You are on page 1of 12

International Journal of Pavement Engineering

ISSN: 1029-8436 (Print) 1477-268X (Online) Journal homepage: http://www.tandfonline.com/loi/gpav20

Evaluation of rutting potential for asphalt concrete


mixes containing copper slag

Hossam F. H. Abdelfattah, Khalid Al-Shamsi & Khalifa Al-Jabri

To cite this article: Hossam F. H. Abdelfattah, Khalid Al-Shamsi & Khalifa Al-Jabri (2016):
Evaluation of rutting potential for asphalt concrete mixes containing copper slag, International
Journal of Pavement Engineering, DOI: 10.1080/10298436.2016.1199875

To link to this article: http://dx.doi.org/10.1080/10298436.2016.1199875

Published online: 20 Jun 2016.

Submit your article to this journal

Article views: 13

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


http://www.tandfonline.com/action/journalInformation?journalCode=gpav20

Download by: [McMaster University] Date: 23 June 2016, At: 22:58


International Journal of Pavement Engineering, 2016
http://dx.doi.org/10.1080/10298436.2016.1199875

Evaluation of rutting potential for asphalt concrete mixes containing copper slag
Hossam F. H. Abdelfattah, Khalid Al-Shamsi and Khalifa Al-Jabri
Department of Civil and Architectural Engineering, Sultan Qaboos University, Muscat, Oman

ABSTRACT ARTICLE HISTORY


Copper slag (CS) is a by-product of the copper extraction process, which can be used as coarse and/or fine Received 14 January 2016
aggregate in hot mix asphalt (HMA) pavements. This study used CS as a replacement of the fine aggregate Accepted 4 May 2016
with a percentage of up to 40% by total aggregate weight. The objective of this study was to evaluate the KEYWORDS
effect of CS on the rutting potential of the asphalt concrete mix using two methods. One method is based Asphalt concrete; copper
on the Dynamic modulus |E*| testing result. Actual pavement temperature data from a test section were slag; rut depth; dynamic
used with the developed |E*| master curves. EverStressFE finite element program was used to perform a modulus; M-E PDG; flow
linear elastic load-deformation analysis for a pavement section and to determine the vertical resilient strain number
in a 40-mm HMA surface layer. The M-E PDG permanent deformation model was used with and Excel Visual
Downloaded by [McMaster University] at 22:58 23 June 2016

Basic for Applications code to predict the accumulated rutting for different CS mixes for 10 million ESALs.
The other method used the data from the flow number (FN) test. Based on the |E*| approach, the results
indicated that adding 5% CS in the mix increased the predicted rutting from 0.59 to 0.98 mm at 10 million
ESALs (increase by 68%). When 40% CS was used, rutting increased by more than 700% compared with the
control mix. After analysing the FN results with the Francken model, the results indicated a decrease in FN
as CS content is increased, indicating higher rutting potential. The decrease in FN ranged from 9% for 5%
CS to 95% for 40% CS. The mixes containing up to 10% CS satisfied the minimum FN criteria for rutting. A
calibration process for the M-E PDG distress prediction models that allows the use of waste and by-product
materials such as CS should be considered in the future.

1. Introduction reverberatory CS as conventional coarse and fine aggregate for HMA


Waste and by-product material generation is increasing world- concrete (Gorai et al. 2003). Shahu et al. (2013) investigated the
wide. Due to the disposal cost of these materials, there has been use of CS, fly ash and dolime mix as a base course for flexible pave-
an increased demand for recycling some of these materials in ments. The study evaluated the mixes using unconfined compressive
applications such as highway pavements (FHWA 2012). However, strength, triaxial and durability tests. The study recommended a
the use of these materials in pavements should not compromise mix of 20% fly ash and 80% CS to be stabilised with 15% dolime
the expected performance of the pavement. as the optimum mix in terms of unconfined compressive strength.
Slags are molten by-products of high-temperature ­processes, Huang et al. (2007) reviewed different waste materials for
and are primarily used to separate the metal and non-metal con- use in asphalt pavements. Steel slag was recommended for
stituents contained in the bulk ore. Copper slag (CS) c­ ontains replacement of coarse aggregate, with one drawback. The high
materials such as iron, aluminium, calcium oxide, silica and oth- specific gravity of the slag could lead to mix density, implying an
ers. Slag is discharged from the furnace at more than 1000 °C. increase in transport cost. Ahmedzade and Sengoz (2009) used
When it is cooled slowly, it forms a dense, hard crystalline prod- steel slag as a coarse aggregate in HMA with AC-5 and AC-10
uct, while quick solidification, by pouring the molten slag into binders. Mechanical characteristics of all mixtures were evalu-
water, forms a granulated amorphous slag. Approximately 2.2 ated by Marshall stability, indirect tensile stiffness modulus, creep
to 3 t are generated for every ton of production. As an asphalt stiffness and indirect tensile strength tests. The results indicated
concrete material, there has been limited use of CS as fine or improvement of the mechanical properties of asphalt mixtures.
coarse aggregate in the United States (RMRC 2012). In Oman, The asphalt mixture performance tester performs three tests
approximately 60,000 t of granulated CS is produced every year. for evaluation of rut resistance of the mix: the dynamic mod-
Researchers have reported the use of CS in hot mix asphalt ulus (|E*|), the flow number (FN) and the flow time tests. The
(HMA) concrete. Fine CS has been used in HMA pavements in dynamic modulus |E*| of HMA is one of the fundamental
California and granulated CS has been incorporated into asphalt inputs in the Mechanistic Empirical Pavement Design Guide
mixes in Georgia to improve stability. Although it is rarely used, (M-E PDG), developed under NCHRP Project 1-37A (NCHRP
Michigan Department of Transportation Specifications consider 2004) and published by AASHTO (AASHTO 2015). The M-E

CONTACT  Hossam F. H. Abdelfattah  hossam@squ.edu.om


© 2016 Informa UK Limited, trading as Taylor & Francis Group
2    H. F. H. Abdelfattah et al.

PDG (AASHTO 2015) uses a laboratory-derived relationship results from both approaches were used to evaluate the rutting
to estimate the rate or accumulation of plastic deformation potential for the CS mixes.
using a repeated load permanent deformation triaxial test. The
relationship is adjusted to match rut depth measured on the
3.  Material properties and mix design
roadway through field calibration. |E*| expressed as complex
modulus master curve is used to estimate the vertical resil- A 12.5-mm nominal maximum aggregate size (NMAS) blend
ient strain in the HMA layer. M-E PDG rut model estimates gradation was composed of four crushed limestone aggregate
the plastic axial strain based on the elastic strain (AASHTO 2015). sizes: 20, 10, 3–5, 0–3 mm, in addition to mineral filler. A pen-
FN test is a variation of the repeated-load permanent defor- etration grade 60/70 binder was used. Aggregate bulk specific
mation test used by researchers in 1970s. It uses a haversine load gravity was 2.812, 2.805, 2.795, 2.790 and 3.388 for the 20, 10,
application of 600 kPa peak load for 0.1 s followed by 0.9 s rest 3–5, 0–3 mm and CS, respectively. The filler apparent specific
period. The permanent axial deformation is measured at the end gravity was 2.750. The chemical composition of aggregates and
of the rest period and converted to strain by dividing it by the CS is shown in Table 1. The main compounds in all aggregates are
original gauge length. The FN is defined as the load repetition Calcium and Magnesium Oxides in addition to Silicone Dioxide
value when shear deformation in the asphalt layer, under con- in the mineral filler, while the main compounds in CS were Ferric
stant volume, starts. It is equivalent to the start of the tertiary flow and Silicone Oxides.
with the mix. This stage is preceded by primary and secondary CS was used as a replacement of the 0–3-mm size aggre-
stages of the permanent deformation versus number of loading gate with a percentage of 5, 10, 15, 20, 30 and 40%, by total
cycles (Rodenzo 2010, Bonaquist 2013, FHWA 2013). aggregate weight. The blend gradations for all six mixes in
Downloaded by [McMaster University] at 22:58 23 June 2016

The FN has been correlated with rutting resistance of mixtures addition to the control (0% CS) are shown in Figure 1 together
from various full-scale pavement tests. Recently, a tentative crite- with the upper and lower limits of the gradation band for
rion for FN has been developed under NCHRP Project 9-33 for an Omani specification for dense wearing course class B (DGRLT
average rut depth of 7 mm, corresponding to 95% reliability that 2010). Few sieves fell marginally outside the lower limits for
the rut depth will be less than 12 mm. The recommended mini- the 30 and 40% CS mixes. The mixes were designed accord-
mum FN requirement is presented as a function of the expected ing to Marshall Mix design procedure (AASHTO T245).
ESALs, as part of the evaluation of the performance of asphalt Table 2 presents the mix design results and the standard lim-
concrete mixes (NCHRP 2011, Bonaquist 2013, FHWA 2013, ). its for Omani specification for dense wearing course class
B (DGRLT 2010), as well as the Asphalt Institute (AI 1995)
specifications for heavy traffic surface or base mix.
2.  Objectives and scope
For all mixes, the optimum binder content was selected to
The objective of this study was to evaluate the rutting resist- satisfy the maximum stability, and flow and air voids spec-
ance of asphalt mixes containing CS as a fine aggregate with ification ranges. In case the air voids were higher than the
up to 40%, by total aggregate weight. The evaluation is based limits for different binder contents for a certain mix, then the
on the results of two tests: the dynamic modulus and FN. The binder content producing the lowest air voids was selected.
|E*| approach uses the developed master curves together with The binder content ranged from 3.7 to 5.5%, by total mix
a linear elastic load-deformation analysis of a pavement section weight. Some criteria were not satisfied according to Oman
to determine the vertical elastic strain in the asphalt concrete or AI specifications. Stability, for example, was not satisfied
layer. Actual pavement temperature records were used. An Excel for some mixes according to Oman specification for wearing
Visual Basic for Applications (VBA) code was written to per- course. One the other hand, it was satisfied according to AI
form the analysis using the M-E PDG performant deformation criteria. AV% were slightly higher than AI limits for all mixes
model. In the other approach, daily air temperatures for Muscat except 30% CS. VMA was also lower than the AI minimum
International Airport were analysed for 20  years, and the FN value. However, it should be realised that the rutting predic-
tests were conducted at the resulting high Performance Grade tive capabilities of the Marshall criteria for mix performance
(PG) temperature according to AASHTO TP79. The analysed is questionable (FHWA 2013). Therefore, the objective of the

Table 1. Chemical composition of aggregate and CS.

Aggregate
Chemical compound 20 mm 10 mm 3–5 mm 0–3 mm Mineral filler CS
LOIa 44.19 43.41 44.58 44.12 23.64 –
CuO – – – – – 0.34
SiO2 4.06 6.47 2.51 3.31 27.79 33.93
Fe2O3 0.76 0.61 0.53 0.67 5.19 53.4
Al2O3 0.41 0.62 0.34 0.49 3.14 1.61
CaO 22.92 23.39 28.94 29.44 20.55 2.32
MgO 19.7 19.0 14.62 14.52 7.56 1.54
Na2O 6.39 5.89 7.18 6.87 5.81 3.29
K2O 0.03 0.04 0.02 0.02 0.06 0.47
Analysis (%) 98.46 99.43 98.72 99.44 93.65 96.56
a
Loss on ignition.
International Journal of Pavement Engineering   3

100
90
80

Percentage passing
70
60
50
40
30
20
10
0
0.0 0.1 1.0 10.0 100.0
Sieve size, mm

Control 5% 10% 15% 20%

30% 40% UL LL

Figure 1. Aggregate blend gradation.

Table 2. Optimum asphalt content for different mixes and specifications.


Downloaded by [McMaster University] at 22:58 23 June 2016

Mix (CS%) Binder (%) Stability (kN) Flow (mm) BSGa AVb (%) VMAc (%) VFAd (%)
0% (control) 3.7 22.5 3.6 2.506 4.0 13.4 62
5% 3.8 16.9 3.2 2.512 5.5 14.3 63.2
10% 5.0 13.3 3.2 2.550 5.5 14.9 64.0
15% 5.5 10.6 3.6 2.538 5.6 16.4 65.5
20% 5.5 10.4 3.4 2.543 5.8 16.4 66.0
30% 5.5 8.5 2.7 2.567 6.8 17.8 62
40% 5.4 8.5 2.8 2.650 5.5 16.7 67.4
Spec.e 3.5 – 5.5 14, min 2 – 4 – 3.5 – 5.5 15, min 63 – 75
Spec.f – 8, min 2 – 3.5 – 3–5 Variableg 65 – 75
a
BSG = bulk specific gravity;
b
AV = air voids %;
c
VMA = voids in mineral aggregate %;
d
VFA = voids filled with bitumen %;
e
Oman wearing course class B;
f
Asphalt Institute wearing and base course, for heavy traffic;
g
(from 13 to 15% for AV% from 3 to 5%).

paper was to investigate and quantify the expected mix per- extracted from the middle of the compacted samples using a
formance in terms of rutting. coring machine (Figure 2(a) and (b)). The ends of samples were
then trimmed using a concrete saw. The end smoothness and per-
pendicularity were verified in accordance with AASHTO TP62.
4.  Sample preparation Two compaction trials were needed to reach the required air
voids (AV%) for the cored samples. The resulting air voids of the
Samples were prepared to a target height of 160 mm according
101.6 mm (4 in) test samples were in the range of 7.1 to 7.5%.
to AASHTO T312 using the Superpave Gyratory Compactor.
Samples were placed inside an in-house fabricated environmen-
Samples were prepared at the optimum binder contents shown in
tal chamber to reach the testing temperature. A dummy sample
Table 2 for different mixes. Mixing and compaction temperatures
with a temperature probe placed vertically in the middle of the
were determined from the mid-value of the mixing and com-
sample was monitored with a data acquisition system. The test-
paction viscosity range for the binder, and were found to be 157
ing was initiated when the dummy sample reached the testing
and 144 °C, respectively. Samples were conditioned according to
temperature.
AASHTO R30. After mixing, the loose mix was placed for 4 h of
conditioning at 135 °C. Afterwards, it was placed for 30 min at
the compaction temperature, then compacted. 5.  Dynamic modulus
To prepare a sample for testing, two small vertical cuts were
5.1. Testing
induced on two opposite sides on the outer perimeter of a
compacted sample. This allowed the sample to be inserted in a The dynamic modulus testing was conducted according to
three-sided wooden mould fabricated to hold the sample in place AASHTO TP62 using an MTS 810 testing machine. Two sam-
during coring (Figure 2(a)). The mould is composed of three ples were tested at 25, 30, 40, 50 and 60 °C. At each temperature,
parts joined together with four bolts. One part has a 150-mm the sample was tested at 0.1, 0.5, 1, 5, 10 and 16 Hz frequency. A
inner diameter. The other two opposite parts are straight with fixture different from the axial deformation system proposed in
adjustable distance between them to allow for small variations AASHTO TP62 was manufactured and utilised (Figure 3). Two
in the thickness of the vertical cuts. The 101.6-mm samples were 2.0-mm range extensometers, with a maximum error of 0.29%
4    H. F. H. Abdelfattah et al.

Figure 2. Test sample preparation: (a) coring the compacted sample inside the wooden mould; (b) cored sample and extra material.
Downloaded by [McMaster University] at 22:58 23 June 2016

0 -1.142

Axial deformation (mm)


-1.144
-200
-1.146
Load (N) -400 -1.148
-1.15
-600
-1.152
-800 -1.154
-1.156
-1000
-1.158
-1200 -1.16
19 19.2 19.4 19.6 19.8 20 20.2
Time (s)
Load Axial Deformation

Figure 4. Load deformation for 20% CS mix at 60°C and 10 Hz frequency.

0.1 Hz to maintain a uniform number of sampling points. |E*|


master curves were developed from the average results for each
Figure 3. Axial deformation fixture and extensometers mounted on the sample. mix. Recorded pavement temperature data at 20-mm depth from
an instrumented test section were analysed. Reduced frequencies
were estimated for the mid-value for different groups of pave-
of the displacement reading, were used for vertical displacement ment temperature range. A linear elastic finite element (FE) anal-
measurement. The extensometers were mounted between two ysis was performed on a pavement section to estimate the vertical
aluminium rings with 130-mm inner diameter and 20-mm thick- elastic strain in the HMA layer. The M-E PDG permanent defor-
ness. On each ring, two fixed strips, with 10-mm height, extend mation equation (AASHTO 2015) was used to predict the rut
from the rings to the surface of the sample. A third strip moves depth in the HMA layer. It is to be noted that this equation was
with a tightening screw to be in contact with the sample surface. developed and calibrated for conventional mixes. However, in the
The three strips are equidistant on the ring perimeter and are absence of calibration for waste or by-product material such as
manufactured with an inner surface diameter of 101.6 mm (4 in). CS, the equation was used for comparative analysis between the
Three removable guiding rods connect the upper and lower rings mixes containing different percentages of CS. The next sections
together at a gauge length of 101.6 mm (from the centre of the describe the process in more details.
strips), and provide alignment of the extensometer fixtures
parallel to the sample vertical axis. Once the rings are fixed on 5.2.1.  Dynamic modulus master curves
the sample, the guiding rods are removed and the extensometers The average |E*| testing results were fitted to the M-E PDG mas-
are attached to the fixture on opposite sides of the sample. ter curve equations (FHWA 2013):

log(Max) − log(Min)
5.2.  Results, analysis and discussion log|E ∗ | = log (Min) + (1)
1 + e𝛽+𝛾 log (𝜔r )
Figure 4 shows the testing result for load and average axial defor- ( )
mation for 20% CS mix at 60  °C and 10  Hz frequency. Data ΔEa 1 1
log(𝜔r ) = log (𝜔) + −
collection interval varied from 0.00048 s for 16 Hz to 0.1 s for 19.14714 T Tr (2)
International Journal of Pavement Engineering   5

Table 3. |E*| Master curve fitting parameters for different CS mixes.


CS (%) 0% 5% 10% 15% 20% 30% 40%
Max, ksi 3,390.95 3,348.80 3,321.24 3,251.61 3,254.07 3,163.08 3,245.76
Min, ksi 42.27 23.17 20.45 20.360 33.47 28.57 16.80
β –1.3298 –1.1996 –0.9799 –0.9589 –0.6579 –0.6204 –0.4714
γ –0.9438 –0.8808 –0.811 –0.7912 –0.8834 –0.8322 –0.7782
ΔEa 167,623.5 166,313.7 147,659.6 160,103.5 158,263.2 154,835.4 110,173.2
Adj. R2 0.950 0.961 0.957 0.951 0.962 0.955 0.955
Se/Sy 0.23 0.20 0.21 0.22 0.20 0.22 0.21

25,000

20,000
Fit |E*| (MPa)

15,000

10,000

5,000
Downloaded by [McMaster University] at 22:58 23 June 2016

0
1.E-06 1.E-04 1.E-02 1.E+00 1.E+0 2 1.E+0 4 1.E+06
Reduced frequency (Hz.)
0% 5% 10% 15% 20% 30% 40%

Figure 5. Dynamic modulus |E*| master curves at 25 °C reference temperature.

Table 4. Yearly average for the daily air temperature at Muscat Int’l Airport (20 years analysis period).

Yearly Average, °C
Statistic computed over 20-year period Daily high Daily mean Daily low
Max 34.1 29.1 25.0
Min 32.1 27.5 23.2
Avg 33.1 28.3 23.9
St. Dev 0.6 0.4 0.5
CV% 1.7 1.6 2.1

where, |E*| = dynamic modulus; ωr = reduced frequency in Hz; of approximately 2 years (75, 267 data points). However, the anal-
Max = limiting maximum modulus in ksi; Min = limiting minimum ysis period for rutting accumulation was 20 years. To indicate the
modulus in ksi; and β and γ = fitting parameters; ω = loading fre- variability in temperature throughout a 20-year period of time,
quency at test temperature; T and Tr are test and reference tempera- the daily air temperature from Muscat International Airport was
tures in °K, respectively; ΔEa = activation energy (fitting parameter). analysed. The results of the yearly average for the daily high,
The maximum limiting |E*| was estimated from the mix mean and low air temperatures estimated for a period of 20 years
properties (VMA and VFA) using the Hirsh model (FHWA are shown in Table 4. The results indicate a small variation in
2013). Sigmaplot statistical software (SigmaPlot 2006) was used the yearly averages for either of the high, mean, or low daily
to perform the non-linear regression on |E*| results to produce air temperatures over the 20-year period, as indicated by the
the master curves at a reference temperature of 25 °C. Table 3 small coefficient of variation (CV%). This analysis justified the
shows the model fitting parameters for different mixes. The min- use of the available pavement data by repeating it every 75,267
imum value for adjusted R2 was 0.95, and that for the standard 15-­minute periods until the end of the analysis period.
error ratio (Se/Sy), i.e. ratio of standard error of estimate to the The recorded data from the pavement monitoring station
standard deviation of logarithm of |E*| values, was 0.23. These (for approximately 2  years) indicated that the maximum and
statistics are considered excellent goodness of fit (Rodenzo 2010). minimum recorded HMA pavement temperature data at 20-mm
Figure 5 shows the resulting master curves for different mixes. depth were 61.56 and 15.33 °C, respectively. To reduce the com-
The figure shows the increase in |E*| as the reduced frequency putation effort, the pavement temperatures were grouped into
increases. It also shows the gradual decrease in modulus with 4 °C intervals starting with 12 to 16 °C interval and ending with
the increase in CS content. 60 to 64  °C interval, resulting in 13 temperature groups. The
middle value of the group was used to represent each group. For
5.2.2.  Pavement temperature and reduced frequency every mix with a different CS percentage, the mid-range value
Pavement temperature data at a depth of 20 mm was available for for each of the 13 temperature groups was used to compute the
15-min intervals from a pavement monitoring station for a period reduced frequency (ωr) using Equation (2), knowing the load
6    H. F. H. Abdelfattah et al.

frequency (explained in next section). The reduced frequency


was then used to compute the modulus values from the master
curve for this particular mix for all temperature groups.

5.2.3.  FE modelling and analysis


An actual pavement cross section built on a road in Oman
consisting of 40-mm asphalt surface, 120 bituminous base,
350 mm of subbase, on top of subgrade was used for the ­analysis.
EverStressFE finite element (FE) software (2009), developed
for the Washington State DOT, was used to perform a load-­
deformation elastic analysis and estimate the vertical elastic
strain (εr) in the HMA surface layer at a depth of 20 mm. The
modulus of the 40-mm surface layer was kept variable as per |E*|
values. For the other layers, the material properties were assigned
to each layer as per the laboratory testing data from the road
project, and were kept constant in the analysis. The average CBR
for subgrade and subbase was 30 and 70%, respectively. Using
the AASHTO correlation charts (AASHTO 1993), these values
were converted to modulus values of 103, 131 MPa, respectively.
Downloaded by [McMaster University] at 22:58 23 June 2016

The Marshall stability for the bituminous base was 7.1 kN, and
was converted to a modulus value of 2413 MPa.
The load was taken to be an 80 kN single axle with dual wheels,
with a wheel load of 20 kN, and a tire pressure of 827.4  kPa.
The tire spacing was taken to be 318 mm centre to centre with Figure 6. Finite element model for the pavement section.
a tire width of 280 mm. The wheel spacing, tire pressure, and
tire width were obtained from Part 2 of NCHRP 1-37A report
(NCHRP 2004). modulus of the HMA surface layer to the values ­corresponding
Al-Qadi et al. 2008 recommended, based on a study of full- to the mid temperature values described in the previous section.
size FE section (3 × 3 × 5 m), infinite boundary to be used at The runs were made for the 0% CS mix, and the vertical elastic
least 900  mm from the center of the load and to a depth of strain (εzz) under the wheel load at 20-mm depth in the HMA
1100 mm to reduce the compressive stress in the subgrade to layer was compared for both.
less than 1%. They also cited the suggested infinite boundary At 14 °C, the HMA layer modulus was 22,376 MPa, resulting
by Applied Research Associates Inc. (ARA 2004) to be between in low strain and a difference of 0.9 μstrain. On the other hand, at
760 to 1200 mm, which is considered close enough to the load 62 °C, the HMA layer modulus was 4320 MPa, resulting in higher
to maximise the benefit from using them without altering the strains with a difference of 3.3 μstrain between the two programs.
solution results. The coefficient of variation (COV) ranged from 12% down to 4%,
Figure 6 shows the FE model with one half of one wheel from the low strain end to the high strain end, respectively. These
(shaded area). The section width was selected to represent one- values agree with the reported variability among using different
half of a 1.8-m axle, and as recommended by Al-Qadi et al. (2008) pavement analysis programs (ARA 2004), which reported 12%
and ARA (2004), the boundary conditions were modelled as COV for subgrade compressive strains and 15% for AC tensile
infinite using infinite elements on the two edges of the roadway strains.
away from the two wheel loads comprising half the axle, i.e. at For the analysis, a speed of travel was assumed to be 96.6 km/h.
Y = 0 and X = 900 mm, as well as the bottom of the subgrade. The travel speed was converted to time of load at a depth of
The boundary at X = 0 and Y = 900 mm are lines of symmetry 20 mm into the pavement using the formula used by NCHRP
for the model: halfway through the loading axle, and halfway 1-37A (Al-Qadi and Elseifi 2008):
between the two tires on one side of the axle, respectively. The
Leff
total section depth was taken as 1010 mm with infinite elements t= (3)
at the bottom of the subgrade. The model uses the symmetry to 17.6 V
reduce the computation time. In the X–Y plane, the model uses Where, Leff is the effective length over which the load is spread
6 elements in X and Y directions in the coarse mesh area (load- (in), and V is the vehicle speed (mph). Leff was computed to be
ing area) and 12 elements in the fine mesh area. In the vertical 126.3 mm (4.97 in) based on a tire length of 86.3 mm (3.4 in),
direction, six elements were used for the top two layers, and eight as computed from a rectangular tire imprint area and tire width.
for the subbase and subgrade. The interface condition between The frequency (ω) was estimated as the inverse of time of loading
layers was assumed as fully bonded. as used in the M-E PDG (Al-Qadi and Elseifi 2008).
For verification of the model results, the model was compared A total of 13 runs of the program, based on the temperature
with results from the well-known BISAR multilayer linear-elastic groups, were performed for each of the 7 mixes. The elastic strain
program. A single circular load was used the verification purpose at 20-mm depth in the HMA surface layer was obtained from the
in both programs, as it is the only tire imprint that BISAR can FE analysis under the centre of one of the wheel loads on either
handle. Thirteen runs were made for each program by varying the side the 80 kN axle, as shown in Figure 7. The results indicate
International Journal of Pavement Engineering   7

400

Vertical Elastic Strain in HMA layer


350
300
250

(ε r, microstrain)
200
150
100
50
0
10 14 18 22 26 30 34 38 42 46 50 54 58 62 66
HMA layer temperature (oC)
0% 5% 10% 15% 20% 30% 40%

Figure 7. Vertical elastic strain for the HMA layer at the mid-group temperature value for different mixes.

an increase in vertical resilient strain in the HMA surface layer (national) calibration. These parameters allow for local (state)
as the temperature increases. At 62 °C, the strain increased by calibration for different local conditions.
Downloaded by [McMaster University] at 22:58 23 June 2016

68, 151, 158%, as CS content increased to 5, 10, and 15%, as The program iterates one iteration for every 15-min temper-
compared with the control mix, respectively. With the increase ature and 14.27 ESALs, and continues for a total of 700,800 iter-
in CS to 20, 30 and 40%, the strain increased by 261, 311 and ations (20 years). The concept of strain hardening implemented
624%, compared with the control mix, respectively. A noticeable in M-E PDG (AASHTO 2015) was applied to estimate the equiv-
increase occurs between 15 and 20% slag. The strain at 40% CS alent number of repetitions each time the temperature shifted
is approximately 1.8 times that at 30%. from one of the temperature groups to another in successive iter-
ations. This was done because the model shown in Equations (4)
5.2.4.  Rut depth results to (7) estimates the accumulated permanent strain for a number
A total repetition of 10 million ESALs was assumed to be uni- of repetitions at a given temperature.
formly distributed over a 20-year analysis period resulting in After each iteration, the rut depth in the HMA surface layer
14.27 ESALs for every 15 min. An Excel VBA code was written was calculated as the accumulated permanent strain multiplied
to estimate the accumulated plastic strain at 20-mm depth in the by the layer thickness of 40 mm. The resulting rut depth for the
HMA surface layer based on the 15-min pavement temperature HMA surface layer for different CS mixes is shown in Figure 8.
data. The code would allocate each pavement temperature in one Table 5 shows the rut depth values at 3 and 10 million ESALs
of the thirteen 4 °C groups, and calculate the accumulated plastic for different CS mixes. The percentage increase in rut depth, as
strain for this temperature value from the field-calibrated model compared with the control mix, is also shown. It is to be noted
in M-E PDG, Equations (4) to (7) (AASHTO 2015): that, based on the permanent deformation model, this percent-
age is the same at both ESALs. Rut depth increased, compared
𝜀p(HMA) with the control mix, by 68% for the 5% CS mix and by more
= 𝛽1r kz ∗ 10−3.35412 ∗ n0.4791𝛽2r ∗ T 1.5606𝛽3r (4) than 700% for the 40% CS mix. The results agree with the rel-
𝜀r(HMA)
ative values of resilient axial strain; a significant increase in rut
depth was noticed between 15 and 20%. Similarly to the elastic
( ) strain results, the rut depth at 40% was approximately 1.8 times
kz = C1 + C2 ∗ D ∗ 0.328196D (5)
that at 30%.

C1 = −0.1039 ∗ (HHMA )2 + 2.4868 ∗ HHMA − 17.342 (6)


6.  Flow number
(7)
2
C2 = 0.0172 ∗ (HHMA ) − 1.7331 ∗ HHMA + 27.428 6.1. Testing
Where εp(HMA)  =  accumulated permanent plastic strain in the The FN test was performed on two samples according to
HMA layer, in.; εr(HMA) = resilient elastic strain calculated from AASHTO TP79 using the MTS 810 system. The load was applied
the linear elastic load-deformation analysis in this study, in.; for 0.1 s every 1.0 s. A seating load of 235.6 N producing 30 kPa
n = number of axle-load repetitions; T = pavement temperature, was applied and the repeated axial load was applied producing
°F; kz = depth confinement factor; D = depth below the surface, 4712.4  N load (600  kPa axial stress). The test was conducted
in.; and HHMA = total HMA thickness, in; β1r, β2r and β3r are local at the high adjusted PG temperature, which is based on rut-
calibration parameters, set to 1.0 for global calibration. ting damage. The transfer function developed by Mohseni and
The M-E PDG rut model computes the accumulated plastic Carpenter (2005) for 50% reliability was used to determine the
strain in the HMA layer based on the resilient strain, temper- high adjusted PG temperature (LTPPBind 2005):
ature and number of load repetitions at that temperature. The
calibration parameters β1r, β2r and β3r were set to 1.0, as for global PGd = 48.2 + 14DD − 0.96DD2 −2RD (8)
8    H. F. H. Abdelfattah et al.

5
4.5
4
3.5

Rut depth (mm)


3
2.5
2
1.5
1
0.5
0
0 2,000 4,000 6,000 8,000 10,000
ESALs (x 1000)
0% 5% 10% 15% 20% 30% 40%

Figure 8. Rut depth accumulation for different CS mixes.

Table 5. Rut depth development for different CS mixes.


Downloaded by [McMaster University] at 22:58 23 June 2016

Rut depth (mm)


CS % in the mix (by total aggregate weight) 3 million ESALs 10 million ESALs Increasea in %
0 0.33 0.59 0
5 0.55 0.98 68
10 0.87 1.56 166
15 0.92 1.64 180
20 1.27 2.26 286
30 1.51 2.69 358
40 2.72 4.83 724
a
compared with the control mix.

0 -5.2
-500 -5.3

Axial deformation, mm
-1000
-5.4
-1500
Axial load, N

-2000 -5.5
-2500 -5.6
-3000 -5.7
-3500
-5.8
-4000
-4500 -5.9
-5000 -6
37 38 39 40 41 42
Load time, s
Load Deformation

Figure 9. Sample axial load and deformation for 0% CS.

where, PGd = PG Damage at a Rut Depth; DD = 20-Year Degree- Figure 9 shows a sample testing result of the vertical
Days >10 °C (×1000 °C); and RD = Rut Depth (5–13 mm). c­ ompressive load and axial deformation for one of the sam-
The mean daily air temperature for a period of 20  years ples containing 0% CS. The data sampling period was 0.01 s.
was obtained from the weather station at Muscat International The zero deformation was considered as the start of the first
Airport. The data were used to compute the degree days above load pulse. The permanent strain (εp) is computed as the dif-
10 °C. The average 20-year degree days was found to be 6669 °C ference between the end deformation in the 0.9 s unload-
with a standard deviation of 161 °C. AASHTO TP79 requires a ing period and the zero deformation, divided by the sample
target rut depth of 12.5 mm, and a reliability of 50%. This resulted height. Figure 10 shows the resulting permanent axial stain
in a damage-based PG at 50% reliability (PGd) of 73.9 °C. LTPP versus the number of loading cycles for one of the 10% CS
pavement temperature algorithms (LTPPBind 2005) was used samples. The figure clearly shows the three stages of cumu-
to adjust the selected PG for a 20-mm depth into the pavement, lative permanent strain versus loading cycles, i.e. primary,
resulting in a testing temperature of 71.5 °C. secondary and tertiary zones.
International Journal of Pavement Engineering   9

0.06 1.50E-03

1.30E-03
0.05

Permanent strain rate, per cycle


1.10E-03

Permanent strain (εp)


0.04
9.00E-04

0.03 7.00E-04

5.00E-04
0.02
3.00E-04
0.01
1.00E-04
FN = 117
0 -1.00E-04
0 50 100 150 200 250
Cycle
Permanent strain Permanent strain rate

Figure 10. Permanent axial strain and strain rate versus number of cycles for 10% CS.
Downloaded by [McMaster University] at 22:58 23 June 2016

Table 6 Francken model fitting and nonlinear regression parameters.


CS %, and sample number A B C D Adj. R2 Se/Sy
0%-1 8.84E-04 0.6838 1.33E-06 0.0514 0.997 0.04
5%-1 1.56E-03 0.5235 3.89E-06 0.0228 0.993 0.09
10%-1 1.15E-03 0.5914 1.25E-05 0.0336 0.997 0.06
15%-1 5.46E-03 0.4837 2.43E-05 0.1172 0.997 0.05
20%-2 4.69E-03 0.5157 4.61E-05 0.1095 0.997 0.04
30%-2 3.63E-03 0.6975 8.81E-05 0.2390 0.999 0.02
40%-2 2.19E-03 0.7425 2.07E-05 0.2427 0.999 0.02

200
180 10 to <30million ESAL FN = 181.91e-0.086 (CS%)
160 R² = 0.886
Flow number (FN)

140
120
100
80
3 to <10 million ESAL
60
40
20
0
0 5 10 15 20 25 30 35 40
Copper slag, %

Figure 11. FN for different percentage of CS at 71.5 °C.

6.2.  Analysis and discussion of results Table 6 shows the non-linear regression parameters, the
adjusted R2, and the standard error ratio (Se/Sy) for one sam-
For computing the FN value, the Francken model was fit to the
ple from each mix, for conciseness. The minimum value for
permanent strain data versus number of cycles (AASHTO TP79):
adjusted R2 was 0.992, and the maximum value for the stand-
( ) ard error ratio was 0.09. The obtained results for all samples
𝜀p = AnB + C eDn − 1 (9)
indicated excellent goodness of fit, in addition to satisfying
where, A, B, C and D are the fitting parameters of the model the precision statement in AASHTO TP79 for the range of
which is a composite model combining a power model to char- values for a 12.5 NMAS.
acterise the primary and secondary stages, and an exponential The first derivative or permanent strain rate per cycle
model fitting the tertiary stage (Rodenzo 2010). (dεp/dn) for a 10% CS sample is also shown in Figure 10. FN
10    H. F. H. Abdelfattah et al.

was determined at the point where the first derivative is mini- Field studies in the form of test sections for CS mixes are
mum or the second derivative changes from negative to positive. recommended to provide field validation of the findings. In the
FN was found to be 117 for the sample in Figure 10. Similarly, absence of calibration parameters in the M-E PDG permanent
FN was computed for all other samples. Figure 11 shows the deformation model for waste and by-product material such as CS,
average FN for different percentages of CS. The figure shows a and despite its general agreement with experimental results of this
decreasing trend for FN with the increase in CS content. A trend study, predicted rut depth may be refined if a calibration process,
equation for the relation between the percentage of copper slag that incorporates test data of waste and by-product materials sim-
(CS%) and FN is shown in the figure with an adjusted R2 of ilar to those developed in this research, is considered in the future.
0.886. FN is an indicator of the rut resistance of the mix. Lower
FN indicates a lower resistance to rutting. With the increase in Funding
CS content, the FN decreasing trend agrees with the increase This work was supported by Sultan Qaboos University (SQU) through the
in rut depth, as predicted by the M-E PDG rut model. At 10% Internal Grant Program, project [IG/ENG/CAED/11/04].
copper slag, FN decreased by 40% compared with the control
mix. A significant drop in FN occurred from 10 to 15% CS by
another 47%, compared with the control mix. The decrease, as References
compared with the control mix, ranged from 85 to 95% for 20, AASHTO, 1993. AASHTO guide for design of pavement structures.
30 and 40% CS mixes. The minimum recommended FN criteria Washington, DC: American Association of State Highway and
Transportation Officials.
for different ESALs (AASHTO TP79) are shown as solid lines in AASHTO, 2015. Mechanistic-empirical pavement design guide, a manual
Figure 11. Based on these results, mixes containing up to 10% of practice. 2nd ed. Washington, DC: American Association of State
Downloaded by [McMaster University] at 22:58 23 June 2016

CS satisfy the minimum requirement criteria to resist rutting for Highway and Transportation Officials.
3 to 10 million ESALs. Ahmedzade, P. and Sengoz, B., 2009. Evaluation of steel slag coarse
aggregate in hot mix asphalt concrete. Journal of Hazardous Materials,
165, 300–305.
Al-Qadi, I., et al., 2008. Dynamic analysis and in situ validation of perpetual
7. Conclusions pavement response to vehicular loading. Transportation Research
The main objective of this study was to evaluate the effect of Record: Journal of the Transportation Research Board, 2087, 29–39.
Al-Qadi, I. and Elseifi, M.A., 2008. Frequency determination from
CS aggregate on the rutting potential of HMA surface layer. CS vehicular loading time pulse to predict appropriate complex modulus
was used to replace the fine aggregate with a percentage of up in MEPDG. Journal Association of Asphalt Paving Technologists, 77,
to 40%, by total aggregate weight. The evaluation of the mix 739–772.
performance was based on the results of the dynamic modulus Applied Research Associates (ARA), Inc., 2004. Guide for mechanistic-
and the FN tests. empirical design of new and rehabilitated pavement structures. In:
Final Document Appendix RR: Finite Element Procedures for Flexible
The use of CS reduced the dynamic modulus |E*| values. The Pavement Analysis. Champaign, IL: ARA, ERES Division.
linear elastic load deformation analysis indicated progressive Bonaquist, R., 2013. Evaluation of flow number (Fn) as a discriminating
increase in vertical resilient strain in the HMA surface layer. In HMA mixture property. SPR# 0092-09-01. WI: Wisconsin Highway
the analysed pavement section, the HMA surface layer vertical Reproach Program. Available from: http://wisdotresearch.wi.gov/wp-
strain increased by 68% when 5% CS was used in the mix. The content/uploads/WisDOT-WHRP-project-0092-09-01-final-report.pdf
[Accessed 9 January 2016].
strain increased by more than 600% when 40% CS was used. A Directorate General of Roads and Land Transport (DGRLT), 2010.
noticeable increase in strain occurred when CS increased from Highway design specifications. Muscat:  Ministry of Transport and
15 to 20%. The strain at 40% CS was approximately 1.8 times Communications, Sultanate of Oman.
that at 30%. EverStressFE, 2009. Computer software for 3-D finite analysis of flexible
The results of the rutting analysis predicted by the M-E PDG pavement structures version 1.0. Orono, ME: University of Maine.
Federal Highway Administration (FHWA), 2012. User guidelines for
permanent deformation model agreed with the relative values of waste and by-product materials in pavement construction. Available
resilient axial strain. Rutting increased progressively as CS con- from: https://www.fhwa.dot.gov/publications/research/infrastructure/
tent was increased. A noticeable increase in rut depth was noticed structures/97148/intro.cfm [Accessed 10 December 2015].
between 15 and 20% CS content. The predicted rut increased Federal Highway Administration (FHWA), 2013. Asphalt mixture
from 0.59 to 0.98 mm over 10 million ESALs (increase by 68%) performance tester (AMPT), TechBrief FHWA-HIF-13-005. Available
from: http://www.fhwa.dot.gov/pavement/asphalt/pubs/hif13005.pdf
for 5% CS and by more than 700% for 40% CS. Similarly, to the [Accessed 11 July 2015].
resilient strain results, the rut depth at 40% was approximately Gorai, B., Jana, R.K., and Premchand, 2003. Characteristics and utilisation
1.8 times that at 30%. of copper slag – a review. Resources, Conservation and Recycling, 39,
The average 20-year degree-days above 10  °C for Muscat 299–313.
Airport was found to be 6669  °C. For a target rut depth of Huang, Y., Bird, R.N., and Heidrich, O., 2007. A review of the use of recycled
solid waste materials in asphalt pavements. Resources, Conservation and
12.5 mm, and a reliability of 50%, this resulted in a damage-based Recycling, 52, 58–73.
high PG temperature of 73.9 °C. The FN results were fit to the LTPP Bind, 2005. LTTPBind version 3.1. MD: Pavement Systems LLC. 
Francken model with excellent goodness of fit. The FN results Available from: http://www.pavesys.net/services/
were shown to decrease as CS content was increased in the mix. Mohseni, A. and Carpenter, S., 2005. Development of SUPERPAVE
The decrease in FN ranged from 9% for 5% CS to 95% for 40% CS high-temperature performance grade (PG) based on rutting damage
(with discussion and closure). Journal Association of Asphalt Paving
content. A significant decrease in FN occurred when CS content Technologists, 74, 197–254.
was increased from 10 to 15%. The mixes containing up to 10% NCHRP, 2004. Guide for mechanistic-empirical design of new and
CS satisfied the rutting criteria for 3 to 10 million ESALs. rehabilitated pavement structures. Washington, DC: Transportation
International Journal of Pavement Engineering   11

Research Board, National Cooperative Highway Research Program, AASHTO Designation: TP62, 2004. Standard method of test for determining
NCHRP 1-37A Final Report. dynamic modulus of hot mix asphalt concrete mixtures. AASHTO
NCHRP, 2011. A manual for design of hot mix asphalt with commentary. Standards, 2004th ed. Washington, DC: AASHTO.
Washington, DC: Transportation Research Board, National Cooperative AASHTO Designation: T312, 2004. Standard method of test for preparing
Highway Research Program, NCHRP Report 673. and determining the density of hot-mix asphalt (HMA) specimens by
Recycled Materials Resource Centre (RMRC), 2012. User guidelines for means of the Superpave gyratory compactor. AASHTO Standards,
by-products and secondary use of materials in pavement construction. 2004th ed. Washington, DC: AASHTO.
Available from: http://rmrc.wisc.edu/ug-mat-nonferrous-slags/ [Accessed AASHTO Designation: T245, 2004. Standard method of test for resistance to
23 August 2015]. plastic flow of bituminous mixtures using Marshall Apparatus. AASHTO
Rodenzo, M.C., 2010. Rutting criteria for asphalt mixtures based on flow Standards, 2004th ed. Washington, DC: AASHTO.
number analysis. Thesis (PhD). Arizona State University. AASHTO Designation: R30, 2004. Standard practice for mixture
Shahu, J.T., Patel, S., and Senapati, A., 2013. Engineering properties of conditioning of hot-mix asphalt (HMA). AASHTO Standards, 2004th
copper slag–fly ash–dolime mix and its utilization in the base course ed. Washington, DC: AASHTO.
of flexible pavements. Journal of Materials in Civil Engineering, 25 (12), AASHTO Designation: TP79, 2014. Standard method of test for determining
1871–1879. dynamic modulus and flow number for asphalt mixtures using the asphalt
SigmaPlot, 2006. SigmaPlot Version 10.0 user’s guide. Chicago, IL: Systat mixture performance Tester (AMPT). AASHTO Standards, 2014th ed.
Software. Washington, DC: AASHTO.
Downloaded by [McMaster University] at 22:58 23 June 2016

You might also like