

3,000

2,500

2,000
y = 0.98x – 36.12
R 2 = 0.99
1,500

1,000 Line of Equality

500

0
0 500 1,000 1,500 2,000 2,500 3,000 3,500
Measured Uniaxial Dynamic Modulus (ksi)

0°C 10°C 20°C 40°C


(a)
Backcalculated IDT Dynamic Modulus (ksi)

3,500

3,000

2,500

2,000
y = 1.00x – 60.18
R 2 = 0.98
1,500

1,000 Line of Equality

500

0
0 500 1,000 1,500 2,000 2,500 3,000 3,500
Measured IDT Dynamic Modulus (ksi)

0°C 10°C 20°C 40°C


(b)

FIGURE 8   Measured versus backcalculated (from resilient modulus and creep)
dynamic modulus in (a) uniaxial test mode and (b) IDT test mode.
Lee, Kim, Choubane, and Upshaw 11

the uniaxial test mode increased by 0.2%, considered to be negligible from the enhanced dynamic modulus have been constructed by
for all practical purposes. taking these additional frequencies into account to allow for some
As briefly mentioned, another potential and possibly more cru- overlap of the modulus values between the different temperatures.
cial advantage of enhancing the backcalculated dynamic modulus by The updated master curves are shown in Figures 9 and 10, for the
using the creep data is that since the creep test was conducted for uniaxial and IDT modes, respectively, along with the master curves
1,000 s, it allows for estimating the dynamic modulus at frequencies from the measured dynamic modulus. Figure 11 shows the shift
lower than the lowest frequency at which the actual dynamic modulus factors that were calculated in obtaining the different master curves.
tests were conducted (0.1 Hz). Therefore, the enhanced creep compli- Compared with those shown in Figures 6 and 7, the backcalculated
ance was converted to the dynamic modulus at 0.005 Hz, 0.01 Hz, master curves shown in Figures 9 and 10 show better agreement
and 0.05 Hz, in addition to all frequencies where the actual dynamic with those from the measured dynamic modulus, especially at lower
modulus test was conducted (0.1 Hz to 10.0 Hz). The master curves frequencies. This clearly illustrates the advantage of enhancing the

3,500

3,000
Dynamic Modulus (ksi)

2,500

2,000

1,500

1,000

500

0
0.000001 0.0001 0.01 1 100
Reduced Frequency (Hz)

Measured Uniaxial Dynamic Modulus Backcalculated Uniaxial Dynamic Modulus

(a)

10,000

1,000
Dynamic Modulus (ksi)

100

10

1
0.000001 0.0001 0.01 1 100
Reduced Frequency (Hz)

Measured Uniaxial Dynamic Modulus Backcalculated Uniaxial Dynamic Modulus

(b)

FIGURE 9   Master curves at 10°C from measured versus backcalculated


(from resilient modulus and creep tests) uniaxial dynamic moduli in
(a) arithmetic and (b) logarithmic scale.
12 Transportation Research Record 2296

3,500

Dynamic Modulus (ksi) 3,000

2,500

2,000

1,500

1,000

500

0
0.000001 0.0001 0.01 1 100
Reduced Frequency (Hz)

Measured IDT Dynamic Modulus Backcalculated IDT Dynamic Modulus

(a)

10,000

1,000
Dynamic Modulus (ksi)

100

10

1
0.000001 0.0001 0.01 1 100
Reduced Frequency (Hz)

Measured IDT Dynamic Modulus Backcalculated IDT Dynamic Modulus

(b)

FIGURE 10   Master curves at 10°C from measured versus backcalculated


(from resilient modulus and creep tests) IDT dynamic moduli in (a) arithmetic
and (b) logarithmic scale.
Lee, Kim, Choubane, and Upshaw 13

Log (Shift Factor)


0
0 5 10 15 20 25 30 35 40 45
-1

-2

-3

-4

-5
Temperature (°C)
IDT Dynamic Modulus Test IDT Creep and Resilient Modulus Tests
Uniaxial Dynamic Modulus Test Uniaxial Creep and Resilient Modulus Tests

FIGURE 11   Shift factors for constructing master curves.

creep compliance and thus the backcalculated dynamic modulus by this study, would allow for evaluating the dynamic modulus of asphalt
using the creep test data. mixtures from existing resilient modulus and creep test data, and the
construction of the master curves for input into the new design guide.

Summary and Conclusions


Acknowledgments
Because of the difficulty associated with the dynamic modulus test
in the uniaxial mode, Florida DOT has been conducting most of The authors acknowledge James Greene, Gregory Sholar, Charles
its testing on field cores in the IDT mode. In addition, because the Holzschuher, Kenneth Green, and Sandy Mihocik for their diligent
dynamic modulus is not required in calculating the energy ratio efforts and invaluable recommendations provided throughout the
of the tested mixture, the dynamic modulus tests were not typically research.
conducted in previous research studies. Instead, Florida DOT pri-
marily conducted the Superpave IDT tests, resilient modulus, creep,
and strength tests, on the IDT specimens. To allow for the historical References
IDT test data to be used in the new design guide, a methodology
was necessary to estimate the dynamic modulus and to construct the  1. AASHTO Guide for Design of Pavement Structures. AASHTO, Washing-
dynamic modulus master curve from the above tests. ton, D.C., 1993.
In this study, a methodology was presented that allows for back-   2. Roque, R., B. Birgisson, C. Drakos, and B. Dietrich. Development
and Field Evaluation of Energy-Based Criteria for Top-Down Cracking
calculating the fundamental viscoelastic material property, that is, Performance of Hot Mix Asphalt. Journal of the Association of Asphalt
creep compliance from the random loading and its corresponding Paving Technologists, Vol. 73, 2004, pp. 229–260.
response. The creep compliance in regard to power function was back-   3. Zhang, Z., R. Roque, B. Birgisson, and B. Sangpetgnam. Identification
calculated from the commonly used uniaxial and IDT resilient mod- and Verification of a Suitable Crack Growth Law for Asphalt Mixtures.
Journal of the Association of Asphalt Paving Technologists, Vol. 70, 2001,
ulus test data by using the algorithms derived from the visco­elastic
pp. 206–241.
constitutive relations. Then, the dynamic modulus master curves   4. ARA, Inc., ERES Division. Guide for Mechanistic–Empirical Design
were generated from the measured and backcalculated dynamic of New and Rehabilitated Pavement Structures. NCHRP 1-37A, TRB,
modulus values. Although the master curve from the backcalculated National Research Council, Washington, D.C., 2004.
dynamic modulus showed reasonable agreement with those from the   5. Lacroix, A., A. A. M. Khandan, and Y. R. Kim. Predicting the Resilient
Modulus of Asphalt Concrete from the Dynamic Modulus. In Transpor-
measured dynamic modulus, deviations were noticed in the lower tation Research Record: Journal of the Transportation Research Board,
frequency (or higher temperature) ranges. In order to overcome this No. 2001, Transportation Research Board of the National Academies,
problem, the backcalculated dynamic modulus was enhanced with Washington, D.C., 2007, pp. 132–140.
the creep test data. The master curves generated with these enhanced   6. Lacroix, A. T., Y. R. Kim, and S. R. Ranjithan. Backcalculation of
dynamic modulus showed a better agreement with those from the Dynamic Modulus from Resilient Modulus of Asphalt Concrete with an
Artificial Neural Network. In Transportation Research Record: Jour-
measured dynamic modulus. nal of the Transportation Research Board, No. 2057, Transportation
Although the MEPDG could not endorse the use of the resilient Research Board of the National Academies, Washington, D.C., 2008,
modulus test protocol as the primary means of characterizing the pp. 107–113.
asphalt concrete modulus in the design of flexible pavements, it has   7. Lee, H. S., and J. Kim. Backcalculation of Dynamic Modulus from
Resilient Modulus Test Data. Canadian Journal of Civil Engineering,
been a primary mixture test, and a considerable amount of laboratory Vol. 38, No. 5, 2011, pp. 582–592.
testing has been done to date. The practical nature of the proposed back­   8. Wineman, A. S., and K. R. Rajagopal. Mechanical Response of Polymers:
calculation methodology, coupled with the approach developed in An Introduction. Cambridge University Press, New York, 2000.
14 Transportation Research Record 2296

  9. Findley, W. N., J. S. Lai, and K. Onaran. Creep and Relaxation of Non- 17. Kim, Y. R. Modeling of Asphalt Concrete. ASCE Press and McGraw-Hill,
linear Viscoelastic Materials. Dover Publications, Inc., New York, New York 2009.
1976. 18. Park, S. W., and Y. R. Kim. Fitting Prony-Series Viscoelastic Models
10. Roque, R., W. G. Buttlar, B. E. Ruth, M. Tia, S. W. Dickison, and with Power-Law Presmoothing. Journal of Materials in Civil Engineer-
B. Reid. Evaluation of SHRP Indirect Tension Tester to Mitigate Crack- ing, ASCE, Vol. 13, No. 1, 2001. pp. 26–32.
ing in Asphalt Concrete Pavements and Overlays. Final Report. FDOT 19. Park, S. W., and R. A. Schapery. Methods of Interconversion Between
B-9885. University of Florida, Gainesville, 1997. Linear Viscoelastic Material Functions. Part I—A Numerical Method
11. Roque, R., W. G. Buttlar, B. E. Ruth, and S. W. Dickison. Short-Loading- Based on Prony Series. International Journal of Solids and Structures,
Time Stiffness from Creep, Resilient Modulus, and Strength Tests Using Vol. 36, 1999, pp. 1653–1675.
Superpave Indirect Tension Test. In Transportation Research Record 1630, 20. Roque, R., and W. G. Buttlar. The Development of a Measurement and
TRB, National Research Council, Washington, D.C., 1998, pp. 10–20. Analysis System to Accurately Determine Asphalt Concrete Properties
12. Kim, J., R. Roque, and B. Birgisson. Obtaining Creep Compliance Param- Using the Indirect Tensile Mode. Journal of the Association of Asphalt
eters Accurately from Static or Cyclic Creep Tests. ASTM International, Paving Technologists, Vol. 61, 1992, pp. 304–332.
ASTM STP, Vol. 1469, 2005, pp. 177–197. 21. Buttlar, W. G., and R. Roque. Development and Evaluation of the Stra-
13. Kim, J., G. Sholar, and S. Kim. Determination of Accurate Creep tegic Highway Research Program Measurement and Analysis System
Compliance and Relaxation Modulus at a Single Temperature for Visco­ for Indirect Tensile Testing at Low Temperatures. In Transportation
elastic Solids. Journal of Materials in Civil Engineering, ASCE, Vol. 20, Research Record 1454, TRB, National Research Council, Washington,
No. 2, 2008, pp. 147–156. D.C., 1994, pp. 163–171.
14. Hondros, G. The Evaluation of Poisson’s Ratio and the Modulus of 22. Oh, J. H., and E. Fernando. Comparison of Resilient Modulus Values Used
Materials of a Low Tensile Resistance by the Brazilian (Indirect Tensile) in Pavement Design. Final Report, FDOT BDL76. Texas Transportation
Test with Particular Reference to Concrete. Australian Journal of Applied Institute, College Station, 2011.
Science, Vol. 10, No. 3, 1959, pp. 243–268. 23. Kim, Y. R., Y. Seo, M. King, and M. Momen. Dynamic Modulus Testing
15. Lee, H. S., and J. Kim. Determination of Viscoelastic Poisson’s Ratio and of Asphalt Concrete in Indirect Tension Mode. In Transportation Research
Creep Compliance from the Indirect Tension Test. Journal of Materials Record: Journal of the Transportation Research Board, No. 1891, Trans-
in Civil Engineering, ASCE, Vol. 21, No. 8, 2009, pp. 416–425. portation Research Board of the National Academies, Washington, D.C.,
16. Kim, J., and R. C. West. Application of the Viscoelastic Continuum 2004, pp. 163–173.
Damage Model to the Indirect Tension Test at a Single Temperature.
Journal of Engineering Mechanics, ASCE, Vol. 136, No. 4, 2010, The Characteristics of Asphalt Paving Mixtures to Meet Structural Requirements
pp. 496–505. Committee peer-reviewed this paper.

S-ar putea să vă placă și