Está en la página 1de 31

Hirsch Model for Estimating the Modulus of Asphalt

Concrete

Donald W. Christensen
1
, Jr., Terhi Pellinen
2
, and Ramon F.
Bonaquist
3


Abstract

The purpose of this paper is to present a new, rational and
effective model for estimating the modulus of asphalt concrete
using binder modulus and volumetric composition. The model is
based upon an existing version of the law of mixtures, called the
Hirsch model, which combines series and parallel elements of
phases. In applying the Hirsch model to asphalt concrete, the
relative proportion of material in parallel arrangement, called the
contact volume, is not constant but varies with time and
temperature. Several versions of the Hirsch model were evaluated,
included ones using mastic as the binder, and one in which the
effect of film thickness on asphalt binder modulus was
incorporated into the equation. The most effective model was the
simplest, in which the modulus of the asphalt concrete is directly
estimated from binder modulus, VMA, and VFA. Models are
presented for both dynamic complex shear modulus (|G*|), and
dynamic complex extensional modulus (|E*|). Semi-empirical
equations are also presented for estimating phase angle in shear
loading and in extensional loading. The proposed model was
verified by comparing predicted modulus and phase angles to
values reported in the literature for a range of mixtures.

Key Words: Superpave, asphalt concrete, dynamic modulus,
shear modulus, models, law of mixtures.

Introduction

The purpose of this paper is to describe in detail the
development and use of a new model for predicting the modulus of
asphalt concrete, which is based upon an existing model for

1
Senior Engineer and
3
Chief Operating Officer, Advanced Asphalt Technologies LLC, Sterling VA
2
Assistant Professor, Purdue University, West Lafayette IN
The oral presentation was made by Dr. Christensen


composite behavior called the Hirsch model. This model was
developed as part of NCHRP Projects 9-25 and 9-31, the
objectives of which are to evaluate existing Superpave
requirements for volumetric composition and modify them if
warranted. The Hirsch model for hot mix asphalt concrete
(HMAC) was developed during the initial phases of these NCHRP
projects to serve as a tool for analyzing the effect of changes in air
voids, voids in mineral aggregate (VMA) and other volumetric mix
factors on the modulus of asphalt concrete and related mechanical
properties. Although there are a number of models for estimating
the modulus of HMAC mixtures, these models are for the most
part empirical, they rely on obsolete semi-empirical asphalt binder
rheological data, and do not adequately utilize volumetric
composition in their formulation.
This paper presents the background needed to understand the
conceptual development of the modified Hirsch model, and how
this model has evolved to its current form. Several alternate
formulations of the model are presented: one that potentially
accounts for the stiffening effect of mineral filler within the mastic
of the mixture; one that includes a term for the effect of film
thickness; and a simple model based directly on volumetric
properties. These three models for predicting asphalt concrete
modulus were evaluated using a substantial set of data from the
Federal Highway Administrations (FHWAs) Accelerated
Loading Facility (ALF) project, the MNRoad Project, and the
WesTrack Project (1). These data were fit to the models using
non-linear least squares. All three versions exhibited nearly
identical accuracy, but model parameters relating to both mastic
modulus and film thickness effects were not significant.
Therefore, the simplest form of the Hirsch model is recommended
for general use by pavement engineers and researchers in
estimating modulus values of asphalt concrete mixtures.
This paper presents two versions of the Hirsch model for
asphalt concreteone for estimating the dynamic complex shear
modulus |G*|, and one for estimating the dynamic complex
modulus in extension, |E*|. Semi-empirical models are presented
for estimating phase angles for both loading methods based upon
the Hirsch model. This paper includes fairly extensive verification
of the Hirsch model, based upon a number of comparisons of
predicted moduli values with measured values as published in the


literature. This paper is based upon appendices to NCHRP Project
9-25 and Project 9-31 Interim Reports, but has been edited for
publication as a research paper (2, 3).

Background

The proposed method for estimating HMAC modulus is based
upon a widely used model for predicting properties of composite
materialsthe law of mixtures. As generally formulated, the law
of mixtures represents the mechanical response of two separate
phases in parallel (4):

E
c
= v
1
E
1
+ v
2
E
2
(1)

Where E refers to the modulus, or some other material property, v
refers to the volume fraction of a given phase, the subscript c refers
to the composite, and the subscripts 1 and 2 refer to different
phases present in the composite. A less common form of the law
of mixtures can be derived for phases in series (4):

1/E
c
= v
1
/E
1
+ v
2
/E
2
(2)

Where the variables are as defined above. In the early 1960s, T. J.
Hirsch developed a variation of the law of mixtures for modeling
the mechanical behavior of asphalt concrete (5). The Hirsch model
combines parallel and series arrangement of the phases (4):

1/E
c
= v
1s
/E
1
+ v
2s
/E
2
+ (v
1p
+v
2p
)
2
/(v
1p
E
1
+v
2p
E
2
) (3)

Where v
1s
and v
2s
refer to the volume fractions of phases 1 and 2,
respectively, in series arrangement, and v
1p
and v
2p
refer to the
volume fractions of phases 1 and 2, respectively, in parallel
arrangement. An equivalent expression for the Hirsch model can
be formulated if it is assumed that the relative proportions of phase
1 and phase 2 are the same in the series and parallel portion of the
model:
( )
|
|
.
|

\
|
+
+
|
|
.
|

\
|
+ =
2 2 1 1 2
2
1
1
1
1
1
E v E v
x
E
v
E
v
x
E
c
(4)


Where x is the ratio of phases in parallel arrangement to the total
volume. When x = 1, the Hirsch model produces results identical
to completely parallel phases (Equation 1), whereas when x = 0, it
represents a pure series arrangement (Equation 2). This model is
therefore quite flexible, and can be used to represent a wide range
of composite behavior. Hirsch found that for several portland
cement concrete mixtures, x had a value of about 0.5. These three
composite models are shown in Figure 1.

V
1
V
2

V
1
V
2

V
V
V V
1s
2s
1p 2p

(a) Parallel phases (b) Series phases (c) Hirsch model

Figure 1. Schematic representation of composite models for
parallel, series, and Hirsch (combination) arrangement of
phases.

Equation 3 is the basis for the proposed equation developed by
the research team for predicting asphalt concrete modulus from
binder modulus and volumetric properties. Asphalt concrete tends
to behave like a series composite at high temperature, but more
like a parallel composite at low temperature, and so the Hirsch
model should be appropriate for estimating the modulus of asphalt
concrete. However, for this model to be useful in modeling the
modulus of asphalt concrete, the relative proportions of the series
and parallel phases must be time and temperature dependent. The
aggregate phase in the parallel portion of the model is important in
characterizing the behavior and performance of HMAC, as it
represents that portion of the aggregate particles in intimate contact
with each other; this portion of the aggregate is therefore called the
aggregate contact volume, Pc. In general, as the aggregate contact
volume increases, so will the modulus, strength, and resistance to
permanent deformation. High values of Pc indicate a very


effective structure producing good strengths and stiffness, typical
at low temperature. Low values of Pc tend to occur at high
temperatures, and indicate mixtures with low strength and
stiffness.
Various modifications of the Hirsch model were devised and
evaluated in developing the final version. Four of the most
important are shown in Figure 2. In many of the early
formulations of the model, it was assumed that the binder in
asphalt concrete was the mastic (mineral filler plus binder) rather
than binder alone. In Figure 2(a), the bulk of the aggregate is in a
series arrangement with a parallel combination of the aggregate
contact volume and mastic. Better results were obtained with the
arrangement shown in 2(b), where the aggregate contact volume is
combined in parallel with a series arrangement of the bulk
aggregate and mastic. In this figure, Vc refers to the aggregate
contact volume, Va refers to the aggregate volume exclusive of
the contact volume, Vm is the mastic volume, and Vv is the air void
volume. For the configuration shown in 2(a), the appropriate
equation for dynamic modulus would be as follows:

( )
1
2
' 1 '

(

+ =
VmEm PcEa
Va
Ea
Va
Ec (5)
Where:
Ec = composite (asphalt concrete) modulus
Va = volume fraction of aggregate, excluding contact
volume and mineral filler
Ea = aggregate modulus
Pc = aggregate contact volume, as volume fraction
Vm = volume fraction of mastic
Em = mastic modulus

For the configuration shown in Figure 2(b), the equation for the
composite modulus is given by the following equation:
( )
1
2
'
1

|
.
|

\
|
+ + =
Em
Vm
Ea
Va
Vc VcEa Ec (6)





Va
Vm Vc Vv

Va
Vm
Vc
Vv

(a) Series Version (b) Parallel Version
Vc
Vv Va,
Vm
Vap Vvp Vmp
Vas
Vms
Vvs

(c) Dispersed Version (d) Alternate Version

Figure 2. Schematic representation of four alternate versions
of modified Hirsch model.

Where the variables are as defined previously. The arrangement in
2(b) can in fact be generalized by using an exponent m rather than
1 in the series portion of the model:
( ) ( )
m
m m m
VmEm Ea Va Vc VcEa Ec
1
1 1
' 1 + + =

(7)
The exponent m can take any value from 1 to 1, though in this
case meaningful models would result only for values between 1
and about 0.2. For the case of perfect spheres of aggregates
within the mastic matrix, m = -0.5; this arrangement is shown in
Figure 2(c). Thus, Equation 7 represents a very flexible form of
the Hirsch model.
In order to use the models represented in Figures 2(a), (b) and
(c), the modulus of the mastic must be estimated. The NCHRP 9-
25 research team has evaluated two approaches to calculating the
modulus of the mastic based upon the binder modulus and the
volume fractions of binder and mineral filler. The first is a version


of the Einstein equation developed by Shashidhar and his
associates (6)

( )
Eb
Vf CVf
AVf
Em
(

+
+
=
' ' 1 1
' 1
(8)
Where:
A = K
E
-1
K
E
= generalized Einstein coefficient
Vf = volume fraction filler in mastic = Vf/Vm
C = (1-Vf
max
)/Vf
max
2

Vf
max
= maximum volume fraction of filler in mastic

The Einstein equation is rational in form and appears to be
reasonably accurate. However, the predictions are quite sensitive
to the maximum volume fraction of filler (Vf
max
). As the volume
fraction of filler in the mastic approaches this maximum, the
predicted modulus of the mastic becomes extremely high; when
Vf = Vf
max
, the predicted value of Em is infinite. In reality, the
binder would most likely never actually incorporate this amount of
filler, as this represents a hypothetical maximum, at which the air
void content of the mastic would be exactly zero. Instead, a certain
amount of excess filler would occur throughout the mix as free
filler. However, this is difficult to account for within the
framework of a relatively simple method for predicting mixture
modulus.
An alternative approach involves the use of the generalized law
of mixtures:
( )
n
n n
Eb Vb Ea Vf Em
1
' ' + = (9)

Where:
Vb = volume fraction binder in mastic = Vb/Vm
Eb = modulus of binder
n = exponent with values from 1 to +1

For dispersed systems, n will range from slightly less than 0 to
about 0.5. A preliminary evaluation based upon typical binder-
filler systems has indicated that the predictions of Equation 9 are
of the same magnitude as those of Equation 8 when n = -0.2.
Equation 9, because of its relative simplicity and robustness, is
attractive for use in the various models. Results using this


equation compared favorably with those produced using the
Einstein equation. Therefore much of the analyses of the various
forms of the Hirsch model therefore used Equation 9 for estimating
mastic modulus.
Preliminary analyses showed that these three versions of the
Hirsch model did not exhibit good accuracy. The version of the
model that was found to exhibit consistently good accuracy is
shown in Figure 2(d). This alternate formulation of the Hirsch
model is very similar to the original model. The only difference in
this version is that the series and parallel sub-units of the model are
combined in parallel, rather than in series. This in effect places
more emphasis on the parallel sub-unit of the model. This version
of the Hirsch model produced the best results, and has the
additional advantages of being relatively simple and very similar to
the original version of the model as formulated by Hirsch.
Mathematically, it can be expressed using the following equation:


( )
( )
1
2
2
'
'
'

+
+ + +
+ + =
VmsEm
Vvs Vms
Ea
s Va
Vvs Vms s Va
VmpEm pEa Va Ec

(10)
Where the variables are as defined previously, but the addition of
the subscripts p and s denote parallel and series phases,
respectively (see Figure 2(d)). As with the standard version of the
Hirsch model, this alternate formulation can be stated in somewhat
simpler terms by using the contact volume Pc to represent the
proportion of parallel to total phase volume:

( ) ( )
( )
1
2
'
1 '

(

+
+ + + =
VmEm
Vv Vm
Ea
Va
Pc VmEm Ea Va Pc Ec (11)
Various types of functions where used to describe the
contact factor for use in Equation 11. Eventually, the research
team found that the following equation for the contact factor
provided the best results:



1
1
'
'
2
0
P
P
VMA
Em VFM
P
VMA
Em VFM
P
Pc
|
.
|

\
|

+
|
.
|

\
|

+
= (12)
Where VFM is the fraction of aggregate voids filled with mastic,
VMA is voids in the mineral aggregate, exclusive of mineral filler,
and P
0
, P
1
, and P
2
are empirically determined constants. Equation
12 is in fact a type of logistic function, which produces a sigmoidal
response in log-log space typical of the behavior of many
viscoelastic materials. The first constant in Equation 12, P
0
, is
directly related to the contact factor at high temperatures and/or
low frequencies; P
1
is an exponent related to the rate of change of
the contact factor with respect to the binder modulus Em; P
2
is
related to the location of the contact factor, which is directly
related to the overall stiffness of the asphalt concrete mixture.
As will be discussed later in this paper, preliminary analysis of
the mastic version of the Hirsch model demonstrated that it was
accurate, but also seemed to indicate that it was not necessary to
consider the stiffening effects of mastic. Instead, a simpler version
of this model, which treats asphalt concrete as a simple three-phase
system of aggregate, asphalt binder, and air voids, seemed
appropriate:
( ) ( )
( )
1
2
1

(

+
+ + + =
VbEb
Vv Vb
Ea
Va
Pc VbEb VaEa Pc Ec
(13)
Where Eb represents the binder modulus, and Vb represents the
effective binder volume. Note that the aggregate volumes in
Equation 13 now represent true aggregate volume including the
volume of mineral filler. The corresponding equation for the
contact factor is identical to Equation 12, except that true VMA is
substituted for VMA (voids plus binder volume plus mineral filler
volume).
A final modification of the Hirsch model was devised to
determine if film thickness could be incorporated into the
expression for modulus. The equation for this version of the
Hirsch model is essentially identical to Equation 13, but the
effective binder modulus, Eb, is substituted for the true binder
modulus:


( ) ( )
( )
1
2
'
1 '

(

+
+ + + =
VbEb
Vv Vb
Ea
Va
Pc VbEb VaEa Pc Ec
(14)
The effective binder modulus is calculated by assuming that
the binder modulus at the aggregate surface is equivalent to the
glassy modulus, but then decreases to the normal binder modulus
value over a certain characteristic distance, t
T
, representing the
transition zone between the aggregate and binder. If the binder
film thickness is called t
F
, the equation for effective binder
modulus is then given using the following equation:

( ) Eb t E t t
EbE t
Eb
T g T F
g F
+
= ' (15)
Where E
g
is the glassy modulus of asphalt cement binder, assumed
to be 1 GPa (145,000psi).
The reader should keep in mind that the above functions for the
Hirsch model, though stated in terms of the extensional modulus E,
can be just as easily formulated in terms of the shear modulus G.
Furthermore, the modulus values used in any of these equations
can be determined from creep, stress relaxation, or dynamic
modulus tests. The Hirsch model is a rational, though semi-
empirical method of predicting asphalt concrete modulusthat is,
its structure is logical and based upon the law of mixtures, but its
use in practice requires calibration with measured data. It should
not be confused with a constitutive equation, which rigorously
delineates the relationships among stress, strain, and material
properties (such as modulus) in two or three dimensions. The
Hirsch model can however be used to estimate modulus values that
are used in various such constitutive equations.

Data used in Refining the Hirsch Model

In order to evaluate the various versions of the Hirsch model,
and refine the most promising of these models it was necessary to
establish a database of modulus values for a wide range of
mixtures. Such a data set was created, based upon dynamic
modulus measurements made at Advanced Asphalt Technologies,
LLC, (AAT) and at the Arizona State University (ASU) as part of
NCHRP Project 9-19 (4). The AAT data set consists of dynamic


shear modulus and phase angle data collected using the SST, using
the frequency sweep tests. The ASU data set consists of dynamic
compression modulus and phase angle data. The mixtures tested
were the same for both data sets, and originated from FHWAs
Accelerated Loading Facility (ALF) project, the MnROAD
Project, and the WesTrack Project; Pellinen has thoroughly
summarized these projects and the materials used in both data sets
(1). Inclusion of both shear and compression data allowed the
research team to develop versions of the Hirsch model for both
cases, which is essential since the relationship between |G*| and
|E*| for asphalt concrete mixtures is not at all straightforward nor
well documented.
A summary of the database is given as Table 1. The database
includes results from testing on 18 mixtures using eight different
binders and 5 different aggregate sizes and gradations. A total of
206 observations are included in the dataset for each type of
measurement (shear and compression). A wide range of
volumetric compositions is also represented, although one
shortcoming in this data is the lack of mixtures with low air voids
and low VFA. The data set is however extensive, and suitable for
initial development of the Hirsch model. It is possible that larger
and more robust data sets could be used in the future to further
refine this method for estimating asphalt concrete modulus.

Method of Analysis
The general approach used in evaluating the various versions
of the Hirsch model, and refinement of the most promising was
non-linear least squares (7). Christensen has presented a detailed
description of the use of this technique in analyzing IDT creep data
(8). In general terms, this procedure uses an iterative, numerical
procedure to calculate the values for parameters in a function so
that the sum of the square of the error terms is minimized.
Graphical techniques were used to identify and eliminate several
records that were clearly outliers, and to determine if other
potential predictor variables existed which were not included in the
model. All of the outliers eliminated from the data were high-
temperature measurements made using the SST at very low stress
levels. These measurements were considered to be unreliable,
since the very low stress levels used were difficult to measure and
appeared to be quite noisy.


After refinement of the final versions of the Hirsch model,
simple curve fitting was used to develop a function for phase
angle. Such equations are sometimes needed for calculation of
storage and loss modulus values.

Table 1. Summary of Database of Mixture Modulus Data

Factor ALF MNRo
ad
WesTrack T
otal
Design
Method
Marshall Marsha
ll
Superpave N
/A
Binders AC-5, 10,
20
SBS-
modified
PE-
modified
AC-20
120/15
0-PEN
PG-64-22 8
Aggregate
Sizes and
Gradations
19-mm
dense
37.5-mm
fine
9.5-
mm fine
19-mm
fine
19-mm
coarse
5
Mixtures 7 5 6 1
8
Total Data
Points
78 59 69 2
06
For Complete Database
Voids, Vol.
%
5.6 to 11.2
VMA, Vol.
%
13.7 to 21.6
VFA, % 38.7 to 68.0
Temperatur
es, C
4, 21, 38
Frequencies 0.1 and 5
|G*|, MPa 20.0 to 3,880
|E*|, MPa 183 to 20,900
Phase angle,
degrees
8 to 61

Results
The three most promising versions of the Hirsch model were
all variations of the alternate formulation shown graphically in


Figure 2(d), and represented in its basic form in Equation 10. The
three versions of this model are called here the mastic version
(Equation 11), the simple version (Equation 13), and the transition
zone version (Equations 14 and 15). In all cases, Equation 12 was
used to characterize the contact factor Pc.
Model for |G*| The results of the non-linear least squares
analyses of these three models for dynamic complex shear
modulus (|G*| as measured using the SST) are summarized in
Table 2. The value for parameter P
0
could not be determined
reliably using the least-squares procedure, probably because the
database did not include enough values at high temperatures and
low frequencies. However, it is important to include at least an
estimate for this parameter, since it represents the limiting asphalt
concrete modulus at high temperatures and/or low frequencies.
Comparing model predictions with published master curves for
asphalt concrete, it was estimated that an appropriate value for this
constant is about 3.
All three versions of the Hirsch model exhibited identical r
2

values of 96.8 percent, suggesting that the more complex versions
of the model, intended to account for mineral-filler effects and the
effect of film thickness, are no more effective than the model
treating asphalt concrete as a simple three-phase system. For the
mastic model, the law of mixtures exponent for the mastic, n
(Equation 9) has an unrealistic value of 6.9, and a standard error
of 281 percent, indicating that this parameter does not contribute
significantly to the accuracy of the model. For the transition zone
model, similar problems are observed in the estimate of the
transition zone thickness; t
T
is estimated at only 0.03 microns, and
has a huge standard error of estimate of almost 4,000 percent. It is
therefore concluded that the most effective version of the Hirsch
model is the relative simple version represented by Equation 13,
which can be given in terms of the complex shear modulus, |G*|,
VFA, and VMA:



Table 2. Summary of Least-Squares Estimation of Parameters
for |G*| Hirsch Model.

Mastic Model
Equation(s
):
11 & 12
Parameter Estimat
e
Std.
Error
Ea, lb/in
2
635,000 7.2 %
P
0
3 N/A
P
1
0.678 2.4 %
P
2
396 10.4 %
n (Eqn. 9) 6.9 281%
t
T,
microns N/A N/A
r
2
96.8 % N/A

Simple Model
Equation(s
):
13 & 12
Parameter Estimat
e
Std.
Error
Ea, lb/in
2
601,000 7.1 %
P
0
3 N/A
P
1
0.678 2.4 %
P
2
396 6.9 %
n (Eqn. 9) N/A N/A
t
T,
microns N/A N/A
r
2
96.8 % N/A

Transition Zone Model
Equation(s
):
14, 15 & 12
Parameter Estimate Std.
Error
Ea, lb/in
2
601,000 7.2 %
P
0
3 N/A
P
1
0.678 2.4 %
P
2
452 8.7 %
n (Eqn. 9) N/A N/A
t
T,
microns 0.03 3,760 %
r
2
96.8 % N/A


( )
( )
1
* 000 , 601
100 1
1
000 , 10
* 100 1 000 , 601 *

(
(

+
(

|
.
|

\
|
+ =
binder
binder mix
G VFA
VMA VMA
Pc
VMA VFA
G VMA Pc G

(16)
Where |G*|
mix
is the complex shear modulus for the mixture, and
|G*|
binder
is the complex shear modulus for the binder, both in units
of lb/in
2
, and VFA and VMA are both given as percentages. The
binder modulus can either be determined experimentally using the
dynamic shear rheometer (DSR) or similar device, or can be
estimated from one of several available mathematical models. It
should be at the same temperature and loading time selected for the
mixture modulus, and in consistent units. The contact factor is
given by the following function:

678 . 0
678 . 0
*
396
*
3
|
|
.
|

\
|
+
|
|
.
|

\
|
+
=
VMA
G VFA
VMA
G VFA
Pc
binder
binder
(17)
Figure 3 shows the |G*| values predicted using Equations 16
and 17 versus measured values. There is generally good
agreement.
Model for |E*| Non-linear least squares analysis was also
performed using the |E*| data gathered by Pellinen at Arizona State
University as part of NCHRP Project 9-19 (1). In the initial
calibration of the model for |E*|, performed during Phase I of
NCHRP Project 9-25, the data set used coincided with that
developed for the |G*| model. However, subsequent evaluation of
the model showed that results at extreme high and low
temperatures were not always accurate. Therefore, as part of Phase
I of NCHRP Project 9-31, an expanded data set was created, which
included additional data at 9 and 54C. This resulted in a
somewhat greater range of values for both |E*| and phase angle in
the data set. Furthermore, in the expanded data set replicate
measurements were averaged, partly because of the greatly
expanded size of the resulting data set, but also to minimize



1000
10000
100000
1000000
1000 10000 100000 1000000
Measured |G*|, psi
P
r
e
d
i
c
t
e
d

|
G
*
|
,

p
s
i

Figure 3. Predicted Versus Measured Shear Complex
Modulus (r
2
= 96.8 percent); Solid Line Represents Equality.

variability due to experimental error. It was felt that this approach
would provide a better estimate of the accuracy of the model. The
final values for the Hirsch model parameters for dynamic complex
modulus in extension were estimated to be as follows:
Ea: 4,230,000 psi ( 6.5 %)
P
0
: 649 ( 9.0 %)
P
1
: 19.7 ( 20.7 %)
P
2
: .575 ( 3.0 %)

The r
2
value for the |E*| data (98.2 %) is slightly higher than
for shear data. Also, note the much higher value for Ea, which is
expected as compression moduli for most materials, including
asphalt concrete, are almost always much higher than shear
moduli. The other parameters are similar to the values determined
for the shear model. The predicted and measured values for |E*|
are shown in Figure 4; residuals as a function of predicted |E*| are
shown in Figure 5. Applying appropriate rounding to the
coefficients listed above, the pertinent equation for compression
modulus is as follows:



( )
( )
1
* 3 000 , 200 , 4
100 1
1
000 , 10
* 3 100 1 000 , 200 , 4 *

(
(

+
(

|
.
|

\
|
+ =
binder
binder mix
G VFA
VMA VMA
Pc
VMA VFA
G VMA Pc E

(18)
1.0E+04
1.0E+05
1.0E+06
1.0E+07
1.0E+04 1.0E+05 1.0E+06 1.0E+07
Measured |E*|, psi
P
r
e
d
i
c
t
e
d

|
E
*
|
,

p
s
i

Figure 4. Predicted and Measured values for Complex
Modulus in Compression (r
2
= 98.2 percent); Solid Line
Represents Equality.

Note that the binder shear modulus in Equation 18 is multiplied
by 3 as an estimate of the extensional modulus: |E*|binder
3|G*|binder. This is based on an assumption of incompressibility,
that is, that Poissons ratio is 0.5. The following function is used
for estimating the contact volume for use in conjunction with
Equation 18:



-0.5
-0.4
-0.3
-0.2
-0.1
0.0
0.1
0.2
0.3
0.4
0.5
1.0E+04 1.0E+05 1.0E+06 1.0E+07
Predicted |E*|, psi
R
e
s
i
d
u
a
l

(
L
o
g

|
E
*
|
)

Figure 5. Residuals (Predicted-Measured) Values of Log |E*|
for Hirsch Model.


58 . 0
58 . 0
* 3
650
* 3
20
|
|
.
|

\
|
+
|
|
.
|

\
|
+
=
VMA
G VFA
VMA
G VFA
Pc
binder
binder
(19)
Prediction of Phase Angle As discussed previously, in addition
to a need for estimating modulus in shear and compression, there is
also a need to estimate the phase angle from compositional data.
For example, the fatigue models developed during the Strategic
Highway Research Program (SHRP) used the loss modulus as a
predictor variable, which is a function of both the complex
modulus and the phase angle (9). The research team found a good
empirical relationship between the log of the contact factor Pc
(Equation 12) and the phase angle. Plots showing these
relationships, which includes the empirically determined equations
for phase angle as a function of log (Pc), are shown in Figures 6
and 7. The following function can be used to estimate phase angle
using Pc as determined from shear data (r
2
= 82.9 %):
( ) 6 . 9 log 39 log 5 . 9
2
+ = Pc Pc (20)



y = -9.4633x
2
- 38.834x + 9.6031
R
2
= 0.8294
0
10
20
30
40
50
60
70
-3.0 -2.5 -2.0 -1.5 -1.0 -0.5 0.0
Log (Pc)
P
h
a
s
e

A
n
g
l
e
,

D
e
g
r
e
e
s

Figure 6. Phase Angle as a Function of Log (Pc), Shear Data.

y = -20.56x
2
- 54.619x
R
2
= 0.8906
0.00
10.00
20.00
30.00
40.00
50.00
-3.0 -2.5 -2.0 -1.5 -1.0 -0.5 0.0
Log (Pc)
P
h
a
s
e

A
n
g
l
e
,

D
e
g
r
e
e
s

Figure 7. Phase Angle as a Function of Log (Pc), Compression
Data.

The corresponding function for estimating from compression
data is given by the following Equation (r
2
= 89 percent):



( ) Pc Pc log 55 log 21
2
= (21)

These relationships, though not highly accurate, are useful for
estimating phase angle from asphalt concrete volumetric
composition. It should be kept in mind that the measurement of
phase angle is particularly difficult, so the amount of scatter in
Figures 6 and 7 is not surprising. There are two significant
differences in these plots. The phase angle in shear (Figure 6) has
a value of 9.6 when Pc is zero; this indicates that in shear
measurements, the phase angle is about 10 degrees even at very
low temperatures and high frequencies. For the extensional data,
this is not truethe phase angle is zero when Pc is zero. It is
possible that the non-zero phase angle when Pc is zero for shear
data is an artifact, caused by friction in the horizontal bearing used
in the SST system (there are no bearings in the compression tests,
outside of those inherent in the actuators). The other significant
difference is that for the compression data, there is a clear
maximum in phase angle, whereas the shear data shows no
maximum. In general, it appears that the phase angle data in
extension are somewhat more reasonable than that determined in
shear.

Verification of the Model
In order to verify the final version of the Hirsch model in as
independent a manner as possible within the limited time available
during the initial phases of NCHRP Projects 9-25 and 9-31, the
models for shear modulus (Equations 16, 17 and 20) were used to
estimate shear modulus for data as reported by Alavi and
Monismith in research related to the original SHRP program (10).
These measurements were made using a cylindrical shear test, a
completely different technique than that upon which the model was
developed. Alavi and Monismith used one aggregate (19-mm
nominal maximum size) and one binder, but a range of asphalt and
air void contents. Binder modulus values were estimated using the
model developed by Christensen and Anderson for SHRP binders
(11). Some volumetric information was estimated because of the
limited data reported by Alavi and Monismith. The predicted and
measured values of shear modulus are shown in Figure 8; the
agreement is good, though the Hirsch model appears to slightly
under-predict modulus values at higher levels. In Figure 9,


predicted and measured phase angles are plotted for these same
data. The agreement here is not as good, especially at higher phase
angles. In Alavi and Monismiths data, the phase angle value, after
peaking, decreases to very low values. In the SST shear data, on
the other hand, the phase angle increases at a decreasing rate to a
maximum value without any subsequent decrease. The research
team believes this is an inherent difference in the measurement
methods. Phase angle predictions using the Hirsch model are
probably most accurate at low temperatures and/or high
frequencies.
As a second verification of the model, the complex modulus in
compression (|E*|) was predicted using both the Hirsch model and
using Andrei and Witczaks equation, and compared to the
measured values reported by Alavi and Monismith (12, ,10). This
comparison is shown graphically in Figure 10. The values for the
Hirsch model are in excellent agreement, whereas Andrei and
Witczaks equation slightly under-predicts at higher modulus
values. Although this comparison is too limited to make broad
generalizations, it suggests that the Hirsch model is in general
agreement with Andrei and Witczaks equation, and is at least as
accurate. Hirsch model predictions for |E*| are probably in better
agreement with Alavi and Monismiths data, compared to the |G*|
predictions, because the experimental technique used by Alavi and
Monismith for the shear measurements was substantially different
than that used by AAT in making the SST measurements. In
contrast, Pellinens compression moduli were probably determined
using methods giving results comparable to Alavi and
Monismiths.
Another comparison useful for verification of the Hirsch model
is shown in Figure 11, which is a master curve for mixture V0W1
from Alavi and Monismiths study (10). This figure shows
predicted and measured values for |E*| and phase angle as a
function of reduced frequency at 40C. This figure confirms that
the frequency dependence and general shape of the functions for
complex modulus and phase angle as predicted by the Hirsch
model are reasonable and in good agreement with experimental
values. However, it appears that the phase angle predictions do not
vary quite as strongly with frequency as the measured values.



10
100
1000
10000
100000
10 100 1000 10000
Predicted |G*|, MPa
M
e
a
s
u
r
e
d

|
G
*
|
,

M
P
a

Figure 8. Measured Complex Shear Modulus and Values
Predicted Using the Hirsch Model; Solid Line Represents
Equality (data from Refereence 10).
0
10
20
30
40
50
60
70
0 10 20 30 40 50 60 70
Predicted Phase Angle, Degrees
M
e
a
s
u
r
e
d

P
h
a
s
e

A
n
g
l
e
,

D
e
g
r
e
e
s

Figure 9. Measured Phase Angle s and Values Predicted Using
the Hirsch Model; Solid Line Represents Equality (data from
Reference 10).




100
1000
10000
100000
100 1000 10000 100000
Measured |E*|, MPa
P
r
e
d
i
c
t
e
d

|
E
*
|
,

M
P
a
Witczak Model
Hirsch Model
Equality

Figure 10. Predicted and Measured |E*| Values; Solid Line
Represents Equality (data from Reference 10).

A more thorough verification of the Hirsch model was
performed using data from a recent sensitivity study by Witczak
and his associates (13). In this study, |E*| measurements were
made on a range of mix variations based upon an Arizona
Department of Transportation mixture. This asphalt concrete used
a 25-mm nominal maximum aggregate size blend with an PG 64-
22 asphalt. The basic mix design was varied using four different
target air void levels (1.5, 4, 7 and 10 percent) and four binder
content levels (3.9, 4.55, 5.2, and 5.9 percent). In addition to
measured |E*| values, modulus values were also predicted using
Witczaks equation. In Figure 12, |E*| values predicted using the
Hirsch model are compared to measured values reported by
Witczak and his team. The standard error in this case is 41 %,
which is slightly better than the standard error for Witczaks model
(45 percent), but about double that for the experimental error for
these data (20 percent). Although not as good as actual
measurements, the accuracy of the model predictions is probably
suitable for many practical design and analysis applications.
Estimated modulus values in fact are probably similar in reliability
to measured values when only limited replicate tests can be
performed, or when laboratory personnel are not experienced in
making modulus measurements on asphalt concrete specimens.


100
1000
10000
100000
1.E-04 1.E-02 1.E+00 1.E+02 1.E+04 1.E+06
Reduced Frequency at 40 C, Hz
|
E
*
|
,

M
P
a
0
10
20
30
40
50
60
70
80
90
P
h
a
s
e

A
n
g
l
e
,

d
e
g
r
e
e
s
Meas. |E*|
Pred. |E*|
Meas. Phase
Pred. Phase

Figure 11. Master Curve for Mixture V0W1 as Reported by
Alavi and Monismith (10), and as Predicted Using the Hirsch
Model.

A second comparison with these data is shown in Figure 13,
which is a plot of modulus values predicted using the Hirsch model
against values determined using Witczaks equation. In this case,
the agreement is extremely close. However, such close agreement
between these models will not always occur; in this case, it is
probably fortuitous and brought about by close agreement between
the SHRP binder data used in the Hirsch model and the routine
data used in Witczaks equation. Given the close agreement
between these models, and the similar standard errors compared to
the measured |E*| values, this confirms the finding that the Hirsch
model and Witczaks equation are of similar accuracy and in
reasonably close agreement. As pointed out earlier, the Hirsch
model is simpler and more rational, and makes use of |G*| data as
produced using standard Superpave binder tests..



10
100
1000
10000
10 100 1000 10000
Measured |E*|, ksi
P
r
e
d
i
c
t
e
d

|
E
*
|
,

k
s
i

Figure 12. Predicted and Measured Dynamic Modulus Values
Using Data from NCHRP 9-19 Sensitivity Study; Solid Line
Represents Equality (13).

10
100
1000
10000
10 100 1000 10000
Predicted |E*| (ksi), Witczak Model
P
r
e
d
i
c
t
e
d

|
E
*
|

(
k
s
i
)
,

H
i
r
s
c
h

M
o
d
e
l

Figure 13. Dynamic Modulus Values Predicted Using the
Hirsch Model and Using Witczaks Equation, NCHRP 9-19
Sensitivity Data; Solid Line Represents Equality (13).

A final verification of the Hirsch model was done using
bending beam data and mixture creep compliance data published
as part of the SHRP project (14). Mixture creep compliance values


were determined on a range of field cores using the indirect tensile
creep tests. Using reported binder creep modulus values as
determined using the bending beam rheometer, and reported
volumetric composition, the Hirsch model was used to predict
creep modulus values, which were compared to values reported by
Lytton and associates. The results of this comparison are shown in
Figure 14. The Hirsch model appears to under-predict the modulus
values at high stiffness levels, but at this point the mixture creep
modulus is very high, in the range of 2 to 6 millionpsi, and
approaching the glassy limit. At values below about 2 millionpsi,
the predicted values are in reasonably good agreement with the
measured values. Comparison of Figure 14 with Figure 12 show
that these comparisons are in agreement; the under-prediction seen
with the Hirsch model at very high modulus values is probably a
function of the estimated constant glassy modulus value
(4,200,000psi). From a practical perspective, once an asphalt
concrete mixture is exhibiting modulus values in this range, it will
be extremely stiff and brittle, and would be probably subject to
high levels of thermal cracking and fatigue cracking if used in a
pavement in this condition. This discrepancy should therefore not
be considered a significant problem.

100
1000
10000
100 1000 10000
Measured Creep Modulus, ksi
P
r
e
d
i
c
t
e
d

C
r
e
e
p

M
o
d
u
l
u
s
,

k
s
i5 seconds
100 seconds
equality

Figure 14. Predicted Creep Modulus Values Compared with
Measured Values, Using SHRP A-005 Low-Temperature Data;
Solid Line Represents Equality (14).



Comparison of |E*| and |G*| Values
A final important comparison can be made between the
measured moduli in shear and compression for the entire data set
used to develop the Hirsch model. Recall that the shear moduli
were measured at AAT using the SST frequency sweep procedure,
while the compression moduli were determined entirely
independently at Arizona State University, though the same
materials were used. This allows a direct comparison of
independently determined modulus values. A plot of |G*| versus
|E*| for these data is shown in Figure 15. Two important
observations can be made concerning this figure. First, the R
2

value for this relationship, 92.9 %, is actually lower than the values
for the predicted and measured moduli for both shear and
compression data. This indicates that there is a large amount of
noise in the modulus measurements. Because of the relatively
poor precision of modulus determinations on asphalt concrete, it is
possible that modulus predictions using the Hirsch model (or other
similarly accurate procedure) might be nearly as accurate as
independent modulus measurements made on the same mixture.
Thus for many practical applications, using modulus values
predicted using an accurate model is probably just as effective as
using measured values, and of course much less time consuming
and expensive.
A second important observation concerning Figure 15 is the
relationship between |G*| and |E*|. Paving engineers might be
tempted to try to convert compression moduli to shear moduli
using the simple relationship G = E/(1+), where is Poissons
ratio. If a value of 0.4 is assumed for , this suggests that |G*|
|E*|/2.8. The line representing this equation is included on Figure
15, and clearly shows that this approximation is not at all accurate.
Although the SST is certainly a far from ideal test system, the
agreement between SST-based predictions and Alavi and
Monismiths data (Figure 5) suggests that the relationship shown
in Figure 15 is for the most part real. The inaccuracy of simple
conversions between shear and compression moduli is probably
due to several factors, including non-linearity. However, non-
linearity is generally not significant at low temperatures and high
frequencies, but as seen in Figure 15, the simple conversion is not
accurate even at high modulus values. This discrepancy is
probably caused in large part by anisotropy in the mechanical


y = 0.0603x
1.0887
R
2
= 0.9291
1000
10000
100000
1000000
10000 100000 1000000 10000000
Measured |E*|, psi
M
e
a
s
u
r
e
d

|
G
*
|
,

p
s
i
|G*| = |E*|/2.8

Figure 15. Comparison of Measured Modulus in Shear (|G*|)
and Compression (|E*|), for ALF, MnROAD, and WesTrack
Data.

behavior of asphalt concrete. Other researchers have found similar
discrepancies between HMAC modulus values made using
different loading geometries (9) Therefore, engineers should use
caution in selecting modulus test data or predicted modulus values
for use in pavement design and analysis. For example, Bonnaures
fatigue equations were developed on the basis of flexural modulus
data (15, 16); for best accuracy, the modulus values used in
conjunction with his fatigue equation should therefore be values
based on flexural measurements, or if these are not available,
extensional data. Engineers should not rely on conversions
between shear, compression, and flexural data based upon linear
elastic theory and assumptions of homogeneity and anisotropy.
Although not as elegant, empirically determined relationships such
as that illustrated here and reported by other researchers are likely
to be more reliable.

Conclusions

A relatively simple version of the Hirsch model for
composite behavior has been developed for estimating the


complex modulus and phase angle of asphalt concrete in
shear and compression.
The Hirsch model is both simpler and more rational than
existing models for predicting modulus, and requires as
input only asphalt concrete volumetric composition and
SHRP binder data, and so is ideal for use in examining the
relationships among HMAC volumetrics, modulus, and
related aspects of pavement performance
The model appears to be in good agreement with Andrei
and Witczaks model, and is of similar accuracy.
Modulus values predicted using the Hirsch model are
potentially nearly as accurate as measured modulus values.
For many pavement design and analysis procedures,
predicted modulus values can be effectively used, and can
be determined much more quickly and cheaply. Using
predicted modulus values should be considered when
reliable measurements by experienced lab personnel are not
available.
The relationship among modulus values determined using
different test methods are complex and cannot be
accurately predicted using simple linear elastic theory and
assumptions of homogeneity and anisotropy. Engineers
should be careful to select the appropriate modulus value
for their intended purpose, whether that value is measured
or predicted using one of the available models.

References

1. T. K. Pellinen, Investigation of the Use of Dynamic Modulus as
an Indicator of Hot-Mix Asphalt Performance, A Dissertation
Presented in Partial Fulfillment of the Requirements for the
Degree Doctor of Philosophy, Arizona State University, May
2001, 803 pp.
2. D. W. Christensen, NCHRP Project 9-25: Requirements for
Voids in Mineral Aggregate for Superpave Mixtures, Interim
Report to the National Cooperative Highway Research
Program, Advanced Asphalt Technologies, LLC, September
2001.
3. D. W. Christensen, NCHRP Project 9-31: Air Void
Requirements for Superpave Mix Design, Interim Report to the


National Cooperative Highway Research Program, Advanced
Asphalt Technologies, LLC, April 2002.
4. R. Nichols, Composite Construction Materials Handbook,
Englewood Cliffs, N. J.: Prentice-Hall, Inc., 1976, pp. 259-262.
5. T. J. Hirsch, Proceedings of the American Concrete Institute,
Vol. 59, 1962, p. 427.
6. N. Shashidhar, S. P. Needham, B. H. Chollar, and P. Romero,
Prediction of the Performance of Mineral Fillers in SMA,
Journal of the Association of Asphalt Paving Technologists,
Vol. 68, 1999, p. 222.
7. S. C. Chapra and R. P. Canale, Numerical Methods for
Engineers, New York: McGraw-Hill, Inc., 1988, 812 pp.
8. D. W. Christensen, Analysis of Creep Data from Indirect
Tension Test on Asphalt Concrete, Journal of the Association
of Asphalt Paving Technologists, Vol. 67, 1998, pp. 458-489.
9. A. A. Tayebali, J. A. Deacon, and C. L. Monismith,
Development and Evaluation of Surrogate Fatigue Models for
SHRP A-003A Abridged Mix Design Procedure, Journal of
the Association of Asphalt Paving Technologists, Vol. 64,
1995, pp. 340-364.
10. S. H. Alavi and C. L. Monismith, Time and Temperature
Dependent Properties of Asphalt Concrete Mixes Tested as
Hollow Cylinders and Subjected to Dynamic Axial and Shear
Loads, Journal of the Association of Asphalt Paving
Technologists, Vol. 63, 1994, pp. 152-175.
11. D. W. Christensen and D. A. Anderson, Interpretation of
Dynamic Mechanical Test Data for Paving Grade Asphalt,
Journal of the Association of Asphalt Paving Technologists,
Vol. 61, 1992, pp. 67-98
12. D. Andrei, Witczak, M.W., and Mirza, W., Development of a
Revised Predictive Model for the Dynamic (Complex) Modulus
of Asphalt Mixtures, NCHRP 1-37A Inter Team Technical
Report, University of Maryland, March 1999.
13. M. W. Witczak, M. Bari, and M. M. Quayum, Sensitivity of
Simple Performance Test Dynamic Modulus |E*|, NCHRP 9-19
Subtask C4b Report, Tempe, AZ: Arizona State University,
Department of Civil and Environmental Engineering,
December 2001.
14. R. L. Lytton, J. Uzan, E. G. Fernando, R. Roque, D. Hiltunen,
and S. M. Stoffels, Development and Validation of


Performance Prediction Models and Specifications for Asphalt
Binders and Paving Mixes, Report SHRP-A-357, Washington,
D.C.: Strategic Highway Research Program, 1993.
15. F. P. Bonnaure, A. H. J. J. Huibers, and A. Boonders, A
Laboratory Investigation of the Influence of Rest Periods on
the Fatigue Characteristics of Bituminous Mixes,
Proceedings, the Association of Asphalt Paving Technologists,
Vol. 51, 1980, p. 104.
16. F. Bonnaure, G. Gest, G. Gravois, and P. Uge, A New Method
of Predicting the Stiffness of Asphalt Paving Mixtures,
Proceedings, the Association of Asphalt Paving Technologists,
Vol. 46, 1977, p. 64.

También podría gustarte