Sunteți pe pagina 1din 26

RELIABILITY-BASED PAVEMENT PERFORMANCE MODEL

BY

MURANA, ABDULFATAI ADINOYI

M.Sc./ENG/06840/2006-07

AN M.Sc SEMINAR PRESENTATION SUBMITTED TO THE DEPARTMENT OF CIVIL

ENGINEERING, FACULTY OF ENGINEERING, AHMADU BELLO UNIVERSITY, ZARIA,

NIGERIA

SUPERVISORS: ENGR. (DR.) A.T.OLOWOSULU

DR. I. ABUBAKAR

SEPTEMBER, 2010

1
ABSTRACT
In Mechanistic-empirical (M-E) pavement design, the Monte Carlo method has proven to
be an effective means of determining reliability. The Pavement Performance models proposed by
Craus et al., Finn et al., FHWA-ARE and that for rutting were tested. Craus and Finn Pavement
Performance models result in best fit for the damage reliability relationship in terms of reliability
values as a result of increased axle load application. Craus and Finn Pavement Performance
models are not that sensitive to increase in axle load application compared to FHWA-ARE model
for all levels of number of Monte Carlo simulation studied. Craus and Finn Pavement
Performance models show non-conservative estimation of the fatigue life in pavement and the
modulus term for the asphalt layer included in their equation is to capture the relationship between
stiffness and fatigue cracking. The Finn Pavement Performance model shows the most promise in
terms of developing a quick predictive model for pavement reliability relationship between
stiffness and cracking. FHWA-ARE Pavement Performance variability is most affected by the
inputs closer to the pavement surface i.e. asphalt modulus. Axle weight has an overwhelming
effect on the output variability in terms of fatigue and rutting therefore deserve careful
characterization. The results from the parametric study demonstrated that the minimum number of
Monte Carlo simulation cycles that should be used for most practical design scenarios to provide
enough sufficient repeatability for damage reliability relationship is 2, 000 cycles.

INTRODUCTION
1.1 Preamble

The empirical-mechanistic (E-M) based method of pavement design is based on the mechanics of
materials, which relates input such as a wheel loads to output such as pavement response. The
response is then used to predict pavement distress (including cracking) and performance based on
laboratory experiments and field testing (Huang 1993). In the E-M based methods of pavement
design, a number of failure criteria related to specific distress and pavement performance must be
established based on theory and field observations (Aliand and Tayabji 1998). As one of the
important factors of E-M based pavement design, fatigue cracking has been reported as the most
prevalent form of structure distress of flexible (i.e., asphalt concrete) Pavements (Finn 1973).
Among the many factors causing flexible pavement cracking, traffic loading, subgrade
characteristics, and the environmental factors are the primary elements. Many flexible pavement

2
design methods consider traffic load induced fatigue cracking as a major design criterion (Huang
1993). In view of the fact that predicted pavement distress varies a great deal at the end of the
designed servicing period, it is more reasonable to introduce probabilistic approaches to pavement
design and management, since there is significant variability in predicting traffic loading,
environmental conditions, and construction quality (Lu sun et al., 2003).
Lemer and Moavenzadeh (1971), and Darter and Hudson (1973) were among the first to
introduce the reliability concept to pavement design and management. Reliability concepts were
also incorporated in the Texas flexible pavement design systems (Irick et al. 1987; Uzan et al.
1990) and in the AASHTO Design Guide (AASHTO 1993) (Lu sun et al., 2003).
Flexible pavement fatigue cracking is usually controlled by the maximum tensile stress at
the bottom of the asphalt layer. A number of predictive models of fatigue cracking have been
developed over the past three decades to characterize traffic load induced fatigue cracking. To
predict fatigue cracking of flexible pavement, damage needs to be cumulated according to certain
rules. The most popular rule of these rules is the Miner’s law. Cumulated damage is interpreted as
degree of fatigue deterioration of flexible pavement due to traffic loading (Lu sun et al., 2003).
It is well known that rutting increases at an increasing rate during the initial years of
operation, and then stabilizes with time (Monismith and Tayebali 1988; Roberts et al. 1996).
Research has shown that the pavement response under traffic loading consists of both recovery
and irrecovery components (Perl et al. 1983; Sousa and Weissman 1994; Uzan 1996). The
recovery part is elastic in nature, while the permanent part is related to plasticity. In general,
elastic strain remains constant, while plastic strain decreases with increasing load application
numbers (Jian-Shiuh et al., 2004).
The role of reliability in pavement design is to quantify the probability that a pavement
structure will perform, as intended, for the duration of its design life. Many of the parameters
associated with pavement design and construction exhibit natural variability (Timm et al, 1999).
Determining reliability requires quantifying the variability of the input values (such as
layer thickness and modulus) and then using those values to estimate the variability of the output
(expected pavement life) (Bruce, 2001). Development of analytical (mechanistic) methods for
resilient response model of pavements may be traced back to Burmister, (1943), who presented
a method to determine stresses, strains and displacements in a two layer elastic system based on
axi-symmetric analysis of vertical load using stress function (Huang, 1993).

3
1.2 Statement of Problem
A mechanistic-empirical (M-E) pavement analysis and design has been proposed for use
in Nigeria (Olowosulu, 2005). A brief summary of the process is described below: Part 1
consists of the development of input values, which include traffic, climate and material. Part 2
of the design process is structural/performance analysis. It also shows the step-by-step
procedure of M-E, starting with an assumed initial layer thickness through selection of the
optimum layer thickness. The analysis is an iterative trial-and-error solution. Initially with
assumed layer thickness, the critical stresses and strains are computed using the ELYSM 5
computer program. These are then compared with relevant failure criteria. When any criterion is
exceeded, the thicknesses are adjusted. This procedure is repeated until all failure criteria are
satisfied.
There is statistical variation in the input parameters. Consequently, there is variability in
the calculated stresses and strains that lead to variations in the number of allowable loads. There
is also variability in the number of expected loads during the design period. Finally, there is
variability in regard to the transfer functions that predict pavement life. The component of
concern was the pavement performance model that predicts pavement life in flexible pavement
thickness design. There are several pavement performance models proposed by researchers in
which there is need to propose that for Nigeria.
1.3 Aim and Objectives
1.3.1 Aim:
To propose the transfer functions that predicts pavement life for Nigerian Empirical
Mechanistic Pavement Analysis and Design System.
1.3.2 Objectives:
 To assess the different fatigue transfer functions and proposes the one that best fit
NEMPADS.
 To generate variability of design parameters using Monte Carlo method with the aid of
MATLAB.
 Assess the effects of axle loads application on pavement analysis and design reliability.
 To establish the minimum number of Monte Carlo simulation required to provides
enough sufficient repeatability for damage reliability relationship.

4
1.4 Location of Study
The study area belongs to region one (i.e. Kano and Kaduna states) of Claros et al., 1986
designation. The Master Test Section (MTS) number for this location was 1-1. It has a route
number A236. The distance in kilometer from the first node is 35 km. the study route is from
node number 241 and node name Zaria (A126) to node number 394 and name A236/state road.
This zone has mean annual rainfall of less than 1, 000 mm with a rainy season of two to three
months. Also, wMAAT for this region were 24 for Kaduna and 22 for Jos.

LITERATURE REVIEW
2.1 Transfer Functions
The empirical component of M-E design is pavement life equation, known as a transfer
function. Transfer function use pavement responses calculated by the mechanistic model and
predict the life of pavement in terms of fatigue cracking or rutting. It acts as a chain between the
pavement reactions and appeared damages in the pavements (Ameri and Khavandi, 2009). It
relates the pavement responses determined from mechanistic models to pavement performance as
measured by the type and severity of distress (rutting, cracking, roughness, and so forth)
(Thompson and Nauman, 1993).
It is generally recognized that the allowable number of traffic load repetitions is closely
related to tensile strain at the bottom of the asphalt layer. A universal form of the fatigue law used
to predict fatigue-cracking life of flexible pavements (Finn 1973; Finn et al. 1973, 1977) is
𝑁 = 𝑘1 ∈−𝑘2 𝐸 −𝑘3 (2.1)
where ϵ = maximum tensile strain at bottom of asphalt layer; E = resilient modulus (i.e.,
stiffness) of the asphalt layer; ki (i = 1,2,3) are parameters of fatigue law; and N = total number of
load repetitions to failure.
2.2 Reliability Analysis
One of the methods cited by Harr, (1987) is the exact methods which includes numerical
integration and Monte Carlo simulation. Briefly, the Monte Carlo method involves artificially
reproducing each input distribution, entering the values into the function, and obtaining the
output distribution. The primary advantage of an exact method is that the complete probability
distribution of the dependent random variable is determined (Harr, 1987).

5
2.2.1 Techniques for incorporating the variation of the input parameters
There are two generally accepted techniques for accommodating the variation of the input
parameters in the design model. These are the Monte-Carlo (Jooste, 1999) and Rosenblueth
(Eckmann, 1997) techniques.
The Monte-Carlo simulation technique randomly generates huge numbers of input data
sets from the known distributions of the input parameters while adhering to the distribution
characteristics of the individual input parameters. These input data sets serve as input to the
structural analysis model and by running the structural analysis model successively using the
different input data sets, a distribution of the resilient pavement response parameters is generated.
The distribution of the pavement response parameter in turn serves as the input to the pavement
performance model (Jooste, 1999).
Monte Carlo simulation is a type of simulation that relies on repeated random sampling
and statistical analysis to compute the results. Monte Carlo simulation can be considered as a
methodical way of doing so-called what-if analysis. In Monte Carlo simulation, we identify a
statistical distribution which we can use as the source for each of the input parameters. Then, we
draw random samples from each distribution, which then represent the values of the input
variables. For each set of input parameters, we get a set of output parameters. The value of each
output parameter is one particular outcome scenario in the simulation run. We collect such output
values from a number of simulation runs. Finally, we perform statistical analysis on the values of
the output parameters, to make decisions about the course of action. We can use the sampling
statistics of the output parameters to characterize the output variation (Samik, 2008).
The Rosenblueth technique is actually a point estimate approximation technique whereby
the continuous distribution of a particular input parameter is approximated by a discrete
distribution of two adequately chosen values of that input parameter. The criteria for selecting the
two discrete values are that the first three statistical moments of the continuous and discrete
distributions must be equal (Eckmann, 1997).
2.2.2 Methods for incorporating variation and reliability in pavement design
Figure 2.1 below shows a simplified diagrammatic representation of a mechanistic-
empirical design procedure. The two highlighted blocks of the diagram represents the components
of the process where measured data are input. Every single input parameter that is measured
empirically and entered into the system has a certain variation associated with it because of the

6
natural variability of the parameter and error in the measurement technique which is hopefully
small. Assuming that the variation in the model is a true reflection of the actual variation of the
physical system, there are therefore two entry points for introducing variability in the design
process namely in the input data which characterize the system and in the performance models
which model the distress or deterioration of the system in response to loading.

1. System geometry input


2. Load characterization
3. Material input
parameters: Resilient
properties Strength
properties

Structural analysis
model: Pavement response
F and ,

Pavement performance
model:Transfer function

Pavement bearing
capacity estimate

N
o
Adequate ?
Yes

Final pavement
design

Figure 2.1: Schematic diagram of a mechanistic-empirical design procedure

METHODOLOGY
3.1 Analysis of Pavement Structure
In mechanistic design method, the pavement is idealized as a layered elastic structure
consisting of various sub-layers of asphalt concrete surfacing, granular base, sub-base, and the
sub-grade. Materials are assumed to be homogeneous and isotropic. The layers are horizontally
infinite with each layer characterized by its resilient modulus MR and Poisson’s ratio μ. Horizontal
tensile strain εt at the bottom of the asphalt concrete layer and vertical compressive strain εz on the
subgrade are identified as the causative factors for fatigue and rutting failures, respectively (Shell

7
1978; Thickness 1981). Based on the algorithm for the analysis of a layered elastic system for
Nigeria (Claros et al., 1986), a computer program NEMPADS was developed by a researcher
(Olowosulu, 2005) for computation of stresses in a pavement structure using the concept of
standard computer programs ELSYM5 developed by other organization.
NEMPADS uses a modified version the Monte Carlo method described by Timm et al.
(1999). A flow chart representation of the NEMPADS procedure is shown in Figure 3. The steps
in the NEMPADS Monte Carlo simulation are as follows:
1. Randomly select input values from their respective probability distributions.
2. Calculate the damage using Miner’s Hypothesis.
3. Perform enough cycles to generate a repeatable output distribution.
4. Determine the number of cycles that resulted in Damage < 1.
5. Calculate reliability according to Equation 3.1:
𝑛𝑢𝑚𝑏𝑒𝑟 𝑜𝑓 𝑐𝑦𝑐𝑙𝑒𝑠 𝑤ℎ𝑒𝑟𝑒 𝐷𝑎𝑚𝑎𝑔𝑒 < 1
𝑅𝑒𝑙𝑖𝑎𝑏𝑖𝑙𝑖𝑡𝑦 = 100 × 3.1
𝑡𝑜𝑡𝑎𝑙 𝑛𝑢𝑚𝑏𝑒𝑟 𝑜𝑓 𝑐𝑦𝑐𝑙𝑒𝑠

3.2 Characterization of Pavement Material


3.2.1 Pavement layer thickness
Pavement layer thicknesses were taken as normally distributed with coefficient of
variation of 5%, 8% and 15% for asphalt concrete, granular base and granular subbase
respectively. The respective pavement layer thicknesses for asphalt concrete, granular base,
granular subbase and subgrade were 2.5, 5.5, 2.8, and 300 inches respectively.
3.2.2 Layer modulus
The resilient modulus (MR) is a measure of the elastic property of a soil recognizing
certain non-linear characteristics. Resilient modulus can be used in a mechanistic analysis using
multi-layer elastic systems for prediction of cracking, rutting, etc (Claros et al., 1986). The
resilient modulus of all the materials (asphalt concrete, granular base, granular subbase, and
subgrade) were taken to be log-normally distributed (Timm et al., 1999).
3.2.2.1 Subgrade materials
Soils usually display stress dependent resilient behavior characterized by a resilient
modulus MR. Soils were collected from several locations on various National Highways, and the

8
resilient modulus test was conducted in a triaxial device equipped for repetitive loading in
which the confining pressure and deviator load are varied. The load and deformation were
recorded during the recommended 200 repetitions at each load setting and confining pressure.
The subgrade modulus obtained after adjusting the modulus for representative stresses
by the Nigerian overlay design methodology research for the route selected for study was 26,
000 PSI. Based upon literature review (Timm et al., 1999), the modulus of the unbound
materials was taken to be log-normal distribution and a practical Coefficient of Variation of
40% was adopted for the unbound materials of the selected route.
3.2.2.2 Granular base and subbase
The resilient modulus of granular materials is stress dependent, and it varies both in
radial and vertical directions because of different levels of confinement and traffic loading
(Animesh and Pandey, 1999). The promising factors which represent the in-situ conditions are
water content, density, load duration, and stress state (Claros et al., 1986).
The Nigerian overlay design methodology estimated the elastic modulus of bases and
subbases to be 90, 000 PSI and 45, 000 PSI respectively (Claros et al., 1986). Based on the
conclusion of Timm et al. (1999) and other researchers, a log-normal distribution was also used
for the granular base and subbase materials. A practical modulus Coefficient of Variation of
30% was adopted for granular base and subbase materials respectively for the route selected
Timm et al. (1999).
3.2.2.3 Asphalt concrete
For asphalt concrete samples the dynamic indirect tensile test (ASTM D-4123) was used
to estimate modulus. The most significant factors that represent the in-situ conditions are
temperature, load duration, and stress state. The estimate of the elastic modulus of asphalt
concrete from the resilient modulus test using the standard test AASHTO T-274 for the study
area was taken to be 900,000 PSI (Claros et al., 1986). Based upon the several research carried
out, the practical modulus Coefficient of Variation of 20% was adopted for asphalt concrete of
the selected route.
3.2.3 Poisson’s ratio
Poisson’s ratio is the ratio of transverse strain to axial strain when a material is axially
loaded. Yoder and Witczak (1975) cite that, for most pavement material, the influence of many
factors on Poisson’s ratio is generally small. According to Pavement Evaluation Unit of

9
Nigerian Federal Ministry of Works and Housing, asphalt concrete is highly dependent upon
temperature, where Poisson’s ratio varies between 0.15 at cooler temperatures (less than 30 0 F)
to 0.45 at warmer temperatures (1200 F plus), with a typical value of 0.35, cement stabilized
bases tends to increase Poisson’s ratio value towards 0.30 from sound (crack free) to value of
0.15 as a result of degree of cracking in stabilized layer, with a typical value of 0.20, granular
base/subbase uses lower Poisson’s ratio value of 0.30 for crushed material and high Poisson’s
ratio value of 0.40 for unprocessed rounded gravels/sands, with a typical value of 0.35, and
subgrades Poisson’s ratio value depends on the type of subgrade soil i.e. Poisson’s ratio value of
0.30 is use for cohensionless soils while Poisson’s ratio value of 0.50 is for very plastic clays
(cohesive soils), with a typical value of 0.40.
3.2.4 Traffic input
Currently, there is a choice between using Equivalent Single Axle Loads (ESALs) or load
spectra. An ESAL is defined as an 80 kN (18 kip) dual tire axle load. The load spectrum consists
of a combination of single and dual tires in single, tandem, or tridem axle configurations (Bruce,
2001). The traffic was modelled as an equivalent number of standard single axle loads of 80 kN
with a contact stress of 550 kN/m2 and a loaded radius of 152 mm.
3.3 Layered-Elastic Analysis Output
The LEA model calculates normal stresses, strains, and deflections as well as shear
stresses at any point in the pavement structure. In NEMPAD, critical strains are used to determine
damage and reliability. The critical strains are the tensile strain at the bottom of the asphalt layer
and the compressive strain at the top of the subgrade (Olowosulu, 2005).
3.4 Monte Carlo Simulation and Reliability Formulation
When a distribution is characterized by a well-known function (e.g., normal or
lognormal), it is possible to work directly with equations to artificially generate the distribution
Timm et al. (1999). Box and Muller (1958) have shown that if U11 and U12 are two
independent standard uniform variates, then
S11 = sqrt (-2 × log (U11)) × sin (2 × pi × U12) Equation 3.2
S12 = sqrt (-2 × log (U11)) × cos (2 × pi × U12) Equation 3.3
are a pair of statistically independent standard normal variates. Therefore, a pair of random
numbers from a normal distribution (N(μ,σ)) may be obtained by:
H1 = [M1 + (D1*S11) Equation 3.4

10
H2 = [M2 + (D2*S12) Equation 3.5
Equations 3.2 to 3.5 can then be repeated for layers 3 and 4. For lognormally distributed
modulus values, equations 3.2 and 3.5 can again be used to generate S11 and S12. For a
lognormal variable E and transformed variable Y = ln(E), equations 3.6 and 3.7 can be used to
calculate the standard deviation and mean of the transformed variable, respectively.
D1 = sqrt (log (CV^2+1)) Equation 3.6
M1 = log (M) - ((D1^2)/2) Equation 3.7
Finally, two E values (log-normally distributed) are calculated by:
E1 = exp (M1 + (D1*S11)) Equation 3.8
E2 = exp (M1 + (D1*S12)) Equation 3.9
The various values obtained from the above concepts were incorporated into the existing
computer program, NEMPADS. The program enables the designer to generate the horizontal
tensile strain at the bottom of the existing asphalt concrete layer and vertical compressive strain at
the top of the subgrade so as to set levels of input variability and evaluate their effects on the
design reliability using Microsoft Excel.
3.5 Pavement Performance Model
3.5.1 Fatigue
Claros et al. (1986) used a version developed by Craus et al., (1984) that was calibrated
using data from the AASHO Road test for thin pavements and a failure criterion of thirty percent
class II cracking. This model includes a modulus term for the asphalt layer in order to capture the
relationship between stiffness and fatigue cracking.
The Nigerian version is shown in Equation 3.10.

15
𝜀𝑡 −3.291 𝐸 0.854
𝑁𝑓 = 9.727 × 10 ( −6 ) ( 3) (3.10)
10 10
where
Nf = Number of allowable 8200 kg ESAL applications,
εt = Horizontal tensile strain at the bottom of the asphalt layer, and
E = dynamic modulus of the asphalt concrete in PSI
A fatigue model developed by Finn et al. (1977) and similar the one developed by Craus et
al., (1984) was also evaluated which was the version used by Olowosulu, (2005). This model

11
(equation 3.11) also includes a modulus term for the asphalt layer in order to capture the
relationship between stiffness and fatigue cracking.
𝜀𝑡 −3.291 𝐸 0.854
𝑁𝑓 = 1.219 × 1016 ( ) ( 3) (3.11)
10−6 10
where
Nf = Number of allowable 8200 kg ESAL applications,
εt = Horizontal tensile strain at the bottom of the asphalt layer, and
E = dynamic modulus of the asphalt concrete in PSI
The FHWA-ARE equation which was also developed using the AASHO road test data
(FHWA report, 1975) was also evaluated. It was observed that this equation is considered a
conservative estimation of the fatigue life in pavement. The equation has the following
expression:

−15
1 5.16
𝑁𝑓 = 9.73 × 10 ( ) (3.12)
𝜀𝑡
3.5.2 Rutting
Equation 3.13 shows the rutting transfer function used in Nigerian Overlay Pavement
Design. This equation was also calibrated for NEMPAD as described by Claros et al (1986).
1 4.7036
𝑁𝑟 = 1.36 × 10−2 ( ) (3.13)
𝜀𝑣
where
Nr = Number of allowable 8, 200 kg ESAL application.
εv = Vertical compressive strain at the top of the subgrade
3.5.3 Miner’s hypothesis
In the simplest case (single load configuration and no seasonal variations in material
properties), the damage over the life of the pavement can be characterized by equation 3.14:
𝑛
𝐷𝑎𝑚𝑎𝑔𝑒 = 3.14
𝑁
where
Damage = an index indicating the expected level of damage after n load applications
(Damage ≥ 1 indicates pavement failure)
n = applied number of loads
N = number of loads required to cause failure (based on empirical transfer functions).

12
Table 3.1: Reliability values for fatigue models with 1, 000 cycles

ESALs CRAUS ET AL. FINN ET AL. FHWA-ARE


337750 99.6 99.7 85.4
506625 99.5 99.6 74.7
675500 98.9091 99.2 67.8636
844375 97.84 99 60.2
1013250 95.15 98.04 53.4
1182125 92.35 96.52 48.5
1351000 89 93.8 43.7727
1519875 85.4 91.4 40.7
1688750 81.2 88.95 36.9
1857625 77.8 85.9 34.7
2026500 73.7 82.9 32

Table 3.2: Reliability values for fatigue models with 1, 500 cycles

ESALs CRAUS ET AL. FINN ET AL. FHWA-ARE


337750 99.6667 99.7333 85.7
506625 99.5333 99.6667 75.3
675500 98.92 99.2667 67.9
844375 97.9 99 60.8636
1013250 95.2 98.1 54.6364
1182125 92.4 96.6 49.6818
1351000 89.2 93.9 43.96
1519875 85.7 91.45 40.7727
1688750 81.75 89.2 37.0909
1857625 78.0909 86.25 34.7273
2026500 74.4667 83.25 32.1818

Table 3.3: Reliability values for fatigue models with 2, 000 cycles

ESALs CRAUS ET AL. FINN ET AL. FHWA-ARE


337750 99.75 99.8 85.8667
506625 99.6 99.7 75.3333
675500 98.95 99.4 68.05
844375 97.9545 99.05 61.12
1013250 95.28 98.1364 54.9
1182125 92.6 96.6818 49.85
1351000 89.3333 94.08 44
1519875 85.8667 91.7333 41.08
1688750 81.8667 89.2667 37.4
1857625 78.1 86.4667 35
2026500 74.55 83.3333 32.56

13
Table 3.4: Reliability values for fatigue models with 2, 200 cycles

ESALs CRAUS ET AL. FINN ET AL. FHWA-ARE


337750 99.76 99.8 85.9545
506625 99.6364 99.72 75.3636
675500 99.0667 99.4 68.24
844375 98.3 99.0667 61.35
1013250 95.3636 98.3 54.92
1182125 92.6667 96.8 49.88
1351000 89.3636 94.1333 44.05
1519875 85.9545 91.76 41.2
1688750 81.9545 89.3182 37.55
1857625 78.16 86.5455 35.15
2026500 74.5909 83.4091 32.7

Table 3.5: Reliability values for fatigue models with 2, 500 cycles

ESALs CRAUS ET AL. FINN ET AL. FHWA-ARE


337750 99.7727 99.8182 86.08
506625 99.64 99.7273 75.4
675500 99.1 99.4091 68.3333
844375 97.84 99 61.12
1013250 95.4667 98.3333 55
1182125 92.6818 97.0667 50.2
1351000 89.44 94.1364 44.5333
1519875 86.08 91.8182 41.6667
1688750 82.04 89.4 37.6
1857625 78.2 86.64 35.7333
2026500 74.64 83.48 33.0667

Table 3.6: Reliability values for rutting model with different cycles.

ESALs 2, 500 Cycles 2, 200 Cycles 2, 000 Cycles 1, 500 Cycles 1, 000 Cycles
337750 99.7 99.5333 99.5 99.48 99.4545
506625 98.9 98.8 98.8 98.75 98.9
675500 98.4 98.3333 98.25 98.16 98.1364
844375 97.36 97.4667 97.45 97.3636 97.36
1013250 97 96.8 96.8 96.7273 96.68
1182125 96.1 95.9333 95.9 95.7727 95.68
1351000 95.3 95.2 95.0667 94.9545 94.88
1519875 94.9 94.75 94.6 94.4545 94.4
1688750 94.5 94.05 93.9333 93.7727 93.68
1857625 93.8 93.4 93.2 93.1364 93.12
2026500 93 92.65 92.4091 92.4 92.32

14
Table 3.7: Reliability values for Craus et al. model with different cycles

ESALs/CYCLES 1000 1500 2000 2200 2500


337750 99.6 99.6667 99.75 99.76 99.7727
506625 99.5 99.5333 99.6 99.6364 99.64
675500 98.9091 98.92 98.95 99.0667 99.1
844375 97.84 97.9 97.9545 98.3 98.3333
1013250 95.15 95.2 95.28 95.3636 95.4667
1182125 92.35 92.4 92.6 92.6667 92.6818
1351000 89 89.2 89.3333 89.3636 89.44
1519875 85.4 85.7 85.8667 85.9545 86.08
1688750 81.2 81.75 81.8667 81.9545 82.04
1857625 77.8 78.0909 78.1 78.16 78.2
2026500 73.7 74.4667 74.55 74.5909 74.64

Table 3.8: Reliability values for Finn et al. model with different cycles

ESALs/CYCLES 1000 1500 2000 2200 2500


337750 99.7 99.7333 99.8 99.8 99.8182
506625 99.6 99.6667 99.7 99.72 99.7273
675500 99.2 99.2667 99.4 99.4 99.4091
844375 99 99 99.05 99.0667 99.1
1013250 98.04 98.1 98.1364 98.3 98.3333
1182125 96.52 96.6 96.6818 96.8 97.0667
1351000 93.8 93.9 94.08 94.1333 94.1364
1519875 91.4 91.45 91.7333 91.76 91.8182
1688750 88.95 89.2 89.2667 89.3182 89.4
1857625 85.9 86.25 86.4667 86.5455 86.64
2026500 82.9 83.25 83.3333 83.4091 83.48

Table 3.9: Reliability values for FHWA-ARE et al. model with different cycles

ESALs/CYCLES 1000 1500 2000 2200 2500


337750 85.4 85.7 85.8667 85.9545 86.08
506625 74.7 75.3 75.3333 75.3636 75.4
675500 67.8636 67.9 68.05 68.24 68.3333
844375 60.2 60.8636 61.12 61.35 61.5333
1013250 53.4 54.6364 54.9 54.92 55
1182125 48.5 49.6818 49.85 49.88 50.2
1351000 43.7727 43.96 44 44.05 44.5333
1519875 40.7 40.7727 41.08 41.2 41.6667
1688750 36.9 37.0909 37.4 37.55 37.6
1857625 34.7 34.7273 35 35.15 35.7333
2026500 32 32.1818 32.56 32.7 33.0667

15
Table 3.10: Reliability values for rutting model with different cycles

ESALs 1000 1500 2000 2200 2500


337750 99.4545 99.48 99.5 99.5333 99.7
506625 98.7273 98.75 98.8 98.8 98.9
675500 98.1364 98.16 98.25 98.3333 98.4
844375 97.36 97.3636 97.45 97.4667 97.5
1013250 96.68 96.7273 96.8 96.8 97
1182125 95.68 95.7727 95.9 95.9333 96.1
1351000 94.88 94.4545 94.6 94.75 94.9
1519875 94.4 94.4545 94.6 94.75 94.9
1688750 93.68 93.7727 93.9333 94.05 94.5
1857625 93.12 93.1364 93.2 93.4 93.8
2026500 92.32 92.4 92.4091 92.65 93

ANALYSIS AND DISCUSSION OF RESULT


The number of Monte Carlo simulations was set at 1000, 1500, 2000, 2200 and 2500
cycles respectively in each of the phase of study.
Phase one allow the designer to know which of the fatigue model that best fit and the
effect of axle load on pavement structure. The reliability values obtained on this phase of study
has been presented in Tables 3.1 - 3.5 for the fatigue models at different cycles and axle load
application. Table 3.6 presented the reliability obtained at different number of Monte Carlo
simulation for the rutting model with different axle load application.
The second phase allows the designer to ascertain the required number of Monte Carlo
simulation that could yield approximate better result. The reliability values obtained in each of the
study for this phase have been presented in Tables 3.7 - 3.10. The performance of the selected
pavement structure was evaluated using Monte Carlo simulations. Tables 3.1 - 3.5 depicts the
number of simulations carried out and the corresponding reliabilities of the pavements for the
various cases considered. Figures 4.1 - 4.5 illustrate the effects of each input parameter’s
variability on output variability in terms of each of the fatigue model studied at different cycles
ranging from 1, 000 to 2, 500 cycles.

16
120

100

80
Reliability %

60 Craus et al.
Finn et al.
40
FHWA-ARE

20

0
0 337750 675500 1013250 1351000 1688750 2026500
Equivalent Single Axle Load (ESALs)

figure 4.1: Graph of ESALs Vs Reliability values for fatigue with


1, 000 cycles

120

100

80
Reliability %

60 Craus et al.
Finn et al.
40
FHWA-ARE

20

0
0 337750 675500 1013250 1351000 1688750 2026500
Equivalent Single Axle Load (ESALs)

figure 4.2: Graph of ESALs Vs Reliability values for fatigue with


1, 500 cycles

17
120

100

80
Reliability %

60 Craus et al.
Finn et al.
40
FHWA-ARE

20

0
0 337750 675500 1013250 1351000 1688750 2026500
Equivalent Single Axle Load (ESALs)

figure 4.3: Graph of ESALs Vs Reliability values for fatigue with


2, 000 cycles

120

100

80
Reliability %

60 Craus et al.
Finn et al.
40
FHWA-ARE

20

0
0 337750 675500 1013250 1351000 1688750 2026500
Equivalent Single Axle Load (ESALs)

figure 4.4: Graph of ESALs Vs Reliability values for fatigue with


2, 200 cycles

18
120

100

80
Reliability %

60 Craus et al.
Finn et al.
40
FHWA-ARE

20

0
0 337750 675500 1013250 1351000 1688750 2026500
Equivalent Single Axle Load (ESALs)

figure 4.5: Graph of ESALs Vs Reliability values for fatigue with


2, 500 cycles

4.1 FHWA-ARE fatigue model


Studying figures 4.1 - 4.5, it is clear that the FHWA-ARE fatigue model resulted in the
best fit for the damage reliability relationship in terms of sensitivity to increased axle load
application. FHWA-ARE fatigue is highly sensitive to increase in axle load application than the
other two models for all levels of number of Monte Carlo simulation studied. This affirms the
conservative estimation of the fatigue life in pavement as was observed by Claros et al. (1986).
4.2 Craus and Finn fatigue model
From figures 4.1 to 4.5, it can be concluded that both Craus and Finn fatigue models
result in best fit for the damage reliability relationship in terms of reliability values as a result of
increased axle load application. Craus and Finn fatigue models are not that sensitive to increase
in axle load application (Tables 3.1 to 3.5) compared to FHWA-ARE model for all levels of
number of Monte Carlo simulation studied. This indicates that both equations are good predictor
for NEMPADS fatigue when considering high level of reliability. Both models show non-
conservative estimation of the fatigue life in pavement and include a modulus term for the
asphalt layer in order to capture the relationship between stiffness and fatigue cracking (Claros et
al., 1986).

19
101
100
99
98
Reliability %

97
96
rutting
95
94
93
92
0 337750 675500 1013250 1351000 1688750 2026500
Equivalent Single Axle Load (ESALs)

figure 4.6 : Graph of ESALs Vs Reliability value for rutting with


2, 500 cycles

74.8

74.6

74.4
Reliability %

74.2

74
Craus et al.
73.8

73.6
0 500 1000 1500 2000 2500 3000
Number of cycles

Figure 4.7 : Reliability value Vs No. of Cycles for 2026500 ESALs using
Craus et al. Model

20
83.6
83.5
83.4
Reliability %

83.3
83.2
83.1
Finn et al.
83
82.9
82.8
0 500 1000 1500 2000 2500 3000
Number of cycles

Figure 4.8 : Reliability Vs No. Of Cycles for 2026500 ESALs using Finn et
al. Model

33.2

33

32.8
Reliability %

32.6

32.4

32.2 FHWA-ARE

32

31.8
0 500 1000 1500 2000 2500 3000
Number of cycles

Figure 4.9 : Reliability Vs No. Of Cycles for 2026500 ESALs using


FHWA-ARE Model

21
93
92.9
92.8
Reliability %

92.7
92.6
92.5
Rutting
92.4
92.3
92.2
0 500 1000 1500 2000 2500 3000
Number of cycles

Figure 4.10: Reliability Vs No. of Cycles for 2026500 ESALs using


Rutting Model

Figures 4.7, 4.8, 4.9 and 4.10 illustrates the effect of number of Monte Carlo simulation
cycles required on different axle load application for Craus fatigue, Finn fatigue, FHWA-ARE
fatigue and NEMPAD rutting models respectively.
Studying these figures, the designer noticed an increased in reliability values as the
number of Monte Carlo cycles increases from 1, 000 to 2, 000 and recorded no much
difference in reliability values for 2, 200 and 2, 500 Monte Carlo simulation cycles. So, it can
be observed that Monte Carlo simulation cycles less than 2, 000 cycles would not provides
enough sufficient repeatability for damage reliability relationship (Bruce, 2001) and Monte
Carlo simulation cycles of 2, 000 up to 5, 000 (Timm et al., 1999) as the case may be could
enough to provide sufficient repeatability for damage reliability relationship.

5.1 Conclusions
 Axle weight has an overwhelming effect on the output variability in terms of fatigue and
rutting therefore deserve careful characterization.
 The minimum number of Monte Carlo simulation cycles that should be used for most
practical design scenarios to provide enough sufficient repeatability for damage reliability
relationship is 2, 000 cycles.

22
 FHWA-ARE fatigue variability is most affected by the inputs closer to the pavement
surface i.e. asphalt modulus.
 Monte Carlo simulation is an effective means of incorporating reliability analysis into the
M-E design process for flexible pavements.
 Reliability may be defined as the probability that the allowable number of loads exceeds
the expected actual number of loads (R = P [N>n]) which is in consistent with other
definitions of reliability.
 The NEMPADS (Finn) fatigue model shows the most promise in terms of developing a
quick predictive model for pavement reliability relationship between stiffness and
cracking is to be considered.

5.2 Recommendation
Many more simulations are required to encompass the full range of pavements types and
CV values for both thickness and modulus. A neural network may prove useful in developing a
comprehensive predictive model for pavement reliability. A quick method of calculating
reliability will greatly speed up the process of evaluating preliminary pavement designs.

REFERENCES
Aliand, H. A., and Tayabji, S. D. (1998): Evaluation of mechanistic empirical performance
prediction models for flexible pavements: Transportation Research Record 1629,
Transportation Research Board, Washington, D.C., 169–180.
Ameri M. And Khavandi Alireza, (2009): Development of Mechanistic-Empirical Flexible
Pavement Design in Iran: Journal of Applied Sciences, Vol. 9, No. 2, pg. 354-359.
American Association of State Highway and Transportation Officials (AASHTO), 1993:
Guide for Design of Pavement Structures: Washington, DC.
Animesh Das and Pandey, B. B. (1999): Mechanistic-Empirical Design of Bituminous Roads:
An Indian Perspective Journal of Transportation Engineering, Vol. 125, and No. 5.
Box, G.E.P and Muller, M.E., (1958): A Note on the Generation of Random Normal Deviates:
Annals of Mathematical Statistics, 29, 1958, Pp. 610-611.

23
Bruce A. Chadbourn (2001): Development of a Quick Reliability Method for Mechanistic-
Empirical Asphalt Pavement Design: Minnesota Department of Transportation, 1400
Gervais Ave. Maplewood, MN 55109-2044.
Burmister, D.M., (1943): The Theory of Stresses and Displacements in Layered Systems and
Application to the Design of Airport Runways proceedings: Highway Research Board,
Vol. 23.
Claros, G., Carmicheal, R.F., and Harvey, J., (1986): Development of Pavement Evaluation
Unit and Rehabilitation Procedure for Overlay Design Method: Vol. 2, Overlay Design
Manual, Texas Research and Development Foundation for the Nigeria Federal Ministry
of Works and Housing, Lagos.
Craus, J., Yuce, R., and Monismith, C. L. (1984): Fatigue behaviour of thin asphalt concrete
layers in flexible pavement structures: Proc., Assoc. of Asphalt Paving Technologists,
Association of Asphalt Paving Technologists, 53, 559–582.
Darter, M. I., and Hudson, W.R., (1973): Probabilistic Design Concepts Applied to Flexible
Pavement System Design: Report 123-18, Centre for Transportation Research,
University of Texas at Austin.
Douglas Hubbard (2007): how to measure anything: finding the value of intangibles in business:
pg. 46, John Wiley & Sons.
Eckmann Bernard, (1997): New tools for rational Pavement Design: Proceedings of the 8th
international conference for Asphalt pavements, University of Washington, Seattle, 1997,
pp. 25-42.
Finn, F. N. (1973): Relation between cracking and performance: Special Rep. No. 140, Highway
Research Board, Washington, D.C.
Finn, F., Saraf, C., Kulkarni, R., Nair, K., Smith, W., and Abdullah, A. (1977): The uses of
distress prediction subsystems for the design of pavement structures: Proc., 4th Int. Conf.
on Structural Design of Asphalt Pavements, Ann Arbor, Mich., International Society for
Asphalt Pavements (ISAP), Lexington, Ky., 1, 3–38.
Harr, Milton E., (1987): Reliability-Based Design in Civil Engineering: McGraw-Hill, Inc.
Hellekalek P. (1998): Good random number generators are (not so) easy to find: Dept. of
Mathematics, Salzburg University, Hellbrunner Straûe 34, A-5020 Salzburg, Austria
Mathematics and Computers in Simulation 46 485-505.

24
Heukelom W. and Klomp A. J. G. (1962): Dynamic Testing as a Means of Controlling
Pavements during and after Construction: Proceedings of the International Conference on
the Structural Design of Asphalt Pavements, Ann Arbor, Michigan, U.S.A.
Huang, Y.H., (1993): Pavement Analysis and Design: Prentice-Hall, Inc., New Jersey.
Irick, P., Hudson, W. R., and McCullough, B. F. (1987): Application of reliability concepts to
pavement design: Proc., 6th Int. Conf. on the Structural Design of Asphalt Pavements,
Ann Arbor, Michi.
Jian-Shiuh Chen, P.E., Chih-Hsien Lin, Erwin Stein and Jurgen Hothan, (2004): Development of
a Mechanistic-Empirical Model to Characterize Rutting in Flexible Pavements: Journal
of Transportation Engineering, Vol. 130, No. 4, pg. 519-525.
Jooste F J. (1999): The influence of variability on routine pavement design: Preprint of Paper
Presented At the 7th Conference on Asphalt Pavements for Southern Africa. 1999.
Jordaan, G J. (1992): Towards Improved Procedures for the Mechanistic Analysis of Cement-
treated Layers in Pavements: Proceedings of the 7th International Conference on the
Structural Design of Asphalt Pavements, Nottingham, England, 1992.
Kulkarni, Ram B., (1994): Rational Approach in Applying Reliability Theory to Pavement
Structural Design: Transportation Research Record No. 1449, Transportation Research
Board, pp. 13-17.
Lemer, A. C., and Moavenzadeh, F. (1971): Reliability of highway Pavements: Highway
Research Record 362, Highway Research Board, Washington, D.C., 1–8.
Lu Sun, W. Ronald Hudson, P.E., and Zhanming Zhang (2003): Empirical-Mechanistic method
based stochastic modelling of fatigue damage to predict flexible pavement cracking for
transportation infrastructure management: Journal of Transportation Engineering, Vol.
129, No. 2, March 1, 2003.
Monismith, C. L., and Tayebali, A. A. (1988): Permanent deformation (rutting) considerations in
asphalt concrete pavement sections: J. Assoc. Asphalt Paving Technol., 57, 414–463.
Monismith C. L., Seed H. B., Mitry F. G. and Chan C. K., (1967): Prediction of Pavement
Deflections from Laboratory Tests: Proceedings of the Second International Conference
on the Structural Design of Asphalt Pavements, 1967.
Olowosulu, A.T., (2005): A Framework for Mechanistic-Empirical Pavement Design for
Tropical Climate: Journal of Civil Engineering, Vol. 5, No. 1 pp 44-51.

25
Perl, M., Uzan, J., and Sides, A. (1983): Visco-elasto-plastic constitute law for a bituminous
mixture under repeated loading: Transportation Research Record 911, Transportation
Research Board, Washington, D.C., 20–27.
Roberts, F. L., Kandhal, P. S., Brown, E. R., Lee, D. Y., and Kennedy, T. W. (1996): Hot mix
asphalt materials, mixture design, and construction, 2nd Ed., NAPA Education
Foundation, Lanham, Md.
Samik Raychaudhuri, (2008): Introduction to Monte Carlo simulation: Proceedings of the 2008
Winter Simulation Conference Oracle Crystal Ball Global Business Unit 390 Interlocken
Crescent, Suite 130 Broomfield, C.O. 80021, U.S.A.
Shell International Petroleum Company Limited (1978): Shell Pavement Design Manual –
Asphalt Pavements and Overlays for Road Traffic: London.
Sousa, J. B., and Weissman, S. L. (1994): Modelling pavement deformation of asphalt-aggregate
mixes: J. Assoc. Asphalt Paving Technol., 63, 224–257.
Theyse H. L. (2000): The Precision and Accuracy of Mechanistic-Empirical Pavement Design:
CSIR Built Environment, Pretoria, Gauteng, South Africa
Thickness design (1981): Asphalt pavements for highways and streets: Manual Series No. 1
(MS-1), the Asphalt Institute.
Thompson, M. R., and Nauman D. (1993): Rutting Rate Analyses of the AASHO Road Test
Flexible Pavements: In Transportation Research Record 1384, TRB, National Research
Council, Washington, D.C., 1993, pp. 36–48.
Timm, D.H., Newcomb, D.E., Birgisson, B., and Galambos, T.V., (1999): Incorporation of
Reliability into the Minnesota Mechanistic-Empirical Pavement Design Method: Final
Report, MN/RC-1999-35, Minnesota Department of Transportation.
Uzan, J. (1996): Asphalt concrete characterization for pavement performance prediction: J.
Assoc. Asphalt Paving Technol., 65, 573–607.
Uzan, J., Zollinger, D. G., and Lytton, R. L. (1990): The Texas flexible pavement systems
(TFPS) mechanistic-empirical model: FHWA/TX-91/455-1, Texas Transportation
Institute, College Station, Tex.
Yoder, E.J., and Witczak, M.W., (1975): Principles of Pavement Design: New York: John
Wiley and Sons.

26

S-ar putea să vă placă și