You are on page 1of 47

Road wear from Heavy

Vehicles
- an overview
Report nr. 08/2008
NVF committee Vehicles and Transports

Frfattare:

Titel:
Serie:
Upplaga:
Utgivningsort:

Mattias Hjort
Mattias Haraldsson
Jan M. Jansen
Road Wear from Heavy Vehicles an overview
NVF-rapporter
100
Borlnge, Sverige

Tryck:

Vgverket 2008

ISSN:

0347-2485

NVF-rapporterna kan bestllas via respektive lands sekretariat per telefon, fax,
e-post eller post. Se kontaktuppgifterna p nst sista sidan.
En uppdaterad rapportfrteckning finns p frbundets nordiska hemsida,
http://www.nvfnorden.org.

Road Wear from Heavy Vehicles


- an overview

Report nr. 08/2008


NVF committee Vehicles and Transports

Sammanfattning

Denna rapport r en sammanstllning ver hur olika fordonsberoende parametrar fr


tunga fordon inverkar p vgslitaget. Exempel p sdana parametrar r axellast och
dckstorlek. Studien baseras huvudsakligen p Cebons Handbook of Vehicle-Road
Interaction, DIVINE projektet, och COST 334. Syftet har varit att sammanstlla olika resultat utan
att gra ngon egen utvrdering av resultaten.

Yhteenveto

Tm raportti on yhteenveto raskeiden ajoneuvojen eri ajoneuvoperusteisten parametrien


vaikutuksesta teiden kulumiseen. Tllaisia parametreja ovat esimerkiksi akselipaino ja
rengaskoko. Tutkimus perustuu pasiallisesti Cebon's Handbook of Vehicle-Road Interaction ksikirjaan, DIVINE -projektiin ja COST 334:n. Tarkoituksena on yksinomaan ollut kert yhteen
tulokset ilman sen enemp analysointia.

Summary

This report is an overview on how different studies of road wear have reported the effect of
various vehicle dependent parameters, for heavy vehicles, such as axle load, tyre properties etc.
The intention with this overview has only been to collect the different results, without any kind of
evaluation. The overview is mainly based on Cebons Handbook of Vehicle-Road Interaction, the
DIVINE project, and COST 334.

Table of contents
1. Introduction

10

2. Pavement distress modes

12

2.1 Introduction

12

2.2 Pavement construction

12

2.3 Pavement distress modes

13

2.4 Actual road wear

15

2.5 Pavement sustainability

16

3. Vehicle Effects

17

3.1 Introduction

17

3.2 Relative effects of different axle loads

17

3.3 Dynamic axle loading

19

3.4 Effects of different axles: single axles, tandem axles and tri-axles 19
3.5 Tyre specific effects

22

3.6 Suspension effects

25

4. Experimental results from the literature

26

4.1 Overview

26

4.2 Experimental findings: static axle load

27

4.3 Experimental findings from using different tyres

27

4.4 Experimental findings regarding suspensions and dynamic loading 33


4.5 A model taking all factors into account
5. Calculation models used in the Nordic countries
5.1 Classification of special heavy transport in Denmark

33
36
36

5.2 The Swedish Road Administration method of calculating


the equivalent number of standard axles

39

6. Acknowledgements

43

7. References

44

1 Introduction
This report within NVF Fordon och Transporter is intended as an overview on how different
studies of road wear have reported the effect of various vehicle dependent parameters, such as
axle load, tyre properties etc. The intention with this overview is only to collect the different
results, without any kind of evaluation. Chapter 2 gives a short introduction to pavement
construction, and the various modes of damage that occurs. Chapter 3 describes the vehicle
dependent parameters and the models that are normally used for predicting vehicle induced road
wear. The range of reported values in the literature of these vehicle parameters are given in
chapter 4. Two examples of calculation models used in Denmark and Sweden are described in
some detail in chapter 5.
Some major works have been conducted in recent years, and form the basis of this report. The
primary sources are

Handbook of Vehicle-Road Interaction [1] (1999) by David Cebon.

DIVINE (Dynamic Interaction between Vehicle and INfrastructure Experiment) [2], OECD
project 1992-1998. An experimental study focusing on how dynamic loads are affecting
road deterioration.

COST 334 Effects of Wide Single tyres and dual Tyres [3] (1996-2000). Full scale
tests were carried out, in addition to literature studies, in order to describe all tyre
parameters that affect the road wear, and how their combined effect can be determined.

A general description of the field of Interaction between vehicle/climate and the road, including a
discussion on the further need of research, has recently been written in Swedish. :

Interaktion mellan fordon/milj och vg [4] (2004) by Ulf Isacsson, KTH

That report has also been used as a basis for this report.
.
The present increase in deployment of mechanistic design, especially in USA, of course
enhances the world wide research and development within the issue of this report. Some very
interesting results have been presented recently from USA, but they have not been discussed in
context of this report. These new results are described very briefly below.
From implementation Mechanistic-empirical design in USA, the 3 major approaches about axle
load versus road wear are:
1. Development of a load damaging factor.
The dynamic effect of road unevenness adds a factor to the road wear calculated from static
axle loads. This effect is caused by an interaction between the dynamic of the vehicle and the
wave length and amplitude of the road surface roughness, and depending on the travel
speed. Thereby the damaging factor results, such that the effect on road wear will relate to
class of road, i.e. different effect on motorways and minor roads as well as rural and urban
roads.
2. Effect of axle loads, repetitions and spacing on dissipated creep strain energy, DCSE, in
relation to development of cracks in pavement materials.

10

Traditional road wear is viewed in relation to strain-fatigue caused by load repetitions,


however new research focus on dissipated creep strain energy induced in the pavement
materials by the actual strain work under the repeated axle loads. In this relation the spacing
between the load repetitions are taking into account, because with narrow spacing complete
strain release doesnt happen between the load applications. This effect change the
traditional way of perceiving road wear always as repeated single loads toward regarding
road wear differently from axle groups.
This is an important effect to obey, when calculating road wear applied by modern lorries and
semitrailers often having dual, triple or maybe quad-axle configurations
3. Investigation of axle load spectras from regular lorry traffic
The variation in axle loads from different axle load levels are examined to establish a
representative value for axle loads based on weigh-in-motion, WIM, registrations of regular
traffic loads exposure on pavements. The results indicate the use of lognormal distribution
for description of normal axle load distribution.
Especially inference of axle loads from WIM registration must be investigated in relation to
calculation of road wear from regular traffic flow.

11

2. Pavement distress modes


2.1 Introduction
Pavement wear is a process in which several different deterioration processes act and interact,
influenced by a variety of factors. (These factors include environmental factors such as
temperature and moisture, but only the traffic-related factors will be considered in this report. For
the effects of environmental factors, see e.g. [1, 4]). The focus of this report is instead on the
influences of various vehicle parameters, such as tyre type (single / dual / wide single), tyre size,
wheel load, inflation pressure. To describe the effect a vehicle has on road deterioration, a
distinction has to be made into different modes of distress. The different distress modes are
described in section 2.3

2.2 Pavement construction


A typical flexible (or asphalt) pavement can be divided into four layers

Figure 2.1: A typical asphalt pavement construction


The top layer is normally an asphalt layer. In British English, the word 'asphalt' refers to a mixture
of mineral aggregate and bitumen. Bitumen is a mixture of organic liquids that are highly viscous,
black, sticky, entirely soluble in carbon disulfide, and composed primarily of highly condensed
polycyclic aromatic hydrocarbons. Bitumen is the residual (bottom) fraction obtained by fractional
distillation of crude oil. It is the heaviest fraction and the one with the highest boiling point.
The base layer is an unbound layer which typically consists of unstabilized aggregates. The
aggregate bas could also be stabilized with asphalt, portland cement, or another stabilizing
agent.
The subbase is mostly a local aggregate material. Also, the top of the subgrade is sometimes
stabilized either cement or lime.

12

Rigid pavements are mostly found in major highways and airports, and like flexible pavements
they are designed as all-weather, long lasting structures to serve modern day high speed traffic.
The rigid pavement has a top layer consisting of a concrete slab of thickness of 10-30 cm. The
load transmission mechanisms of these two pavement types are different. The rigid pavement
distributes the load over a large area, while flexible pavements, when the traffic load is applied
on top of the surface layer, a localized deformation occurs under the load, as shown in Fig. 2.1.
The load is distributed over a small area at the surface, but as the depth increases, the same
load is distributed over a larger area. Therefore, the highest stress occurs at the surface and the
stress decreases as the depth increases. Thus, the highest quality material needs to be at the
surface, and as the depth increases, lower quality materials can be used. When the load is
removed the pavement layers rebound. A very small amount of deformation, however, could
stay permanently which could accumulate over many load repetitions causing rutting in the
wheel path. [7]. The design life of a flexible pavement may be in the range of 15-20 years, while
it is common for a concrete pavement to be designed with a service life of 30-40 years.

2.3 Pavement distress modes


Pavement wear or pavement distress is the degradation of pavement quality due to loading by
traffic and/or climate. For flexible pavements, the following distress modes (visible distress
together with the deterioration process causing it) are relevant (definitions and descriptions are
based on ref [3]):
Cracking

fatigue cracking: being cracking in the bituminous or cement bound material originating
at the bottom of the respective layers, due to fatigue of the material by a great number of
repetitions of bending due to wheel loads (Fatigue defined in this way is used as a
parameter in pavement design. This does not include surface cracking and cracking due
to thermal cycling, although these are also due to fatigue because of repeated stress
cycles.)

thermal cracking: being cracking in the bituminous material due to tensile stresses
caused by temperature changes

surface cracking: being cracking in the bituminous material originating at the surface of
the pavement, due to fatigue of the material by a great number of shear loadings of the
pavement surface by the tyre (Ageing of bituminous materials plays an important role
here, too.)

reflective cracking: being cracking of the (top) bituminous layers (often in a composite
structure) as a result of cracks or joints in bound layers below.

Rutting
Rutting is the development of depressions in the pavement surface along the wheel paths,
typically with a width of several decimetres and a length of tens to thousands of meters. The
three main categories are

Primary rutting: rutting due to permanent deformation of bituminous layers, (Permanent


deformation can be due to (post)compaction or (plastic and viscous) deformation caused
by shearing stresses.)

13

Secondary rutting: rutting due to permanent deformation in the subgrade or in granular


layers below the asphalt layers.

Rutting due to abrasion of the pavement surface by studded tyres.

Other distress modes

Ravelling: being the loss of stones in the surface of the pavement as a result of failure of
the bond between the aggregate and the binder by a great number of shear loadings in
combination with ageing of the material. The primary case, is however insufficient quality
of the pavement material.

Roughness: being (longitudinal) unevenness of the pavement, mostly due to several


combined factors (rutting, cracking, potholes, uneven settlements, etc.)

Potholes: resulting either from local collapse due to structural defects, or from frost
acting on water ingress (often through cracks). Potholes are not necessarily caused by
loading but mainly due to insufficient quality of the pavement

The more important of these distress modes are shown in Figure 2.2.

Figure 2.2. Various modes of pavement distress. (from Ref. [3]). For explanation on
nomenclature, see Sec.2.2. HMA is short for Hot Mix Asphalt.)
The main distress modes, especially from the point-of-view of traffic loading are:

14

1. Fatigue cracking. This occurs mainly on relatively weak / thin pavements (Visible
cracking in thick pavements is likely to originate (at least partly) at the surface.)
2. Primary rutting. This occurs mostly on main roads with thick bituminous layers.
3. Secondary rutting. This occurs mainly on relatively weak / thin pavements.
It is clear that these different distress modes react differently to changes in the influencing
factors. Take e.g. the influence of tyre type, where the stress and strain conditions near the
surface of the pavement are strongly influenced by the contact stresses and their distribution in
the tyre-pavement interface, whereas the stresses and strains deeper in the structure are mainly
influenced by the total load. Therefore, a change in contact stress distribution due to a change in
tyre type can generally have most influence on the upper layers2.
Most design methods for flexible pavements since the 1960s are based on the prevention of
fatigue cracking and secondary rutting. In relatively weak /thin pavements this does not always
succeed, but in thick pavements it generally succeeds. Then, primary rutting may become the
dominant distress.
Permanent deformation of bituminous layers is usually not considered as a part of structural
design. With proper bituminous mixture design the tendency for permanent deformation can be
decreased. However, the bituminous mixture design is always a compromise between many
properties (including price) and small changes in mixture composition during the manufacturing
may worsen the properties of bituminous mixtures. It is possible to manufacture mixtures, which
will not deform easily, by using modified bitumens but they are much more expensive and thus
their use is limited.
Ravelling and surface cracking, being the most superficial distress modes, may be influenced by
any differences in contact stress distributions between different tyre types.

2.4 Actual road wear


The road wear from actual commercial vehicles of course depends on the degree of loading and
type of goods carried by the vehicle. The actual road wear exposed by the traffic on pavements
can only be determined by weighing of actual lorries, busses, semi- trailers etc. in traffic flow.
Maximum permissible gross weight and axle loads are not always reached by regular commercial
vehicles, because a great proportion of the freight is volume based.
Besides distinction in type of vehicle classes the road wear from typical vehicles is also different
from traffic on rural and urban roads. The actual road wear exposed by the traffic flow very much
also depends on the traffic composition.
The discussion in this report therefore must be seen in relation to actual axle loads from vehicles
in regular traffic flow.

However, exceptions to this rule may occur, depending on structure and material quality (e.g. a very
critical stress-sensitive granular layer below a rather thick AC layer may be the main cause of rutting
increase due to a slight increase in stresses).

15

2.5 Pavements sustainability


The road wear from the traffic will be exposed to road pavements, and depending on the ability of
the pavement materials and design the pavement will develop distresses over time.
With time the pavement will develop described distress modes in various degrees until the
pavement reaches an unacceptable condition and will be rehabilitated by resurfacing and
possibly strengthening. Therefore road wear is a very important parameter to focus at for design
and maintenance of road pavements.
Prescription of thresholds for unacceptable condition normally relates to functional and structural
properties of the pavement. Classical functional criterias related to the road surface are building
of rut depth and deterioration of the evenness and the structural criteria is proportion of cracked
area in wheel tracks, related to the bound pavement materials e.g. asphalt.

16

3 Vehicle effects
3.1 Introduction
This section describes the influence of properties of different vehicle components on exposed
road wear. It is the intention of this chapter to explain the concepts, and to briefly describe what
kind of road damage the different factors leads to. Experimental findings concerning these factor
are reported in Chapter 4.
The basic parameter is the axle load. The effect of different axle load sizes is described in
section 3.2, explaining the concept of Load Equivalency Factors. It is important to realise that the
actual forces on the road are not equal to the static axle loads, but vary because of vehicle
dynamics. This is the topic of section 3.3, where dynamic loads are described. Also, the effect of
axle loads may be influenced by neighbouring axles, as explained in section 3.4. Another factor
to be taken into account is the lateral wander of the traffic, elaborated in section 3.5. In section
3.6, the load sharing between twinned tyres is discussed. The influence of different tyres are
described in section 3.7, introducing the concept of super single tyres. Finally, different kind of
suspensions are described in section 3.8.

3.2 Relative effects of different axle loads


For pavement design, but also to determine the pavement wear effect of different tyres, the
pavement wear effects of different axle loads have to be determined. Generally this is described
by a Load Equivalency Factor (LEF), where an axle load is said to be equivalent (producing
equal pavement wear) to a number of applications of a reference (standard) axle load. The most
well-known of such a LEF is the so called fourth power law which is expressed mathematically
as follows:

N ref
Nx

W
= x
W
ref

(3.1)

where Wx and Wref are axle loads and Nx and Nref are the corresponding number of load
applications.
The exponent 4 in the fourth power law was found in the AASHO Road Test, carried out in USA
between 1958-1960.. However, it was not strictly constant in that test but varied from about 3.6 to
4.6. Later experimental and theoretical research has indicated greater variability in the exponent,
but has not been conclusive. As an example, it was found in the OECD FORCE project that the
exponent depends also on the extent of distress, the exponent being smaller in earlier phases
than in later phases of failure.
It must be understood that the fourth power law includes all distress modes. The most important
at the AASHO road test were rutting (caused by subgrade deformation) and roughness
(unevenness) of the road. Cracking had a minor effect and deformation of bituminous mixtures
was not important.
When individual distress modes are considered, different exponent values are found. E.g. COST
334 reports that cracking of bituminous layers has a value of 4 7, permanent deformation of the
subgrade has an exponent of perhaps 3 4 and permanent deformation of bituminous layers a
value of 1 2. As these values depend on many factors (a.o. material variations) and are not
fully known, the stated values should be regarded as best estimates [3].

17

For use in pavement design, where the actual spectrum of axle loads has to be converted to an
equivalent total number of standard axle loads, COST 334 concludes that the precise exponent
value is not very important. For exponent values between 2 and 6, most actual axle load spectra
were found to translate to roughly the same equivalent number of standard axles. (For low
exponents, the multitude of smaller axle loads contribute to the bulk to the total equivalent
number. For high exponents, the few overloaded axles contribute the most.) It was thus
concluded that the overall value of 4 is well suited [3].
For detailed studies into pavement wear effects the exponent values for the individual distress
modes should be distinguished. This is especially the case when conclusions should be drawn
from accelerated tests at high load values.
Besides being used within road design, the fourth power law has also been used for evaluating
the road damaging potential of abnormal vehicles. E.g., in Denmark, a formula based on the
fourth power law is currently used for classifying the road damage of Heavy Abnormal
Transports, and is the base for taking decisions on allowed routes for heavy transports. This
formula is described in some detail in section 5.1.
However, according to Cebon [1], the validity of the fourth power law is questionable,
particularly for current axle loads and axle group configurations; tyre sizes and pressures; road
construction; and traffic volumes: all of which are significantly different from the conditions of the
AASHO road test.
The Divine project comes to a similar conclusion, and write that the use of the fourth power law
may not be appropriate in all situations unless the environment, traffic, pavement type and
pavement construction methods are the same as, or very similar to, those in the AASHO Road
Test [2].
Note that the fourth power law was derived from measurements on heavy vehicles. To apply the
law also on light vehicles, such as passenger cars, would imply a vast extrapolation outside the
range of vehicle loads used in the AASHO experiments. The Swedish Road Administration
advice against comparisons between heavy vehicles and passenger cars, with respect to the
concept of an equivalent number of vehicles [8].
An example on how to apply the fourth power law on measured axle loads is taken from
Denmark. From weighing of actual axle loads at weighing stations or weigh-in-motion bridges,
the load equivalent of vehicle classes can be obtained for ordinary traffic using the fourth power
law. Results from these measurements, compared with the ESAL resulting from maximum
allowed weights of these vehicles, are shown in table below:
Table 3.1: Load equivalent of vehicle classes obtained from weighing the actual axle loads
Road wear in 10 T ESAL by
Typical permissible Gross
Typical Vehicle classes
4th-power law
weight
From
From max.
measured
load
loads
2-axle lorry
0.3
1,4
18000 kg
Lorry with trailer
1,1
3,4
38000 kg
Semitrailer
0.9
2,6
42000 kg
Busses
0,4
1,4
18000 kg
For more details regarding this study, we refer to Jan M. Jansen, Vejdirektoratet, Denmark.

18

3.3 Dynamic axle loading


When a vehicle is not moving, the vertical (axle, wheel and tyre) loads it imparts on the
pavement, due to the force of gravity, are constant. These are the static loads. When the vehicle
is moving along a road, however, unevenness of the road will cause the vehicle to move up and
down. This will cause a dynamic variation of the loads on the pavement, above and below their
static values.
The magnitude of this dynamic variation depends also on the vertical dynamics of the vehicle,
including such factors as the mass and stiffness distribution of the vehicle structure, payload
mass distribution, suspension and tyres, and on the road surfaces longitudinal profile and the
speed of the vehicle. The variation generally increases with both speed and nature road
unevenness.
The magnitude of dynamic loads is mostly expressed as the Dynamic Load Coefficient (DLC),
defined by the OECD as the ratio of the RMS (root mean square) dynamic wheel load to the
mean wheel load. The RMS of the dynamic wheel load is essentially the standard deviation of the
probability distribution of the total wheel load. The mean value reflects the static wheel load. So,
the DLC is the coefficient of variation of the total wheel load.
This is reported by OECD to range between

5 10% for well-damped air suspensions and soft, well-damped steel leaf suspensions.

20 40% for less road-friendly suspensions (OECD 1992).

Dynamic loading increases pavement wear. Because of the power-law dependency of pavement
distress on axle loads, the loads above the static load increase the pavement wear more than the
decrease in wear due to the loads below the static load.
Besides load magnitude, also frequency content is important for pavement wear. Most heavy
vehicles have dynamic wheel loads either in the 1.5 4 Hz range, associated with bounce
(up/down) and pitch (rotating forward/backward) motions of the vehicle body, or in the 8 15 Hz
range, associated with axle-hop vibration. Axle hop vibrations are more significant if the
pavement is rough and the vehicle speed is higher than approximately 40 km/h.
As stated before, the tyre characteristics (vertical spring compliance and damping) influence the
dynamic vehicle loads. Therefore, these should be considered when establishing pavement wear
effects of different tyres.

3.4 Effects of different axles: single axles, tandem axles and tri-axles
According to COST 334, tandem axles and tri-axles (see the definitions in Table 3.1) generally
cannot be treated by summation of the effects of their constituting individual axles, because of
two reasons:

The load spreading of thick pavements may be such that the responses (stresses and
strains) due to neighbouring axles in a tandem or tri-axle configuration may substantially
increase the responses under the axle considered. Due to the non-linearity of the
performance relations, such increased responses will lead to much more pavement wear
than the summed responses of individual axles.

Due to the visco-elastic nature of bituminous materials, stresses and strains caused by an
axle load need some time to relax after the axle has passed. When another axle arrives
within that period, some residual stresses and strain will still be present, which may

19

compound with the stresses and strains caused by the new axle, resulting in higher total
values. The effects of this mechanism are not well understood.
For axle load limitations, this is reflected in maximum allowed tandem axle (and tri-axle) loads
which are less than twice (or three times) the allowed single axle load. (Two axles at more than
1.8 m spacing are not considered a tandem axle but a double axle and are treated as two
single axles.)
Cebon concludes the opposite, and writes that It is generally concluded that for equal damage to
flexible and rigid pavements, tandem and triaxle groups can carry more weight than the same
number of widely spaced single axles, because the primary response fields of nearby axles
overlap.
The complicated relation between inter-axle spacing and relative road damage is illustrated in
Figure 3.1, taken from Cebon, where the effect of tandem axle group spacing on theoretical
fatigue damage in a rigid pavement is shown. The figure shows that there is an optimum axle
spacing, and using either a wider or more narrow axle spacing will result in increased road
damage. Moreover, this optimum axle spacing depends to a large extent on the thickness of the
concrete slab.

Figure 3.1: The effect of tandem axle group spacing on theoretical fatigue damage in a rigid
pavement (From Cebon, p.301)
For pavement design purposes, however, the loads of tandem axles and tri-axles are mostly
converted to a number of equivalent standard axle loads (Nesal) by summing the contributions of
the individual axles [3]. These individual contributions are then calculated using the Load
Equivalency Factor described in Section 3.2., resulting in:

N esal =

nr of axles

20

Waxle

Wstandard axle

(3.2)

Table 3.2: Definitions of different axle combinations


Single axle

a single axle with more than 1.8 m spacing from other axles.

Double axle

a configuration of two axles, with more than 1.8 m spacing. (This is not a
tandem axle, and the individual axles of a double axle are considered
separately.)

Tandem axle

a configuration of two axles, with less than 1.8 m spacing between the axles.
(Often the suspension of a tandem axle is such that the load on the tandem
axle is shared rather equally between the constituent axles.) The maximum
load is dependent on axle spacing and suspension, and is different for motor
vehicles or for trailers and semitrailers. (The EC also distinguishes a bogie,
being two axles with shared suspension and less than 1.3 m spacing).

Tri axle

a configuration of three axles, with relatively short longitudinal distance


between the axles. (Often the suspension of the tri-axle is such that the load
on the tri-axle is shared rather equally between the constituent axles.)

As an example on how different axle configurations are used in regulations, the maximum
allowed loads in Sweden are shown in the table below:
Table 3.3: The maximum allowed loads (in metric tons) in Sweden for different axle
configurations and road classifications (BK1, BK2 and BK3). From [4].
BK 1

BK 2

BK3

a) Non driven axle

10

10

b) Driven axle

11,5

10

a) Inter-axle distance < 1,0 m

11,5

11,5

11,5

b) 1,0 m Inter-axle distance < 1,3 m

16

16

12

c) 1,3 m Inter-axle distance < 1,8 m

18

16

12

d) 1,3 m Inter-axle distance < 1,8 m,

19

16

12

20

16

12

a) The distance between the outer axles < 2,6 m

21

20

13

a) The distance between the outer axles 2,6 m

24

22

13

1. Axle load

2. Tandem axle load

and the driven axle has dual tyre fitment and air suspension, or
equal suspension, or if the driven axles has dual tyre fitment
and the load on each axle does not surpass 9,5 tons
e) Inter-axle distance 1,8 m
3. Tri axle load

21

3.5 Tyre specific effects


The current trend (in Europe) is smaller tyre diameter and higher tyre pressure. A smaller tyre
diameter enables lower vehicle floors, which increases the volume that is possible to transport. A
higher tyre pressure might have a positive effect on fuel consumption. Also, wide single tyres are
beginning to replace the traditional dual tyres which can be explained by a lower weight,
reductions in fuel consumption, and a lower cost of tyre wear. Although beneficial for
transporters, the effect of these trends might be an increase to road wear since they imply a
smaller contact area between a tyre and the road. This area, henceforth called footprint is an
important road wear factor. The larger the footprint, the less the load distributed on every road
area unit. The difference in road wear between single and dual tyres is thus not caused by the
differences in tyre types as such. (COST 334, section 4.1)
Single tyres are normally used on axles with loads below 8 ton, while dual tyres are the most
common for axle loads above 8 ton. However, there exist wide base single tyres exist for axle
loads of more than 10 ton. The use of single or dual tyres also depends on the axle formation.
The more axles constituting the formation (tandem or tri-axle), the more common is the use of
single tyres. The wide base single tires, is more commonly used in Europe. For example, as of
1997, wide-base singles comprised 30 percent of the tires in France and Britain compared to 2
percent in the U.S. [10]. The difference in size between one of the most extreme wide base
single tyres and an ordinary dual tyre is shown in Fig. 3.2. There are many different types of
wide-base tires available in varying sizes and inflation pressures. In addition to simply enlarging
the size of the tire, there have been recent developments in tread and load-bearing technology
that allow for lower inflation pressures and even greater surface area than the initial wide-base
tires that were introduced in the early 1980s [10].

Figure 3.2: Wide-base Single Tire (445/50R22.5) compared to standard dual tyres
(275/80R22.5). (Picture from Ref. [10], with permission from the author)
However, although wide base single tyres might increase road damage through its smaller
footprint and higher inflation pressure, it improves roll-over stability of the vehicle. This in turn has
consequences for suspension design. As is discussed below, suspension stiffness is necessary
for roll-over stability but increases road damage. It is therefore at least a theoretical possibility
that the increased road damage induced by the change from dual tyres to super single tyres can
be compensated for by reducing the spring stiffness. Also, the damping needed to reach the

22

minimum road damage is about 25 percent lower for wide single tyres than for dual tyres, mainly
because of a lower unsprung mass. This improves the possibility to design an optimal
suspension. (Cebon p. 453).
Another important aspect to take into account when comparing single and dual tyre assemblies is
the concept of unequal load sharing of the dual tyres. When comparing dual and single tyre
assemblies at equal wheel load, generally the assumption is made that the wheel load is shared
equally between both tyres of the dual assembly. However, in practice this might not be true. A
number of reasons could cause an unequal load division (load imbalance) between both tyres:
1. differences in vertical stiffness between both tyres, because of

differences in inflation pressure (mainly due to poor maintenance)

different tyre structure ( due to e.g. different brands)

2. differences in vertical compression between both tyres, because of

differences in diameter between both tyres

bending of the vehicle axle

transverse unevenness of the pavement surface

These reasons are illustrated in Figure 3.3.

Figure 3.3: Causes unequal load sharing between tyres in a dual tyre assembly (load
imbalance). From Ref. [3].
Axle- and tyre configuration also affect road damage in an indirect way, an effect that is denoted
lateral wander.
In practice, not all wheels will pass at the same lateral position in a road section. Vehicles
generally follow a slightly zigzagging course between the bounds of the traffic lane, which is
called lateral wander. Therefore, the wheel positions of consecutive vehicles will be transversely
distributed over the pavement.

23

Detailed measurements and analysis of this distribution are reported by Blab [5]. He showed that
the probability distribution of the vehicle positions is a Laplace distribution, in stead of the normal
distribution that is often assumed. For a certain vehicle width and lateral wheel spacing, the
probability distribution of the wheel (centre) positions is a Laplace distribution, too. However, the
number of hits by a tyre per cm pavement width is approximately normally distributed, due to the
summation over various vehicle widths, wheel spacings, dual and single wheels, and various tyre
widths. The difference is shown in Figure 3.4.

(a)

(b)

Figure 3.4: The difference between probability distribution of the wheel positions (a), and the
number of hits by a tyre per cm pavement width (b). From Ref. [5].
Lateral wander distributes pavement loading, and hence pavement wear, over a larger area of
the pavement. This prolongs the pavement service life. The effects of lateral wander are different
for the different distress modes. They also may differ between dual tyres and wide base singles.
COST 334 reported that the distress reduction factor of lateral wander on primary rutting
compared to non-wandering loading were in the range 0.67 0.87, using different road structures
and a range of tyres. Generally, the effects of lateral wander for the dual tyres that were tested
were all very similar and close to those of a 385 mm wide single tyre. The effect of lateral wander
for a 495 mm wide single tyre were smaller (about 30%) than for the other tyres. They also
concluded that the beneficial effects of lateral wander for all tyres increase with decreasing
pavement thickness.

24

3.6 Suspension effects


As mentioned above (Section 3.3) vertical movements in the vehicle create a dynamic axle load
that can be higher as well as lower than the static load. A road-friendly suspension decreases the
magnitude of the vertical movements that are triggered by unevenness of the road, which in turn
means a lower dynamic axle load, for a given static load The dynamic axle load is affected by the
type of suspension on the vehicle and its ability to reduce movements in various frequency
ranges. In general, the dynamic load increases with spring stiffness. The dynamic load is also
sensitive to the damping ratio, the speed of which the amplitude of movements is reduced. EU
(directive 92/7/EEC) defines a road-friendly suspension as a one with a sprung mass frequency
no greater than 2 Hz and a damping ratio above 20 percent of critical damping. It is also required
that the Coulomb damping ratio does not exceed 50 percent of the viscous damping.
From a road wear perspective, the purpose of the suspension is to reduce motions to decrease
the dynamic load. But the suspension characteristics have implications for road safety as well so
there are certainly important trade offs. Reducing the stiffness in order to decrease the dynamic
load also reduces the resistance to roll over.
The traditional truck suspension is constructed using spring leaves. The characteristics of this
type of suspension vary considerably under different road and driving conditions. The friction
between the leaves might result in a locked suspension in case the pavement is smooth and the
speed moderate. This is caused by sliding friction between spring leaves and has been found to
adversely affect suspension performance. In the locked cases the suspension is stiff and the
dampening abilities poor. Newer steel spring suspensions have been designed to remedy this
problem by minimizing the contact area between individual leaves. With this construction
stiffness is lower compared to the traditional design and a complete lock is not possible. The
dampening effect is however reduced considerably why it is necessary to use hydraulic dampers.
In recent years, air suspension has been developed. These provide a low stiffness and a smooth
deflection characteristic. Air suspensions are used in combination with hydraulic dampers.
Hydraulic dampers (shock absorbers) are common in modern trucks in combination both with leaf
and air springs. Their force generating characteristics depend on the amplitude and frequency of
the imposed motion. A semi-active damper is able to dissipate energy at a continuously, variable
and controllable rate. It can be switched off when it is required to feed power into the suspension.
(Cebon)

25

4. Experimental results from the literature


4.1 Overview
The experimental results regarding vehicle influence on road deterioration are indeed wide
spread. This is likely because of the many different vehicle features contributing to the road wear,
which in turn is highly dependent on the kind of road, and the type of damage that is considered,
which makes it hard to deduce the effects from a single feature alone.
David Cebon made a summary of some of these effects in his book from 1999, (see ref. [1], page
321). The results, shown here in Fig. 5, are based on 19 references from 1965 1992, with the
majority of the work carried out during the 1980s. In that summary no distinction between
different forms of distress modes, or different pavement thicknesses has been made.

Figure 4.1: Summary of literature on the effects of various vehicle features on road damage.
From Ref. [1].
From the chart in Fig. 4.1, Cebon draws the following conclusions:
1. Applying a tandem suspension load to a single axle can be expected to increase road
damage by a factor of up to 25 (first bar of chart)3.
2. Replacing dual tyres with wide-base single tyres may increase road damage by a factor of
up to 10 (second bar of chart).
3

Cebon points out that this is not a particularly realistic scenario, but the included it for comparison purposes.

26

3. Unequal static load sharing between axles in a tandem suspension may increase road
damage by a factor of up to 3 (third part of chart).

4. The fourth bar summarises the literature on the road-damaging effects of dynamic tyre
forces. The average increase in damage caused by dynamic forces, compared to static
forces alone is approximately 10% 40% (mean damage on the fourth bar of the chart,
as calculated by the road stress factor and/or by neglecting spatial repeatability). This is
small compared with the effects of tyre type and unequal static load sharing shown in the
second and third bars. If instead a high degree of spatial repeatability is assumed, the
relative increase in peak road damage caused by dynamic forces is in the range 2 14
(peak damage in the fourth bar of the chart), which is comparable with the effects of
tyres and unequal static load sharing.
The rest of this chapter compares results from the main references used in this report, with
respect to vehicle influence on road deterioration.

4.2 Experimental findings: static axle load


As stated in Section 3.2, the concept of number of equivalent standard axles, and the so called
fourth power law, has been a highly debated topic. There has been many different reports in the
literature regarding the exponent n in the load equivalence equation. This relation, comparing the
damage caused by axles of different loads is

N ref
Nx

W
= x
W
ref

(4.1)

where Wx and Wref are axle loads and Nx and Nref are the corresponding number of load
applications.
Cebon concludes that for flexible pavements, values of 1.3 6 has been reported in the
literature, while for composite and rigid pavements, values are thought to be as high as 8 33.
COST 334 makes a distinction between individual distress modes, and reports that cracking of
bituminous layers has a value of 4 7, permanent deformation of the subgrade has an exponent
of perhaps 3 4 and permanent deformation of bituminous layers a value of 1 2.

4.3 Experimental findings from using different tyres


The COST 334 project, conducted research, in addition to a literature study, to draw conclusions
on the effect from tire fitment on pavement wear. Specifically, wide single tires and dual tyres
were compared. The results of the project were finalised in 2001, see Ref. [3].
In that project roads were divided into two categories:

Primary roads: Defined as the network of principal roads in a country or state, generally
comprising motorways (autoroutes, autostrade, etc) and other principal roads, state
owned or otherwise. This network provides the major links between large urban areas
and key national long-distance routes.

Secondary roads: Defined as the network of secondary roads in a country or state,


generally comprising those roads owned by state, regional or local authorities, and acting
as links between primary routes, but excluding some rural roads.

27

The road damage from a specific tyre and assembly is described with respect to a reference tyre
in terms of Tyre Configuration Factor (TCF). This concept is introduced in that project, and the
interpretation is straightforward: if a tyre has a TCF value equal to three, that means that for the
same axle load, this tyre is three times as aggressive as the reference tyre. Different formulas for
the TCF were derived for different pavement thicknesses and different distress modes. A tyre
specific part taking the tread width, contact length and the tyre pressure ratio (with respect to the
recommended pressure) into account is combined with a factor taking the tyre fitment into
account. This last factor comprises factors for tyre characteristics regarding dynamic force
transmissibility, and for potential load imbalance between the tyres of a dual assembly. The TCF
formulae are shown below in table 4.1.
Table 4.1: The TCF formulae from COST 334.
Total factor for translation to
real world conditions
Pavement Tyre specifications
thickness

Dual
tyres

Wide base
single tyres

Single
tyre

Distress mode primary rutting


Medium

(width/470) -1.68 * (length/198) -0.85 * (pres. ratio) 0.81


or
(width/470)-1.65 * (pres. ratio) 1.42 * (diameter/1059) -1.12

1.01

0.97

1.00

Thick

(width/470) -1.68 * (length/198) -0.85 * (pres. ratio) 0.81


1.01
or
(width/470) -1.65 * (pres. ratio) 1.42 * (diameter/1059) -1.12

0.99

1.00

Distress mode secondary rutting


Thin

(total width/570) -2.57 * (pres. ratio)1.58

0.97

0.97

1.00

Medium

(total width/570) -2.57 * (pres. ratio)1.58

0.98

0.97

1.00

Thick

about equal to 1

Distress mode fatigue


Thin

(total width/570) -2.88 * (length/198) -3.13


or
(total width/570) -2.44 * (diameter/1059) -2.47

0.94

0.97

1.00

Medium

(total width/570) -1.36 * (length/198) -1.40


or
(total width/570) -1.23 * (diameter/1059) -1.14

0.96

0.97

1.00

Thick

about equal to 1

Below TCF values for different tyres and assemblies reported in COST 334 are presented, using
a reference tyre of width 235 mm put in dual fitment. In Fig. 4.2 4.4 the graphical overviews of
the TCF values of the individual tyres are presented. Dual tyres are represented in green,
standard single tyres in red and wide base single tyres in blue. The TCF values for primary
roads are represented by solid bars, the TCF values for secondary roads are striped. Note that
for primary roads, only primary rutting is considered, while for secondary roads, it is the average
of primary rutting, secondary rutting and fatigue cracking that is used. In most cases the TCF-

28

values for each of the distress modes considered for secondary roads, are of the same order.
The only exception concerns steering axles, where secondary rutting accounts for roughly 60%
of the total TCF-value.

Figure 4.2: TCF of common current and possible future tyres for towed axles.

Figure 4.3: TCF of common current and possible future tyres for driven axles.

29

Figure 4.4: TCF of common current and possible future tyres for steering axles.
Based on the presented results, COST 334 makes the following conclusions:

The values in the graphs show that there is not one unique answer to the question
whether the common current and (possible) future wide base singles are better or worse
with respect to pavement damage.

Replacement of duals by wide base singles, both on towed or driven axles, generally
results in more pavement damage, for the observed range of common current and
possible future tyres. This effect is more pronounced on secondary roads.

Replacement of single tyres on steering axles by wide base singles, however, results in a
reduction of pavement damage.

Cebon cites several sources, exhibiting a wide span of results (see Cebon p. 321). The earliest
reports, from 1965 and 1978, indicate that the pavement damage from wide single tyres are up to
7-10 greater compared to dual tyres. Later work, from 1991 and 1992, report an increased road
damage between 1.1 4 for wide based single tyres. Another report (by M. Huhtala, VTT), from
1988, concludes that wide based single tyres are likely to cause 3.5 7 times more damage than
dual tyres, and that the worst conditions are for thinner asphalt layers. That study also reports
that a wide-based single tyre is only 1.5 times more damaging than an unevenly inflated dual pair
with 500 kPa in one tyre and 1000 kPa in the other.
Regarding these studies, Cebon notes that the large pavement damage factors in the studies
that he cites comes from the use of the fourth power law applied on measured (or calculated)
strains in the road under dual and wide single tyres. Quoting Cebon, This raises the important
question of whether a fourth power is appropriate, or, whether it may bias the results
excessively.
Cebon concludes that various experimental and theoretical studies have indicated that single and
wide based single tyres can cause up to 10 times more fatigue damage on thin flexible
pavements, compared to dual tyres carrying the same static load. Moreover, tyre contact
conditions are less important for rutting of thicker flexible pavements for which wide single tyres

30

are only 1.5 2 times more damaging than dual tyres, and that the tyre type has little influence
on fatigue damage of rigid pavements.
A TFK report from 1989 [9] presents equivalence factors for different tyre configurations, shown
in Fig. 4.5.

Figure 4.5: The load equivalents for different tyre configurations and tyre dimensions as a
function of axle load on two different asphalt thicknesses: 80 mm (left diagram) and 150 mm
(right diagram). A dual fitment with 12R22,5 tyres with axle load 10 tons has been used as a
reference and has the value 1. For all of the 5 configurations, three different tyre air pressures
are indicated, where the bold line corresponds to recommended air pressure, and the upper and
lower line corresponds to +20% and -20% of recommended air pressure respectively. From Ref.
[9].
A recent study from 2005 [10] concluded that a wide-base single tire (445/50R22.5) caused
similar, if not identical, pavement response as a conventional dual tire assembly. In fact, the two
configurations for the three measured responses (asphalt strain, base stress, subgrade stress)
produced, statistically, the same results.
Regarding the influence from using different tyre pressures: COST 334 reports that the ratio of
actual to recommended inflation pressure was shown to be influential for the cases of primary
rutting on thick (and probably medium) pavements and secondary rutting on thin and medium
pavements. An inflation pressure 10% higher than that recommended for the actual tyre load
results in about 15% increase in pavement wear.
Cebon reports (see Cebon p. 304-305) that several studies has indicated that fatigue damage
due to tensile strain at the bottom a thin asphalt pavements is likely to increase rapidly with
average contact pressure, while the inflation pressure has little effect on subgrade rutting. Based
on asphalt pavement strain measurements, it has been reported that a 40% increase in tyre
pressure would increase fatigue damage by 26%. Also, laboratory measurements on a 225 mm

31

thick asphalt road surface model have shown that rut depth development was approximately
linearly related to the average contact pressure (independent of load). Another study on high
type pavements found that overinflation of conventional dual tyres by 170 kPa nearly doubles
flexible pavement fatigue. Similar overinflation of wide base single tyres was even more critical,
increasing fatigue by a factor of four. In contrast, the tyre inflation on rigid pavements had a
moderate influence on fatigue.

4.4 Experimental findings regarding suspensions and dynamic loading


According to Cebon a quite simple modification of a trailer steel suspension might reduce road
damage by about 5 percent, see Figure 4.6, bar 1. First, the spring stiffness is reduced by half.
Then rubber blocks are inserted between the spring leaves to reduce inter-leaf friction, that
otherwise gives rise to hysteresis. However, this also eliminates the dampening effect of the
spring leaves, so hydraulic dampers have to be added to the suspension system.
Replacing the standard steel suspension on a tractor/trailer combination by an air suspension
might reduce road damage by 5-6 percent, and optimising the suspension, a further reduction of
9% in road damage could be obtained, see Figure 4.6, bar 2. Experiments showed that adding a
computerized dampening system reduces road damage with an additional 5-6 percent, see
Figure 4.6, bar 3 (Cebon).

Figure 4.6: Comparison of three suspension for a principal road (from Cebon p. 561)
The DIVINE project reports on a comparison between air and steel suspensions conducted as an
accelerated test on an indoor pavement. In both cases the static load was 49 kN and wide single
tyres were used. In brief the test showed that the steel suspension produced a 15 percent
increase in road roughness (IRI) and 10 percent more cracking. Rutting was generally low during
the test and no difference in mean rut depth between air and steel suspensions could be
observed. Looking at the maximum rut depths though, it was concluded that if the maximum rut
depth in the range 11-12 mm is taken as a critical level then the air suspension would achieve
45-65 percent more load cycles than the steel suspension to produce the same rutting distress.
Cebon summarises the general conclusion about the effects of suspension types on dynamic tyre
forces (Cebon, p. 123-124):

32

All studies found dynamic tyre forces to increase with speed and road roughness

It has been noted that reducing suspension stiffness generally reduces tyre forces

Centrally pivoted tandem axle suspensions such as walking beams and single-point
suspensions were always found to generate the highest dynamic loads because of their
lightly damped pitching modes at around 8-10 Hz. It has been noted, however, that these
suspensions can be improved considerably by suitable use of hydraulic dampers.

Four spring tandem suspensions were generally found to generate smaller dynamic
loads than walking beams. Torsion-bar and air suspensions generated the lowest loads.

It has been noted that modern single-spring parabolic suspensions with good hydraulic
damping are not significantly worse than stiff air suspensions. Moreover, air
suspensions without hydraulic dampers could generate significantly higher dynamic loads
than leaf spring suspensions.

Triaxle suspensions were found to generate smaller dynamic loads than tandem
suspensions in several studies.

It has been reported that varying the axle spacing of an air spring tandem suspension had
negligible effect on the dynamic loads, whereas the Dynamic Load Coefficient generated
by a four-spring tandem suspension varied considerably with axle spacing, depending on
the speed and road roughness.

4.5 A model taking all factors into account


Cebon describes a quantity known as the road stress factor, proposed by Eisenmann in 1975,
which uses the assumption that road damage depends on the fourth power of the instantaneous
(dynamic) wheel force. The equation for the road stress factor that accounts for the effects of
wheel configuration and tyre pressures is written:

= ( I II Pstat )
where

(4.2)

= the road stress factor for each tyre

= the dynamic road stress factor (takes suspension type into account)
I = parameter to account for the tyre configuration (single or dual tyres)
II = parameter to account for the tyre contact pressure
Pstat = the static (average) tyre force
Sometimes an additional factor is included to also account for the type of axle group (single,
tandem or triaxle). Cebon concludes that considerable research effort has gone into quantifying
these parameters for a variety of suspensions and tyre contact conditions.
The road stress factor has also been the subject of considerable criticism (also from VTI [11]).
According to Cebon, The Road Stress Factor approach incorporates all of the uncertainties
inherent in the fourth power law, which has itself been the subject of considerable criticism. It has
three other highly questionable features:
I. It assumes that the strain in the road surface is directly proportional to the instantaneous
wheel force and neglects the sensitivity of road surface response to the speed and
frequency of the applied loads and to the structural response characteristics of the road.

33

II. In using the mean value of the damage criterion, it implicitly assumes that road damage is
spread randomly over the surface and does not account for any concentration of damage
which may occur in the vicinity of particular roughness features.
III. It assumes that each suspension system on a vehicle is dynamically independent and
does not influence the tyre forces, and hence road damage, generated by other axles.
Thus suspensions are compared through analysis of the wheel loads generated by
individual axles of axles groups, rather than through analysis of road damage done by the
whole vehicle.
The opinion on the use of the road stress factor seems to be split among the researches. Cebon
reports that Ullidtz notes that accounting for dynamic loads using a number of equivalent
standard axles calculated using a fourth power law (as per the Road Stress Factor), would result
in completely erroneous results. He also quotes Morris, and according to him the road stress
factor is a plausible rule of thumb that can serve as a bench-mark for comparison with more
analytical approaches.
COST 334 used the road stress factor approach to propose a formula for the damage
contribution of a single passage of an axle on primary roads. The so called Axle Wear Factor
(AWF) is a dimensionless factor relating the damage contribution of a specific tyre at a given axle
load and axle configuration to the damage contribution of a single passage of a reference tyre
with a reference axle load (10 tons).

P
AWF = TCF
10
where

(4.3)

TCF = the Tyre Configuration Factor developed in COST 334. The factor depends
on the total tread width (2 x single tyre tread width for dual fitment) and the diameter
of the tyre, and thus takes the tyre fitment (single or dual) into account. The
following formula is recommended: TCF = (tread width/470)-1.65 * (diameter/1059) -1.12
P = Axle load in tons.

The following remarks regarding this formula are made in COST 334:

34

Only primary roads are considered, so that only primary rutting is taken into account,
which implies a power of 2 in the load equivalency factor.

The value of the factor for Axle Configuration is assumed to be equal to unity (1). It is
generally accepted (OECD, 1983) that tandem or triple axles with axle spacings below 1.4
m, cause (slightly) more damage that two, or three passages, respectively, compared to a
single axle of equivalent loading. For primary rutting, however, no specific information is
available, and a factor of unity appears to be reasonable.

The value of the factor for Suspension Configuration is also assumed to be unity (1).
Strictly, this value is valid only for those axles having air suspension, but since this is the
case for most of the heavy goods vehicles under consideration, the assumption is again
reasonable.

The traction effects of the drive axle on the pavement are ignored.

Correction factors for load imbalance and dynamic effects in the TCF formula should be
ignored, since they add only about 1% of additional precision to the calculation of TCF.

35

5 Calculation models used in the Nordic countries


Different calculation models have been used in the Nordic countries in order to compare the
relative road wear due to heavy transports. In this chapter two of them are described.

5.1 Classification of special heavy transports in Denmark


Vejdirektoratet in Denmark is currently using a model to classify special transports with heavy
loads, which has been used since 2002. It is described in Ref. [6]. The model, which can be used
on the entire Danish road network, is focusing on the relative road damage a heavy transport
induced on a road, in relation to the damage produced by a standard heavy vehicle. The relative
road wear is specified in the ESAL10 number, where the reference heavy vehicle axle of 10 ton
load has ESAL10=1.
The parameters that are included in the model for calculating ESAL10 are
Axle configuration

Number of axle groups (axles that are separated by more than 1,8m are divided into
different groups)

Number axles within each group

Distance between the axles within a group

The wheel load for each axle

Tyre configuration

Tyre width

Distance between tyres on the same axle (single or dual fitment, or something in
between)

Tyre air pressure

Suspension type

The formula is

Ai Bij C i Dij pij

ESAL10 =
mi
5
i =1 j =1

mi

(5.1)

where
n = number of axle groups
mi = number of axles within group i
Ai = Constant that accounts for both the number of axles within group i, and the distance
between the axles within group i. There is a distinction between roads with a strong base
(where only primary rutting is considered), and those with a weaker base (where also

36

secondary rutting, a permanent deformation below the pavement, occurs). For the weaker
roads, Ai = the number of axles within the group, independent of the inter-axle distance. For
stronger roads, it is assumed that a group of axles with a short inter-axle distance is
producing less damage to the road, compared to the same group of axles with a larger interaxle distance. Thus, on this kind of roads, Ai is lower than the number of axles. A smaller
inter-axle distance leads to a smaller value of the constant.
Table 5-1
Ai

Strong road

Weak road

Number of Inter-axle

All inter- axle

axles

distances

distance (m)
1.0

1.2

1.4

1.6

1.8

1.05 1.25 1.40 1.55 1.75 2.00

1.50 1.90 2.20 2.50 3.00

1.70 2.35 2.80 3.30 4.00

1.90 2.80 3.40 4.05 5.00

2.00 3.20 4.00 4.75 6.00

2.20 3.60 4.60 5.45 7.00

2.35 4.00 5.15 6.10 8.00

Bij = Constant that takes the tyre fitment on each axle into account. If dual fitment is used (or if
the distance between tyres on an axle is less than the tyre width + 8 cm), Bij = 1. If the tyre
fitment is single (if the distance between tyres on an axle is more than 2,5x the tyre width),
Bij is given a number that is larger than one, and which depends on the tyre width. A wider
tire leads to a lower value on the constant. If the distance between the tyres falls in between
the definitions of single and dual fitment, Bij is supposed to be determined by interpolation
between the single and dual values of the constant, taking the distance between tyres into
account. The values of this constant for single fitment are:
Table 5-2
Tyre width (mm) Bij
256

3,99

315

2,91

365

2,19

425

1,68

445

1,25

Ci = Contant that takes the type of suspension into account. The following suspension types are
distinguished between:

37

Table 5-3
Suspension type

Ci

Leaf spring

1,0

Parabolic spring ?
Coil spring
Rubber

0,95

Hydraulics
Air
Oil
No suspension

1,4

Dij = Constant that accounts for the air pressure of the tyres on a specific axle. A differentaion is
made between single and dual fitments, and whether primary rutting or deformations below
the pavement is considered. Dij is almost linear with respect to the air pressure, as shown in
the figure below.
The constant Dij
1.8
Primary rutting
Subpavement deformation: Single fitment
Subpavement deformation: Dual fitment

1.6

1.4

Dij

1.2

0.8

0.6

0.4

7
8
Tyre air pressure [bar]

10

Figure 5.1: The constant that accounts for tyre pressure


Pij = The load (in metric tons) on one tyre on a specific axle.
Note that the road damage described by ESAL10 is the road damage per wheel track. I.e., the
number of wheels on the axles are not explicitly taken into account. The number of wheels, are
indirectly used in the calculation of ESAL10 , since the wheel load Pij, depends on the number of
wheels on the axle, and the number of wheels on an axle may also affect the tyre fitment
constant Bij. According to the model, however, two identical vehicles driven side by side on a
road produce the same damage to the road as one of the vehicles would do by itself.

38

5.2 The Swedish Road Administration method of calculating the


equivalent number of standard axles
In Sweden, the equivalent number of standard axles per heavy vehicle is a crucial parameter for
road dimensioning. In ATB VG 2005 [12], which is the Swedish Road Administrations common
technical description, containing the demands on construction and maintenance and of roads, the
equivalent number of standard axles per heavy vehicle is called the B-factor. The B-factor is the
average number for all the heavy traffic using the road. The standard axle (see Fig. 5.2) has a
load of 100 kN, and is supposed to have twin mounted wheels with the load evenly distributed.
Each wheel is assumed to have a circular contact patch between tyre and road surface, and the
contact area is exposed to a load with a inflation pressure of 800 kPa.

Figure 5.2: Schematic picture of the standard axle used in Sweden,


specified by ATB VG 2005. From Ref. [12].

Historically, since 1994 a factor 1.3 has been used. However, in ATB vg 2005, section C6,
methods for determining the B-factors are given:

To determine the B-factor measurements of gross vehicle weights and axle loads shall be
carried out. The measurements must be performed during a period of at least 7 days. The
B-factor shall then be calculated from these measurements. These measurements could
be performed using the B-WIM (Bridge Weigh In Motion) technique described in VV
Publication 2003:165. In that report a number of measurement results are presented
which could serve as a support for choosing the B-factor.
For calculation method of the ESAL (equivalent number of standard axles) for each heavy
vehicle type , ATB VG 2003 is referred to, which shows an example on how an axle of
16 tons (apparently not making any difference between single, tandem or triaxles) results
in 6.7 standard axles of 10 tons, using the fourth power rule. The corresponding formula
of the ESAL for each heavy vehicle type that is recommended by ATB VG 2003 can
then be written

39

W
ESAL =

100

(5.2)

where W is the weight of the specific axle in kN.

If measurements cannot be carried out, ATB VG 2005 suggests that four or five of the
most common heavy vehicle classes are used, with weights estimated from local
experience. Then the fourth power rule is applied on each axle.

In a recent report from the Swedish Road Administration [14], concerning B-WIM measurements
during 2004-2005, another way of using the fourth power rule for calculating ESAL is
demonstrated. In section 4.5 of that report, as an illustration on the effects of overloading, four
examples on how ESAL values should be calculated are shown. No formula is given, but from
private communication with Tomas Winnerholt of the Swedish Road Administration, it was
clarified that the fourth power rule together with factors taking the axle type into account has
been used. More specifially:

i
W
ESAL = i k i
n =1 10

i=

number of axles or axle groups

Wi =

axle (group) weight for axle (group) i (ton)

ki =

effect reduction factor for axle (group) i

(5.3)

k = 1 for single axle


k = (10/18)4 = 0,0952 for tandem axle
k = (10/24)4 = 0,0302 for tri-axle
The factors has been chosen with the maximum allowed loads (in metric tons) in Sweden for
different axle configurations (se Table 3.3) in mind, so that a single axle of weight 10 ton, or a
tandem axle of weight 18 ton, or a triple axle of weight 24 ton, each results in ESAL=1. This is
the current approach on how the Swedish Road Administration is calculating ESAL values. It is
based entirely on the legal loads and axle distances. E.g., if the interaxle distance is only
marginally longer than what has been defined by Swedish regulations as a tandem axle, the axle
will be treated as a single axle.
It should be noted that this approach, taking the entire weight of a tandem or tri-axles axle to the
power of four, is different from the more common method described in Section 3.2, in which loads
of tandem axles and tri-axles are converted to an ESAL by summing the contributions of the
individual axles.
The following examples are shown in Reference [14]:

40

Figure 5.2.1: 60 tons equal 3.1 ESAL


For a configuration as in Figure 5.2.1, the calculation is:

6
18
18
18
ESAL = * 0.0952 + * 0.0952 + * 0.0952 + = 3.1
10
10
10
10

Figure 5.2.2: 66.2 tons equal 4.85 ESAL


For a configuration as in Figure 5.2.2, the calculation is:

8.7
20.9
21.2
15.4
ESAL =
= 4.85
* 0.0952 +
* 0.0952 +
* 0.0952 +
10
10
10
10

41

Figure 5.2.3: 20 tons equal 2 ESAL

Figure 5.2.4: 20 tons equal 3.1 ESAL

For a configuration as in Figure 5.2.3, the calculation is:


4

10 10
ESAL= + = 2
10 10
For a configuration as in Figure 5.2.4, the calculation is:
4

13 7
ESAL = + = 3.1
10 10

42

6. Acknowledgements
The authors would like to thank Leif Sjgren for contributing with material to this report. The
members of NVF Fordon och Transporter are also thanked for valuable comments and
suggenstions.

43

7. References
[1] D. Cebon, Handbook of Vehicle-Road Interaction. Swets &Zeilinger Publishers, Lisse, The
Netherlands (1999).
[2] Dynamic Interaction between Vehicle and Infrastructure Experiment, DIVINE Technical
Report, DSTI/DOT/RTR/IR6(98)1/FINAL, OECD (1998).
[3] COST 334 Effects of Wide Single tyres and dual Tyres Report, Chapter 4, version 29,
November 2001
[4] U. Isacsson, Interaktion mellan fordon/milj och vg, KTH, (2004)
[5] R. Blab Die Fahrspurverteilung als Einflussgre bei die bemessung des
Straenoberbaus; Mitteilungen des Institutes fr Strassenbau und Strassenerhaltung TU Wien,
ISTU, Heft 5; Institut fr Strassenbau und Strassenerhaltung TU Wien; 228 pp.; Wien, AT (1995)
[6] Jan M. Jansen, Srtransporters vejslid Klassificering av kretjer. Rapport nr 269,
Vejdirektoratet. 2002
[7] The Handbook of Highway Engineering, CRC Press, Taylor & Francis Group. 2006
[8] Private communication with Leif Sjgren, VTI.
[9] TFK rapport 1989:5 Optimalt Dckval fr Tunga Fordon, Fltmtningar av Vgpknning,
Transportekonomi och Vgkostnader. L. Djrf, M. Huhtala, M. Johansson, E. Samuelsson.
[10] A. L. Priest, D. H. Timm, and W. E: Barrett, NCAT report 05-03 Auburn University:
Mechanistic comparison of wide-base singe vs. standard dual tyre configurations. June 2005
[11] G. Magnusson, H.E. Carlsson, and E. Ohlsson, The influence of heavy vehicles springing
characteristics and tyre equipment on the deterioration of hte road, VTI (Translated by TRRL as
WP/V&ED/86/16, 1986), Report Number 270, 1984.
[12] ATB VG 2005. VV Publication 2005:112.
[13] BWIM-mtningar 2002 och 2003 Slutrapport. VV Publication 2003:165
[14] BWIM-mtningar 2004 och 2005 Projektrapport. VV Publication 2006:136

44

NVF
Vejdirektoratet
Niels Juels Gade 13
Postboks 9018
DK-1022 Kbenhavn K
Danmark
Telefon +45 7244 33 33 telefax +45 33 32 98 30
E-post: nvf@vd.dk
NVF
c/o Vgfrvaltningen
Postbox 33
FIN-00521 Helsingfors
Finland
Telefon +358 204 22 2575 telefax +358 204 22 2471
E-post: nvf@finnra.fi
NVF
c/o Landsverk
Box 78
FO-110 Torshavn
Frerne
Telefon +298 340 800 telefax +298 340 801
E-post: lv@lv.fo
NVF
c/o Vegagerdin
Borgartun 7
IS-105 Reykjavik
Island
Telefon +354 522 1000 telefax +354 522 1009
E-post: hreinn.haraldsson@vegagerdin.is
NVF
c/o Vegdirektoratet
Postboks 8142 Dep
NO-0033 Oslo
Norge
Telefon +47 22 07 38 37 telefax +47 22 07 37 68
E-post: publvd@vegvesen.no
NVF
c/o Vgverket
SE-781 87 Borlnge
Sverige
Telefon +46 243 757 27 telefax +46 243 757 73
E-post: nvf@vv.se
NVF-rapporterna kan bestllas via respektive lands sekretariat per telefon, fax,
e-post eller post. Se kontaktuppgifterna p nst sista sidan.
En uppdaterad rapportfrteckning finns p frbundets nordiska hemsida,
http://www.nvfnorden.org.
45

You might also like